You are on page 1of 565

Geometrie, Physical,

and Visual Optics

Michael P. Keating, Ph.D.


College of Optometry
Ferris State University
Big Rapids, Michigan

Butterworths
Boston London Singapore Sydney Toronto Wellington
Copyright 1988 by Michael P. Keating.
All rights reserved.

No part of this publication may be reproduced, stored in a


retrieval system, or transmitted, in any form or by any means,
electronic, mechanical, photocopying, recording, or otherwise,
without the prior written permission of the publisher.

Library of Congress Cataloging-in -Publication Data


Keating, Michael P.
Geometric, physical, and visual optics.
Bibliography: p.
Includes index.
1. Optometry. 2. Optics. I. Title.
RE951.K43 1988 617.7'5 87-25699
ISBN 0-409-90106-7

British Library Cataloguing in Publication Data

Keating, Michael P.
Geometric, physical and visual optics.
1. Optics 2. Optics , Physiological
I. Title
535'.0246177 QC355.2
ISBN 0-409-90106-7

Butterworth Publishers
80 Montvale Avenue
Stoneham, MA 02180

10 9 8 7 6 5 4 3 2 1

Printed in the United States of America


For all who have been supportiveespecially family, friends,
colleagues, teachers, and my wife Mary Jean.
Preface

This basic textbook, written primarily for optometry intuition and conceptual understanding, so that the
students, contains an integrated approach to geomet- numbers mean something to the reader. Chapters 2
ric, physical, and introductory visual optics. This book through 4 emphasize concepts that are sometimes lost
is nontraditional in the integration, sequencing, and in the rush to get to the equations and calculations of
conceptual development of the material. The nontra- Chapters 5 and beyond. I have included many worked
ditional aspects include an early emphasis on image examples, but I strongly encourage students to first
formation, the use of the vergence-dioptric power work out their own solutions before checking mine.
approach from the beginning, the relation of vergence Due to the needs of optometry students, the physical
to the geometric properties of wavefronts, and the optics chapters are more qualitative than the geomet-
interchangeability of the wavefront representation ric optics chapters.
with the ray representation. This approach has worked There are optical effects everywhere we look, and I
extremely well for me and for my students over a have tried to incorporate them into this book so that
15-year period. In particular, the integration of visual the world becomes optically alive to the reader. This
optics makes the optometry students feel that basic is, perhaps surprisingly, true of physical optics as well
optics is highly relevant to their profession. I wrote as geometric optics. Visual optics in particular is illus-
this book so that others can share in the benefits. trated by the many characters that appear in the book.
The mathematical level of the book assumes a In many cases, subsequent chapters in the book
knowledge of algebra and trigonometry. While some build on previous chapters. For example, Chapter 7 is
introductory knowledge of calculus is helpful, it is not a long chapter, but in fact most of the concepts have
necessary for the level of this text. Since the advent of already been developed in the previous chapters, and
calculators and microcomputers, basic matrix algebra Chapter 7 goes fairly easily despite its length. Chapter
is being increasingly incorporated into high school and 3 is very short, but the concepts are fundamental and
undergraduate algebra courses. After paraxial image often not dealt with very extensively in optics texts.
formation by spherical systems is covered (Chapters 1 There is no doubt that parts of optics are conceptu-
through 11), I do take advantage of matrices in some ally difficult because many of the concepts require
chapters (particularly 12, 16, and a little in 18). Ap- formal operational reasoning (as opposed to concrete
pendix A covers the needed matrix algebra. operational). That is part of the fun and challenge of
For optometry students, the sections on astigmatism teaching it well.
are a very important part of this book. Chapter 15 While this book was written primarily for optome-
gives an integrated treatment of the on-axis aspects, try students, I believe that it offers benefits for other
while Chapter 16 treats the more difficult off-axis students interested in the vergence approach to optics
aspects. and vision. This might include physics undergraduate
There is an emphasis in the book on developing students as well as perceptual psychology students.

xiii
Acknowledgments

To begin, I want to thank Dean Jack Bennett of the addition, I want to thank Associate Dean Jerald
Ferris College of Optometry for his enthusiastic sup- Strickland of the University of Houston College of
port of this book. Next, I want to thank my Ferris Optometry for his role in making the visiting pro-
colleagues, Dr. C. Allyn Uniacke, for his very willing fessorship possible.
help, advice, and critiques; Drs. Vince King and I also want to thank Dr. Norman Wallis, executive
Gerald Lowther for their encouragement and advice; director of the National Board of Examiners in Op-
Barbara Swanson and Doug Nadeau from the Ferris tometry, for his encouragement of this project, and
graphics department for their work on the figures; and Dr. James Carroll of the Pennsylvania College of
a number of my students, Marie Walters, Mike McNa- Optometry for his role in initially getting me involved
mara, Carol McMannan, Sue Ward, Dean Luplow, with optometry.
Kerry Kondal, and Chris Theodoroff, who have I particularly need to thank my wife Mary Jean,
proofed and critiqued various chapters. and my children Mikala, Kristen, and Kevin for their
The goal of writing this book originated as a result support and understanding during the many extra
of a sabbatical year I spent as a visiting professor at hours that it took to write this book.
the University of Houston. Thus, I need to thank both While others have critiqued various sections of this
Ferris State University and the University of Houston work, I clearly am the person solely responsible for
for their financial support during the sabbatical. In any mistakes.

XV
CHAPTER ONE

Optics, Light,
and Vision

1.1 The Search for Solutions Patterns Science is the search for patterns, and pat-
terns are not just pictures, but predictions. Thus, the
"That's one small step for man. One giant leap for ability to discern patterns, to organize and reorganize
mankind." So stated Neil Armstrong, Jr., as he the fruits of vision and evidence, is indispensable.
stepped down onto the moon's surface. The problem
was how to send men to the moon and bring them Context To see an object within its framework means
back alive. The solution required commitment and to see it in context. All problem solving requires
financial resources, as well as scientific and technical trained powers of observation and an attitude of mind
knowledge. ready for the unexpected and eager of examine every
How was the knowledge acquired? How do we clue that chance presents.
proceed in developing new knowledge and solving new
scientific problems? In The Search for Solutions, Modeling Problem solving requires that models be
Horace Judson lists nine problem solving techniques used at every stage from definition to solution. As
of the scientific method. These are: investigation, simple representations of complex forms, processes,
evidence, trial and error, patterns, context, modeling, and functions, they reduce risk, cost, and complexity
prediction, adaptation, and theory. Judson's capsul- to manageable size.
ized summaries of these techniques are: Prediction One of our oldest endeavors is to learn to
avoid the destructive aspects of the environment, to
Investigation Each of us discovers our world by in- capitalize on the positive factors, and to plan ahead.
vestigationtasting, watching, smelling, listeningso Prediction is the search for laws on which to base such
that eventually we turn bits and pieces of experience forecasts.
into generalizations and categories.
Adaptation Frequently, problems change as they are
Evidence We need information to solve problems. being solved. As a consequence, solutions have to
For the scientist, gathering evidence is just the first adapt to complex new realities. Feedback is the sys-
step. Each clue then must be tested and retested, tematic adaptive response to situations in which the
verified and interpreted. problem and solution are in constant changing rela-
tionship.
Trial and error Trial and error is as basic as learning
itself, and errors are essential to the process. Problem Theory The universal human need to tell stories, and
solvers must organize and direct a plan of attack, to explain events, forms the foundation of theory
order priorities, determine the essential elements to making. Theories enable both the scientist and the
combine, and then decide how a desired result is to be layperson to see significance in seemingly random
obtained. events and objects.
1
2 Geometrie, Physical, and Visual Optics

These techniques are just as essential to optics as Some aspects of quantum theory are presented in later
they are to the space program. The optics that this chapters.
book addresses is a simplified model of the complex Historically, optics has been divided into two sub-
optical world. As stated by Judson, the modeling step areas: geometric optics and physical optics. Geometric
reduces complexity to a manageable size. optics deals with the image-forming properties of len-
There have been many scientific advances that, like ses, mirrors, and prisms, and as such, is very import-
the lunar landing, were the result of the well-estab- ant to visual optics. Physical optics deals with the
lished developmental plan, but the most revolutionary physical character and behavior of light and its interac-
advances frequently occur unexpectedly. In those par- tion with matter. Physical optics can be further sub-
ticular cases, the people involved in looking for the divided into such areas as wave optics, quantum op-
general patterns and context discover fundamental and tics, and Fourier optics.
important scientific phenomena for which they were
not specifically looking. Patterns, context, adaptation,
and prediction are all important in this process. This
book is based on experimental investigations, evi- 1.3 An Overview of the Human Visual Process
dence, and trial and errors of many people. Hopefully,
the reader will also have the opportunity to recreate Modern man is well aware of the optical images
and experience firsthand some of these experimental formed by slide and motion picture projectors as well
investigations. as with camera television images. In developed coun-
tries, even elementary school children are aware that
the human eye forms an optical image on the retina.
Typically, modern man is less aware that his perceived
image can differ significantly from his retinal image.
1.2 The Scope of Optics and Its Subfields This difference is due to the physiology and perceptual
psychology of the human visual system. The study of
The main body of optics deals with light and its vision is formally called physiological optics. It is
behavior. As such, optics is a branch of physics. The interdisciplinary in that it includes the physiology and
impact of optics on modern man is enormous. Mi- perceptual psychology of the visual system as well as
croscopes opened the ultrasmall world of bacteria, visual optics.
histology, and biological cells. Telescopes opened the Figure 1.1 shows a flow chart of the human visual
ultralarge world of planets, stars, and galaxies, and process. The process begins with a primary light
enabled humans to study distant terrestrial objects source that serves as the initial generator of the light.
including enemy armies and wild animals. Cameras Some examples are the sun, fire, tungsten filament
opened an incredible world of information capture and light bulbs, flourescent lights, light emitting diodes
provided an aesthetic medium that produces beautiful (LEDs), and lasers. The light from the primary source
results in the hands of an artistic photographer. Inter- illuminates an object. On a molecular scale, the object
ferometers contribute to the ultraprecision world of absorbs the incident light and then reradiates some of
modern manufacturing. Spectacle and contact lenses it. The reradiated light diverges away from the object
enable many of us to overcome the visual handicap of and is incident on the eye. The object that absorbs and
optical defects in our eyes. reradiates is called a secondary source of light. Some
As time passed, the borders of optics pushed out to examples include sailboats illuminated by sunlight,
include nonvisible parts of the electromagnetic spec- actors and actresses illuminated by floodlights, a ra-
trum, especially the infrared and ultraviolet regions. coon illuminated by a flashlight, as well as most of the
Optics combined with electronics to produce televi- other objects that we see. The cornea and the crystal-
sion, photocopiers, and night vision scopes. Lasers, line lens of each eye converge the incident light to
fiber optics, thin films, spatial filtering, computerized form a small inverted image of the object on the retina
image analysis and enhancement, and robot vision are of each eye. Light is the carrier of information from
areas of current activity and development. the secondary source to the retina of the eye.
Theories involving the nature of light, its speed, The retina is composed of ten layers. The rod and
and its interaction with matter have another impres- cone layer absorbs light, and the energy gained is used
sive legacy. These theories were intimately involved to generate neural signals. The neural signals are
with two of the greatest conceptual revolutions of processed in some of the intermediate layers of the
modern man: Einstein's special theory of relativity and retina so that the signals leaving the retina as a whole
the quantum theory of matter. Relativity is a fascinat- are different from the signals that left the layer of rods
ing topic, but it lies outside the scope of this book. and cones. The neural signals then proceed via the
Optics, Light, and Vision 3

Primary
light
source

neural signal
Lateral Geniculate
Bodies

Primary
Visual
Cortex
Perceived
Image FIGURE 1.1. Flow chart for the human
visual process.

optic nerve fibers to the lateral geniculate bodies of than monocular visual performance. In particular,
the thalamus. There the fibers synapse, and the signals binocular depth perception is better than monocular
are processed further and sent to the primary visual depth perception.
cortex, which is adjacent to the back of the skull. The
primary visual cortex has a mass of interconnections to
and from other parts of the brain, and the resulting
electrophysiological activity somehow results in the 1.4 Perceptual Aspects of the Visual Process
visually perceived image. The whole neural process is
amazing in that the neural signals that travel the visual Normal individuals perceive one three-dimensional
pathway are basically the same types of signals that image. The distal stimulus for this image is the illumi-
travel the other neural pathways, such as the auditory nated world. The proximal stimuli are the two-dimen-
pathway; yet, one results in a perceived image and the sional retinal images of the right and left eyes. Our
other results in perceived sound. normal visual experience is coordinated with our
Figure 1.2 shows a schematic representation of the motor experience. Not only do we see the coffee cup
human visual system as viewed from above. The com- on the left side of the desk top, we can also reach out
ponents represented are the two eyes with their re- and pick it up. This coordination typically leads to a
spective optic nerves, the optic chiasm, the lateral seeing is believing philosophy.
geniculate bodies, and the primary visual cortex. On the other hand, vision has many illusionary
Humans can function quite well monocularly. How- aspects to it. Two examples are given to illustrate
ever, in general, binocular visual performance is better some of the illusionary aspects of vision. The first
example is a binocular example in that it depends on
information from both eyes. The second example oc-
curs when viewing with only one eye, so it is a
monocular example.
The first example is the hole-in-the-hand illusion.
People with normal binocular vision may perceive this
illusion as follows. Take a sheet of paper and roll it
into a small tube. Hold the tube close to your right eye
and look through it at a distant object. Place your left
hand, palm inward, against the left outer edge of the
tube at a point near the far end of the tube. Your left
eye is looking at the palm of your hand while your
right eye is looking through the tube at the distant
object. The perceived three-dimensional image is that
of looking through a hole in your hand at the distant
object. Here the visual system has taken two separate
FIGURE 1.2. Schematic representation of the human visual retinal images and processed them into one three-
system (top view). dimensional perceived image.
4 Geometrie, Physical, and Visual Optics

The second example is the moon illusion. Unlike The cornea is followed by a chamber filled with a
the first example, the moon illusion can be observed transparent watery liquid called the aqueous humor.
monocularly as well as binocularly. When a full moon Near the rear of the aqueous humor chamber is the
first rises above the horizon, it is frequently perceived iris. The iris is opaque but has an opening at its center
as being much larger than when it is higher in the sky. called the pupil. The iris contains involuntary muscles,
This illusion occurs despite the fact that the retinal which enable it to constrict or dilate, thus changing the
image size of the moon is the same regardless of its size of the pupil.
elevation in the sky. The difference in the perceived Behind the iris and the aqueous humor lies the
size is due to the difference in the visual information crystalline lens. The crystalline lens is a transparent
or cues between the two different situations. The high-protein material that is surrounded by fibers con-
horizon apparently provides the visual cues that fool necting it to the ciliary muscle. Contraction of the
the perceptual system. ciliary muscle results in changes in the focusing power
These illusions, as well as many others, are dis- of the eye. This process, referred to as accommoda-
cussed in a number of perceptual psychology books. tion, enables the eye to view near objects as well as
This book discusses visual optics and characteristics of distant objects. As the crystalline lens ages, it becomes
the retinal image. You should keep in mind that the less flexible, and consequently a person's ability to
retinal images are the stimuli for the perceived image accommodate is slowly lost with time.
and not the perceived image itself. The crystalline lens is followed by a dark chamber,
which is filled with a transparent gelatinous substance
called the vitreous humor. The retina is the rear
boundary of this chamber.
The rods and cones in the retina absorb some of
1.5 The Eye the incident light and convert the energy into a neural
signal. The rods and cones are not distributed uni-
Insofar as image formation is concerned, the human formly. The region of the retina most sensitive to
eye can be compared to a camera. A camera has a detail is the foveal area, which has many closely
dark interior chamber so that the desired image is not packed cones and only a few rods. Away from the
washed out by stray light. A camera also has a vari- fovea the cones become progressively less dense and
able aperture to let in more or less light and an the number of rods increases. The peripheral region of
adjustable focus in order to image objects at different the retina, dominated by the rods, is best at light and
distances. The human eye has each of these three motion detection, while the foveal area, dominated by
features. cones, is best at form detection, color detection, and
Figure 1.3 is a simplified diagram of the eye. The resolution of fine detail. There are no rods or cones at
cornea is the curved transparent front surface of the the place where the optic nerve leaves the retina, so
eye and constitues the major converging element of this area is blind.
the eye. The tear film on the front of the cornea fills in
the irregularities in the corneal surface and is thus
important to the optical quality of the cornea.
1.6 Electromagnetic Radiation

Visible light is electromagnetic radiation that the reti-


nal rods and cones are capable of absorbing, with the
subsequent generation of a neural signal in the visual
pathways. Visible light constitutes only a small part of
the electromagnetic spectrum. From long to short
wavelengths, the electromagnetic spectrum includes
radio waves, television waves, radar waves, mi-
crowaves, infrared radiation, visible light, ultraviolet
radiation, x-rays, and gamma rays.
The emitters and absorbers of electromagnetic
radiation are the electrically charged particles of mat-
ter. The outer atomic electrons are the usual emitters
and absorbers of visible light. The atomic and molecu-
lar structure of a material determines the intrinsic
FIGURE 1.3. Simplified diagram of the human eye. absorption properties of the electrons in the material.
Optics, Light, and Vision 5

All components of the electromagnetic spectrum 1.7 Wavelengths and Color


propagate through a vacuum with a speed c of about
3 x 108 m/s. (The actual value is 299,792,458 m/s.) Human vision is wonderful in that we not only see
The speed of electromagnetic radiation in a vacuum objects, but we see them in color. Color vision de-
is one of the fundamental constants of the world. The pends on the fact that there are three different types
speed c is independent of both the speed of the source of cones in the retina, each type of which absorbs
of the radiation and the speed of the detector of the maximally at a different wavelength of incident light.
radiation. These are two of the experimental founda- Three neural signals are then generated and processed
tions of the special theory of relativity. and ultimately produce the perceptual color response.
The propagation of electromagnetic radiation The optical stimulus for color is the wavelength
through space can be described by wave equations. distribution of the incident light. Light that has an
The vacuum wavelengths of electromagnetic radiation equal mixture of all visible wavelengths is the stimulus
range from 10"13 m to 105 m. The interaction of for a white, gray, or black response depending on the
electromagnetic radiation with matter is wavelength- amount of light coming from the object and its back-
dependent. Consequently, wavelengths are one of the ground. Light that consists of a single wavelength is
ways used to classify the components of the spectrum. the stimulus for a pure color response. Light of a
The vacuum wavelengths for visible light range from single wavelength is thus called monochromatic light.
about 400 nm to 700 nm. (One nanometer (nm) equals The color response for monochromatic light ranging
10~ 9 m). from long to short wavelengths is: red, orange, yellow,
The ability of waves to propagate around corners is green, blue, (indigo), violet. Indigo is actually a deep
referred to as diffraction. (We are very familiar with blue and not considered a separate color. (The name
the diffraction of sound waves, which, incidentally, are ROY G BIV, is often used to remember the wave-
not electromagnetic waves.) The ability of a wave to length ordering of the spectrum.) Table 1.1 gives the
bend around a corner depends on its wavelength. The color responses as a function of incident vacuum
longer the wavelength, the more the wave bends. The wavelengths. Note that in the electromagnetic spec-
shorter the wavelength, the less it bends. Thus, the trum, infrared is next to visible red and ultraviolet is
diffraction of electromagnetic radiation is more appar- next to visible violet. You should keep in mind that as
ent for the longer wavelength components of the the wavelength is changed, one color smoothly blends
spectrum. This is evident in the names of the longer into another color. Hence, 590 nm light looks oran-
wavelength components (radio waves, television gish-yellow, while 560 nm light looks greenish-yellow,
waves, microwaves) vs the shorter wavelength compo- etc.
nents (x-rays, gamma rays). Monochromatic light that consists of an incident
The 400 nm to 700 nm wavelengths of visible light wavelength of 530 nm typically generates a pure or
are relatively small; hence, diffraction of light is often saturated green response. Incident light that has equal
negligible. In these situations we talk of straight line amounts of all wavelengths together with an extra
or rectilinear propagation of the light. Nevertheless, amount of 530 nm wavelength generates a desaturated
particularly when light propagates through a small green response. When the amount of 530 nm light is
opening, such as the spaces in a bird's feather, diffrac- increased, the response moves toward the saturated
tion can become a dominant effect. For a small pupil green. When the amount of 530 nm light is decreased,
size, diffraction introduces some blur into the retinal the response further desaturates and moves toward
image of the human eye. Diffraction is discussed in white. However, a desaturated green response can
later chapters of this book together with other wave also be generated by combinations of longer and
effects, such as interference and polarization.
The quantum theory of matter attributes wave
propagation properties to all particles including elec- TABLE 1.1
trons, protons, and neutrons. The quantum theory Color Response vs Vacuum Wavelength
also states that electromagnetic radiation is composed
of zero rest mass particles called photons. Typical light Color Response Vacuum Wavelength (nm)
levels consist of stupendous numbers of photons. A
60-watt light bulb puts out about 1019 photons/s. Red 780-620
Orange 620-590
Millions of photons can be added to or removed from Yellow 590-560
the 1019 level without any visible effect. Therefore, we Green 560-490
are not aware of the particle nature of light. The wave Blue 490-450
nature of light is simply a manifestation of the wave Indigo 450-430
Violet 430-380
propagation properties of the individual photons.
6 Geometrie, Physical, and Visual Optics

shorter wavelengths that do not contain 530 nm and At the other extreme are weak absorbers. Weak
that by themselves do not generate a green response. absorbers are usually transparent, but even a weak
In addition, light containing 530 nm together with absorber can become opaque if the thickness is great
longer wavelengths stimulates a color perception that enough. Water absorbs red light weakly. In typical
tends toward yellow or orange. Thus, light of wave- quantities, such as in fishbowls, bathtubs, and swim-
length 530 nm is not intrinsically green, but instead is a ming pools, water is transparent for all wavelengths
stimulus to the green response. including red. However, in oceans illuminated by sun-
Benham's top provides a neat example of the light, the red light fails to penetrate deeper than
psychophysiological dependence of color perception. 30 meters.
The top has a black and white pattern on it. When the Clear glass is evenly transparent at normal thick-
top is rotated at the correct frequency, flashing col- nesses. Tinted glass, as in stained glass windows, has a
ored spots appear on it. Somehow the varying light reduced transmission for the absorbed wavelengths.
levels from the rotating black and white pattern are
neurally processed to give the perceived color re-
sponses.
The psychophysiological dependence of color per- 1.9 Reflection
ception can also be noted by viewing the same color
against different backgrounds. The appearance of the Transmission of light through a medium is decreased
color changes slightly but definitely with the changing by any absorption that may be present. In addition to
surroundings. absorption, the amount of transmitted light can be
While color is clearly recognized as a perceptual reduced by surface reflection. The word reflection
response, it is still convenient to tag or label mono- comes from the Latin word reflectere, which means to
chromatic light by the color responses listed in Table bend back. Reflection is a surface or boundary
1.1. In other words, monochromatic light of vacuum phenomenon. In general, whenever light is incident on
wavelength 530 nm is referred to as green light, mono- a surface or a boundary between two different med-
chromatic light of vacuum wavelength 650 nm is re- iums, some of the incident light is reflected or bent
ferred to as red light, and monochromatic light of back. The percent of the light reflected, and the
vacuum wavelength 460 nm is referred to as blue light. wavelength dependence of that percentage, depends
on the materials involved and the angle of incidence of
incident light.
On a molecular level, the reflection process
involves an interaction of the incident light with the
1.8 Absorption electrons in the atoms and molecules of the surface
layers. There is a connection between a material's
Different materials have different absorption strengths absorption properties and its reflection properties.
for visible light. In many cases, the absorption Strong absorbers, such as metals, are strong reflectors.
strength of a material is also a function of the incident Weak absorbers, such as water and clear glass, are
wavelength. A material may absorb the long (or red) weak reflectors.
wavelengths more than the short (or blue) wave- Reflection is classified as either specular or diffuse.
lengths. Wavelength-dependent absorption is called The word specular means mirrorlike. Specular reflec-
selective absorption. tion occurs when the surface is smooth and is the
The total amount of light absorbed in a medium reflection involved in the formation of images by
depends on the absorption strength and on the dis- mirrors or other smooth surfaces such as a pond of
tance that the light travels in the medium. A strong water. Diffuse reflection occurs when the surface is
absorber is usually opaque, but when made thin rough.
enough it can transmit a significant percentage of the Let us consider specular reflection first. Under
incident light. rectilinear propagation conditions, the direction that
Gold, like other metals, is a strong absorber and is light is traveling can be represented by straight lines
opaque for typical thicknessess. However, a thin film called rays. Figure 1.4a shows a ray incident on a
of gold, deposited by vacuum evaporation techniques, smooth surface. The line perpendicular to the surface
transmits a significant percentage of light. The trans- is called the normal to the surface. The angle of
mission of the thin gold film is actually a selective incidence of the ray is defined as the angle 6j that the
transmission. Gold absorbs more strongly in the red ray makes with the normal. The angle of reflection is
part of the spectrum. Thus, the transmitted light is defined as the angle 6S that the reflected ray makes
greenish-blue. with the normal. The law of reflection states that the
Optics, Light, and Vision 7

normal normal

surface surface
a) b)

I normal
point w observer's eye
source |

surface

FIGURE 1.4. Reflections from smooth sur-


faces. a. One ray. b. Parallel rays. c. Position
of the reflected image.

angle of reflection 9S is equal to the angle of the smooth. At some stage a reflected image starts to
incidence 0i5 or mathematically, become visible. Initially, the image is a degraded or
e, = e, (l.i) poor quality image instead of a perfectly clear image,
as in specular reflection from a perfectly smooth sur-
The law of reflection is easy to determine and was face. The reflection from the partially polished surface
known by the ancient Greeks. The law of reflection is has a specularlike component that produces the image
independent of the wavelength of the incident light. plus a diffuselike component that degrades the image.
Figure 1.4b shows three parallel rays incident on Many naturally occurring surfaces give this type of
the smooth surface. The law of reflection applies to mixed reflection.
each ray so the three reflected rays are still parallel. The percent of light reflected can be selective. Gold
Figure 1.4c shows three rays diverging from a point absorbs reds more strongly than the blues and, con-
source of light. Again, the law of reflection applies to sequently, reflects reds more strongly than blues. This
each ray. A scaled drawing easily shows that the can be observed by looking at the color of the image
reflected rays appear to be diverging away from a of a blue sky formed by reflection off a smooth gold
point below the surface. This is the image point that surface. In contrast to gold, silver absorbs and reflects
an observer looking at the reflected light would see. uniformly across the visible spectrum. Transparent
Flat bathroom mirrors as well as smooth water sur- materials also reflect uniformly across the visible
faces form reflected images in this manner. spectrum.
Now let us consider diffuse reflection. Figure 1.5
shows four parallel rays incident on a rough surface.
Here the law of reflection still applies for each ray;
however, the incident angles of each ray are different
and, consequently, the reflected light is diffused in all
directions. In diffuse reflection, the information car-
ried by the incident light beam is lost and no reflected
image of the original source can be seen. Instead, the
diffusing surface itself becomes a secondary source.
Suppose we take a rough metal surface that gives
perfectly diffuse reflection and start polishing it FIGURE 1.5. Reflection from a rough surface.
8 Geometrie, Physical, and Visual Optics

Wow!
What a
beautiful
blue
sky!

FIGURE 1.6. Some effects of atmospheric light


scattering.

1.10 Scattering are smaller than the wavelength of light is called


Rayleigh scattering.
In an optically homogeneous material, a light beam is The moon has no atmosphere. Therefore, if one
attenuated either by absorption in the medium or by were to stand on the moon and look at the sky away
reflection at the boundaries between the mediums. from the sun, the sky would appear black because
However, when the material is not optically homoge- there is no scattering of light. Can you guess how a
neous, the light beam can also be attenuated by lunar sunset would appear?
scattering at each of the nonhomogeneous particles. Clouds contain either water droplets or ice crystals
(As a model for homogeneous and nonhomogeneous suspended in the air, while fogs contain suspended
materials, you might think of vanilla ice cream as water droplets. The suspended water droplets are
homogeneous and chocolate chip ice cream as non- gigantic relative to the air molecules. Each water
homogeneous.) droplet scatters light predominately by surface reflec-
The amount of scattering depends on the relative tions. The scattering from the water droplets is essen-
properties of the nonhomogeneous material and on tially wavelength-independent so the scattered light
the number of nonhomogeneous particles. Scattering looks white. Since multiple scattering can occur in
occurs throughout the nonhomogeneous material, and clouds and fogs, some of the scattered light emerges in
just as for absorption, the attenuation of a light beam the forward direction. A frequent effect of this type of
is a function of the total thickness of the scattering scattering is that the forward scattered light also enters
medium. the eye causing a contrast reduction in the retinal
Scattering can also be selective or nonselective. image of distant objects. The distant objects then
The blue sky is a familiar example of selective scatter- seem to be in a haze or may even become invisible.
ing. Air molecules are very weak light scatterers. They Materials that diffuse or scatter transmitted light
do, as does any object much smaller than visible light are called translucent materials. They are sometimes
wavelengths, scatter blue light more than red light beneficial as in supplying softer lighting, and some-
(i.e., shorter wavelengths more than longer wave- times detrimental as in dense fogs.
lengths). Thus, when we look at the sky away from the
sun, it appears blue because it is the scattered light
that is reaching our eye. In Figure 1.6, the observer at
A looking toward C sees the scattered light, which is 1.11 The Speed of Light in a Medium
predominately blue. The observer at B looking toward
C sees the setting sun. The light reaching B's eye has When light enters a transparent medium from a vac-
lost more blue light than red light due to scattering by uum, it slows down. The speed of the light in the
the air molecules as well as by suspended dust in the medium is characteristic of that medium and of the
lower atmosphere, so the observer at B sees a reddish wavelength of the incident light. The index of refrac-
or orangish setting sun. The predominately bluish tion n of the medium is a ratio of the speed c of the
scattering characteristic of molecules or particles that light in a vacuum to the speed v of the light in the
Optics, Light, and Vision 9

medium, or travel in the water is different from that in the air. The
rays indicating the direction of travel are perpendicu-
n = c/v. (1.2)
lar to the wavefront. In both air and water the rays are
Equation 1.2 can be rearranged to give straight lines, but at the air-water boundary the rays
are bent. The law of refraction states that the angle of
v = c/n. (1.3)
incidence fy, made with the normal to the surface, is
For incident light of wavelength 589.3 nm, the indices related to the angle of refration , made with the
of refraction of some common transparent materials normal to the surface, by the equation
are: water, 1.33; plastic, 1.44 to 1.49; crown glass,
n1 sin 6j = n2 sin 0 r , (1.4)
1.523; other glasses, 1.5 to 1.9; diamond, 2.4. The
index of refraction of air at 20C and 1 atmosphere of where nl is the index of refraction of the medium in
pressure is 1.0003. The wavelength dependence of the which the light is incident, and n2 is the index of
index of refraction of transparent materials is slight refraction of the medium in which the light leaves
and is ignored until Chapter 20. (Figure 1.8).
When the angles 6j and are small, Eq. 1.4 can be
simplified by its small angle approximation,
n ^ n ^ . (1.5)
1.12 Refraction
For light in air (n=1.00) incident on a flat water
When you stick a straight object such as a ballpoint (n = 1.33) surface at an incident angle of 6, Eq. 1.5
pen into some water, the object appears to be bent at gives the angle of refraction as:
the water's surface. Actually, the object remains (1.00)6 = (1.33),
straight, but the light traveling through the surface is
= 6/1.33 = 4.51.
bent. The bending of light at a surface is called
refraction. The word refraction comes from the Latin For light going from air into water the rays get bent
word refractus which means to break back. toward the normal. In fact, Eq. 1.4 or 1.5 shows that
Refraction occurs because of the speed change that bending toward the normal occurs whenever the light
light undergoes when it changes mediums. Consider is entering the higher index medium. When light
light waves in air incident on a flat water surface. travels from the higher index medium to the lower
Assume that the index of refraction, n, of the air is index medium, then Eq. 1.4 or 1.5 shows that the rays
1.00, and that the index of refraction, n, of the water get bent away from the normal.
is 1.33. For a distant point source the incident wave- The phenomenon of refraction was studied by the
fronts are flat, as shown in Figure 1.7a, and are ancient Greeks. Ptolemy of Alexandria made accurate
perpendicular to the direction of travel, as shown by measurements on the refraction angles in 130 AD but
the rays in Figure 1.7b. failed to determine a describing equation. Later,
The lower edge of the incident wavefront enters the about 1000 AD, the Moslem scholar known as
water first, and hence slows down first. Thus, in the Alhazen extended the currently known knowledge of
same amount of elapsed time, that part of the wave- optics and studied refraction, but also failed to de-
front in the water travels a smaller distance than that termine a describing equation. By 1611 AD, Johannes
part of the wavefront in air. The net effect is that the Kepler had determined the equation (Eq. 1.5) that
wavefront becomes skewed so that its direction of describes refraction for small angles, but it was not

FIGURE 1.7. Refraction at a flat air-water inter-


face. a. Wavefronts. b. Wavefronts and rays.
10 Geometrie, Physical, and Visual Optics

normal normal

FIGURE 1 8. Refraction angles, a. n 2 >n,. b.


n2 <n,.

until 1621 AD that Willebrord Snell discovered an slightly different angles that shadows made at high
equation that described refraction for large as well as noon in deep holes that were located many miles apart
small angles. In 1637 AD, Ren Descartes published from each other.
the law of refraction in the form given in Eq. 1.4. When more than one light source is present, then
There is a historical controversy about whether De- the partial shadows can be formed as well as full
scartes independently discovered the law of refraction shadows. A full shadow is called an umbra while a
or whether he had seen Snell's previous but un- partial shadow is called a penumbra.
published work. As a result of this controversy, the Shadows provide some relief on hot summer days,
law of refraction is variously called Descartes' Law but they have an even more important use. Shadows
(especially in French speaking countries) and Snell's provide information or cues about the three-dimen-
Law (especially in English speaking countries). The sionality of solid objects, and these cues are used by
determination of the long hidden law of refraction was our visual system as an aid in depth perception.
a crucial and necessary step in the development of The law of rectilinear propagation breaks down in
high quality optical instruments. those situations in which diffraction of light waves
becomes important. The law of rectilinear propagation
also breaks down when a medium is not optically
isotropic. (Isotropie means uniform in all directions.)
1.13 Rectilinear Propagation, Shadows, The index of refraction of air is temperature-depen-
and Mirages dent. On occasion, this temperature dependence can
lead to observable effects. The air directly above a hot
In most situations, light exhibits a definite straight line surface on a sunny summer day is warmer than the air
or rectilinear propagation. In nature, rectilinear prop- at higher levels above the surface. Because of this, the
agation is observed when shafts of sunlight propagate index of refraction of the air smoothly changes as a
in straight lines through holes between adjacent clouds. function of altitude above the surface, so the air is not
(Such shafts spreading across the sky are called cre- optically isotropic.
puscular rays.) Light rays propagating through the hotter air near
Shadows are another manifestation of rectilinear the surface bend upward while the rays at higher levels
propagation. You are not doubt familiar with your above the surface remain straight. The result is a
own shadow and well aware that you can count your mirage. The person can see both the object and a
fingers by looking at the shadow of your hand, or even shimmering inverted image of it (Figure 1.9). This
use your hands to make shadows that resemble those type of mirage is called an inferior mirage because the
of wolves, elephants, etc. You can also tell the direc- image appears below the object. Automobile drivers
tion of the primary light source by noting the relative are familiar with the inferior mirage as the shimmering
locations of the shadow and the object that is blocking "water" on the road ahead. The shimmering water is
the light. actually an inferior mirage of the sky above. Next time
It is interesting to note that the ancient Greeks you see this inferior mirage, check for the inverted
believed the Earth was round, not flat as was believed image of an automobile or truck in front of you.
by the later Europeans. As one of the supports for this There are many other types of mirages depending
belief, the Greeks cited the shape of the Earth's on the index profile. These include the superior mir-
shadow seen on the moon during a lunar eclipse. They age and the fata morgana mirage. To pursue mirages
contended that the shadow indicated a spherical Earth would unfortunately lead us away from the goals of
and not a flat Earth. To make a reasonable estimate of this book, just as desert mirages have allegedly misled
the radius of the Earth, the Greeks also used the thirsty travelers.
Optics, Light, and Vision 11

u
hot road

FIGURE 1.9. a. Curved rays. b. The in-


ferior mirage.

Problems 5. Light in water (n = 1.33) is incident on a water-air


interface at 4.0 incident angle. Use the small
angle version of Snell's Law to find the angle of
1. Light travels at a speed of 1 . 8 x l 0 8 m / s in a refraction. Does the ray bend toward or away
transparent medium. What is the refractive index from the normal?
of the medium?
6. Light in air is incident on glass (n= 1.56). When
2. The refractive index of water is 1.33. What is the the angle of incidence is 43, what is the angle of
speed of light in water? reflection?
3. Light in air is incident on air-glass (n = 1.523) 7. What is an isotropic medium?
interface. When the angle of incidence is 52, what
is the angle of refraction? Does the light bend 8. What is a homogeneous medium?
toward or away from the normal? 9. White light is incident on a red leaf. Does the light
4. Light in air is incident on an air-glass (n = 1.616) that is diffusely reflected from the leaf contain an
interface at an angle of 3.5. Use the small angle abundance of short wavelengths or an abundance
approximation to Snell's Law to find the angle of of long wavelengths?
refraction. Now use the exact version of Snell's 10. Specify the vacuum wavelengths that are in the
Law to find the angle of refraction. How close are middle of the following regions of the visible light
the answers from the two methods? spectrum: orange, yellow, green, and blue.
CHAPTER TWO

The Geometric
Behavior of Light

2.1 Point Sources, Wavefronts, and Rays wavefront. Figure 2.3 presents the rays for the situa-
tion shown in Figure 2.2. Since the medium is not
The geometric optics theory of image formation is uniform in the upper half of Figure 2.3, the wavefronts
built on the concept of point sources and point images. there are not spherical. Since the wavefronts are not
Consider an isolated point source in a uniform spherical, the rays are curved instead of straight even
medium. The isolated point source emits light waves though they are drawn in perpendicular to each wave-
equally in all directions and the waves travel away front. These curved rays are the type that can occur in
from the source at the same speed. The light waves mirages, in atmospheric refraction, or in gradient
leaving the point source are represented by the wave- index materials. (The acronym currently used for
fronts in Figure 2.1a. Since the wavefronts propagate gradient index is GRIN.) In the lower half of Figure
away from the source at the same speed in all direc- 2.3 the medium is uniform, so the wavefronts there
tions, they have a spherical shape. Each of the diverg- are spherical in shape. Consequently, the rays form
ing spherical wavefronts is centered on the point straight lines pointing away from the point source.
source. Rays and wavefronts provide interchangeable infor-
Figure 2.1b shows a point source embedded in the mation. When the positions of the wavefronts are
surface of a wall. The embedded point source emits known, then one can always draw the rays. Converse-
light in the geometrically allowed directions. In the ly, when the positions of the rays are known, then one
case of the embedded point source, the diverging can always draw the wavefronts. In discussing the
wavefronts are again parts of concentric spheres cen- optics of image formation, it is sometimes conceptual-
tered on the point source. ly easier to use the rays, while at other times it is
If the medium is not uniform, then the wavefronts conceptually easier to use the wavefronts.
will not propagate away at equal speeds. Figure 2.2 In most situations the rays represent the direction
shows a situation in which the lower half of the in which the electromagnetic energy propagated by the
medium is uniform but the upper half is not. The wave is traveling. The exception occurs in light propa-
index of refraction in the upper half of the medium gation through certain anisotropic crystals. The crystal
decreases as a function of vertical distance above the case is not of major concern in this book.
point source. While the wavefronts in the lower half of The situations of major interest in this book are
the medium are spherical, the wavefronts in the upper the situations involving an isotropic homogeneous
half of the medium lose their sphericity because the medium. Therefore, the wavefronts diverging from a
waves propagating upward are moving faster than the point source are spherical in shape, and the corres-
waves propagating outward or downward. ponding rays are straight lines pointing away from the
A ray is defined as a trajectory orthogonal to the point source. The collection of rays or wavefronts
wavefronts. When given the wavefronts, one can draw coming from a single point source is called a homocen-
the rays by having them pass perpendicular to each tric (or monocentric) bundle. A cross section of a
13
14 Geometric, Physical, and Visual Optics

ure 2.5a are all parts of concentric spheres. Each of


the wavefronts is centered on the same point. The
converging light leaving the system forms the point
image and then diverges away from it. Since the
converging wavefronts are spherical, the rays leaving
a) b)
the system are straight lines that point toward the
image point (as shown in Figures 2.5b and 2.5c). The
FIGURE 2.1. a. Diverging waves from an isolated point source. rays remain straight and after passing through the
b. Diverging waves from a point source embedded at B. image point diverge away from it.
When the apertures in the optical system are circu-
lar, then a screen placed in front of the image point
would have a circular illumination patch on it (e.g., at
position A in Figure 2.5c). The circular illumination
patch is called a blur circle. The diameter of the blur
circle is specified by the outer rays. As the screen is
moved back to the image position (position B in
Figure 2.5c), the blur circle decreases in size and
becomes a point. When the screen is moved past the
FIGURE 2.2. Wavefront propagation in a medium that is
nonuniform above the dashed line and uniform below. image position (e.g., toward position C in Figure
2.5c), the blur circle increases in size.
As discussed in Chapter 1, theories are actually
models of the real world. The concept of a point
image occurs in the model represented by the theory
of geometric optics. In the real world only approxi-
mate point images occur. The optical system may not
be perfect so that the converging wavefronts leaving
are not perfectly spherical and a perfect point image is
not formed. Such deviations from perfection are called
FIGURE 2.3. The rays for propagation in a nonuniform (top) aberrations and are discussed in Chapter 20.
region vs uniform (bottom). Even more fundamentally, light propagates
through space in a wavelike manner and waves exhibit
diffraction effects. All optical systems have some aper-
tures in them. Even a single lens has a finite size and
therefore is not only a lens but also an aperture. The
a) finite size apertures cause some diffraction of the light
waves, which prevents the formation of a perfect point
FIGURE 2.4. Uniform propagation from a point source, a. image. Nevertheless, in many cases the apertures are
Spherical diverging wavefronts and rays. b. Rays only.
much much larger than the wavelength of the light,
and the diffraction effects are so small that the image
homocentric bundle and its point source is called a can be considered a point image. Consequently, the
pencil (Figure 2.4). theory of geometric optics is tremendously useful in
describing the image formation process even though it
neglects diffraction. Note that the behavior of a sys-
tem with smaller and smaller apertures eventually
2.2 Converging Wavefronts, Point Images, begins to deviate from the geometric optics de-
and Blur Circles scription.
For the present, assume that diffraction is neglig-
The ideal imaging system takes incoming spherical ible and that the wavefronts leaving an optical system
diverging wavefronts and changes them into outgoing are spherical wavefronts. This assumption holds well
spherical converging wavefronts. The outgoing wave- for regions near the axis of the optical system, or more
fronts then converge to a point as shown in Figure mathematically, for regions in which the paraxial ap-
2.5a. This point is called the image point. Figure 2.5a proximation (literally near the axis) is valid. To be
represents the imaging process in terms of the wave- classified as paraxial, a ray must be near the axis and
fronts, Figure 2.5b shows the rays as well as the make a small angle with the axis. When this condition
wavefronts, and Figure 2.5c shows the rays only. is met, we can use the small angle version of the law of
The outgoing converging wavefronts shown in Fig- refraction for paraxial rays (Eq. 1.5).
The Geometric Behavior of Light 15

point image

point source

FIGURE 2.5. Ideal imaging process. The optical sys-


tem is represented by the box. a. Wavefronts. b. Wave-
fronts and rays. c. Rays only.

2.3 Diverging Wavefronts, Plane Waves, waves coming from it can be considered plane waves,
and Optical Infinity the object point is said to be at optical infinity.
Figure 2.7a shows another representation of the
Everyone who has driven a car or ridden a bicycle is plane wave case. The wavefronts on the right are
intuitively familiar with the concept of curvature. One effectively flat in the localized region under consider-
must slow the car or bike down in order to negotiate a ation because they have traveled a sufficiently large
sharp or highly curved turn. On the other hand, one distance from the point source. Figure 2.7b shows the
can speed through a gradual or relatively flat turn that corresponding rays. When the wavefronts are flat,
has a small curvature. then the rays drawn in perpendicular to the wavefronts
Consider the diverging wavefronts shown in Figure are parallel to each other. These parallel rays can be
2.6. Each of the wavefronts is part of a sphere that is thought of as being very slightly divergent. However,
centered on the point source. The wavefronts nearest in the plane wave region the degree of divergence is so
the point source have a high curvature while the slight that the rays are effectively parallel to each
wavefronts farthest from the point source have a other. Even though these rays are then drawn as
smaller curvature. In fact, the wavefront farthest from parallel, one should keep in mind that they originated
the point source is part of an extremely large sphere at the point source at optical infinity.
and, in the localized region shown, is almost flat. Figure 2.8a shows converging spherical wavefronts
From Figure 2.6, it is clear that the spherical leaving an optical system. Each of the converging
diverging wavefronts lose curvature as they move out wavefronts is spherical and centered on the image
away from the point source. In a localized region far point. The converging wavefronts gain curvature as
away from the point source the wavefronts are very they travel away from the optical system toward the
flat (as shown on the right in Figure 2.6). In the local image point. In particular, the closer the wavefronts
regions where the wavefronts become very flat, they are to the image point, the higher the wavefront
are referred to as plane waves (plane meaning flat). curvature.
When the object point is far enough away so that the Suppose the light leaving an optical system is only

observer's
eye

))
FIGURE 2.6. Diverging spherical wave-
front. Each wavefront is centered on the point
source.
16 Geometrie, Physical, and Visual Optics

>. is neglected, and the index of air is assumed to


equal 1.
point source large i As discussed in the previous sections, the wave-
distance \ fronts associated with point objects and point images
Or) are spherical. From calculus, the curvature of a sphere
is equal to the reciprocal of the radius of the sphere.
plane waves In air (n = 1), the curvature of the spherical wavefront
a) is directly related to the degree of convergence or
| divergence of the light at the position of the wave-
%_ front.
1
The quantitative measure of the degree of converg-
distance ^
V ence or divergence of light in air at a particular
3 ^ position is called the vergence of the light. For light in
air, the vergence at a position is equal to the curvature
FIGURE 2.7. Plane waves that originated at the point source. of the wavefront at that position. The unit for ver-
a. Wavefronts. b. Wavefronts and rays. gence is the diopter (D), which is dimensionally equal
to the reciprocal of a meter (m). Let V stand for the
vergence of the light at a particular position, Q stand
for the curvature of the wavefront at that position, and
q stand for the radius of curvature of the wavefront.
image point Mathematically,
V=Q, (2.1)
and for a sphere,

wttt 1 | | | >
m\
111 > (2.2)
image point formed far away -
Therefore,
b)
FIGURE 2.8. a. A converging bundle leaving a system, b. Very q in meters. (2.3)
slightly converging bundle leaving a system. q
In order to distinguish converging light from di-
verging light, a negative vergence value is used for
very slightly convergent. In this case, the convergent diverging light and a positive vergence value is used
wavefronts leaving the system have only the minutest for converging light. Equations 2.1 to 2.3 can be put in
amount of curvature, and the image point, which is absolute values and then used to get the magnitude of
located at the center of curvature of the converging the vergence. The appropriate plus or minus sign can
wavefront, is exceedingly far away. Then we say that then be inserted depending on whether the light is
the image point is at optical infinity, and it is conven- converging or diverging.
ient to approximate the very slightly converging wave- Alternatively, one can use q as a directed distance
fronts by plane waves. The rays representing the light instead of just a magnitude and establish a plus-minus
are really very slightly convergent, but for all intents sign convention for q. The sign convention for q is as
and purposes they can be considered parallel to each follows. Always measure q from the wavefront to the
other at least in the region immediately behind the center of curvature of the wavefront. When the di-
optical system. rected distance from the wavefront to the center of
curvature of the wavefront is in the same direction
that the light is traveling, then q is a positive value. If
the directed distance from the wavefront to the center
2.4 The Concept of Vergence of curvature of the wavefront is opposite to the direc-
tion that the light is traveling, then q is a negative
The index of refraction of a vacuum is, by definition, value. With these sign conventions, q is always nega-
exactly 1. The index of refraction of air is just slightly tive for a diverging wave and positive for a converging
larger (typically 1.0003). For now, the small difference wave.
The Geometric Behavior of Light 17

labeled A is a distance of 5 cm or 0.05 m from the


point source. The directed distance from this wave-
front to the point source is opposite to the direction
that the light is traveling, so the radius q is negative.
Therefore, q equals -0.05 m. Then, from Eq. 2.3,
a) b) 1
V= = -20.00 D.
FIGURE 2.9. a. A diverging wavefront propagating to the (-0.05 m)
right, b, A converging wavefront propagating to the right.
The wavefront at B is a distance of 25 cm or 0.25 m
from the point source, so q is -0.25 m. Then,
In Figure 2.9a, a diverging wavefront is propagat-
ing to the right. The center of curvature of the diverg- 1
V= = -4.00D.
ing wavefront is at point A, which is to the left of the (-0.25 m)
wavefront. The directed distance from the wavefront
In Figure 2.11, the center of curvature of each of
to its center of curvature is opposite to the direction
the converging wavefronts is at the point image posi-
that the wavefront is traveling; therefore, q is nega-
tion. Since the light is converging, the vergence of
tive. In Figure 2.9b, a converging wavefront is prop-
each wavefront is positive. The wavefront on the left is
agating to the right. The center of curvature of the 1 m from the point image. The directed distance from
converging wavefront is at point B, which is to the the wavefront to the point image is in the direction
right of the wavefront. The directed distance from the that the light is propagating; therefore, q equals +1 m.
wavefront to point B is in the same direction that the From Eq. 2.5.
light is propagating, so q is positive. The values q, Q,
and V in Eqs. 2.1 to 2.3 are all negative for diverging 1
light and all positive for converging light. Note, the V= = +1.00D.
( + lm)
fundamental assignment was that diverging light has a
negative vergence and that converging light has a The next wavefront is 0.5 m from the point image, and
positive vergence. from Eq. 2.5,
Consider the diverging waves shown in Figure 2.10.
Since the wavefronts are diverging, the vergence of
V= = +2.00D.
each wavefront is going to be negative. The wavefront (+0.5 m)

-100.D-20.D -10.D -5.D -4.D -2.D -1.D

point y
source * /

0 1. 5. 20. 50. 100. FIGURE 2.10. Vergence values for


distance in centimeters - wavefronts diverging from a point source.

+ 1.D + 2.D + 4.D +5.0D + 10.D +20.D

W//
i J_
/ / / / /

L
[ / , point

m
I \ image
\
\ 1
100. 50. 25. 20. 10. 5.r1
0 FIGURE 2.11. Vergence values for converging
distance in cm to the point image wavefronts.
18 Geometrie, Physical, and Visual Optics

2.5 Vergence: Conversion Factors the vergence gets closer to zero as the wavefronts
move further from the point source. This is a result of
Frequently, the distances of optical interest are ex- the flattening of the diverging waves as they move
pressed in centimeters (cm) or in millimeters (mm), away from the point source (Figure 2.10). For large
instead of in meters. The distances can be converted enough distances, the vergence is effectively zero and
to meters and then Eq. 2.3 is used to calculate ver- the waves are considered plane waves. The distance is
gence. However, sometimes it is quicker and intuitive- then designated by the infinity symbol (00).
ly easier to use the conversion factor explicitly in the In terms of Eq. 2.5, when the absolute value of q,
vergence equation. labeled \q\, is a very large number, then V is a very
Suppose the radius q is expressed in centimeters. small number. As \q\ gets larger and larger, V gets
Since there are 100cm/m, Eq. 2.3 becomes closer and closer to zero. In the limit of\q\ going to 00,
V goes to zero.
V = In Figure 2.11, the center of curvature of each of
^/(lOO cm/m)' ^2'4^ the converging wavefronts is at the point image posi-
When algebraically simplified, Eq. 2.4 results in tion. Since the light is converging, the vergence of
each wavefront is positive. The wavefront on the left is
(100 cm/m)
q in cm. (2.5) 100 cm from the point image. The directed distance
from the wavefront to the point image is in the
Equation 2.5 shows that the conversion factor can be direction that the light is propagating, therefore, q is
written in the numerator. + 100 cm. From Eq. 2.5,
Again consider the diverging waves shown in Fi- lOOcnWm
gure 2.10. The wavefront labeled A is a distance of
( + 100 cm)
0.05 m or 5 cm from the point source, so q equals
- 5 cm. Then, from Eq. 2.5, The next wavefront is 50 cm from the point image, and
from Eq. 2.5,
w 100cm/m ~
V = -7z 7- = -20.00 D, 100
( - 5 cm) V = (+50) = +2.00D.
which is the same value as before.
Similarly, the wavefront at B is a distance of 0.25 m The vergence values for the other wavefronts are
or 25 cm from the point source, so q equals - 2 5 cm. found similarly and are listed in Table 2.2. Note that
Then, the vergence of the converging light increases as the
light moves closer to the point image position. This
100
V= = -4.00D occurs because the curvature of the converging wave-
(-25) fronts is increasing as the light moves closer to the
Table 2.1 shows the vergence values for diverging image position.
wavefronts as a function of the distance (absolute When the radius q is expressed in millimeters, then
value) from the point source. Note in Table 2.1 that q can be converted to meters and the reciprocal taken
to get the vergence (Eq. 2.3). Alternatively, the con-
version factor can be put in algebraically, in which
case it again ends up in the numerator, or
TABLE 2.1
Vergence Values for Diverging Light 7 (1000 mm/m) .
v , q 111 111111. Vz',u/

Distance from Point Source \q\ Vergence (V) (in air)


(cm) (D)
TABLE 2.2
1 -100.00 Vergence Values for Converging Light
5 -20.00
10 -10.00
20 -5.00 Distance to Point Image (cm) Vergence (D)
25 -4.00
50 -2.00 100 + 1.00
100 -1.00 50 +2.00
200 -0.50 25 +4.00
500 -0.20 20 +5.00
1000 -0.10 10 + 10.00
00 0 5 +20.00
The Geometric Behavior of Light 19

As an example, consider a converging wavefront the point source. From Eq. 2.8,
with a 20 mm radius of curvature. (Note, by the sign
40
convention chosen the vergence must be positive since v
the light is converging.) From Eq. 2.6, ~F3) = _ 3 - 0 8 D
The difference between the approximate value
1000 mm/m (-3.08 D) and the exact value (-3.03 D) is 0.05 D. In
V = =+50.00 D.
(+20 mm) most visual optics cases this difference is negligible.
For a diverging wavefront 2 mm away from the point
source,
1000 2.6 Upstream and Downstream Vergence
V= =-500.00 D.
(-2) Changes
The metric system conversion factors are multiples This section discusses examples of how to compute
of 10 and are all extremely easy to use. If the radius of vergences either upstream or downstream from the
the wavefront is given in inches or feet, then the location of a known vergence value. As mentioned in
conversion factors are not convenient multiples of ten. the last section, one can use the absolute values of the
There are 2.54 centimeters per inch or 1/2.54 inches radius of curvature q of the wavefronts and get the
per centimeter. The reciprocal of 2.54 is 0.394. In one correct sign for the vergence by checking whether the
meter there are 100 cm and, consequently, 39.4 inches. light is diverging or converging. The latter method has
The conversion factor of 39.4 in/m can be used to get the advantages that one does not have to worry about
the relationship, the correct plus-minus sign for q. More importantly, it
(39.4 in/m) helps one to think in terms of whether the light is
q in inches. (2.7) converging or diverging. All the examples in this
section are worked using the absolute value of the
As an example, consider the vergence of a diverg- radius of curvature.
ing wavefront located 13 inches from a point source.
EXAMPLE 2.1
39 4 Consider diverging light traveling to the right as
V= = - 3 03 D shown in Figure 2.12. The vergence of the light at
point A is -10.00 D. What is the vergence of the
The conversion factor of 39.4 is not as easy to work light at point B, which is 15 cm downstream from
with as the metric system units particularly when a A?
calculator is not available. For cases in which precision A good problem solver tries to anticipate or
predict what characteristics the answer will have.
is not needed, the conversion factor of 39.4 in/m is From the sketch, the wavefronts at B will still be
approximated by 40 in/m. With this approximation, diverging and will be flatter than the wavefronts at
A. Therefore, the expected vergence at B will be
(40 in/m) minus and will be a number smaller in magnitude
V q in inches. (2.8)
than 10.00D (i.e., we expect a number like - 8 D
or - 4 D as opposed to a number like - 1 2 D or
Now reconsider the example immediately above. -16 D). Now that we have our expectations, let's
There, a diverging wavefront was 13 inches away from do the calculation.

-10.0

\ \
point source <


v
"~Y~ ltm
FIGURE 2.12. Downstream vergence example.
20 Geometrie, Physical, and Visual Optics

The magnitude of the radius of curvature of the is shorter than the radius of the wavefront at C, or
wavefront at A is k D l = kcl-3cm,
100cm/m and
kAl = VA | qO| = 50 cm - 30 cm = 20 cm.
Then
100cm/m 100 cm/ml
I*AI = 10.00 cm. |v D | = = 5.00D.
(-IO.OOD)I (20 cm)
From Figure 2.12, the radius of curvature of the Since the light is diverging, the vergence is negative
wavefront at B is greater in magnitude than that at and
A, or VD = - 5 . 0 0 D ,
kBl = kAl + 15cm, which agrees with our anticipated results.
and
EXAMPLE 2.3
I qB | = 10 cm + 15 cm = 25 cm. The vergence of light at point C is +18.12 D. What
is the vergence of the light at point B, which is
The magnitude of the vergence VB of the wavefront
36.75 cm upstream? Try doing a quick sketch to
at B is then
convince yourself that VB should be positive and
100 cm/m less than +18.12 D. Then verify that the vergence
|v B | = = 4.00D, at position B is +2.37 D.
(25 cm)
Since the light is diverging, VB is minus, or EXAMPLE 2.4
VB = -4.00 D. Consider converging light traveling to the right
(Figure 2.14). The vergence of the light at point A
The answer qualitatively agrees with our antici- is +8.00 D. What is the vergence at point B, which
pated results that the wavefront would be flatter. is 5 cm downstream from A? The magnitude of the
radius of curvature of the wavefront at A is
EXAMPLE 2.2 100 cm/ m
In Figure 2.13, the vergence of light at position C is kAl = 12.5 cm.
(8.00 D)
-2.00 D. What is the vergence at position D, which
is 30 cm upstream from position C? The wavefronts are still converging at point B, so
Since diverging wavefronts lose curvature as the wavefront there is steeper than at point A
they propagate, the wavefront at D is more highly (Figure 2.14). Therefore, we expect the vergence
curved than the wavefront at C. (The conventional at B to be a number like + 1 0 D , + 1 5 D , or +20 D,
terminology is that the wavefront at D is steeper as opposed to + 6 D , + 4 D , or + 2 D .
than the wavefront at C.) Therefore, we expect From the sketch, the magnitude of the radius of
that the vergence at D will be more minus than the curvature at B is
vergence at C (i.e., we expect numbers like - 5 D \qB\ = 12.5 cm - 5 cm = 7.5 cm.
or - 1 0 D as opposed to - I D or - 0 . 5 D).
Then
The magnitude of the radius of curvature of the
wavefront at position C is 100 cm/ml
|vJ = = 13.33 D.
(7.5 cm)
100 cm/m Since the light is converging at point B, the ver-
= 50 cm.
kcl (-2.00) gence is positive and
From Figure 2.13, the radius of the wavefront at D VB = + 1 3 . 3 3 D.

-2.D
\

V
point source
# image point
l
Y '
1
/ ] /
1*1 I
Y
)
>r~- IQBT
L 30 cm ) ^ 5 cm
Y T
FIGURE 2.13. Upstream vergence example. FIGURE 2.14. Downstream vergence example.
The Geometric Behavior of Light 21

spherical wavefronts,

Q=-, (2.10)

where q is the radius of curvature. Therefore,

v=^. (2.11)

The index of refraction is a dimensionless number;


thus, the units remain the same as before. The ver-
gence V is in diopters when the radius q is in meters.
FIGURE 2.15. Downstream vergence example for point image Alternatively, conversion factors can be used in the
between wavefronts.
numerator when q is expressed in units other than
meters.
As an example, consider a point source in water
EXAMPLE 2.5 (n = 1.33). What is the vergence of the light in the
As a final example in this section, consider light at water at a distance of 28 cm from the point source?
position A with a vergence of +25.00 D (Figure According to the sign convention,
2.15). What is the vergence of the light at a posi-
tion 15 cm downstream? (Try working it out before q = - 2 8 cm,
you read the solution and be sure to try to antici-
pate the results.) and from Eq. 2.11,
The solution is as follows. The magnitude of the
radius of curvature of the wavefront at A is (1.33)(100cm/m)
V=
(-28 cm)
100cm/m
\qA = 4.00 cm. 133
(25.00 D) V=
(-28)'
A quick sketch shows that the light converges to a
point image 4 cm downstream from A and then or
diverges away from that position. Therefore, the V=-4.75D.
wavefront at position B will be diverging.
From the sketch, the magnitude of the radius of Note that the vergence is negative as it should be for
curvature of the diverging wavefront at B is diverging light.
Historically, generalized vergence has been re-
| qB | = 15 cm - 4 cm = 11 cm. ferred to by the name of reduced vergence. The
The magnitude of the vergence at B is ratonale for this terminology and the generalized defi-
nition of vergence will be given in Chapter 8.
lOOcnWm Note that if the light is in air, then n equals one and
1 Bl
(11 cm) the equations in this section then become identical to
Since the wavefront is diverging, the equations in the previous sections. Further, nu-
merical examples of light in media other than air will
VB = -9.09D. be deferred to Chapter 7.

2.7 Generalized or Reduced Vergence


2.8 Extended Sources and Beams
For light propagating in a medium in which the index
Common objects such as trees, people, and books
of refraction differs from one, the definition of ver-
consist of many different point sources. (Conceptual-
gence is generalized as follows: The vergence of light
ly, we might consider each atom in the object as a
at a position A in a medium of refractive index n is
point source.) A primary or secondary source consist-
equal to n times the curvature of the wavefront at ing of laterally separated point sources is called an
position A. Mathematically, extended source. Each point source on the extended
V=nQ (2.9) source puts out its own bundle of light. The collection
of all the bundles put out by the points on the source
where Q is the curvature and V is the vergence. For is called a beam.
22 Geometrie, Physical, and Visual Optics

Let us now consider the beam coming from a slide


or motion picture projector. The beam leaving the
projector is also frequently visible due to light scatter-
ing off dust, smoke, or suspended water droplets in
the air. The beam is diverging or spreading out as it
moves away from the projector. However, the diverg-
ing beam leaving the projector is made up of converg-
ing bundles rather than diverging bundles. Figure 2.17
shows the image formation portion of the projector
system. The bundle leaving the bottom point on the
FIGURE 2.16. A diverging beam made up of diverging bun- object is converged by the projector lens and eventual-
dles.
ly forms the top image point. The bundle leaving the
top point on the object is converged by the lens and
eventually forms the bottom image point. The extreme
The beam from typical primary and secondary ex- top and bottom rays are the beam boundaries. The
tended sources is a diverging beam made up of a beam is clearly spreading out or diverging even though
collection of diverging bundles. Consider the beam each individual bundle is converging. A flashlight
from a common flashlight. At night, the path of the beam never forms an image on a screen because the
flashlight beam is frequently visible due to light scat- individual bundles are diverging. A projector beam
tering off dust or suspended water droplets present in forms an image on a screen because the individual
the air. The beam itself spreads out or diverges as it bundles are converging. The image on the screen is
moves away from the flashlight. The diverging beam is larger than the projector lens because the projector
made up of many individual diverging bundles, each of beam is diverging.
which originated from its own specific point on the Figure 2.18 shows a converging lens that forms an
tungsten filament of the flashlight's bulb (Figure 2.16). extended image that is smaller than the size of the

projector

image

FIGURE 2.17. The light leaving the lens con-


sists of a diverging beam made up of converging
bundles.

-?> image

FIGURE 2.18. The light leaving the lens consists of a


converging beam made up of converging bundles.
The Geometric Behavior of Light 23

lens. The individual bundles leaving the lens are con- the tree at position A. Two bundles of rays are shown,
verging, which is why the image is formed. In this and each bundle forms its respective point image at
case the beam leaving the lens is shrinking in size or position A.
converging as it moves away from the lens. The When the object points all subtend a small enough
individual converging bundles form the image, but the angle directly in front of the lens, then the cross-
converging beam is responsible for the image size sectional shape of each bundle is well approximated by
being smaller than the lens size. a circle. If the screen is moved to position B, which is
The above examples show that a diverging beam closer to the lens, each bundle forms a blur circle
can be made with either diverging or converging bun- instead of a point image. The result is a blurred image
dles, and that a converging beam can be made with on the screen. The blurred image can be simulated by
converging bundles. There is a fourth logical possibi- drawing a small circle at each place that a point would
litythat of a converging beam made with diverging occur in the clear image. The result is still a recogniz-
bundles. Is such a case physically possible? The an- able tree, but fine detail (such as a bird's nest) may be
swer is yes. This situation is treated later in connection lost. If the screen is moved to position C, which is still
with the discussion of the exit pupil of an optical closer to the lens, the blur circles get even larger and
system. (In the meantime, you might contemplate how more information is lost. As the process continues, the
such a system would be set up.) blur circles eventually get large enough so that no
Since both beams and bundles can have converging information about the tree is present on the screen.
and diverging properties, what do we mean by con- Each of these steps is simulated by drawing the blur
verging light? The standard usage is that the term circles larger and larger as in Figure 2.19. A similar
converging light means the individual bundles are sequence would occur if the screen were moved from
converging. The beam may or may not be converging. the clear image position back away from the lens.
The light coming out of the movie projector is then
referred to as converging light because the bundles are
converging. The term diverging light means that the
individual bundles are diverging. 2.10 Pinholes or Image Formation by
Blur Minimization
If the screen is left at a blurred image position, the
2.9 Extended Images and Blur amount of blur can be reduced by placing an aperture
next to the lens and making the aperture smaller. This
Figure 2.19 represents light from a tree being con- reduces the cross-sectional size of each bundle, and,
verged by a lens and forming a clear inverted image of consequently, each blur circle is smaller. The degree

image
object

FIGURE 2.19. Effect of blur circle size on


an extended image.
24 Geometrie, Physical, and Visual Optics

of blur in the image is then reduced. This blur minimi- blur circles, and the illumination pattern on the screen
zation is sometimes called the pinhole effect. resembles an inverted blurred image of the arrow
Figure 2.20a represents light coming from two object.
points on an arrow object. Assume that the arrowhead Figure 2.20c shows the aperture narrowed down
is red and the bottom of the arrow is green. The light further. The effective blur circles decrease further in
is incident on a screen, but because of the divergence size, and a less blurred inverted image appears on the
of the incident bundles, no image is formed and no screen.
information about the arrow is present on the screen. If diffraction did not exist, one could make the
Figure 2.20b shows a circular aperture placed be- aperture so small that only one ray from each object
tween the object and the screen. The aperture limits point would pass through. The illumination distribu-
the size of each bundle, and the illumination pattern tion on the screen would then have one-to-one corres-
on the screen now has a definite red illuminated region pondence with the object. In other words, a clear
on the bottom and a definite green illuminated region image would be formed. This is called the pinhole
on the top. These illumination patches act just like effect because a small aperture, such as a pinhole, is

red

uniform
illumination
green

a)

red
green

green

b)

red green

green

red ^_ green point

green - red point

d) FIGURE 2.20. The pinhole effect.


The Geometric Behavior of Light 25

needed to form a clear image. In nature, pinhole and


images are sometimes accidentally formed by small
holes in curtains or leaves. tan D = f-y.
The pinhole does not provide a perfect image
because of diffraction. When the aperture is still large, Since C = D,
diffraction is negligible. As the aperture is made smal-
ler, the amount of diffraction steadily increases and (2.12)
the light waves start bending as they pass through the
aperture, just as sound waves bend upon passing (The absolute value signs are used in Eq. 2.121 so
^ that
tViOt
through a doorway. As the aperture size is decreased, we do not have to worry about plus and minus signs at
the diffraction eventually overwhelms the reduction in this time.)
blur gained by the pinhole effect. Even so, pinholes Equation 2.12 can be solved for any one of the four
are optically useful. entities contained in it provided the other three are
The image formed by a pinhole is due to blur known. For example, what image size is produced on
minimization and is not due to convergence of the a screen 10 cm behind a pinhole that is placed 20 m
light. Therefore, the position of the image screen from a 5-m tall elephant? From Eq. 2.12,
behind the pinhole is not crucial. As the screen is
moved a clear image remains; however, the image will
change in size.
IH = |o|xg
Figure 2.21 shows an object of size O and an image or
of size I. The horizontal line is drawn through the
center of the pinhole aperture normal to the aperture (5m)(0.1m)
11
plane. The object stretches from the horizontal line to 20 m
the point A. The aperture is a horizontal distance u 115m
from the object. An image screen is placed a horizon- 11
20 '
tal distance from the aperture.
The ray from the top of the object (point A) passes ., 50 cm _
straight through the pinhole and goes to the image
point B. The ray from the bottom of the object travels If the screen is moved further back from the pinhole,
straight through the pinhole along the horizontal line. the image remains clear but is larger. Objects that are
The angles C and D are formed by these two rays, and closer to the pinhole than the elephant would also be
angle C equals angle D since angles on the opposite imaged on the screen, but their size would be larger.
sides of intersecting lines are equal. From the triangles Historically, a device known as the camera ob-
shown, scura, which originally consisted of a hole in the wall
of a darkroom, was used to form images by the
pinhole effect. This principle was used by Aristotle,
tanC=^ Alhazen, Leonardo da Vinci, and Johannes Kepler,
\u\ and by the late 1600s, small hand-held camera ob-
scuras were popular in Europe.
The eye of the cuttlefish nautilus works on the
camera obscura principle. The nautilus eye has no
A lens. It has an open pinhole that fills with seawater on
immersion. A rattlesnake has a typical vertebrate eye.
However, a rattlesnake also has infrared pits that
work on the pinhole principle and supply enough
information to enable the rattlesnake to strike a local-
V
ized infrared target to within an angular accuracy of
\\ / ^ ^ . 1
5.
1
y " The pinhole effect can be used to clear a blurred
image behind a lens, again by minimizing the blur
B
circle sizes. Blurred vision can be the result of an
optical defect of the eye, but might also be the result
of an ocular or neural pathology. A pinhole can be
used to distinguish between these two cases. If one's
FIGURE 2.21. Pinhole imaging parameters. vision is blurred and a pinhole clears it, then the
26 Geometrie, Physical, and Visual Optics

problem is optical. If one's vision is blurred and a blur circles occur on the retina instead of point images
pinhole does not clear it, then the problem is (Figure 2.22b). In this case, the near object appears
pathological. Thus, the pinhole is an important tool blurred to the unaccommodated emmtrope. When
for optometrists and ophthalmologists. the ciliary muscle contracts, converging power is
added to the crystalline lens and the point image is
pulled forward onto the retina. Thus, the accommo-
dated emmetropic eye can clearly see the near object
2.11 Refracting States of the Eye (Figure 2.22c).
The emmtrope can clearly view a range of objects
In the emmetropic eye, the cornea and unaccommo- from far to near by changing the amount of accommo-
dated crystalline lens converge the light from a distant dation. The nearer the object, the more accommoda-
object and form a clear image on the retina (Figure tion is required. When the ciliary muscle contracts
2.22a). Remember that the ciliary muscle around the maximally, maximum accommodation occurs. As an
crystalline lens is relaxed when the eye is unaccommo- object is brought closer and closer to the eye, the
dated. The unaccommodated emmtrope can then see maximum amount of accommodation is eventually
distant objects clearly. reached. If the object continues to be brought closer
When the object is moved closer to the eye, the to the eye, it begins to blur. Just as a person cannot
divergence of the incident light increases. (For exam- sprint at full speed for very long, so too a person
ple, a point source 25 cm away from the eye produces cannot accommodate maximally for very long. If the
a wavefront that when incident on the eye has a object is held at a point requiring a considerable
vergence of -4.00 D; whereas, a distant point source amount of accommodation, fatigue eventually occurs,
produces a wavefront that when incident on the eye accommodation decreases, and the near object ap-
has a vergence of zero.) The unaccommodated em- pears blurred.
metropic eye will converge the light from the near The maximum amount of accommodation available
object, but the converging light runs into the retina declines as a natural part of aging. This is apparently
before the point images are formed. Consequently, due to the fact that the crystalline lens grows in layers
like an onion. The interior or core of the lens hardens
with age and slowly ceases to change its shape when
the ciliary muscle contracts. The symptom of the loss
of accommodative ability is that the aging emmtrope
has to hold things further away to see them clearly.
When an older emmtrope can no longer see near
a) objects clearly, the condition is called presbyopia or
"old age" vision.
The myopic eye is an eye that is too long for the
converging power of the eye (Figure 2.23a). The
unaccommodated myopic eye forms the image of a
distant object in front of the retina somewhere in the
vitreous humor. The light then diverges from this
point and blur circles are formed on the retina. If the
myopic eye accommodates, the point image is formed
even further in front of the retina and the retinal blur
b) gets worse.
When the distant object is moved closer to the
unaccommodated myopic eye, the divergence of the
incident light wave is increased and the image moves
back towards the retina. At some near point the
object will be close enough so that the image point is
formed on the retina. At this point the unaccommo-
dated myopic eye can clearly see the near object
(position A in Figure 2.23b). If the myope now ac-
commodates, objects closer than A can be seen clearly
(position B in Figure 2.23c). The condition of myopia
FIGURE 2.22. Emmtrope, a. Unaccommodated and viewing a is commonly called nearsightedness since an uncorrec-
distant object, b. Unaccommodated and viewing a near object, c. ted myope cannot clearly see distant objects but is
Accommodated for the near object. able to clearly see near objects.
The Geometric Behavior of Light 27

The hyperopic eye is an eye that is too short for the


1} blur
converging power of the eye (Figure 2.24a). The
unaccommodated hyperopic eye converges light from
a distant object, but the image points are not yet
a) formed when the light reaches the retina, so the
retinal image is blurred.
When the hyperopic eye accommodates, the light
inside the eye converges faster and the point images
are pulled forward onto the retina. In this case, the
accommodated hyperopic eye can clearly see a distant
/> object (Figure 2.24b). The hyperopic eye uses up
accommodation for a distant object and needs even
more for a near object. Depending on the amount of
b) hyperopia and the amount of accommodation, the
hyperopic eye may not have enough accommodation
left for a near object (Figure 2.24c). The result is that
the uncorrected hyperope sees distant objects clearly
(by accommodating), but near objects are blurred.
Thus, the condition of hyperopia is commonly called
A
farsightedness.
Unaccommodated hyperopic and myopic eyes
viewing a distant object both produce spherical con-
verging wavefronts inside the eye, and these spherical
c)
converging wavefronts can form point images. The
FIGURE 2.23. Myope, a. Unaccommodated and viewing a optical defect is that the image points do not lie on the
distant object, b. Unaccommodated and viewing a near object at retina. Myopes and hyperopes are collectively called
A. c. Accommodated and viewing a near object at B.
spherical ametropes. (Ametrope literally means not an
emmtrope.) Another optical defect of the eye occurs
when the converging wavefronts are not spherical,
and, consequently, a point image is not formed any-

rza blur where. The resulting condition is called astigmatism.


Stigmatus is the Latin word for point and astigmatism
literally means no point. Astigmatism is discussed in
Chapter 15.
a)
Occasionally, cateracts develop and grow in the
crystalline lens. The cataracts degrade vision by scat-
tering light, which causes the retinal image to become
washed out. One cure for cataracts is to surgically
remove the crystalline lens. The resulting eye has no
accommodative abilities but is correctable with specta-
ZT3 cle lenses, contact lenses, or intraocular lens implants.
An eye without a crystalline lens is called an aphakic
b)
eye {a = without', phakic = Latin for lens).

Problems

blur 1. What is the vergence of the light at a distance of


8 cm from a point source? At 42 cm? At 123 cm?
Draw a sketch of the wavefronts and indicate
whether the wavefronts are gaining or losing cur-
vature as they move away from the point source.
2. What is the vergence of converging light at 83 cm
FIGURE 2.24. Hyperope. a. Unaccommodated and viewing a
distant object, b. Accommodated and viewing a distant object, c. from the image point? At 32 cm? At 13 cm? At
Accommodated and viewing a near object. 4 mm? Draw a sketch of the wavefronts and indi-
28 Geometrie, Physical, and Visual Optics

cate whether the wavefronts are gaining or losing 9. Given a -8.00 D diverging wavefront in air, how
curvature as they move towards the image point. far (in cm) is the wave from its point source? For
3. The light at position A has a vergence of +4.50 D. a -6.00 D wavefront?
What is the vergence at position B, which is 14 cm 10. Given a +7.00 D converging wavefront, how far
downstream from position A? At a point C, which (in cm) is the image point? For a +3.00 D wave-
is 43 cm downstream from position A? front?
4. The light at position A has a vergence of 11. An object is 6 m from a pinhole. A screen is
-15.50 D. What is the vergence at position B, placed 15 cm behind the pinhole. When the object
which is 21 cm downstream from position A? is 2 m tall, what is the size of the image?
5. The light at position A has a vergence of 12. If the screen in no. 11 is moved closer to the
+ 13.50 D. What is the vergence at position B, pinhole, what happens to the image? If the screen
which is 3 cm upstream from position A? is moved further from the pinhole, what happens
6. The light at position A has a vergence of -6.30 D. to the image?
What is the vergence at position B, which is 9 cm 13. If the screen is fixed 15 cm from a pinhole, what
upstream from position A? At position C, which is happens to the image as the object is moved closer
37 cm upstream from A? to the pinhole?
7. A point source is under water (n = 1.33). What is 14. A piece of film 3 in x 5 in is placed 6 in from a
the vergence in the water at a distance of 11cm pinhole. What would be the linear field of view of
from the source? At 21 cm from the source? this pinhole in object space at a distance of 40 feet
8. A point source is in air. What is the vergence of from the pinhole?
the light 6 in from the point source? At 17 in?
CHAPTER THREE

Optical Objects
and Images

3.1 Optically Real vs Physically say that an optical system has an optically real object
Real Objects whenever the light incident on the system is diverging.
The optically real object may or may not be physically
Figure 3.1 shows light waves from a specific source, a real.
tree, incident on a specific optical system, a camera. The location of an optically real object point can be
The tree is a physically real entity. However, the specified in terms of the diverging spherical wavefront
optical stimulus for the camera is provided exclusively that is incident on the optical system (Figure 3.3). The
by the incident divergent light waves. These waves location assigned to the optically real object point for
leave the tree, travel across the space between the tree the system is at the center of curvature of the incident
and the camera, and are then incident on the camera. spherical wavefront. If a physically real object point
Suppose that light waves identical to those coming produced the incident diverging light wave, then the
from the tree can be artificially generated without the center of curvature of the wavefront is located at the
tree present. In that case, the optical stimulus for the physically real object point. However, if the diverging
camera would be unchanged and the camera would light wave was produced by a hologram, lens, or
photograph a "tree" even though no tree is physically mirror, then there is not a physically real object point
present. at the location assigned to the optically real object
A hologram can be used to artificially generate point.
such light waves. Holograms are usually made optical- The rays associated with the incident diverging
ly, but some can also be computer generated. The wavefront will, when extended back, intersect at the
hologram is a recording of a light wave interference center of curvature of the wavefront. Since the center
pattern. When properly illuminated, the interference of curvature is the assigned object point, the incident
pattern artificially generates the light waves identical rays appear to be diverging away from the assigned
to those coming from the tree (Figure 3.2). Not only object position regardless of whether a physically real
does the camera photograph the tree, but a human object is there or not. Figure 3.2b shows the location
observer looking at the waves coming from the holog- for a point on the tree produced by the hologram.
ram sees the tree complete with three-dimensional
depth and parallax effects.
In the aforementioned cases, the optical system
responds identically regardless of whether the incident 3.2 Optically Real vs Physically
diverging light waves are coming from the physically Real Images
real source, or were artificially generated by the holog-
ram. Since optical systems respond to the incident Figure 3.4a shows an optical system with diverging
light, it is useful to separate the concept of an optically light incident on it and converging light leaving it. The
real object from that of a physically real object. We relative distribution of visible electromagnetic radia-
29
30 Geometrie, Physical, and Visual Optics

-eMSSWWu
p observer In analyzing the action of the system, it then
becomes useful to say that the system has an optically
camera real image whenever converging light leaves the sys-
FIGURE 3.1. An observer photographing a tree.
tem. The system has a physically real image only if the
bundles in the converging light are actually allowed to
form point images. In the brick wall case, we would
tion in the object plane is recreated in the image say that the system has an optically real image, but not
plane. The actual electromagnetic radiation at the a physically real image.
image plane constitutes the physically real image. The position assigned to an optically real image
In Figure 3.4b, a brick wall has been placed be- point is at the center of curvature of the converging
tween the back of the system and the position of the spherical wavefront that is leaving the back of the
physically real image. The wall stops the electromag- system (Figure 3.5a). When the brick wall is not
netic radiation from reaching the plane of the phy- present, the bundle actually converges to form the
sically real image. However, the wavefronts leaving physically real image point at the center of curvature
the back of the optical system are still converging and of the exiting wavefront (Figures 3.5b and 3.5c).
still have the same curvature that they had prior to the When the brick wall is present, the light is converging
placement of the brick wall. Furthermore, the angles toward the center of curvature of the exiting wave-
between the different converging bundles leaving the front but is stopped before reaching it (Figure 3.5d).
back of the system are still the same. In effect, the In other words, the optically real image point is the
placement of the brick wall had no effect on the point toward which the light is converging.
converging light waves that are immediately behind The center of curvature of the wavefront that is
the system. leaving the system can be located by drawing in the

> observer
From
primary
light
source hologram camera

a)

apparent
"tree"

hologram
FIGURE 3.2. a. The hologram generates light waves
identical to those coming from the tree. b. The apparent
b) tree as seen by the camera and the observer.

incident spherical
incident spherical wavefront
wavefront

0-
center of
curvature,

L^

FIGURE 3.3. a. Incident spherical wavefront. b. Locating the


a) center of curvature of the incident wavefront.
Optical Objects and Images 31

FIGURE 3.4. a. Converging light forms a physically


real image, b. Converging light is still leaving the system,
but a brick wall stops formation of the physically real
image.

corresponding rays. These rays are of course converg- path of converging bundles and thus prevent the for-
ing and will pass through the center of curvature in mation of the physically real image. The concept of
those cases where the physically real image is formed. the optically real image is then useful in analyzing the
In the brick wall case, the rays are pointing toward the components that formed the converging bundles.
center of curvature when they are stopped by the wall,
but the rays can still be extended straight ahead to
locate the center of curvature. The ray extensions are
the dashed lines in Figure 3.5d. 3.3 Real Images as Objects for
Brick walls are not usually introduced as discussed the Eye
above. However, in a multiple-component optical sys-
tem, lenses or mirrors are frequently placed in the In our everyday lives, we are familiar with the real
images formed by movie and slide projectors. The
light leaving the projector system is converging and a
screen is placed at the position of the clear image. The
light then diffusely reflects from the screen, and the
now diverging bundles travel to the various observers'
eyes (Figure 3.6a). The (physically) real image on the
a) screen serves as the real object for the observers' eyes.
On the other hand, a real image can also be
directly viewed without the use of a screen. When no
screen is present, each bundle of the converging light
leaving the lens system forms a point image and then
diverges away from that point. To view the real image
b) directly, the observer needs to move his eye into the
path of the diverging light (Figure 3.6b). The real
image for the lens directly serves as the real object for
the observer's eye.
The real image for a single lens is inverted relative
to the original object and, depending on the object
distance, may be larger or smaller than the original
object. In Figure 3.6c, the original object is a tree and
the observer is looking at the small inverted real image
of the tree. Since the real image is inverted, it is
particularly easy for the observer to realize that he is
looking at the real image and not at the original
d) object.
The real image is, in effect, floating out in air with
FIGURE 3.5. A converging wavefront leaving the system, a.
Exiting wavefront and its center of curvature C. b. Wavefronts
no solid objects around it and, as such, is called an
converging to C. c. Wavefronts and rays. d. Brick wall does not aerial image. Consequently, the observer may have
change location of C but blocks light from reaching C. depth perception difficulties. The image may be per-
32 Geometrie, Physical, and Visual Optics

eye 5

eye'

aerial image

~^> eye

b)

distant tree

lens
inverted image of tree

FIGURE 3.6. a. A screen makes an image vis-


ible to different observers, b. An aerial image
viewed without a screen, c. Appearance of a real
aerial image to the observer.

ceived as being on the other side of the lens when in In our everyday lives, we are familiar with the
fact the image is between the lens and the observer. images formed by mirrors. However, your image
This misperception is due to the absence of the normal formed by a flat bathroom mirror is not a real image.
perceptual cues when viewing the aerial image. The light leaving the bathroom mirror is diverging, not
Remember that the reason the observer sees the converging. Nevertheless, the mirror's image is clearly
object, whether it is the real tree or the inverted real visible and has some of its own characteristics. For
image of the tree, is that the refracting elements of the example, when you wave your left hand at your image
observer's eye form a real image on the retina. In in the mirror, it waves its right hand back at you.
Figure 3.6c, the inverted real image that the lens Spherical diverging mirrors are now commonly
forms is serving as the real object for the eye and is used for security purposes as well as for giving an
thus the distal stimulus for the observer's visual per- increased field of view in automobile and truck mir-
ception. The proximal stimulus for the observer's vis- rors. Your image formed by a spherical diverging
ual perception is the retinal image formed by the mirror is smaller than you, but it still waves its right
refracting elements of the eye. hand when you wave your left hand. Again, the image
is not a real image because the light leaving the mirror
is diverging and not converging. While the images of
these mirrors are not real, they are certainly visible
3.4 Virtual Images and have the attributes of an image.
Images analogous to those of mirrors can also be
Virtual, adjective Being in essence or in effect formed by lenses (or even by holograms). In general,
though not formally recognized or admitted, for exam- whenever diverging light is leaving an optical system,
ple, He was a virtual saint. we say that the system has a virtual image. Virtual
Optical Objects and Images 33

, observer

center of curvature
of diverging
wavefront leaving
the mirror
point source 2 * FIGURE 3.7. Observer viewing a mirror's image.

images cannot be formed on a screen because converg- assigned to the virtual image point is the point from
ing light is needed to form images on screens. How- which the light appears to be diverging.
ever, virtual images can serve as optically real objects For a flat mirror, the position assigned to the
for our eyes or for other optical systems such as virtual image (i.e., at the center of curvature of the
cameras. In fact it makes no optical difference to an diverging wavefront that leaves the mirror) is where
eye or camera whether it is looking at a real or virtual the image perceptually appears to be. When you stand
image. In each case, the light incident on the eye or 1 m in front of your bathroom mirror, your image
camera is diverging. appears to be 1 m behind it. Note that no light is
The position assigned to a virtual image point is at physically present behind the mirror. You can build a
the center of curvature of the diverging spherical brick wall against the back of your bathroom mirror
wavefront that is leaving the back of the system and it won't affect your image in the mirror.
(Figure 3.7). The rays associated with the exiting Figure 3.8a shows a tree in front of a diverging
diverging wavefront will, when extended back, inter- lens. For each point on the tree, the lens creates a
sect at the center of curvature point that is the as- virtual image point, and its position is assigned to be
signed image position. The ray extensions are the at the center of curvature of the wavefront leaving the
dashed lines in Figure 3.7. In other words, the position back of the lens. An observer looking through the lens

T^> observer

a)

t> observer

b)

distant tree
lens

virtual image

FIGURE 3.8. a. Wavefronts for light passing through a diverg-


c) ing lens. b. Rays showing virtual image location, c. Appearance of
virtual image to the observer.
34 Geometrie, Physical, and Visual Optics

at the large tree sees the virtual image, which is the


small, erect tree.
The virtual image is specified by the diverging
wavefronts that leave the lens. These wavefronts then
travel to the next optical system, which can be an
observer's eye. The object for the observer's eye is FIGURE 3.9. Converging light incident on a system. The cen-
then an optically real object that is specified by the ter of curvature C of the incident wavefront serves as the location
for the virtual object.
diverging wavefronts incident on the eye. The phy-
sically real entities are the diverging light waves that
leave the lens and are incident on the eye. The virtual
image of the diverging lens is the optically real object fronts and in the angles that the individual bundles
for the observer's eye, and is the distal stimulus for the make with each other. Optically, objects for systems
observer's visual perception. The retinal image that are defined in terms of this incoming information.
the eye forms is physically real and is the proximal Consequently, when the incident light is converging,
stimulus for the observer's visual perception. Diverg- we say that the system has a virtual object. The
ing corrective lenses are worn by myopes. position assigned to a virtual object point is at the
center of the curvature of the incident converging
spherical wavefront.
For a convergent wavefront incident on the front
3.5 Virtual Objects surface of a system, the center of curvature is located
behind the front surface (Figure 3.9). The rays as-
When an optical system has diverging wavefronts inci- sociated with the incident converging wavefront are
dent on it, we say that the object is an optically real pointing toward the center of curvature when they hit
object and we assign its position at the center of the front of the system. The ray extensions (the
curvature point of the incident diverging wavefront. dashed lines in Figure 3.9) pass through the center of
As discussed in Section 3.1, there may or may not be a curvature point. Hence, the rays incident on the front
physically real object at that point. The geometric surface of the system are converging toward the virtual
information carried by the incident waves is encoded object position.
in the curvature of the wavefronts and in the angles Normally, the light leaving a source (either primary
that the different bundles make with each other. or secondary) is diverging. Therefore, to get converg-
In general, it makes no difference to the optical ing light incident on an optical system, that light must
system whether the incident light is diverging or con- leave the source and then be converged by some
verging. When diverging light is incident on a strong preceding optical system. The converging light leaving
enough converging system, the exiting light is converg- the first system is then incident on the second system.
ing, and when converging light is incident on the same The second system, with the incident converging light,
converging system, the system simply converges the has a virtual object. The wavefronts leaving the first
light even more. When diverging light is incident on a system and traveling to the second system specify both
diverging system, the system diverges the light even the real image for the first system and the virtual
more, and when converging light is incident on the object for the second system. In effect, the real image
same diverging system, the system still has a divergent for system one is the virtual object for system two
effect on the light. (The exiting light is either less (Figure 3.10).
convergent or may even be divergent depending on Hyperopes are corrected by converging lenses.
the diverging strength of the system.) When a corrected hyperope views a distant object, the
When the light incident on the system is converg- light leaving the correcting lens and incident on the
ing, the object for the system is clearly not real. hyperope's eye is converging. The optically real image
However, just as in the case of incident diverging of the correcting lens serves as the virtual object for
waves, the incident converging waves carry geometric the hyperope's eye. The optical position of this virtual
information encoded in the curvature of the wave- object is at the center of curvature of the wavefront

FIGURE 3.10. The converging light leaving system #1 is


incident on system # 2 . The wavefronts are all centered on C.
Optical Objects and Images 35

incident on the cornea. Since this wavefront is con- at the place from which the exiting light appears to be
verging, the assigned position is behind the eye. How- diverging.
ever, while the optics of the hyperope's eye are anal-
yzed with the virtual object position, remember that
the proximal stimulus for the hyperope's visual per-
ception is the physically real retinal image of the 3.7 The Image for System 1 Is the
hyperope's eye. Just as with an emmtrope or a Object for System 2
myope, the hyperope perceives the object to be in
front of the eye. Figure 3.11 shows four different two-lens systems. In
each case, a physically real object point is located at
position A in front of the first lens ( L J , and a
physically real image point is located at position C
3.6 Object and Image Summary behind the second lens (L 2 ). In each example, the
image for the first lens and the object for the second
The aforementioned definitions for optical objects lens are located at point B. Depending on the exam-
were in terms of the wavefront that is actually incident ple, the image for the first lens is either real or virtual,
on the system. When you encounter the word object, and the object for the second lens is either real or
you should immediately think in terms of what the virtual.
incident wavefront is doing. The aforementioned defi- In Figure 3.11a, the light leaving the first lens is
nitions for optical images were in terms of the wave- converging and a real image is actually formed at
front that is actually leaving the system. When you position B. The light waves that converge to form the
encounter the word image, you should immediately image point at position B then diverge away from
think in terms of what the exiting wavefront is doing. there and are incident on the second lens. Clearly, the
The discussion given in the previous sections of this real image for the first lens serves as the real object for
chapter is summarized below for quick reference. the second lens. Note that the wavefront leaving the
Whenever diverging light is incident on an optical first lens is converging and has its center of curvature
system, the system is said to have an optically real at B, while the wavefront incident on the second lens
object. Note, the object may not be physically real. is diverging and also has its center of curvature at B.
The position assigned to an optically real object point In Figure 3.11b, the first lens is a diverging lens.
is at the center of curvature of the incident diverging The light incident on it is diverging, and the light
wavefront. Consequently, the position assigned to the leaving it is diverging even more. The center of curva-
optically real object is at the place from which the ture of the wavefront leaving the first lens is at B. We
incident light appears to be diverging. say that the first lens has a virtual image, and the
Whenever converging light is incident on an optical location assigned to the image is at B. The diverging
system, the system is said to have a virtual object. This wavefronts that leave the first lens propagate across
can happen only when some previous system has the space between the two lenses and are incident on
converged the light. The position assigned to a virtual the second lens. The second lens has an optically real
object point is at the center of curvature of the object. The center of curvature of the diverging wave-
incident converging wavefront. Consequently, the po- front incident on the second lens is also at B, as are
sition assigned to the virtual object is at the place the centers of curvature of all the diverging wavefronts
toward which the incident light is converging. between the first and second lenses. The position
Whenever converging light leaves an optical sys- assigned to the object for the second lens is at B. In
tem, the system is said to have an optically real image. effect, the virtual image for the first lens serves as the
(A physically real image exists only when the bundles real object for the second lens.
are actually allowed to form their point images.) The In Figure 3.11c, the first lens is a converging lens.
position assigned to an optically real image point is at However, the light incident on the first lens has a high
the center of curvature of the exiting converging wave- enough divergence so that the light emerging from it is
front. Consequently, the position assigned to the opti- still divergent. The first lens did have a converging
cally real image is at the place toward which the effect since the divergence of the exiting light is less
exiting light is converging. than the divergence of the incident light. All of the
Whenever diverging light leaves an optical system, wavefronts traveling from the first lens to the second
the system is said to have a virtual image. The position lens are diverging and have their centers of curvature
assigned to a virtual image point is at the center of located at B. Since the wavefront emerging from the
curvature of the exiting diverging wavefront. Con- first lens is divergent and has its center of curvature
sequently, the position assigned to the virtual image is located at B, we say that the first lens has a virtual
36 Geometrie, Physical, and Visual Optics

^^B

FIGURE 3.11. Object and image positions for


four different two-lens systems. In each case A is
the object for L,, B is the image for L, and the
object for L 2 , C is the image for L 2 .

image located at B. Since the wavefront incident on incident on the second lens. The center of curvature of
the second lens is diverging and also has its center of the wavefront incident on the second lens is also at B.
curvature located at B, we say that the second lens has In effect, the optically real image of the first lens
an optically real object located at B. In effect, the serves as the virtual object for the second lens.
virtual image for the first lens serves as the real object In each of these four cases, the light leaving the
for the second lens. first lens propagates across the space between the two
In Figure 3.lid, the light leaving the first lens is lenses and is incident on the second lens. The light
converging. The converging light is then incident on that leaves the first lens carries information about the
the second lens. All of the converging wavefronts image of the first lens. The light incident on the
traveling from the first lens to the second lens have a second lens carries information about the object for
common center of curvature located at B. Since the the second lens, but the information carried by the
wavefront emerging from the first lens is converging, light leaving the first lens is the same information that
we say that the first lens has an optically real image, the light is carrying when it is incident on the second
and we assign the image position to be at B, which is lens. Since the information transferred is the same, we
the center of curvature of the wavefront leaving the say that the image of the first lens is the object for the
first lens. Note that in this case, no physically real second lens.
image point exists at B. The wavefront incident on the The preceding statement can be extended to other
second lens is converging. We say that the second lens optical systems including mirrors. Consider your bath-
has a virtual object, and we assign the object position room mirror. Use your face as the object for the
to be at the center of curvature of the wavefront mirror. The mirror forms a virtual image of your face.
Optical Objects and Images 37

The mirror's virtual image serves as the object for even try making up your own two- or three-lens
your eye. The object for system two (your eye) is the example and analyzing it.)
image for system one (the mirror).

3.9 Object Space/Image Space


3.8 Conjugate Points and the Principle
of Reversibility The word space has several different definitions and
uses. The word space in object space or image space is
The principle of reversibility states that rays represent- used in an abstract mathematical sense. A definition of
ing the path of light through a system would, upon the mathematical use of the word space is as follows:
reversal of their direction, retrace their original path an aggregate or set of points or things that have a
back through the system. Reversibility at a reflecting common property. Consider the original four cases in
or a refracting surface follows directly from the law of Figures 3.11 again. The light is traveling to the right
reflection (Eq. 1.1) or the law of refraction (Eq. 1.4). and in each case, the assigned object point for lens L2
Suppose in the situations shown in Figure 3.11 that was point B. In the first three cases, point B was the
the light is reversed and made to travel through the postion of an optically real object for L 2 and was in
two lenses from right to left. According to the princi- front of (to the left of) L 2 . In the last case, point B
ple of reversibility, only the direction indicators on the was the position of a virtual object for L2 and was
rays need to be changed. The ray positions and angles located behind (to the right of) L 2 . Remember that in
remain unchanged. all the cases, the object position was taken to be at the
In the reversed (i.e., right to left) situation, lens L 2 center of curvature of the wavefront that is physically
has a real object at C and either a real or a virtual incident on L 2 .
image at B. In the reversed situation for the first three The set of all points, rays, or ray extensions as-
figures (Figures 3.11a, b, and c), the image for L 2 is sociated with the light incident on an optical system
optically real, while in the fourth figure (Figure constitutes the object space for the system. For L 2 , all
3.lid), the image for L 2 is still at B but is now virtual. the points B are in this set, that is, are in the object
In each of the two-lens examples, the two points B space for L 2 . Note, the object space for L2 is a
and C are coupled together. In the left to right mathematical set that includes points both to the left
situation, B was the object point and C the image and to the right of L 2 . The requirement for belonging
point, while in the reversed situation, C was the object to the object space set is not where the point is, but
point and B the image point. Because of the coupling only that the point be associated with the incident
between B and C, they are labeled as conjugate points light.
for lens L 2 . For L 2 , the statement "C is conjugate to For each of the original (left to right) cases shown
B" means, that for light traveling to the right, B is the in Figure 3.11, the image for lens Lj is located at point
object and C is the image, while in the reversed case, B. In the first and last cases, the image for Lr is real
C is the object and B is the image. and point B is located to the right of Ll, while in the
For lens L1? the conjugate points in each case are second and third cases, the image for Lx is virtual and
A and B. For the left to right case, A is the object point B is located to the left of Lj. Remember that the
point and B is the image point, while for the reversed image position was defined as the center of curvature
case, B is the object point and A is the image point. of the wavefront that is physically leaving Lj.
In the left to right situation for the two-lens system The set of points, rays, or ray extensions associated
(both Lt and L 2 ), the original object in each case was with the light leaving an optical system constitutes the
at A and the final image was at C. In the reversed image space for the system. In the original cases of
situation, the system object was at C and the system Figure 3.11, the points B are all in the image space for
image was at A. Thus, the conjugate points for the Lj. Note that the image space set includes points that
two-lens system are A and C. are both left and right of the lens. The requirement for
The ability to determine and use conjugate points is belonging to the image space set is not where the point
an important skill in analyzing optical systems. As is, but only that the point be associated with the light
illustrated by the above discussion, it is crucial to leaving the system.
recognize which lens or lenses are involved. In particu- It may be helpful in initially learning the object
lar, A is conjugate to B for the first lens. However, for space/image space concepts to think about imaging
both lenses together, A is conjugate to C. (You should with a tinted lens. (Sunglasses are common examples
take some time to think about this. In fact you might of tinted lenses.) Figure 3.12 shows four examples of
38 Geometrie,. Physical, and Visual Optics

a)

b)

B *2A red

FIGURE 3.12. The lens in each case is a red tinted lens.


d) The light leaving the lens is red, and the image B is red.

imaging with a tinted lens. In each example, the light In Figure 3.12c and d, a previous lens, which is not
is traveling left to right. shown, has been used to generate white converging
In Figure 3.12a, a white point source is placed in light incident on the red tinted diverging lens. In
front of a red tinted converging lens. The incident Figure 3.12c, the object for the diverging lens is virtual
rays, drawn in black, represent the incident white and is located at the center of curvature of the incident
light. The exiting rays, drawn in red, represent the wavefront. The center of curvature is found by extend-
transmitted red light. Point A is in object space and ing the incident black rays forward until they intersect
point B is in image space. at point A. Point A is shown in black since it is
In Figure 3.12b, the white point source is placed in associated with the incident white light and not with
front of a red tinted diverging lens. The incident rays the exiting red light. The light leaving the lens is still
are black representing the incident diverging light. converging, but not as convergent as the incident light.
The outgoing rays are red and represent the exiting A real image is formed at point B. In this case, point
light, which is more divergent than the incident light. A is in the object space set (or simply, in object
The image is virtual and the associated position is at space), and point B is in the image space set (or
the center of curvature of the exiting wavefront. This simply, in image space), even though they are both to
center of curvature is located by extending the outgo- the right of the lens.
ing red rays backward until they intersect at point B. In Figure 3.12d, the incident white light is converg-
Clearly, point A is in object space. Point B is as- ing, but the degree of convergence is not as high as in
sociated with the red light leaving the lens. Therefore, part c. Consequently, the diverging lens has enough
it is in image space even though it is to the left of the power to make the outgoing red light divergent. Both
lens. the image and the object points are virtual. The object
Optical Objects and Images 39

point is at A, represented in black, and the image lens 1 have a real or virtual image? Is lens 1 a
point is at B, represented in red. Point A is in object converging or a diverging lens?
space even though it is to the right of the lens, and 4. Consider lens 3 in Figure 3.13. Which point (A,
point B is in image space even though it is to the left B, C, or D) is the object position for lens 3? Does
of the lens. lens 3 have a real or a virtual object? Is lens 3
In conclusion, when you see the term object space, converging or diverging? What point serves as the
you should think in terms of the convergence or center of curvature for the wavefront incident on
divergence of the light incident on the system even if lens 3?
the object is virtual and located behind the system.
When you see the term image space, you should think 5. Consider Figure 3.14. Point B is in the object
in terms of the convergence or divergence of the light space for which lens ( 1 , 2 , 3, 4, or 5)? What lens
leaving the system even if the image is virtual and is point B in the image space for?
located in front of the system. Once learned, the 6. Consider Figure 3.14. Point D is in the object
object space/image space concepts serve as a useful space for which lens (1, 2, 3, 4, or 5)? What lens
tool in dealing with multiple-component systems. is point D in the image space for?
7. Consider Figure 3.14. What lens (1, 2, 3, 4, or 5)
is point E in the object space for? What lens is
point E in the image space for?
Problems 8. Consider Figure 3.14. List the converging lenses.
List the diverging lenses.
1. A flat bathroom mirror reflects light without 9. Consider Figure 3.14. For lens 2, what point is
changing the vergence. Suppose you stand in front conjugate to point C? Now consider lens 2 and 3
of the bathroom mirror and look at your image in as one system. For lens 2 and 3 as one system,
the mirror. Is the image formed by the mirror real what point is conjugate to C?
or virtual? The mirror's image serves as the object 10. Consider Figure 3.14. There is light blue water
for your eye. Does your eye have a physically real between lens 1 and 2. There is light green jello
object, an optically real object, or a virtual between lens 2 and 3. There is light red water
object? When you hold your right hand up, does between lens 3 and 4. Draw the wavefronts that
your image in the mirror hold its right or its left have point E as their center of curvature. We want
hand up? to represent point E by the color of the medium
2. Consider lens 2 in Figure 3.13. Which point (A, that these wavefronts are in. What color should
B, C, or D) is the object position for lens 2? Does we use for E?
lens 2 have a real or virtual object? Which point 11. A myope wears a diverging spectacle correction.
serves as the image for lens 2? Is the image for When the myope looks through the spectacle lens
lens 2 real or virtual? Is lens 2 a converging or a at a distant tree, is the optical object for his eye
diverging lens? What point is the center of curva- the tree, or the image of the tree formed by the
ture of the wavefront incident on lens 2? spectacle lens? Is the image of the spectacle lens
3. Consider lens 1 in Figure 3.13. Which point (A, real or virtual? Is the object for the myope's
B, C, or D) is the image position for lens 1? Does cornea real or virtual?

1 2> 2I

<^^ B<C *Z

FIGURE 3.13. Three-lens system refer-


red to in problem nos. 2 through 4.
40 Geometric, Physical, and Visual Optics

FIGURE 3.14. Five-lens system referred


to in problem nos. 5 through 10.

12. A hyperope wears converging spectacle lenses. cornea? The spectacle corrected hyperope has a
When the hyperope looks through the spectacle real image on the retina. For the spectacle lens-
lens at a distant tree, converging light leaves the eye system, what is the retinal image conjugate
spectacle lens and is incident on the hyperope's to?
cornea. Does the hyperope's cornea have a real or 13. Which is the better statement: "Seeing is believ-
virtual object? Where is the center of curvature of ing" or "Sight is an illusion"? Why?
the wavefront that is incident on the hyperope's
CHAPTER FOUR

Thin Lenses and


Ray Diagrams

4.1 How Spherical Lenses Work spherical glass surface in which the air bulges into the
center of the glass surface. The edges of the plane
waves enter the glass surface first and are consequent-
The long glass rod shown in Figure 4.1a has a spheri- ly slowed down before the center section of the plane
cal surface on the front. Plane waves from a distant wave. Thus, when the entire wavefront has entered
point source are incident from the left. In Figure 4.1b, the glass, the edges are lagging behind the center.
the first wavefront is entering the glass. The speed of Within the accuracy of the paraxial approximation, the
the wavefront in the glass is slower than its speed in wavefronts in the glass are spherical diverging wave-
the air. Consequently, the center of the wavefront, fronts.
which enters the glass first, slows down first and starts Figure 4.4 shows plane waves in the glass incident
lagging behind the edges. The result is that the wave- from the opposite direction on the same spherical
fronts in the glass become converging wavefronts. surface shown in Figure 4.3. Here, the middle of the
Within the accuracy of the paraxial approximation, the wavefront enters the air first and thus speeds up first.
converging wavefronts are spherical and form a point Hence, the middle of the emerging wavefront leads
image at position E in Figure 4.1c. the edges, so that the wavefront that emerges into the
Figure 4.2 shows plane waves in the glass incident air is again diverging.
from the opposite direction on the same spherical glass The glass surface shown in Figures 4.3 and 4.4 has
surface. As the first wavefront approaches the surface, the physical property such that, at its center, the air
the edges of the wavefront enter the air (before the bulges into the glass. Note that this surface diverges
center of the wavefront) and speed up. Consequently, plane waves traveling in either direction, i.e., plane
the edges of the wavefront start leading the center of waves initially in the air are diverged when passing
the wavefront that is still in the glass. Within the into the glass (Figure 4.3), and plane waves initially in
accuracy of the paraxial approximation, the exiting the glass are diverged when passing into the air (Fig-
wavefronts in air are converging and spherical and ure 4.4).
form a point image at G. Figure 4.5a shows a converging glass lens made
The spherical glass surface in Figures 4.1 and 4.2 with two spherical surfaces. In the center, each glass
has the physical property of its center bulging out into surface bulges out into the air. Each of the surfaces
the air. This surface converges plane waves no matter contributes to the converging action of the lens. A
which way the waves travel through the surface, i.e., converging lens is thicker in the middle than it is at the
plane waves initially in the air are converged when edges.
passing into the glass (Figure 4.1), and plane waves Figure 4.5b shows a diverging glass lens made with
initially in the glass are converged when passing into two spherical surfaces. In the center, each glass sur-
the air (Figure 4.2). face is indented so that the air bulges into the glass.
Figure 4.3 shows plane waves in air incident on a Each glass surface contributes to the diverging action
41
42 Geometrie, Physical, and Visual Optics

>
plane waves
glass

a)

glass

b)

FIGURE 4.1. a. Plane waves in air incident on a


glass rod with a convex spherical surface, b. In the
glass, the middle of the wavefront lags behind the
c) edges, c. The wavefronts in the glass converge to E.

of the lens. A diverging lens is thinner in the middle


than it is at the edges.
Figure 4.6 shows two more glass lenses made with
spherical surfaces. In each case, the left surface of the
lens is convergent and the right surface of the lens is
divergent. The net converging or diverging effect de-
pends on the converging or diverging power of the
glass
individual surfaces. In Figure 4.6a, the converging
surface (left) has more power than the diverging
b) surface (right). Consequently, the net effect of the
FIGURE 4.2. a. Plane waves in the glass incident on the convex lens is to converge light. The tip-off that the lens in
surface of the previous figure, b. In the air, the wavefront edges Figure 4.6a is a converging lens is that it is thicker in
lead the middle resulting in light converging to G. the middle than it is at the edges.
Thin Lenses and Ray Diagrams 43

a)

glass

b) M

FIGURE 4.3. a. Plane waves in air incident on a con-


cave air-glass interface, b. In the glass, the edges of the
wave lag behind the middle, c. The resulting diverging
wavefronts in the glass.

<e

a) a) b)
FIGURE 4.5. a, A converging lens with two converging sur-
^ faces. b. A diverging lens with two diverging surfaces.

b)
FIGURE 4.4. a. Plane waves in the glass incident on the
concave interface of the previous figure, b. In the air, the middle
of the wavefronts leads the edges resulting in diverging waves.

In Figure 4.6b, the diverging surface (right) has


more power than the converging surface (left). The a) b)
net effect of the lens is a diverging effect. The tip-off FIGURE 4.6. a. A converging lens with one converging and
that the lens in Fig. 4.6b is a diverging lens is that it is one diverging surface, b. A diverging lens with one converging
thinner in the middle than it is at the edges. and one diverging surface.
44 Geometrie, Physical, and Visual Optics

4.2 Optical Axis of incidence Q{ is zero. Therefore, from Eq. 4.1, the
angle of refraction must also be zero. Therefore, a
Figure 4.7 shows a diverging lens. Each surface of the ray incident normal to a surface passes straight
lens is spherical. The center of curvature of the left through the surface without bending.
spherical surface is marked C1? and the center of Consider an incident ray lying on the optical axis of
curvature of the right spherical surface is marked C 2 . a spherical lens. The optical axis is normal to both of
The dashed lines shown in Figure 4.7a complete the the surfaces, so the angles of incidence and refraction
spheres for which C1 and C2 are the respective centers of the ray are equal to zero for both surfaces. Thus, a
of curvature. ray coming in along the optical axis does not bend
Figure 4.7b shows a straight line drawn through the while passing through the lens. The ray leaves the lens
centers of curvature of both spherical surfaces (i.e., still traveling along the optical axis.
through Cj and C 2 ). This line is a symmetry axis for In order for two points to be conjugate points, any
the lens. A clockwise or counterclockwise rotation of ray associated with the object point must, after passing
the lens around this axis produces no change in the through the lens, also be associated with the image
lens image. This line, through the centers of curvature point. Since a ray incident along the optical axis does
of the two spherical surfaces, is called the optical axis not bend, any object point (real or virtual) lying on
of the lens. Figure 4.8 shows the respective centers of the optical axis has a conjugate image point that also
curvature, C1 and C 2 , for the left and right spherical lies on the optical axis.
surfaces of several other lenses together with the
optical axis.
Whenever a line passes through the center of cur-
vature of a sphere, it is normal to the sphere's surface. 4.3 Thin Lenses
Since the optical axis passes through the center of
curvature of each spherical surface of a lens, it is The central thickness of the lenses discussed in the
normal to each surface. For the diverging lenses previous sections is the thickness of the lenses along
shown in Figures 4.7 and 4.8, the optical axis passes the optical axis. When the central thickness of a lens is
through the thinnest portion of the lens, while for the small enough, the converging and diverging properties
converging lenses, the optical axis passes through the of the lens in air are independent of the shape or form
thickest portion of the lens. of the lens as well as the direction in which the light is
The small angle approximation to Snell's law re- traveling through the lens. In this case, the lens is
lated the incident angle of a light ray to the angle of called a thin lens. All converging thin lenses are
refraction of the light ray by the equation, represented by the symbol shown in Figure 4.9. Figure
4.10 shows the symbol for all diverging thin lenses.
n,6i = n 2 9 r . (4.1)
The optical axis of a thin lens is drawn normal to
The incident angle Q{ and the angle of refraction the lens as shown in Figure 4.11. The point on the lens
were defined relative to the normal for a surface. through which the optical axis passes is called the
When an incident ray is normal to a surface, the angle optical center of the lens.

/ N "^
optical
FIGURE 4.7. C, and C 2 are the respective
axis centers of curvature for the spherical surfaces of
the lens. The optical axis is the line that passes
through C, and C 2 .
a) b)

5- <"
>

FIGURE 4.8. C, and C 2 are the respective


C2 Ci centers of curvature for the spherical sur-
faces of each lens, and the connecting line is
the optical axis.
Thin Lenses and Ray Diagrams 45

A
FIGURE 4.9. Thin converging lenses and the shape-independent
representation.

V
<=

FIGURE 4.10. Thin diverging lenses and the shape-indepen-


A dent representation.

V
4^

v v\
optical optical

optical optical
center center FIGURE 4.11. The optical center and optical axis for thin
V lenses.

4.4 Secondary Focal Point (F2) lens. The basic definition of the secondary focal point
is still the same, i.e., the secondary focal point is the
The secondary focal point (F 2 ) for a thin lens is axial image point conjugate to an axial object point at
defined as the on-axis image point that results when a optical infinity. Again, the waves incident on the lens
bundle of plane waves is incident normally on the lens. from the distant axial object point are effectively flat
An alternative definition is that the secondary focal or plane waves. The lens diverges the waves and the
point F 2 for a thin lens is the on-axis image point axial image point, F 2 , is virtual and located at the
conjugate to an on-axis object point at optical infinity. center of curvature of the exiting diverging wavefront.
The two definitions are equivalent since the on-axis This position is found in Figure 4.13b and c by extend-
object point at optical infinity results in incident plane ing the outgoing rays back until they intersect. (The
waves. Since an axial object point is always conjugate extensions are dashed.)
to an axial image point, F 2 is always located on the For incident light traveling to the right, F 2 lies to
optical axis. the right of a converging lens (Figure 4.12) and to the
Figure 4.12 shows the situation for a thin converg- left of a diverging lens (Figure 4.13). However, for the
ing lens. The wavefronts from the distant axial point diverging lens case, keep in mind that F 2 is a point in
source are effectively flat, or plane waves, when they image space. F 2 is associated with the light leaving the
reach the lens. The lens converges the light and a real lens (i.e., the light physically to the right of the lens)
image point is formed at F 2 on the axis behind the and not with the plane waves incident on the lens.
lens. The location of F 2 relative to the lens is a measure
Figure 4.13 shows the situation for a thin diverging of the converging or diverging power of the lens. A

^
> ^
HHri
>I [ TH FIGURE 4.12. a. Plane waves incident on a
converging lens resulting in waves that con-
**J^ verge to the secondary focal point, b. Wave-
b) c) fronts and rays. c. Rays only.
46 Geometrie, Physical, and Visual Optics

->-

IV\M\
^
^F^
TT Wl 1
II 1 1 1 3fe
> ^CH
FIGURE 4.13. a. Plane waves incident on a
J/r*^L/
diverging lens resulting in diverging waves

b)
leaving, b. Wavefronts and rays showing the
virtual secondary focal point, c. Rays only.

strong lens has F 2 located close to the lens, while a lens. In order to get plane waves to leave the diverging
weak lens has F 2 located far away from the lens. lens, the incident light must be converging. The object
point is located at the center of curvature of the
incident wavefront and is virtual. The virtual object
point can be located by extending the incident rays
forward until they intersect. Again, the rays represent-
4.5 Primary Focal Point (F,) ing the light leaving the lens are all parallel to each
other and to the optical axis since one of the rays is
The primary focal point (F x ) is defined as the on-axis traveling along the axis.
object point that results in plane waves leaving the For incident light traveling to the right, Fj lies to
lens. An alternative definition is that the primary focal the left of a converging lens (Figure 4.14) and to the
point Fj is the on-axis object point that is conjugate to right of a diverging lens (Figure 4.15). However, for
an on-axis image point at optical infinity. the diverging lens case, keep in mind that Fx is a point
Figure 4.14 shows the situation for a thin converg- in object space. is associated with the light incident
ing lens. The wavefronts incident on the lens are on the lens, i.e., the light physically to the left of the
diverging, while the wavefronts exiting the lens are lens, and not with the plane waves that are exiting the
plane waves. The rays leaving the lens are effectively lens.
parallel to each other and to the optical axis since one The location of Fx relative to the lens is a measure
of the rays is traveling along the axis. Clearly, for a of the converging or diverging power of the lens. A
thin converging lens, Fj is a real object point. strong lens has Fj located close to the lens, while a
Figure 4.15 shows the situation for a thin diverging weak lens has Fx located far away from the lens.

> 1
^
^~r
^ ^ FIGURE 4.14. a. Wavefronts diverging
* < from the primary focal point result in plane
V 'r waves leaving, b. Wavefronts and rays. c.
b) Rays only.

FIGURE 4.15. a. Converging wave-


fronts incident on a diverging lens resulting
in plane waves leaving. The center of cur-
vature of the incident wavefront is the
primary focal point, b. Wavefronts and
rays. c. Rays only.
Thin Lenses and Ray Diagrams 47

4.6 Equidistance of Focal Points for argument also shows that the primary and secondary
a Thin Lens in Air focal points simply change places when light travels in
the opposite direction through a lens.
When the central thickness of a lens in air is small
enough, then the imaging properties of the lens are
independent of the lens shape and of the direction the
light travels through the lens. This was the thin lens 4.7 Predictable Rays/Converging
assumption. Let us apply this assumption to a converg- Lenses/Off-Axis Object Points
ing lens.
In Figure 4.16a, incident plane waves are traveling Figure 4.17 shows the rays used in defining the pri-
to the right and the secondary focal point, F 2 , is mary and secondary focal points for a thin converging
located 10 cm to the right of the lens. In Figure 4.16b, lens in air. The secondary focal point, F 2 , is an on-axis
plane waves incident on the same lens are traveling to image point conjugate to optical infinity, while the
the left. In this case, since the imaging properties of primary focal point, F l 5 is an on-axis object point
the thin lens in air are independent of the direction in conjugate to optical infinity.
which the light is traveling, F 2 is now located 10 cm to Figure 4.18 shows an extended object located a
the left of the lens. finite distance in front of a thin converging lens. Our
According to the principle of reversibility, rays initial interest is in the off-axis point at the top of the
travel the same path backward through a system; thus, object. The figure shows some of the incident diverg-
only the direction indicators on the rays need to be ing rays from the top object point.
changed. Figure 4.16c shows the principle of re- We cannot predict how most of the rays in Figure
versibility applied to Figure 4.16b. However, in the 4.18 will leave the lens. However, we can use Figure
reversed situation (Figure 4.16c), the light is diverging 4.17 to predict how two of the rays will leave the lens.
away from the point 10 cm to the left of the lens and In Figure 4.17a, the incident rays are all parallel to the
plane waves are leaving the lens. Consequently, the axis and the exiting rays all point toward F 2 . Con-
point labeled F 2 in Figure 4.16b becomes the primary sequently, any incident ray that is parallel to the axis
focal point, Fl9 for the same lens in Figure 4.16c. will point toward F 2 when it leaves the lens. One of
Since the imaging properties of the thin lens in air the incident rays in Figure 4.18 is parallel to the axis;
are independent of the direction in which the light is therefore, it must point toward F 2 when it leaves the
traveling, then by comparing Figures 4.16a and 4.16c, lens. Figure 4.19a shows this ray by itself, and Figure
one can see that the primary and secondary focal 4.19b shows this ray together with the other incident
points are both 10 cm from the lens. This argument rays.
can be repeated for any focal length. Therefore, the In Figure 4.17b, the incident rays all originate at F1
primary and secondary focal points for any thin lens in and the exiting rays are all parallel to the axis. Based
air are always equidistant from the lens. Note, the on Figure 4.17b, we can predict that any incident ray


FIGURE 4.16. a. Plane waves incident
* from the left give a secondary focal point
10 cm right of the lens. b. Plane waves inci-
"2 ^ ^ ^ " \ ' dent from the right give a secondary focal
" point 10 cm to the left of the lens. c. The
10cm % f
principle of reversibility shows that the pri-
10cm mary focal point is also 10 cm from the lens.
b)

>
i1
>
^ s .

F
^5r~Fz
F
<
,>
^
>- ^
s *
FIGURE 4.17. Diagrams used to predict the respective
behavior of rays associated with the focal points of a
a) { 1 b) converging lens.
48 Geometrie, Physical, and Visual Optics

Figure 4.21a shows an isolation of the two rays and


Figure 4.21b shows the two rays along with the other
incident rays.
The two exiting rays indicate that the light leaving
the lens is converging. The conjugate image point is
real and is located at the place where the two rays
intersect. Now that the conjugate off-axis image point
is located, all the other incident rays can be traced
FIGURE 4.18. Diverging rays incident on a lens. through the lens. These rays must also pass through
the image point (Figure 4.22a). (Figure 4.22b also
shows the incident diverging wavefronts and the exit-
ing converging wavefronts.)
that, as far as the lens is concerned, appears to be The object under discussion lies in a plane that is
coming from Fl will leave the lens parallel to the axis. perpendicular to the optical axis of the lens. Under
For the off-axis object point in Figure 4.18, one of the paraxial conditions, the image also lies in a plane
incident rays leaves the object point and passes perpendicular to the axis. Since an axial object point
through before hitting the lens. Hence, this ray will must be conjugate to an axial image point, we can
leave the lens parallel to the axis. Figure 4.20a isolates locate the axial image point by simply drawing a
this ray and Figure 4.20b shows this ray together with perpendicular line from the off-axis image point to the
the other incident rays. axis (Figures 4.21 and 4.22).
Taken together, the two predictable rays shown in Consider a second ray trace example for the same
Figures 4.19 and 4.20 are sufficient to indicate whether thin converging lens. The object is now placed be-
the light leaving the lens is converging or diverging. tween Fj and the lens. Figure 4.23 shows some inci-

\
1

/ K ^

Fi FN

a)
\f FIGURE 4.19. a. Isolation of the predictable ray
incident parallel to the axis. b. Path of predictable ray.

i 1

F > ^ > F2
FIGURE 4.20. a. Isolation of the pre-
b) dictable ray through the primary focal
<1 point, b. Path of the predictable ray.

Fr FIGURE 4.21. a. Isolation of the two


predictable rays and resultant image loca-
tion. b. Predictable rays among the others.
Thin Lenses and Ray Diagrams 49

FIGURE 4.22. a. Once the image point


is known, all exiting rays must go there, b,
Wavefronts and rays.

the lens and, as far as the lens is concerned, appears to


i \
be coming from Fj. We can use Figure 4.17b as a
guide to see that this ray must exit the lens parallel to

^ the axis. Figure 4.24b isolates this ray.


Fi F2
Figure 4.25a shows both of the predictable rays.
The exiting rays indicate that the outgoing light is
diverging. Consequently, the image is virtual and can
t
be located by extending the rays backward until they
FIGURE 4.23. Incident rays for object inside the primary focal intersect (Figure 4.25b). The virtual image is erect
point. relative to the object and larger than the object. (This
is an example of a simple magnifying lens.)
All of the outgoing rays must appear to be coming
dent diverging rays. Again, most of these diverging from the same virtual image point. So, the remaining
rays are not predictable. However, one of the incident rays can now be traced as shown in Figure 4.26a.
rays is parallel to the axis; therefore, we can use Figure 4.26b also shows the incident diverging wave-
Figure 4.17a as a guide to see that this ray must exit fronts and the exiting diverging wavefronts. In Figure
the lens pointing toward F 2 . Figure 4.24a isolates this 4.26b, the exiting wavefronts have less divergence
ray. Another of the incident rays in Figure 4.23 hits than the incident wavefronts.

b)
4= FIGURE 4.24. Predictable rays associated
with the focal points.

^ 1 -*--
F, F
FIGURE 4.25. Virtual image location from the
predictable rays. a. Actual rays. b. Backward ex-
a) tensions to locate the image.
50 Geometrie, Physical, and Visual Optics

FIGURE 4.27. Ray diagram showing some rays bend down and
other bend up, while the nodal ray passes through the optical
center.

Inspection of Figure 4.27 shows that the unbent ray


passes through the lens at the same point that the
FIGURE 4.26. a. Ray behavior for virtual image, b. Rays and optical axis passes through the lens. This point was
wavefronts.
previously named the optical center of the lens.
The latter property is a general property for thin
lenses and gives us a third predictable ray to use.
Namely, any ray passing through the optical center of
Instructional Comment a thin lens passes straight through without bending.
The ray that does not bend is called the nodal ray, and
Ray diagrams of the type discussed can greatly aid an the optical center of the thin lens is the nodal point for
individual in developing optics intuition. These diagrams
the thin lens.
can be drawn to scale on graph paper, and that is
sometimes necessary. However, in developing optics in-
tuition, I believe that the main utility of ray diagrams is
gained from being able to do a quick and dirty freehand
sketch that provides the needed information. Therefore, I
urge beginners to practice such sketches. 4.9 Further Examples: Converging Lenses
A common pitfall of some beginners is to try to memorize
all the possible types of ray diagrams. Let that be In order to let the ray diagram convey the information
anathema. I believe that one should learn to quickly draw intended, the following convention is used. A ray that
the predictable rays and then trust the ray diagram. In is actually representing light physically present at that
this way, much less memory is required and the skill point in space is drawn as a solid line. Extensions of
gained will tend to remain long after one's memorized the ray either forwards or backwards into regions of
results have been forgotten. space where the light associated with the ray is not
If I give you some fish, you eat for a day.
physically present is shown as a dashed line.
If I teach you to fish, you eat for a lifetime. Consider the dashed line in Figure 4.24b. There,
Old Chinese Proverb the ray of interest left the object and was incident on
the lens as if it had come from the primary focal point,
Fj. The ray was drawn in solid from the object to the
lens, but the extension of the ray from the object to Fl
4.8 Nodal Rays was drawn in dashed.
In Figure 4.25a, the two exiting rays, drawn solid,
Figure 4.27 is a repeat of Figure 4.22a in which the are diverging. Figure 4.25b shows the backward exten-
object was located outside the primary focal point, F ^ sions of the exiting rays in order to locate the virtual
and a conjugate real image was formed. Part of the image. These extensions are dashed. The extensions
incident rays are bent down upon passing through the are associated with the exiting light and not with the
lens and part of the incident rays are bent up upon light that is physically incident on the lens.
passing through the lens. Between the bent up rays Figure 4.28a shows a real object placed outside Fx
and the bent down rays, there must be a ray that of the lens. The incident ray parallel to the axis is
passes straight through the lens without being bent. shown in Figure 4.28b, and it exits the lens pointing
Thin Lenses and Ray Diagrams 51

J_ '

F, Fi I F i \
a) b

K -^
F2 F,
c) *
^ <>

f^rN^ FIGURE 4.28. . Object outside the primary focal


point, b. Predictable ray incident parallel to the axis.
c. Predictable ray through the primary focal point, d.
Predictable ray through the optical center, e. Image
e) location by the three predictable rays.

toward point F 2 . The ray incident on the lens that and the location assigned to the virtual object is at the
passes through Fj exits parallel to the axis as shown in center of curvature of the wavefront incident on the
Figure 4.28c. The nodal ray is shown in Figure 4.28d. lens. The location of the virtual object is found by
Finally, all the rays are shown in Figure 4.28e. extending the converging rays forward until they inter-
Figure 4.29a shows converging light incident on the sect. These extensions are shown as the dashed lines.
lens. The light has been converged by a preceding Out of all the incident converging rays, three are
optical system, which is not shown. Since the incident predictable as discussed previously. Figure 4.29b iso-
light is converging, the object for the lens is virtual lates the nodal ray. Since the nodal ray passes straight

>*

^1
I I "-ta
r^ J

a) y\ >

FIGURE 4.29. a. Converging wavefronts and


rays incident on a converging lens. b. The nodal
ray. c. The ray incident parallel to the axis. d.
The ray incident through the primary focal point.
e. Image location by the three predictable rays. / .
Other rays..
52 Geometrie, Physical, and Visual Optics

through the lens without bending, it is drawn entirely the incident light must be converging. Thus, F t must
as a solid line. be a virtual object point. Figure 4.30b shows the rays
Figure 4.29c shows the incident ray that is parallel associated with Fj for a thin diverging lens. Based on
to the axis. This ray exits the lens pointing toward F 2 . Figure 4.30b, any incident ray pointing toward Fj of a
(Recall Figure 4.17a for the general behavior of rays thin diverging lens leaves the lens parallel to the axis.
incident parallel to the axis.) Remember that Fj and F 2 are equidistant from the
Figure 4.29d shows the incident ray that passes lens.
through Fx and exits parallel to the axis. (Recall Figure 4.31a shows the rays from a near extended
Figure 4.17b for the general behavior of rays incident object located in front of a thin diverging lens. Three
through Fj.) of these rays are predictable. Figure 4.31b isolates the
Figure 4.29e shows all three of the incident predict- incident ray parallel to the axis. In accordance with
able rays. The exiting rays, drawn solid and to the Figure 4.30a, this ray appears to be coming from F 2
right of the lens, show that the light leaving the lens is when it leaves the lens. Figure 4.31c isolates the
converging and forms a real image, which is erect and incident ray that is pointing toward F ^ In accordance
smaller relative to the virtual object. with Figure 4.30b, this ray leaves the lens parallel to
Note that in this case, the lens took the incident the axis. Figure 4.3Id shows the nodal ray. Figure
converging light and converged it even more, which is 4.31e shows all three of the predictable rays. The
what one might expect a converging lens to do. Once exiting rays indicate that the light leaving the lens is
the conjugate image point is located, all the other diverging. Figure 4.3If shows that the virtual image
incident rays can then be traced through the system. position is located by extending the exiting rays back-
Some of these other rays are shown in Figure 4.29f. ward until they intersect. The virtual image is erect
and smaller relative to the object.
Figure 4.32 shows a case of converging light inci-
dent on a diverging thin lens. The incident light was
converged by a preceding optical system, which is not
4.10 Predictable Rays/Diverging shown. The location of the virtual object is at the
Lenses/Off-Axis Object Points center of curvature of the incident wavefront. The
location is found by extending the incident converging
The secondary focal point, F 2 , is the axial image point rays forward until they intersect (Figure 4.32a).
that is conjugate to an axial object point at optical Out of all the incident rays in Figure 4.32a, three
infinity. In other words, F 2 is the on-axis image point are predictable. Figure 4.32b isolates the incident ray
that results when plane waves are incident on the lens. parallel to the axis. In accordance with Figure 4.30a,
For a thin diverging lens, F 2 is a virtual image point. this ray appears to be coming from F 2 when it exits the
Figure 4.30a shows the rays associated with F 2 for a lens. Figure 4.32c isolates the incident ray that is
thin diverging lens. Each incident ray parallel to the pointing toward Fl. In accordance with Figure 4.30b,
axis is bent by the lens and appears to be coming from this ray exits the lens parallel to the axis. Figure 4.32d
F 2 when it exits the lens. Figure 4.30a shows that for a shows the nodal ray. Figure 4.32e shows all three of
thin diverging lens any incident ray parallel to the axis the incident predictable rays with their corresponding
must appear to be coming from F 2 when it leaves the exiting rays. The exiting rays are converging to form a
lens. real image that is erect and larger relative to the
The primary focal point, F 1? is the on-axis object virtual object. Note in this case that the light leaving
point that results in plane waves leaving the lens. the lens is not as convergent as the light incident on
Clearly, to get plane waves leaving a diverging lens, the lens.

FIGURE 4.30. Rays associated with the pri-


mary and secondary focal points of a diverging
a) lens.
Thin Lenses and Ray Diagrams 53

FIGURE 4.31. a. Diverging rays incident on a


diverging lens. b. The predictable ray parallel to the
axis. c. The predictable ray pointing toward the
primary focal point, d. The nodal ray. e. Image
location by the three predictable rays. / . Other rays.

FIGURE 4.32. a. Converging rays incident such


that the virtual object is inside the primary focal
point, b. The predictable ray parallel to the axis. c.
The predictable ray pointing toward the primary
focal point, d. The nodal ray. e. Real image location
by the three predictable rays.
54 Geometrie, Physical, and Visual Optics

Figure 4.33a again considers converging light inci- Instructional Comment


dent on a diverging lens. As compared to the previous
example, the incident light is less convergent. The Again, these types of ray diagrams can be done on graph
virtual object is located by extending the incident rays paper, and that is sometimes necessary. However, I be-
forward until they intesect. Figure 4.33b isolates the lieve the greatest benefit is gained by learning to quickly
incident ray parallel to the axis. This ray appears to be draw a freehand sketch that is accurate enough to supply
coming from F 2 when it leaves the lens. Figure 4.33c the needed information. The beginner should avoid
isolates the incident ray that is pointing toward Fj and memorizing these ray diagrams. Instead, by remembering
the three predictable rays, one should be able to draw a
is parallel to the axis when it leaves. Figure 4.33d ray diagram and trust the results.
isolates the nodal ray.
Figure 4.33e shows all three of the predictable rays.
The exiting rays are diverging and the image is virtual.
The position of the virtual image is located by extend- 4.11 Common Features
ing the exiting rays backward until they intersect as
shown in Figure 4.33f. In this case, the virtual image is The predictable rays involve the nodal point and the
inverted and larger relative to the virtual object. Keep focal points. The nodal point is at the optical center of
in mind that the virtual image is associated only with a lens for both converging and diverging lenses. When
the light physically leaving the lens (i.e., on the right plane waves along the axis are incident on the front of
in Figure 4.33f), and that the virtual object is as- a lens, the resulting image point is the secondary focal
sociated only with the light physically incident on the point, F 2 , which is located at the center of curvature
lens (i.e., on the left in Figure 4.33f). (You might of the exiting wavefront. For light incident on the
want to consider a tinted lens and color code the rays front of a lens, the center of curvature of the exiting
for this situation.) wavefront is behind a converging lens and in front of a

->* + > +

>+ >> +

^ ^ ^ \ FIGURE 4.33. a. Converging rays incident

f L^'
s' F< virtual such that the virtual object is outside the pri-

^
1
^ ' object
v
mary focal point, b. The predictable ray inci-
dent parallel to the axis. c. The predictable ray
k pointing toward the primary focal point, d. The
nodal ray. e. The three predictable rays. /. Ray
extensions locating the virtual image.
Thin Lenses and Ray Diagrams 55

FIGURE 4.34. Common features of wavefronts and


rays associated with the secondary focal point.

diverging lens (Figures 4.34a and 4.34b, respectively). point toward Fx for a diverging lens (Figures 4.35c and
For both converging and diverging lenses, the parallel 4.35d, respectively). In both cases, the rays are easily
rays associated with the incident plane waves are drawn in from the wavefront diagrams.
always associated with F 2 when exiting the lens. The
exiting rays point toward F 2 for a converging lens and
point away from F 2 for a diverging lens (Figures 4.34c 4.12 Scaling of Ray Diagrams
and 4.34d, respectively). In both cases, the rays are
easily drawn in from the wavefront diagrams. The vertical scale on a ray diagram is frequently
When plane waves leave the back of a lens, the different than the horizontal scale. Figure 4.36 shows a
original object point is the primary focal point, Fj. To case where the vertical scale spacing is 1 cm per mark
get plane waves leaving a converging lens, the incident and the horizontal scale spacing is 10 cm per mark.
light waves must be diverging. To get plane waves Scale differences are used so that ray diagrams are
leaving a diverging lens, the incident light must be large enough to see. If the 10 cm per mark were used
converging. The location of the primary focal point is for both vertical and horizontal scales, the ray diagram
at the center of curvature of the incident wavefront. would be squashed along the horizontal axis and the
The center of curvature of the incident wavefront is in details would be hard to see. If the 1 cm per mark
front of a converging lens and behind a diverging lens scale were used, the ray diagram would be too large
(Figures 4.35a and 4.35b, respectively). For both con- for convenient use. The different scales enable the ray
verging and diverging lenses, the parallel rays as- diagram to be compact with the details still visible.
sociated with the exiting plane waves are always as- Note, with different scales the angles are distorted and
sociated with Fj when entering the lens. The incident can look large even though they are, in reality, small
rays point away from Fx for a converging lens and angles made by paraxial rays.

^ -V
y

Oei
<Tl
7

F^i
^C
1 r FIGURE 4.35. Common features of wavefronts and
rays associated with the primary focal point.
56 Geometrie, Physical, and Visual Optics

FIGURE 4.36. Horizontal and vertical


scaling differences.

Problems image. Indicate approximately the location of the


image. Is it real or virtual, erect or inverted,
larger or smaller?
1. An extended object is midway between a converg- 5. For a diverging lens and a real near object, locate
ing lens and its primary focal point. Draw a ray the image by drawing a ray diagram. Is the image
diagram to locate the image. Is the image real or real or virtual, larger or smaller, erect or in-
virtual, erect or inverted, larger or smaller? verted?
2. An extended object is midway between Fl and 2Fj 6. A diverging lens has a virtual primary focal point.
for a converging lens. Draw a ray diagram to Draw a ray diagram to locate the image for a
locate the image. Is the image real or virtual, erect virtual object located midway between the pri-
or inverted, larger or smaller? mary focal point and the lens. Is the image real or
3. An extended object is at 4FX for a converging virtual, larger or smaller, erect or inverted?
lens. Draw a ray diagram to locate the image. Is 7. A diverging lens has a virtual primary focal point.
the image real or virtual, erect or inverted, larger A virtual object is located twice as far from the
or smaller? lens as the primary focal point. Find the image by
4. A converging lens with a secondary focal length of drawing a ray diagram. Is the image real or virtu-
g, has a virtual extended object located a distance al, erect or inverted? Is the image larger, smaller,
g from the lens. Draw a ray diagram to locate the or the same size?
CHAPTER FIVE

Thin Lens
Equations

5.1 Thin Lens Vergence Equation Then the exiting vergence V is

The thin lens in air in Figure 5.1a has a diverging V = P + U,


wavefront incident on it and a converging wavefront V = + 6 . 0 0 D + (-9.00D),
leaving it. The degree of divergence or convergence of
or
these wavefronts at the lens is specified quantitatively
by their vergence. The action of the thin lens on V=-3.00D.
paraxial wavefronts can be described by a simple When plane waves are incident on the +6.00 D lens,
vergence equation: U equals zero and the exiting vergence V numerically
V = P + U, (5.1) equals the dioptric power P of the lens:
where U is the vergence of the wavefront incident on V = P + U,
the lens, V is the vergence of the wavefront exiting the V = + 6 . 0 0 D + 0,
lens, and P is called the dioptric power of the lens.
(Equation 5.1 is derived in Chapter 7.) The dioptric or
power P is a characteristic of the lens and is a measure V=+6.00D.
of the lens' converging or diverging power. The units
of the vergence U and V are diopters, so from Eq. 5.1 Note that the exiting vergence depends on both the
the units of P must also be in diopters. P is positive for dioptric power of the lens and the incident vergence.
converging lenses and negative for diverging lenses. The exiting vergence, therefore, equals the dioptric
If the vergence U of the wavefront incident on the power of the lens only when plane waves are incident
lens is -4.00 D and the vergence V of the wavefront on the lens.
exiting the lens is +2.00 D, then from Eq. 5.1 the
dioptric power P of the lens is
P=V-U,
5.2 Object and Image Distances
P = 2.00 D - (-4.00 D),
An object for a thin lens in air is labeled real or virtual
or depending on whether the incident light is diverging or
converging. The position assigned to the object point
P=+6.00D. is at the center of curvature of the incident wavefront
The dioptric power of a lens does not change when that hits the lens. An image for a lens is labeled real or
the incident vergence changes. Suppose the vergence virtual depending on whether the exiting light is con-
U incident on the +6.00 D lens is changed to -9.00 D. verging or diverging. The position assigned to the
57
58 Geometrie, Physical, and Visual Optics

->> +

image

object

L image
-25 cm*k- -50cm

FIGURE 5.1. a. Wavefronts converged by a lens. b. Ex-


b) tended real object and conjugate real image.

image point is at the center of curvature of the wave- vergence V of this wavefront (in diopters) equals the
front that is just leaving the lens. reciprocal of the image distance (in meters).
Both the object point and image point can be on
either side of a lens. For example, a real object point V=-. (5.3)
might be 15 cm in front of the lens while a virtual v v ;

object point might be 15 cm behind the lens. A sign When the distances are not expressed in meters,
convention is used to distinguish between these cases. then conversion factors can be used in the numerator,
The object distance u and image distance v are di- as discussed in Section 2.5. In Figure 5.1, u =
rected distances along the optical axis measured from -25.0 cm and from Eq. 5.2,
the lens to the object and from the lens to the image 100cm/m
TT A _ _
respectively. The positive direction is the direction
along the optical axis in which the light is traveling -25.0 cm
when it leaves the lens. while v = +50 f 0cm and from Eq. 5.3,
In Figure 5.1a, light leaves the lens traveling to the , 100 cm/m ^
right, which thus becomes the positive direction (indi- v =
^; = +2.OOD.
cated by the symbol, >+). The image distance v, 4-50.0 cm
measured from the lens to the image position, is in the In Figure 5.2, the light leaving the lens is again
same direction that the exiting light is traveling and is traveling to the right. The incident light is converging
therefore positive (i; = +50.0 cm). The object distance and the center of curvature of the incident wavefront
w, measured from the lens to the object position, is is 20.0 cm to the right of the lens. The light leaving the
opposite to the direction that the exiting light is lens is diverging and the center of curvature of the
traveling and is therefore negative (u = -25.0 cm). exiting wavefront is 10.0 cm to the left of the lens.
The object distance u is the radius of curvature of Here,
the wavefront incident on the lens and the vergence U u = + 20.0 cm,
of this wavefront (in diopters) equals the reciprocal of
U== * =+5.00D,
the object distance (in meters). u +20.0n cm
1 v= -10.0 cm,
U= (5.2)
and
The image distance v is the radius of curvature of
the wavefront that is just leaving the lens and the V = iv = _J^
-10.0 cm = _ 1 0 . 0 0 D .
Thin Lens Equations 59

^+

u= +20cm

** +

v= -10cm

-*-4-

virtual
object v JK

virtual
image^ 10cm- -20 cm-

FIGURE 5.2. a. Converging wavefronts incident on a di-


v= -10cm u= + 20cm verging lens. b. Diverging wavefronts leaving the lens. c. Exten-
ded virtual object and conjugate virtual image.

5.3 Dioptric Power: Focal Length V is equal to the dioptric power P of the lens, or
Relationships
V = P + U = P.

The secondary focal length, / 2 , is a directed distance From Eq. 5.3,


along the optical axis of a lens and is measured from
the lens to the secondary focal point, F 2 . For plane v-i.
waves incident on the lens, the image point occurs at V

the secondary focal point, or By combining the above three equations

v=f2-
When plane waves are incident on the lens, the inci-
dent vergence U equals zero and the exiting vergence
'-i
Equation 5.4 relates the secondary focal length (in
(5.4)
60 Geometrie, Physical, and Visual Optics

P=+4.00D V 4- Then, by combining the above equations,


i P=-
1
(5.5)
K
Equation 5.5 relates the primary focal length (in me-
F, ters) of the lens to the dioptric power (in diopters).
When the primary focal length is expressed in units
25cm- -25 cm- other than meters, the appropriate conversion factor
can be used in the numerator.
U= -25cm f 2 =+25cm Note, from Eqs. 5.4 and 5.5,
k = -L
FIGURE 5.3. Primary and secondary focal lengths for a which confirms that for a thin lens in air the focal
+4.00 D thin lens.
points are equidistant from and on opposite sides of
the lens.
The focal lengths for the +4.00 D lens in Figure 5.3
meters) of the lens to the dioptric power (in diopters).. are:
(Some authors use Eq. 5.4 as the defining equation for 100
dioptric power of a thin lens in air instead of the more f = - = = +25.0 cm,
+4.00 D
general vergence equation, 5.1.) When the secondary
focal length is expressed in units other than meters, 100
= -25.0 cm.
the appropriate conversion factor can be used in the ^1 P +4.00 D
numerator. In this case, the focal points are both real.
The primary focal length, fl9 is a directed distance The focal lengths for the -3.00 D lens in Figure 5.4
along the optical axis of a lens and is measured from are:
the lens to the primary focal point, F x . When plane r 1 100
waves are leaving the lens, the object is at the primary fl = p = -3.00
. n n nD = " 3 3 . 3 3 Cm,
focal point, or
/.-i- P
-100
-3.00 D
= +33.33 cm.

For plane waves leaving the lens, V= 0 and from Eq. In this case, the focal points are both virtual.
5.1,
v=o = p + u,
or 5.4 Lateral Magnification
P=-U.
The images of extended objects are typically larger or
From Eq. 5.2,
smaller than the object. The ratio of the image size to
the object size is called the lateral magnification (m).
U=
Let O be the object size and I be the image size.

P = -3.00D-
\

F2 Fl

^> OO <JO ~ , OO OO . ^
.

f2= -33.33 cm U = +33.33 cm


FIGURE 5.4. Primary and secondary focal lengths for a
/k -3.00 D thin lens.
Thin Lens Equations 61

Assume O is positive if the object point is above the Equation 5.7 shows that the lateral magnification is
optical axis and negative if the object point is below equal to the linear ratio of the image distance to the
the axis. Similarly, assume I is positive if the image object distance. For this reason lateral magnification is
point is above the axis and negative if the image point sometimes called linear magnification.
is below the axis. Then, Since the object and image distances are recipro-
cally related to the vergences U and V, we can com-
I bine Eqs. 5.7, 5.2, and 5.3 to obtain
(5.6)
U
A positive lateral magnification (m) indicates that the m= (5.8)
image is erect relative to the object and a negative m
indicates the image is inverted relative to the object. Equation 5.8 shows that a knowledge of U and V is
A lateral magnification m greater than 1 indicates the sufficient to compute the lateral magnification.
image is larger than the object and an m less than 1
indicates the image is smaller than the object. Assume
that the object size O is +4.0 cm and the image size I
is -2.0 cm. Then from Eq. 5.6, the lateral magnifica-
5.5 Examples
tion m is,
-2.0 cm It is a good problem solving technique to predict what
m = = -0.5. characteristics an answer should have before doing any
+4.0 cm
calculations. In working these examples, I suggest that
Figure 5.5 shows the nodal ray for a lens that has a a quick ray diagram sketch initially be made. This
real object and a real image. Since the nodal ray is a sketch provides the predictions for what the numerical
straight line, the angle w subtended by the object at answer should be. If the answers do not correlate with
the nodal point is equal in magnitude to the angle w' the predictions, that is a tip-off that something is
subtended by the image at the nodal point. Then, wrong. If this does occur, then double-check the ray
diagram (might have to be more accurate this time)
tan w' = and the numerical calculations.
and
O EXAMPLE 5.1
tan w = .
u A real object is located 100 cm in front of a
Since w == w', the above equations give +5.00 D lens. Where is the conjugate image? Is it
real or virtual, erect or inverted, larger or smaller?
I _ O The primary and secondary focal lengths for a
v u ' +5.00 D lens are, from Eqs. 5.4 and 5.5,
or 1 100cm/m ^ ^
r
I _ v / = p = T I F - + 2 0 0 0 c m
0~ u' 100
= -20.00 cm.
Therefore, from Eq. 5.6,
f.-l
From a ray diagram sketch (Figure 5.6), the predic-
m= (5.7)
tions are: the image is real, inverted, smaller, and

FIGURE 5.5. Geometry relating object and image size.


62 Geometrie, Physical, and Visual Optics

P=+5.00D
-** +

FIGURE 5.6. Ray diagram for a real


object 100 cm in front of a +5.00 D
lens.

located at about 25 cm. Now, let us calculate the EXAMPLE 5.2


numbers. A real object is placed 30 cm from a +5.00 D lens.
Is the conjugate image real or virtual, erect or
u = -100 cm, inverted, larger or smaller? What is the image
distance? (You might try this example yourself
- 100
before looking at the solution.)
The focal points for the +5.00 D lens are
20.0 cm away as found in Example 5.1. From a ray
- cm =
u = -100 -1.00D.
diagram sketch (Figure 5.7), the predictions are:
The negative U value agrees with the fact that the the image is real, inverted, larger, and about 60 cm
light incident on the lens is diverging. from the lens. The calculations follow.
V = P + U, u = - 3 0 cm,
V=+5.00 +(-1.00),
V=+4.00D.
u=I,
u
The positive V value indicates that the light leaving 100
U = = -3.33D.
the lens is converging, which agrees with the ray (-30)
diagram. The negative U value agrees with the fact that the
1 light incident on the lens is diverging.
V = 77,
V = P + U,
100 V = +5.00 + (-3.33),
= +25 cm.
+4.00 D
V=+1.67D.
The image distance of +25 cm agrees well with the The positive V value indicates that the light leaving
prediction. the lens is converging, which agrees with the ray
U diagram.
m = - = -0.25.
+4 1
=
The lateral magnification of -0.25 indicates that V'
the image is inverted and 0.25 times as large as the 100
+60.0 cm.
object. Both items agree with the ray diagram. + 1.67

FIGURE 5.7. Ray diagram for a real ob-


ject 30 cm in front of +5.00 D lens.
Thin Lens Equations 63

FIGURE 5.8. Wavefronts for Figure 5.7.


Vergence values can be calculated for any of
the wavefronts.

The image distance also agrees with the ray dia- point is 30 cm in front of the lens. Therefore, the
gram. wavefront 5 cm in front of the lens is 25 cm from
the object point and has a vergence of -4.00 D.
U -3.33
= -2.0.
V + 1.67 EXAMPLE 5.3
A real object is placed 10 cm in front of a +7.00 D
The lateral magnification indicates that the image is lens. Is the image real or virtual, larger or smaller,
inverted and twice as large as the object. Note that erect or inverted? What is the image distance?
the lateral magnification can also be calculated (Again, try this example yourself before looking at
from Eq. 5.7: my solution.)
v +60 cm The focal lengths for a +7.00 D lens are:
m = - = -^ = -2.0. 100
u - 3 0 cm 1
= +14.3 cm,
^ P +7.00
The vergence U incident on the lens is -3.33 D.
The vergence V leaving the lens is +1.67D. The 1. 100
= P = -14.3 cm.
wavefronts are shown in Figure 5.8. It is instructive +7.00
to calculate the vergence of the light at places other From a ray diagram sketch (Figure 5.9), the predic-
than the lens. For example, what is the vergence of tions are: the image is virtual, erect, larger, and the
the light at a point 40 cm behind the lens? image distance is about 30 cm. The calculations
All of the wavefronts leaving the lens are spheri- follow.
cal and have their centers of curvature at the image
point. The image point is 60 cm behind the lens, so u = - 1 0 cm,
the wavefront 40 cm behind the lens is 20 cm from 100
the image point or has a radius of curvature of U = -10.00 D.
(-10)
20 cm. Therefore, the vergence of the wavefront
40cm behind the lens is then +5.00D (i.e., 100/ The negative U value agrees with the fact that the
20 cm). light incident on the lens is diverging.
What is the vergence of the light at a point 5 cm V = P + U,
in front of the lens? All of the wavefronts incident
on the lens are diverging and have their centers of V = +7.00 + (-10.00),
curvature located at the object point. The object V=-3.00D.

FIGURE 5.9. Ray diagram for real object 10 cm in front


of a +7.00 D lens.
64 Geometrie, Physical, and Visual Optics

FIGURE 5.10. Wavefronts for Figure 5.9. C and B show


positions for vergence values in front of and behind the lens,
respectively.

The negative V value indicates that the light leav- at the real object that is 10 cm in front of the lens.
ing the lens is diverging, which agrees with the ray Since point C is 6 cm in front of the lens, then point
diagram. Then C is 4 cm from the object. Therefore, the wavefront
at C has a vergence of -25.00 D. Note that the
100
= -33.3 cm, vergence of the wavefront incident on the lens was
-3.00 -10.00 D and that the vergence 6 cm in front of the
lens is -25.00 D.
m = = z- = +3.33.

The lateral magnification indicates that the image is EXAMPLE 5.4


erect relative to the object and 3.33 times larger A virtual object is located 50 cm from a +7.00 D
than the object. Note that in this case the light lens. Is the conjugate image real or virtual, larger
leaving the converging lens is diverging (-3.00 D) or smaller, erect or inverted? What is the image
but not as divergent as the incident light (U = distance?
-10.00 D). A plus lens used as a magnifying lens The virtual object means that the incident light
works like this. is converging and the center of curvature of the
Let us consider the vergence of light at a point B incident converging wavefront is located 50 cm be-
that is 10 cm behind the lens. The diverging wave- hind the lens (Figure 5.11). The focal points are
front just leaving the lens has a radius of 33.3 cm located 14.3 cm from the +7.00 D lens (Example
(Figure 5.10). All the other diverging wavefronts 5.3). From a ray diagram sketch (Figure 5.12), the
behind the lens are further from the virtual image, predictions are: the image is real, erect, smaller,
so their radii will be greater than 33.3 cm. There- and located at about 10 cm. The calculations
fore, the radius of the wavefront at point B is follow.
43.33 cm and the vergence of the wavefront is: u = +50 cm.
100 ^ ., ^ Light leaves the lens going to the right. The posi-
V
* = ^4333 = - 2 3 1 D tive object distance agrees with this since the posi-
Note that the vergence of the wavefront leaving the tive direction is the direction in which the light is
lens was -3.00 D and that the vergence 10 cm traveling when it leaves the lens.
behind the lens is -2.31 D. 100
What is the vergence of light at a point C that is U=
6 cm in front of the lens? The center of curvature of (T5)=+200D
all the diverging wavefronts incident on the lens is Here is the best place to check for signs. Since the

P=+7.00D

. .virtual
U^object

FIGURE 5.11. Incident rays for a virtual object located 50cm


from a +7.00 D lens.
Thin Lens Equations 65

+ 7.00D

, virtual
r^-^object

FIGURE 5.12. Real image location for the Figure 5.11.

incident light is converging, U must be positive -4.00 D lens are:


otherwise something is wrong.
100
= - 2 5 . 0 cm,
V = P + U, -4.00
V = +7.00 + (+2.00), 100
= - = +25.0 cm.
V=+9.00D. -4.00
The positive V value indicates that the light leaving Remember that for a minus lens, the focal points
the lens is converging, which agrees with the ray are both virtual and F 2 is located in front of the
diagram. lens. From a ray diagram sketch (Figure 5.13), the
predictions are: the image is virtual, erect, smaller,
100 and located at 15 cm. The calculations follow.
+9.00 = +11.1 cm,
U _ +2 - 5 0 cm,
= +0.22.
V " +9 100
U = = -2.00D.
The lateral magnification indicates that the image is
(-50)
erect and only 0.22 times as large as the object. The negative U value agrees with the fact that the
This agrees with our ray diagram. In this example, light incident on the lens is diverging.
the incident light is already converging (U =
+2.00 D), and the lens converges it even more V = P + U,
(V=+9.00D). V = -4.00 +(-2.00),
V=-6.00D.
EXAMPLE 5.5
A real object is located 50 cm in front of a -4.00 D The negative V value indicates that the light leav-
lens. Is the conjugate image real or virtual, larger ing the lens is diverging, which agrees with the ray
or smaller, erect or inverted? What is the image diagram.
distance? 100
The primary and secondary focal lengths for the = =
^6 "16-67
U -2
m = = - = +0.33.
V o
-*- + The lateral magnification indicates that the image is
P=-4.00D Y
erect and 0.33 times smaller than the object. In this
case, the lens took the incident light ( U = - 2 . 0 0 D )
and diverged it even more ( V = - 6 . 0 0 D ) .
real.
object
EXAMPLE 5.6
A virtual object is located 8 cm from a 4.00D
lens. Where is the conjugate image? What is the
lateral magnification? Is the image real or virtual?
(Remember, you should try this first yourself.)
The focal points are virtual and 25 cm from the
lens (Example 5.5). From a ray diagram sketch
FIGURE 5.13. Ray diagram for a real object 50 cm in front of (Figure 5.14), the predictions are: the image is
a -4.00 D lens. real, erect, larger, and located at about 12 cm. The
66 Geometric, Physical, and Visual Optics

FIGURE 5.14. Ray diagram for a virtual object located


8 cm from a -4.00 D lens.

calculations follows. inverted, larger or smaller? What is the image


u = +8 cm, distance?
The primary and secondary focal lengths for the
-8.00 D lens are:
1 100
The positive U value agrees with the fact that the = -12.5 cm,
J2
PD -8.00
light incident on the lens is converging.
1 100
V = P + U, ^1 -8.00 = +12.5 cm.
V = - 4 . 0 0 + ( + 12.5), From the ray diagram sketch (Figure 5.15), the
V=+8.50D. predictions are: the image is virtual, inverted,
The positive V value indicates that the light leaving larger, and the image distance is about 30 cm. The
the lens is converging, which agrees with the ray calculations follow.
diagram. u = +20 cm,
100 100
v = +8.50 = +11.8cm, U = (+20) = +5.00D.
u
m =
+ 1 2 5 The positive U value agrees with the fact that the
V = T ^ 5 = +147
light incident on the lens is converging.
The lateral magnification indicates that the image is
erect and 1.47 times larger than the object. V = P + U,
V = -8.00 + (+5.00),
EXAMPLE 5.7
A virtual object is located 20 cm from a -8.00 D V=-3.00D.
lens. Is the conjugate image real or virtual, erect or The negative V value indicates that the light leav-

P=-8.00D >
/
-> +


<
j ^ > virtual ot

S ^y* /F,
^y^
y ^y^
/
virtual y ^y^f /
image y y^ /
y
y^s^
^y^
/
/
\ ^1 /

7J
^~
FIGURE 5.15. Ray diagram for a virtual object located
20 cm from a -8.00 D lens.
Thin Lens Equations 67

ing the lens is diverging, which agrees with the ray 5.6 Boundary Role of the Focal Points
diagram.
100
v= - n n = -33.3 cm, Besides the conjugate relationship of the focal points
-3.00 with optical infinity and the use of the focal points in
U = +5 drawing predictable rays, the focal points also serve as
m = V " - 3 = -1.67. boundaries in the imaging process. Consider a real
The lateral magnification indicates that the image is object located in front of a converging lens of dioptric
inverted and 1.67 times larger than the object. power P. When the object is between optical infinity
What is the vergence at a point D that is 5 cm in and the primary focal point F 1? then the negative
front of the lens and at a point E that is 5 cm vergence U of the incident diverging wavefront is not
behind the lens (Figure 5.16)? The light in front of enough to overcome the positive dioptric power P.
the lens is converging. Each incident wavefront has The vergence V of the wavefront leaving the lens is
its center of curvature at the virtual object position positive, the light leaving is converging, and the result-
20 cm behind the lens. So, the wavefront at point D ing real image is a finite distance behind the lens.
has a radius of curvature of 25 cm. Its vergence is When the real object is at F 1? then plane waves leave
the lens, V is zero, and the image is at optical infinity.
U
U = ^ - = +4.00D. When the real object is between Fx and the lens, the
(+25)
Note that this wavefront with a vergence of negative incident vergence U is great enough to over-
+4.00 D at a point 5 cm in front of the lens has a come the positive dioptric power P. The exiting ver-
vergence of +5.00 D when it hits the lens. Also gence V is negative, the light leaving is diverging, and
note that the dashed rays associated with the virtual the resulting virtual image is located a finite distance
image do not have anything to do with the incident in front of the lens. (I suggest that you make some
light. These dashed rays are in image space because quick ray diagrams to check the above assertions.)
they are associated only with the light leaving the In the case of a virtual object for the converging
lens. lens, the incident light is converging and U is positive.
The light behind the lens is diverging, and each Then, from V= P + U, V is always greater than P; and
wavefront has its center of curvature located at the from the reciprocal relations v = 1/V and/ 2 = 1/P, the
virtual image position 33.3 cm in front of the lens.
So, the radius of curvature of the wavefront at image distance v is less than the secondary focal length
point E is 38.3cm (i.e., 33.3cm+ 5cm). The ver- f2. As a result, the real image is trapped between the
gence is lens and the secondary focal point F 2 .
A real object in front of a diverging lens always
y
=(^3)=-2-61D results in diverging light leaving the lens. In the
E
equation V= P + U, P and U are both negative, so V
Note that the diverging wavefront leaving the lens is more negative than P. Again, from the reciprocal
has a vergence of -3.00 D and at a point 5 cm relations = 1 / V and f2 = 1 /P, the image distance v is
downstream it has a vergence of -2.61 D.
negative and smaller in magnitude than f2. Thus, the
virtual image is trapped between F 2 and the lens.
In the case of a virtual object for the diverging lens,
the converging incident light gives a positive value for
the incident vergence U. When the virtual object is
located between the lens and the (virtual) primary
focal point Fj, the positive vergence U is great enough
to overcome the negative dioptric power P. Con-
sequently, V is positive, the exiting light is converging,
and the resulting real image is located a finite distance
behind the lens. When the virtual object is at F l 5 then
plane waves leave the lens and the image is at optical
infinity. When the virtual object is located between Ft
and optical infinity, then the positive U is not great
enough to overcome the negative P. Thus, the exiting
vergence V is negative, the exiting light is diverging,
and the resulting virtual image is located a finite
distance in front of the lens. (Again, I suggest that you
FIGURE 5.16. Wavefronts for vergence calculations at posi- make some quick ray diagrams to check the above
tions in front of or behind the lens (D and E respectively). assertions.)
68 Geometrie, Physical, and Visual Optics

5.7 The Symmetry Points

In developing intuition about the optics of lenses, it is


helpful to consider the special case where the lateral
magnification equals - 1 . 0 . In this case, the image is
the same size as the object but is inverted relative to
the object. Let us start by finding the object distance
that results in m = - 1 . From Eq. 5.8:
U
m=
V
FIGURE 5.17. The symmetry points for a converging (plus)
-1 = u lens.
V
or
V=-U. are located at
u = 2f,= 2(-20.0 cm) = -40.0 cm,
From Eq. 5.1,
and
V = P + U.
v = 2f2 = 2(+20.0 cm) = +40.0 cm.
Then from above, We can use Eq. 5.7 to check the lateral magnifi-
- U = P + U, cation:
v +40
- 2 U = P, m = - = T7T = - I -
u -40
U = - As an alternate check, we can take the object
2'
distance of - 4 0 cm and compute the resulting
or image distance and lateral magnification.
J_ = _ 2 u = -40 cm,
U P 100
U= -2.50 D,
Since u = 1/U and fx = - 1 / P , then -40 " "
u = 2f1. v =P + U,
Thus, when the object distance is twice the primary v = +5.00 + (-2.50),
focal length, the resulting lateral magnification is equal v = +2.50 D
to - 1 . It is easily shown that the image distance in this 100
case is twice the secondary focal length, or v= +40 cm,
+2.50 "
v = 2f2. U
m=
V
The axial points at 2/ and 2/ 2 are conjugate points
-2.50
and are labeled 2FX and 2F 2 , respectively. These m= -1.
+2.50 "
conjugate points are called the symmetry points of the
lens. The symmetry planes contain the symmetry The ray diagram i s shown in Figure 5.17
points and are perpendicular to the optical axis of the
lens. EXAMPLE 5.9
Find the symmetry points for a -20.00 D lens. The
primary and secondary focal lengths are:
EXAMPLE 5.8
Find the symmetry points of a +5.00 D lens. The
/ 2 = =
primary and secondary focal lengths of a +5.00 D P ^2 = _ 5 0 c m
'
lens are: . 1 100 ._.
/ =_
2
= 1 = 100
+20.0 cm, ' p = "^2a = +50cm

P +5.00 Again,
1. 100 u = 2fx = 2 ( + 5 . 0 cm) = +10.0 cm,
= P +5.00
-20.0 cm.
and
From the above discussion, the symmetry points v = 2f2 = 2 ( - 5 . 0 cm) = -10.0 cm.
Thin Lens Equations 69

P=-20.00D

FIGURE 5.18. The symmetry points for a diverging (minus)


lens.

One check is: The image distance is - 2 . 5 cm and the lateral magnifi-
v -10 cation is +1.25. Now set the object at 1 cm in front of
the lens and repeat the calculations. The image dis-
An alternate check follows. tance becomes -1.111 cm and the lateral magnifica-
tion m becomes +1.111. The calculations can be
u = +10 cm, repeated as the object is moved closer and closer to
U= ^ = +10.00 D. the lens. The results are shown in Table 5.1.
Note from the table that the smaller the object
Note, the incident light is converging and the ob- distance becomes, the closer the image distance gets to
ject is virtual. the object distance and the closer the lateral magnifi-
V = P + U, cation gets to +1. In the limit of the object distance
going to zero, the image distance goes to zero and the
V= -20.00 + ( + 10.00), lateral magnification goes to + 1 .
V = -10.00 D. The physical meaning of the +1 limit is that a thin
The light leaving the lens is diverging and the lens has no optical effect when the object distance is
image is virtual. zero. One way to observe this is to lay a thin lens
100 directly on some letters in a book. The letters appear
m the same size with or without the lens over them.
-10 = -10 cm,
Another way to have an object at the lens is to use
U +10D a previous lens to converge light and form an image at
= -1.
V -10D the second lens. The image for the first lens is the
The ray diagram is shown in Figure 5.18. object for the second lens, and the second lens object
distance u2 is zero. Figure 5.19 is a ray diagram for
this case.
5.8 Unit Lateral Magnification Figure 5.19a shows two rays representing the inci-
dent converging light. As usual, the ray parallel to the
What happens to the lateral magnification as an object axis passes through F 2 when it leaves the lens. The
is moved closer and closer to a lens? Let's consider a other ray, passing through F 1? leaves the lens parallel
specific case with a +10.00 D lens. First, set the object to the axis. The two outgoing rays are diverging away
at 2 cm in front of the lens. from the same point on the lens toward which the
u = 2 cm,
100
U = =-50.00 D, TABLE 5.1
Lateral Magnification as Object Distance Goes to Zero
V = P + U,
V = + 1 0 + (-50), u (cm) v (cm) m

V = -40.00 D, -2.000 -2.500 + 1.250


-1.000 -1.111 + 1.111
100 -0.500 -0.526 + 1.053
i; = = -2.5 cm,
-40 -0.250 -0.256 + 1.026
-0.125 -0.127 + 1.013
U -50
m= = +1.25. -0.0625 -0.0629 + 1.0063
-40
70 Geometrie, Physical, and Visual Optics

/ *~ I

y/Fi F2X
FIGURE 5.19. a. Converging rays form-
ing an image at a thin lens. b. Single
Y
notched ray is bent down at the lens, and
double notched ray comes out parallel to
a) b) the axis.

incoming rays converged. The object distance is zero, In Figure 5.21, an object is moved counterclock-
the image distance is zero, and the lateral magnifica- wise around the circle starting from optical infinity,
tion is +1. passing through the lens, and ending back at optical
infinity. The points marked on the circle include the
primary ( F J and secondary (F 2 ) focal points as well
as the symmetry points 2FX and 2F 2 . The A on the
5.9 Object Movement/Image Movement outside of the circle marks the starting position for the
real object at optical infinity. Its real conjugate image
Let us consider what happens to the image as an point is at F 2 and is marked on the inside of the circle
object for a converging lens is moved smoothly by A. As the object is moved counterclock-
through space. Figure 5.20 helps to visualize the situa- wise from optical infinity to 2FX, its image moves
tion. The light incident on the lens is traveling to the counterclockwise from F 2 to 2F 2 . The object position
right. The object positions are represented as points at 2Fj is now marked on the outside by B, and its
on a distorted circle. The points on the left half of the image position at 2F2 is now marked on the inside by
circle represent real object points, while the points on B'. As the object is moved from 2Fj to Fl9 its image
the right half of the circle represent virtual object moves from 2F2 to optical infinity. At F x , the object
points. position is represented on the outside by C, and its
The circle depicts the fact that there are two ways image position at optical infinity is represented on the
to reach optical infinity. A real object can be moved inside by C . As the real object continues its move-
left from the lens until it reaches optical infinity, or a ment, this time from Fj to the lens, its image also
virtual object can be moved right from the lens until it continues its movement, starting from optical infinity
reaches optical infinity. As a real object is moved left, (becoming virtual at this stage) and traveling all the
the incident light is initially diverging but loses nega- way from optical infinity until it finally catches up with
tive vergence until all the waves incident on the lens the object at the lens. The conjugate object and image
effectively become plane waves. As a virtual object is positions at the lens are marked D and D', respective-
moved right, the incident light is initially converging ly. Next, as the object becomes virtual and moves
but loses positive vergence until all the waves incident from the lens to optical infinity (object position D to
on the lens again effectively become plane waves. A), its conjugate image becomes real once again,

-oo-

2F, Fi 2F2

FIGURE 5.20. Distorted circle to help visualize


object movement/image movement relations.
Thin Lens Equations 71

A
-QO-

D'
k

B C A' B'
-
2F, Fi F2 2F2

>f FIGURE 5.21. ', ', C, and D' are the respec-
tive image positions conjugate to the object posi-
tions A, B, C, and D.

moving from the lens to F 2 (image position D' to A'). second lens. The final image position can be found by
Both the object and its conjugate image have com- successively applying V= P + U to each lens.
pleted a full trip around the circle. A special equation can be derived for the total
A similar circle can be constructed for a minus lens lateral magnification mt in the two-lens system. Let Ol
(see the problems). and l1 be the object and conjugate image size for the
first lens, and let 0 2 and I 2 be the object and conju-
gate image size for the second lens. The total lateral
magnification mt is defined as the ratio of the final
5.10 Erect and Inverted Relationships: image size I2 to the original object size 0 1 ? or
Single Lenses x
2
m =
For a single thin lens, a real image is always inverted or
The above equation can be multiplied and divided by
relative to a conjugate real object. One can easily
prove this statement from the lateral magnification 0 2 . After algebraically rearranging, the result is
equation, m = U/V. For a real object, the incident I2 02
light is diverging and U is negative. For a real image, m = x
the exiting light is converging and V is positive. A ' ; cv
negative U value divided by a positive V value gives a However, since the image for lens one is the object for
negative m value, which indicates inversion. lens two, Ij = 0 2 . Therefore,
For a single thin lens, a virtual image is always
erect relative to a conjugate real object. Here, U is JL
m , = O,
again negative, but now V is also negative since the
exiting light is diverging. A negative U value divided The lateral magnifications ^ and m2 for lenses one
by a negative V value results in a positive m value, and two are, respectively,
which indicates the image is erect relative to the
object. m, =
Similarly, one can show that for a single thin lens a
real image is always erect relative to a conjugate _ I,
x
2
virtual object, and a virtual image is always inverted m? =
relative to a conjugate virtual object.
o,
By combining the above three equations, one finds
that
mt = m 2 m 1 . (5.9)
5.11 Two-Lens Systems The total lateral magnification is the product of the
lateral magnifications for each individual lens. This
Consider a lens of dioptric power Px placed at distance statement is easily extended to multiple lens systems
d in front of a second lens of dioptric power P 2 . The containing more than two lenses. In particular, for a
light that leaves the first lens is incident on the second system of n thin lenses,
lens, so the image for the first lens is the object for the m
t = m
n m
n- m 2 m! (5.10)
72 Geometrie, Physical, and Visual Optics

-4.00D
+ 7.50D

FIGURE 5.22. Rays and wavefronts in a two-lens sys-


tem. There are downstream vergence changes between the
24 cm two lenses.

EXAMPLE 5.10 For lens 2:


A +7.50 D lens is placed 10 cm in front of a
u2 = 24 cm - 10 cm = +14 cm,
-4.00 D lens. A real object is placed 30 cm in front
of the +7.50 D lens. Find the final image position
and the total lateral magnification, and state U 2 = ^ = +7.14D.
whether the final image is real or virtual.
For lens 1: The converging wavefront leaving the first lens has
a vergence of +4.17 D. From the figure, the con-
ux = - 3 0 cm, verging wavefronts gain curvature as they cross the
gap between the two lenses so that the vergence of
the converging wavefront incident on the second
, = ^ = -333, lens is +7.14 D.
^ ^ + ^, V2 = P2 + U 2 ,
Vj = +7.50 + (-3.33) = +4.17 D, V 2 = - 4 . 0 0 + 7.14=+3.14 D,

100 100
= +24.0 cm, = +31.8 cm,
+4.17 +3.14
U9 +7.14
Ux = -3.33 m9 = V = +2.27.
mt = V = -0.80. 2 +3.14
x +4.17
All the axial wavefronts are schematically repre-
Figure 5.22 schematically shows the diverging sented in Figure 5.23.
light incident on lens 1 and the converging light The total lateral magnification is
leaving it. The light incident on the second lens is
converging so the object for the second lens is mt = m ^ ! = (+2.27)(-0.80) = -1.82.
virtual. Note from Figure 5.22 that the image for The final image is real, inverted, and 1.82 times
lens 1 is located a distance of 14cm (i.e., 24 c m - larger than the original object.
10 cm) past lens 2. An off-axis ray diagram can be used to confirm

FIGURE 5.23. Rays and wavefronts leaving the sec-


ond lens of the previous figure.
Thin Lens Equations 73

FIGURE 5.24. a. The image of thefirstlens is


the object for the second lens. b. Image location
for the second lens. c. Actual ray paths without
extensions.

the above results. The secondary focal lengths for EXAMPLE 5.11
the two lenses are, respectively, A +12.00 D lens is located 5 cm in front of a
+8.00 D lens. An object is located 5 cm in front of
100 the +12.00 D lens. Where is the final image? Is it
() = ^ = +7.50 = +13.33 cm,
real or virtual? What is the total lateral magnifi-
100 cation?
= - 2 5 . 0 cm.
-4.00
ux = -5 cm,
The first part of the ray diagram is shown in Figure
5.24a. Once the image Ii is located, all the rays for 100
the first lens can be drawn. In particular, we want = 5 = -20.00 D,
to pick those rays that will also turn out to be the
predictable rays for the second lens. Figure 5.24b ^ , + ^
shows the completion of the ray diagram. Figure
5.24c shows just the actual ray paths. V1 = 12.00 + (-20.00) = -8.00 D,
74 Geometrie, Physical, and Visual Optics

P^+12.000 P2=+8.00D

5cm 5cm

12.5^ FIGURE 5.25. Axial rays showing that the virtual


image for the first lens is the real object for the second
u2 lens.

100 100
= -12.5 cm, = +43.7 cm,
+2.29
Uj _ -20.00 U, -5.71
m, = -8.00
= +2.50. m7 = -2.49.
V2 +2.29
Figure 5.25 schematically shows the diverging The total lateral magnification is
light incident on lens 1 and the diverging light
leaving it. The light incident on the second lens is mt = m 2 m i = (-2.49X+2.50) = -6.22.
diverging so the object for the second lens is real. The final image is real, inverted relative to the
From Figure 5.25, it is easy to see that the total object, and 6.22 times larger than the original
distance from the virtual image to lens 2 is 17.5 cm. object (Figure 5.26).
Then,
, = -(12.5 + 5) cm = -17.5 cm,
100 5.12 Two Thin Lenses in Contact
U, = -5.71D.
-17.5
Note that the vergence of the diverging wavefront An important aspect of separated two-lens systems is
leaving the first lens is -8.00 D. The diverging the vergence change of the light as it propagates from
wavefronts lose curvature as they cross the gap lens 1 to lens 2. Therefore, the vergence of the light
between the two lenses so that the vergence of the incident on lens 2 is, in general, not equal to the
light incident on the second lens is -5.71 D.
vergence of the light leaving lens 1.
:P2+U2, Let us now consider the case in which two lenses
V2 = +8.00 + (-5.71) = +2.29 D, are placed in contact with each other. The separation

FIGURE 5.26. Extended ray diagram


for Figure 5.25.
Thin Lens Equations 75

is zero, and the light leaving lens 1 is immediately 100


incident on lens 2. Consequently, for the two thin 2 = +11.8cm.
+8.50 D
lenses in contact with each other, the vergence of the The lateral magnification is
light incident on lens 2 is always equal to the vergence
of the light leaving lens 1, or Uj _ -2.50 D
m = 7 T = +8.50 D = -0.29.
u 2 =v,.
In this case, we can simplify the two-lens problem EXAMPLE 5.13
What single lens acts like a +3.00 D thin lens and a
considerably. For lens 1, -8.00 D thin lens held together?
V ^ P . + U,, From Eq. 5.12,
and for lens 2 P = (+3.00 D) + (-8.00 D) = -5.00 D.
V2 = P2 + U 2 . A -5.00 D thin lens acts exactly like the combi-
nation.
Vj can be substituted into the above equation for U 2
to give EXAMPLE 5.14
A thin lens with a secondary focal length of
V2 = P2 + V,. +50.00 cm is held in contact with a second thin
The substitution for W1 then results in lens. The secondary focal length of the combination
is +16.67 cm. What is the secondary focal length of
^ , + ^ + ,, the unknown lens?
or You might need to be careful here. Equation
2 = (2 + ,) + , . 5.12 says that the dioptric powers are additive. This
does not mean that the secondary focal lengths are
The above equation relates the incident vergence \Jl additive.
to the exiting vergence V2. Furthermore, the above The dioptric power of lens 1 is
equation shows that the relationship is of the form 100
Pi = = +2.00D.
V2 = Pt + U!, (5.11) +50.0
where The dioptric power of the combination is
P ^ P . + Pj. (5.12)
P
' = ^7 = +600D
Equation 5.11 shows that the two thin lenses in con-
tact with each other simply act like one thin lens with Then from Eq. 5.12, the dioptric power of the
a dioptric power of P t . Equation 5.12 states that unknown (second) lens is
dioptric powers are additive for thin lenses in contact P =P -P
with each other. P2 = (+6.00 D) - (+2.00 D) = +4.00 D.
The secondary focal length of the originally un-
EXAMPLE 5.12 known lens is
A +4.00 D thin lens is placed in contact with a
+7.00 D thin lens. A real object is located 40.0 cm f2 = 100/(+4.00D) = +25.0 cm.
in front of the combination. Where is the final The secondary focal lengths of the two thin lenses
image? Is it real or virtual? What is the lateral are, respectively, +50 cm and +25 cm, while the
magnification? secondary focal point of the combination is
From Eq. 5.12, the two thin lenses in contact act + 16.67 cm. Clearly the focal lengths are not ad-
like a single thin lens with a power of ditive.
Pt = (+4.00 D) + (+7.00 D) = +11.00 D.
Note the fact that dioptric powers are additive for
Then
thin lenses in contact results from the fact that there is
ux = -40.0 cm, no vergence change between the two lenses. For two
and separated thin lenses, there will generally be a ver-
gence change between the two lenses; then the diop-
tric powers are not additive.
From Eq. 5.11, The additivity of the dioptric powers is easily exten-
V2 = +11.00 D + (-2.50 D) = +8.50 D. ded to include n thin lenses in contact with each other,
or
The light leaving the combination is converging and
the image is real. The image distance is = + - + + 2 + (5.13)
76 Geometrie, Physical, and Visual Optics

FIGURE 5.27. Projector lens ray


t diagram showing that the image size
depends on the geometric relationship
between two different bundles.

5.13 Bundles, Beams, and Lateral is a property of at least two bundles (Section 5.13).
Magnification Consequently, use of the plane wave approximation
can significantly affect image size calculations.
A movie (or slide) projector forms a large image on a The lateral magnification equation, m = U/V,
distant screen. Figure 5.27 represents the image for- shows that if U is very close to zero, then m is very
mation by a single thin lens projector. The light beam small. From Eq. 5.6, the image size I in terms of m is
leaving such a projector is a diverging beam. (This
I = mO,
beam is frequently visible in movie theaters due to
light scattering off dust particles in the air.) The where O is the object size. A very small m can be
diverging beam is composed of individual converging offset by a large object size O, so that I is not
bundles. The fact that the real image is formed is due necessarily a small value. However, the plane wave
to these individual converging bundles. The fact that approximation (U = zero) leads to m = zero and, con-
the image size is larger than the lens size is due to the sequently, I = zero, which is clearly wrong for large
diverging beam. enough objects. Either the small U value must be kept
The image size is determined not by an individual and not set equal to zero, or an alternate method is
bundle, but rather by the geometric relationship be- needed to calculate image sizes for distant objects.
tween at least two different bundles. (Note that an Let us consider an alternate method. Figure 5.28
axial bundle could also be drawn in Figure 5.27.) The represents a real object located far away from a thin
lateral magnification, defined as the image size divided converging lens. Two bundles are shown: one from the
by the object size, thus, depends on the relationship top of the object and one from the axial object point.
between at least two different bundles, i.e., it is not The object is far enough away so that the plane
considered a property of one, individual, solitary bun- wave approximation can be used and each bundle by
dle. This two-bundle dependence has important impli- itself can be represented by parallel rays. The rays
cations in the next section and in the optics of from the axial point are parallel to the axis since one
spherocylindrical lenses. of the rays coincides with the axis. The rays from the
off-axis point are parallel with each other but not with
the axis. Note that there are two predictable rays in
the off-axis bundle: one through the nodal point of the
lens (the optical center), and the other through the
5.14 Image Sizes for Distant Objects primary focal point Fj. The extended image is formed
in the secondary focal plane of the lens.
When an object is far away from a thin lens, the Figure 5.29 shows only the nodal ray from the
vergence U of the wavefront incident on the lens is off-axis point. The nodal ray forms a straight line
very close to zero. In this case, we frequently use the between the conjugate object and image points. The
plane wave approximation and set the vergence U angle w is the angle that the object subtends at the
equal to zero. Use of the plane wave approximation nodal point of the lens, and the angle w' is the angle
does not significantly affect the calculation of the that the image subtends at the nodal point. Since the
image distance since the image distance is a property nodal ray is a straight line between the conjugate
of a single bundle. The lateral magnification, however, object and image points then the angles w and w', are
Thin Lens Equations 77

FIGURE 5.28. Rays showing two different


incident bundles from a distant object. The
image is inverted and at the secondary focal
point.

equal. From the figure, result is again 8.7 mm. In fact, the percent difference
between the small angle method and the tangent
tanw' = |I|// 2 , method is less than 0.01% in this case.
or Can a large object ever be imaged with zero size?
|l|=/2tanw, (5.14) A typical star is larger than the planet Earth, but so
far away that its subtended angle at the planet Earth is
where w has been substituted for w'. very, very small. Consequently, the subtended angle w
Equation 5.14 is the alternate equation for the is effectively zero, and from Eq. 5.15 the image size I
image size when the object is distant. It avoids the is then zero. As a result, when a person looks at a star
lateral magnification difficulties with the ultrasmall on a clear night, the person's retinal image of the star
lateral magnification values. is effectively a point image. (If you are driving, be
sure to park before checking the stars.) You might
EXAMPLE 5.15 also note some radial streaks of light around the point
A full moon subtends an angle of 0.5 degrees at the image of the star. These radial streaks are due to
earth's surface. A real image of the full moon is diffraction from the structure of the crystalline lens
formed by a +1.00 D thin lens. What is the size of sutures. The twinkling of the stars is due to atmos-
the image? pheric turbulence. If you get a ride on the space
The secondary focal length of a +1.00D lens is shuttle, you can observe the stars without having to
+ 100 cm. From Eq. 5.14, look through the atmosphere.
|I| = ( + 100cm)tan(0.5);
|I| = ( + 100cm)(0.0087)
= 0.87 cm = 8.7 mm.
Often even a large distant object, such as the 5.15 Reversibility and Finding the Object
moon, subtends a small angle at the lens. When this
occurs, the small angle approximation for the tangent Many of the previous examples in this chapter deal
can be used so that Eq. 5.14 results in with finding the image once the object is given. Sup-
pose instead that the image is given and we are asked
|I|=/ 2 w (5.15) to find the object. One approach is to use the thin lens
where w is expressed in radians. When the small angle vergence equation, solve for the incident vergence U,
approximation is used for the full moon example, the and then find the object distance u.

FIGURE 5.29. Nodal ray to determine


the image size for a distant object.
78 Geometrie, Physical, and Visual Optics

Another approach is to use the principle of re- Vrev = P + U rev ,


versibility. Section 3.8 discussed the fact that two
V rev =+9.00 D +(-5.00 D),
conjugate points, A and B, are conjugate no matter
which way the light is passing through the system. or
When A is the object point, then B is the image point.
When the light direction is reversed, then B becomes Vrev = +4.00D.
the object point and A becomes the image point. If Then,
the image point B is known and the conjugate object
100 _ 100
point A is needed, then the direction of the light can
V +4'
be reversed, B considered as the object point, and A
found as the conjugate image. The following example or
uses the reversibility method.
u = +25 cm.
EXAMPLE 5.16 Note that for the reversed case in Figure 5.30b the
A +9.00 D thin lens in air has a real image located positive direction is to the left, so the real image at
20 cm behind it. Where is the conjugate object and +25 cm is to the left of the lens. Consequently, for
is it real or virtual? light traveling in the original direction (to the
Figure 5.30a shows an on-axis sketch of the right), the object is real and located 25 cm to the
given situation and Figure 5.30b shows the reversed left of the lens (Figure 5.30a).
situation. Remember that the dioptric power P
remains +9.00 D no matter which way the light
travels through the lens. From the above discussion, there are at least two
For the reversed situation, the positive direction approaches that one can use in finding the object once
is to the left and a real object is located 20 cm from the image is known. For visual optics purposes, there
the lens. Then are some conceptual advantages to using the re-
= -20 cm. versibility method as given in Example 5.16. In par-
So, ticular, the reversibility method facilitates the under-
100 cm/m standing of the optics of retinoscopy and ophthal-
Urev
(-20 cm) = -5.00D, moscopy.

original

1 ^ + 9.00D
-r ^C i ^
reversed

Vrev = ? Urev = - 2 0 cm

N FIGURE 5.30. Principal of reversibility, a. Original light


b) t traveling right, b. Reversed light traveling left.
Thin Lens Equations 79

Problems 21. A real object is located 40 cm in front of a


+6.50 D thin lens. The +6.50 D thin lens is lo-
For problems 1 through 10, find the image, specify cated 20 cm in front of a -10.00 D thin lens. What
whether it is real or virtual, and give the lateral is the vergence of light that exits the -10.00 D
magnification. Draw an extended ray diagram to con- lens? What is the total lateral magnification?
firm your calculations. Also, calculate the vergence of 22. A real object is located 34.0 cm in front of a lens
the wavefront 5 cm in front of the lens and the ver- of unknown power. A +7.00 D thin lens is located
gence of the wavefront 5 cm behind the lens. 10.0 cm behind the first lens. The second lens has
a real image located 29.0 cm from it. What is the
1. +6.00 D lens, real object 50 cm from the lens. dioptric power of the first lens?
2. +6.00 D lens, real object 25 cm from the lens. 23. A real object is located 13 cm in front of a lens of
3. +6.00 D lens, real object 10 cm from the lens. dioptric power P. A real image is located on a
4. +6.00 D lens, virtual object 10 cm from the lens. screen placed a distance x behind the lens. A
-7.00 D lens is placed between the object and the
5. +6.00 D lens, virtual object 30 cm from the lens. lens at a distance of 5 cm from the lens. Relative
6. -7.00 D lens, real object 40 cm from the lens. to the minus lens, where must the object be
7. -7.00 D lens, real object 10 cm from the lens. placed in order to again have an image on the
8. -7.00 D lens, virtual object 10 cm from the lens. screen? Remember that neither P nor x changes.
9. -7.00 D lens, virtual object 20 cm from the lens. 24. A thin lens of dioptric power P has a distant
object and a real image located on a screen that is
10. -7.00 D lens, virtual object 50 cm from the lens. at distance x behind the lens. A +5.00 D thin lens
11. What is the secondary focal length of a +14.00 D is then placed against the lens of power P. What
lens? Of a -8.00 D lens? thin lens located 14.0 cm in front of the +5.00 D
12. The primary focal length of a lens is -15.0 cm. lens will again result in a clear image on the
What is the dioptric power of the lens? screen? Remember that neither P nor x changes.
13. A 20 mm tall real object is located 28.0 cm from a 25. What single lens is equivalent to a +9.00 D thin
+5.50 D thin lens. What is the image size? lens combined with a -6.00 D thin lens?
14. Give the object distance that results in a lateral 26. For a distant tree, a thin lens held 19.0 cm in front
magnification of -1.0 for a +6.50 D lens. For a of a wall results in an inverted image on the wall.
-4.50 D lens. When a second thin lens of unknown power is
combined with the first, the real image is on the
15. Give the object distance that results in a lateral
wall only when the combination is held 69.0 cm
magnification of +1.0 for a +2.50 D lens. For a
from the wall. What is the dioptric power of the
-8.00 D lens.
second lens?
16. A +5.50 D lens has a real image located 35 cm
27. A distant building subtends an angle of 90 minutes
from the lens. Is the conjugate object real or
at a +2.00 D thin lens. What is the size of the
virtual? What is the object distance?
image (in mm)?
17. A +5.00 D lens has a real image located 12 cm
28. A camera has a single thin lens, and a person
from the lens. Is the conjugate object real or
stands 3.6 m from the camera lens. The person is
virtual? What is the object distance?
250 cm tall and her image is 35.2 mm tall. What is
18. A -7.50 D lens has a virtual image located 9 cm the dioptric power of the camera lens? What is the
from the lens. Is the conjugate object real or secondary focal length of the lens?
virtual? What is the object distance?
29. Section 5.9 contains a distorted circle representa-
19. A +8.20 D lens forms a real image that is inverted tion of the object movement/image movement
and 4 times the size of the object. What is the relation for a plus thin lens. Construct the object
image distance? What is the object distance? movement/image movement circle for a minus
20. A real object is 20 cm from a thin lens. The thin lens.
conjugate image is erect and 3 times larger than
the object. What is the lens power?
CHAPTER SIX

Thin Lens Eye


Models

6.1 Emmtropes The image distance is 16.67 mm and the screen should
be placed that distance behind the +60.00 D lens in
The simplest optical model of the eye is a thin lens and order to simulate Amy's eye (Figure 6.1a).
screen model. The thin lens represents the cornea and A distant object is conjugate to the secondary focal
the crystalline lens while the screen represents the point of a lens. An alternate solution for the emme-
retina. Air is assumed to be between the lens and the tropic lens-screen distance is to recognize that the
screen. A typical dioptric power for the lens represent- secondary focal point of the unaccommodated emme-
ing the unaccommodated human eye is +60.00 D. tropic eye coincides with the retina. Thus, the lens-
An emmtrope is a person who can clearly see a screen distance is just equal to the secondary focal
distant object without accommodating. The light that length of the lens. From Eq. 5.4, the secondary focal
reaches the eye from the distant object consists of length is
plane waves. The unaccommodated emmetropic eye
converges the plane waves and forms a clear image on f= i
the retina. In the thin lens and screen model of the 32 p>
unaccommodated emmetropic eye, the screen, which or
represents the retina, is placed at the image point 1000mm/m_ , 1 f i f i 7 m m
conjugate to optical infinity. /2
(+60 D) -+1667mm
Consider Amy who is an emmtrope and has an
unaccommodated eye with +60.00 D power. What Brian is an emmtrope whose eye is simulated with
should the lens-screen separation be in a model a lens-screen separation of 18.00 mm. What is the
simulating Amy's eye? One method for answering this power of the lens in the model for Brian's unaccom-
question is to find the image position for incoming modated eye? In this example, the image distance for
plane waves. From Eq. 5.1, a distant object is 18.00 mm which is thus the sec-
ondary focal length for the eye. Therefore,
V = P + U,
V = + 6 0 D + 0,
V=+60D,
>-k-
1000 mm/m
and (18 mm) '
P = +55.56 D.
_ 1000 mm/m
V Brian's unaccommodated eye has a dioptric power less
~ (+60 D) ' than the typical value of +60.00 D but is still emme-
i; = +16.67 mm. tropic because the eye is longer than the 16.67 mm
81
82 Geometrie, Physical, and Visual Optics

+ 60.00D
*

FIGURE 6.1. a. Plane waves from a


distant object incident on a thin lens and
screen model of an emmetropic eye. b.
The reversed situation when the retina is
a) Y b) the object, plane waves emerge from the
16.67 mm emmetropic eye.

lens-screen separation corresponding to the +60.00 D the divergence of the incident light increases and the
emmetropic eye. image moves back towards the retina. Eventually, the
We usually think in terms of light entering the eye. image is formed on the retina and the near object
However, consider the situation when an examiner appears clear to the unaccommodated myope. In this
uses an ophthalmoscope or a retinoscope to examine case the near object is located at the far point of the
the eye. The light is initially directed into the eye to myope's eye. It is a characteristic of a myopic eye that
illuminate the retina (just as you turn on the room its far point is a real point located a finite distance in
lights to illuminate a wall). The retina then acts as a front of the eye.
secondary source with the diffusely reflected light Cindy is a myope whose unaccommodated eye is
diverging away from it and emerging from the eye. simulated by a +58.00 D lens and a 20 mm lens-screen
Figure 6.1a shows plane waves from a distant ob- distance. One way to check that the model is simulat-
ject point entering an unaccommodated emmetropic ing a myopic eye is to determine if the light from a
eye model. The image point is on the retina. Accord- distant object point is imaged in front of the retina.
ing to the principle of reversibility, light diverging For incident plane waves, the image is at the
from a point on the retina retraces the same path back secondary focal plane. Then,
out of the eye. (In Figure 6.1a, simply reverse the
arrow direction on each ray.) Therefore, the light f2 = P '
emerging from the unaccommodated emmetropic eye
consists of plane waves with the image point at optical 1000
infinity (Figure 6.1b).
f2 =
+58'
The far point of an eye is the point conjugate to the / 2 =+17.24 mm.
axial retinal point for the unaccommodated eye. For
the unaccommodated eye, an object point at the far The lens-screen distance is 20 mm. Consequently, the
point results in an image point on the retina (light is clear image is formed in front of the screen represent-
incident on the eye), and an object point on the retina ing the retina (Figure 6.2a). While this calculation
results in an image point at the far point (light is shows that Cindy's eye is myopic, it does not de-
emerging from the eye). The far point is also known termine the amount of myopia.
by its Latin name, punctum remotum. One way to determine the amount of myopia is to
The far point of an emmtrope is at optical infinity. take the retina as the object and consider the light that
In the following sections, two types of ametropes, emerges from the eye, as occurs in ophthalmoscopy
myopes and hyperopes, are discussed. The far point and retinoscopy. In the reversed case, the object
for ametropes is not at optical infinity. distance u for the retina is - 2 0 mm (Figure 6.2b).
Then,
u = - 2 0 mm,
1000
6.2 Myopes U= = -50.00 D.
(-20)
Myopes are nearsighted. When a myope views a dis- Since
tant object, the eye converges the light to form a clear V = P + U,
image in front of the retina (i.e., in the vitreous
humor) resulting in a blurred image on the retina. V = + 5 8 + (-50),
When the object is brought closer to the myopic eye, V=+8.00D,
Thin Lens Eye Models 83

P=+58.00D

FIGURE 6.2. a. Rays from a distant point


incident on a myopic eye. b. When the retina
20.0 mm is the object, the waves emerging from a
myopic eye are converging.

and object is at the far point of the eye, Pc be the dioptric


power of the contact lens, U c the vergence of the light
100 = +12.5 cm.
v= -=- incident on the contact lens, and Vc the vergence of
the light exiting the contact lens. From Eq. 5.1,
The light emerging from Cindy's eye has a converg-
PC=VC-UC.
ence of +8.00 D. Note that plane waves emerge from
an emmetropic eye. Cindy's myopic eye needs 8.00 D Then, since
less converging power in order to be emmetropic. The U c = 0,
8.00 D is a measure of Cindy's refractive error. Pc=V c ,
The converging light coming out of Cindy's eye
forms a real image at 12.5 cm in front of the eye Since the light leaving the contact lens is immediately
(Figure 6.2b). The real image point at 12.5 cm is incident on the cornea, there is no vergence change
conjugate to the retina and, consequently, is the far and
point of Cindy's eye. By reversibility, an object placed V = U fp
at the far point results in a clear image on the retina.
(In Figure 6.2b, the position of the rays would be From the above two equations,
unchanged, but the arrows on the rays would be Pc = U fp . (6.1)
reversed and point to the right.) For an object point at
12.5 cm in front of the eye, the light incident on the In the case of the myope Cindy, Eq. 6.1 gives the
eye would have a divergence of -8.00 D. This -8.00 D contact lens power. Remember that Eq. 6.1
-8.00 D, in effect, offsets or neutralizes the 8.00 D of is general and applies to hyperopes as well as to
excess converging power in the eye, resulting in a clear myopes.
image on the retina. Dan is a myope whose unaccommodated eye has a
A contact lens correction is a correcting lens that power of +62.00 D. The far point of Dan's eye is
sits on the cornea. In the thin lens and screen model, located 50 cm in front of the eye. In the thin lens and
the contact correction is simulated by a thin lens screen model, what is the lens-screen distance? What
placed against the front of the thin lens representing contact lens correction does Dan require?
the eye. What contact lens would correct Cindy for When an object is placed at Dan's far point, a clear
distant vision? image will be formed on the retina. In this case,
To get a clear image on the retina of her un- wfp = - 5 0 cm,
accommodated eye, Cindy must have a vergence of
-8.00 D incident on the cornea. Plane waves come and
from a distant object, so the contact lens must be a , = ^ 2.00D.
lens that produces a vergence of -8.00 D from the ** (-50)
incident plane waves. From the thin lens vergence Since
equation, the contact lens must have a power of V = P + U fp ,
-8.00 D.
In the above example, the dioptric power of a then
distance vision contact lens was equal to the vergence V=+62 + (-2)=+60.00D,
of the light incident on the cornea when the object was
at the far point of the eye. This is true in general, and and
is shown algebraically as follows. Let Ufp be the 1000
v = 7T- = +16.67 mm.
vergence of the light incident on the cornea when the ou
84 Geometrie, Physical, and Visual Optics

2D 8D

12.5cm-

FIGURE 6.3. Far point location for a 2 D


50.0 cm and an 8 D myope.

The lens-screen separation for Dan's eye is 16.67 mm. The concept of the far point, and hence U fp , is
This happens to be the same lens-screen separation of particularly useful. Once the far point is known, then
a typical emmtrope. Dan's +62.00 D power is 2.00 D the correcting lens for the eye can be determined even
more than the typical emmtrope, and the Ufp of if the internal dimensions of the eye are not known.
-2.00 D simply neutralizes the +2.00 D error.
Dan's distance vision contact correction must take
incoming plane waves and convert them into light of
-2.00 D vergence. Hence, Dan's contact lens power 6.3 Hyperopes
Pc is -2.00 D, which is the same result that Eq. 6.1
gives. An unaccommodated hyperopic eye is too short to
The amount of myopia is directly quantified by the clearly image a distant object on the retina (Figure
absolute value of U fp , which was 8.00 D and 2.00 D, 6.4a). When an unaccommodated hyperope looks at a
respectively, for Cindy and Dan. An alternative con- distant object, the eye converges the light, but the
ceptual quantification of the amount of myopia is in converging light runs into the retina before the image
terms of the far point location. The closer the far point is formed. Hence, the retinal image is blurred. To get
is to optical infinity, the less the degree of myopia and a clear image on the retina of an unaccommodated
the closer the eye is to being emmetropic. The closer hyperope, the light incident on the eye must be con-
the far point is to the eye, the greater the degree of verging. This means that the far point of a hyperope's
myopia. Dan, the 2.00 D myope, had a far point 50 cm eye is virtual and located a finite distance behind the
in front of the eye, while Cindy, the 8.00 D myope, eye (Figure 6.4b).
had a far point 12.5 cm in front of the eye. Note that Fay has an unaccommodated eye that is simulated
in all cases the far point for a myope is a real point as by a +60.00 D thin lens and a screen located 16 mm
opposed to a virtual point (Figure 6.3). behind the lens. Since the typical +60.00 D eye has a

FIGURE 6.4. a. Rays from a distant point


incident on an unaccommodated hyperopic
eye. b. Converging incident light (virtual far
point) is needed to get an image on the hy-
perope's retina, c. When the retina is the
object, diverging light leaves the hyperopic eye
resulting in a virtual image at the far point.
Thin Lens Eye Models 85

length of 16.67 mm, Fay's eye is clearly too short and lens-screen distance of 16.67 mm. Where is George's
is therefore hyperopic. What is the degree of hy- far point? What distance vision contact lens correction
peropia and where is the far point for Fay's eye? does George require?
Again, as in retinoscopy, we will consider the light A typical emmetropic model eye has a dioptric
emerging from the eye when the retina serves as the power of +60.00 D and a lens-screen distance of
object (Figure 6.4c). Then 16.67 mm. George's lens-screen distance is 16.67 mm,
u = 16 mm, but George only has +54.00 D of power. Hence,
George is lacking +6.00 D of converging power. In
and other words, George has 6.00 D of hyperopia.
1000
U = (-16) 62.50 D. Alternatively, we can determine the amount of
hyperopia by taking the retina as the object and
Since considering the light emerging from the eye. Thus,
V = P + U,
u = 16.67 mm,
V = 60 + (-62.50), and
V=-2.50D,
and
100 Since
v = , ^ = -40.0 cm.
(-2.50) V = P + U,
The light emerging from Fay's eye is diverging and has V = + 5 4 + (-60),
a vergence of -2.50 D. The image point is virtual, V=-6D,
located 40 cm behind the +60.00 D lens, and is the far
point for Fay's eye. then
Plane waves emerge from an emmetropic eye. Fay 100
v = (-6) = 16.67 cm.
has diverging light of -2.50 D emerging from her eye.
Fay's eye is lacking 2.50 D of converging power. This The light leaving George's eye is diverging and has a
2.50 D is the degree of hyperopia. vergence of -6.00 D. Plane waves emerge from an
Since Fay's eye is lacking +2.50 D, light entering emmetropic eye, so the -6.00 D vergence confirms
her eye must be converging and have a vergence of that George is lacking 6.00D of power (i.e., George
+2.50 D in order to produce a clear retinal image. In has 6.00 D of hyperopia). The image conjugate to the
other words, to obtain a clear retinal image, the object retina is virtual and located 16.67 cm behind the
for the eye must be virtual and located 40 cm behind +54.00 D lens. This is the far point for George's eye.
the eye at the far point. Then wfp = +40cm and To get a clear retinal image, the light incident on
Ufp = +2.50D. (In Figure 6.4c, the rays would be George's eye must have a vergence Ufp equal to
unchanged except that the arrows would be reversed +6.00 D. Consequently, the contact lens correction
and point to the right.) necessary for George to clearly see a distant object
A contact lens correction that enables Fay to see a without accommodating is +6.00 D.
distant object without accommodating must take inci- Fay was a 2.50 D hyperope and had a virtual far
dent plane waves and produce exiting light of vergence point located 40 cm behind her eye, while George was
+2.50 D. A +2.50 D contact lens does this. We can a 6.00 D hyperope and had a virtual far point located
also find the power of the contact lens from Eq. 6.1: 16.67 cm behind his eye. Note that the closer the far
Pc = Ufp = +2.50D. point is to optical infinity, the lesser the degree of
hyperopia and the closer the eye is to being emme-
George is a hyperope whose unaccommodated eye tropic. The closer the far point is to the eye, the
is simulated by a +54.00 D thin lens and a typical greater the degree of hyperopia (Figure 6.5).

6D 2.5D

16.67 cm

40.0 cm FIGURE 6.5. Far point location for a 2.5 D


and a 6.0 D hyperope.
86 Geometrie, Physical, and Visual Optics

6.4 Spectacle Lens Corrections leaving the lens and traveling to the cornea must be
spherical and centered on the far point. The wavefront
While contact lens corrections sit on the cornea, spec- leaving the spectacle lens is 13 mm or 1.3 cm closer to
tacle lens corrections are mounted in a frame and sit a the far point than the wavefront incident on the
certain distance in front of the cornea. The distance cornea. Therefore, the radius of curvature of the
from the back of the spectacle lens to the cornea is wavefront leaving the spectacle lens is 10 cm minus
called the vertex distance. Because of the vertex dis- 1.3 cm, or 8.7 cm (Figure 6.6b). The radius is negative
tance, the dioptric power of a spectacle correction can since the wavefront is diverging, and the vergence
differ from the dioptric power of the contact lens leaving the spectacle lens is
correction. The reason for this difference is that the
wavefronts that leave the back of the spectacle lens 100
V= =-11.50 D.
gain or lose curvature as they cross the gap to the (-8.7 cm)
cornea. Since plane waves are incident on the spectacle lens
Helen has a far point that is real and located 10 cm for distance vision (U = 0), the dioptric power of the
in front of her cornea. What contact lens correction lens is:
and what spectacle lens correction does Helen need
V = P + U,
for distance vision? Assume a vertex distance of
13 mm. P=V-U,
First of all, since Helen's far point is in front of her
eye, Helen is a myope and needs diverging light P=-11.50-0,
incident on her cornea. Secondly, Helen's far point is
or
relatively close to her eye, so Helen is a fairly high
myope. When an object is placed at Helen's far point, P = - 1 1 . 5 0 D.
wfp = 10.0 cm, Helen's distance vision spectacle correction worn at
a vertex distance of 13 mm is -11.50 D, while Helen's
and contact lens correction is -10.00 D. Note that the
U f p = -10.00 D. spectacle correction and contact correction were de-
termined by knowing the far point location of Helen's
Helen needs a vergence of -10.00 D incident on her eye and without knowing the dioptric power or length
cornea. From Eq. 6.1, Helen's distance vision contact of the eye.
lens correction is then -10.00 D. Ian wears a +5.00 D distance vision contact lens
The distance vision spectacle lens correction must correction. What distance vision spectacle lens correc-
take incoming plane waves and diverge them in such a tion worn at a vertex distance of 15 mm does Ian
manner that the vergence of the light reaching the need?
cornea is -10.00 D (Figure 6.6a). Since the diverging Ian is hyperopic, and from Eq. 6.1,
wavefronts lose curvature as they travel across the gap
from the spectacle lens to the cornea, the vergence Ufp = Pc = +5.00D,
leaving the spectacle lens must be more minus than resulting in
-10.00 D.
For the proper correction, the wavefront incident 100
= +20.0 cm.
on the cornea must be spherical and centered on the
far point of the eye that is 10 cm in front of the Ian's far point is virtual and located 20 cm behind the
cornea. In fact, all of the wavefronts in the bundle cornea.

fp

~~v
8.7 cm
8
1.3 cm
FIGURE 6.6. a. Plane waves incident
on a spectacle correction for a myopic
eye. The vergence of the wavefront inci-
dent on the cornea must be -10.00 D.
b) y b. Geometry to get vergence of the
10.0 cm wavefront leaving the spectacle lens.
Thin Lens Eye Models 87

uice e
1.5 cm
20 cm

FIGURE 6.7. a. Plane waves incident on a


spectacle correction for a hyperopic eye. The
vergence of the wavefront incident on the cor-
nea must be +5.00 D. b. Geometry to get
v vergence of the wavefront leaving the spectacle
21.5 cm
a) b) lens.

The spectacle lens correction must be a converging 100


lens. However, the converging wavefronts leaving the = - 1 6 . 0 cm.
(-6.25)
back of the spectacle lens gain curvature as they cross The point 16 cm in front of the spectacle lens is the far
the gap between the lens and the cornea. Therefore,
point of Jeremy's eye. Each of the wavefronts crossing
the vergence leaving the spectacle lens must be less
the gap is spherical and centered on the far point. The
positive than the incident vergence of +5.00 D that is
cornea is 1.2 cm further away, so the wavefront inci-
required on the cornea (Figure 6.7a).
dent on the cornea has a radius of magnitude 17.2 cm.
All the converging wavefronts in the bundle be- The radius is negative, and the vergence of the wave-
tween the spectacle lens and the cornea are centered front incident on the cornea is
on the far point. The radius of curvature of the
wavefront incident on the cornea is 20 cm. The radius 100
U
rp = = -5.81D.
of curvature of the wavefront leaving the spectacle (-17.2)
lens is 1.5 cm longer, or 21.5 cm (Figure 6.7b). The Alternatively, one could just use the magnitude of the
vergence of the wavefront leaving the lens is then radius and calculate the absolute value of U fp , or
100 100
V= = +4.65D. |u fp | = = 5.81D.
+21.5 17.2
Since the incident waves from a distant object are Then since we know the light is diverging, we simply
plane waves, the dioptric power of the spectacle lens assign the minus sign to U fp , or
must then be +4.65 D. (Note that as expected the Ufp = - 5 . 8 1 D .
+4.65 D is positive and less than +5.00 D.)
It is important to keep in mind whether the wave- An aphakic eye has no crystalline lens and is thus
fronts traveling across the gap between a spectacle lacking plus power. The far point of an aphake is
lens and the cornea are gaining or losing curvature. similar to that of a high hyperope: virtual and a finite
Simple sketches like Figures 6.6a and 6.7a quickly distance behind the eye. Consider Karen, an aphake,
supply this information. who needs a distance vision spectacle lens correction
The amount of vergence change across the gap of +14.25 D at a vertex distance of 15 mm. Karen
depends on the power of the lens involved. High chooses a frame that gives a vertex distance of 12 mm.
powered lenses, either plus or minus, produce large What lens power does Karen need for the new frame?
vergence changes, while low powered lenses produce For incident plane waves, light of vergence
small vergence changes. Changes smaller than 0.25 D + 14.25 D would leave the spectacle lens, which is
are usually considered clinically insignificant. Changes 15 mm from the cornea. Since the wavefronts gain
of 0.25 D or greater are considered clinically sig- curvature as they move toward the cornea, the ver-
nificant. gence at a distance of 12 mm from the cornea will be
Jeremy has a distance vision spectacle correction of greater than +14.25 D (Figure 6.8a).
-6.25 D. The vertex distance is 12 mm. What is The center of curvature of the wavefront leaving
Jeremy's distance vision contact lens power? the spectacle lens is at the far point of Karen's eye.
In Jeremy's case, the diverging wavefronts leaving The radius of curvature, which is also the image
the spectacle lens become flatter as they travel to the distance for the lens, is
cornea (Figure 6.6a). Therefore, the contact lens 100
power will be less minus than -6.25 D. v = ( + 14.25) = +7.02 cm.
The center of curvature of the wavefront leaving
the spectacle lens is The vergence at 12mm from the cornea (i.e., 3 mm
88 Geometrie, Physical, and Visual Optics

O fp

FIGURE 6.8. a. Downstream vergence


changes involved in a vertex distance change for
v an aphakic correction, b. Geometry to calculate
b) 0.3 cm 6.72 cm resulting differences in spectacle correction
v power.
7.02 cm

behind the spectacle lens, Figure 6.8b) is Again, this statement assumes that plane waves are
100 incident on the lens, so this also is simply another way
V = (7.02-0.3)' of stating that diverging wavefronts lose curvature as
they travel from the spectacle lens to the eye.
100 The gain in plus effectivity or loss in minus effec-
v = (6.72) ' tivity statements may not accurately describe what
happens when diverging or converging light is incident
V = +14.88 D.
on a spectacle lens because there the vergence incident
For distance vision, a spectacle lens placed 12 mm on the lens may be changing in the opposite manner as
from Karen's eye must take incoming plane waves and the effectivity changes.
change them into converging waves with a vergence of What is actually changing in the effectivity cases is
H-14.88 D. From P = V - U, the power of the spectacle the vergence of the wavefronts as they travel from one
lens is then +14.88 D. Note that the +14.88 D lens at position to the other. Therefore, the name vergence
a vertex distance of 12 mm is a change of +0.63 D effectivity is more appropriate than lens effectivity.
from the +14.25 D lens at a vertex distance of 15 mm, However, the lens effectivity terminology is useful for
and this amount is clinically significant. those who are not familiar with vergence.

6.5 Lens Effectivity Terminology 6.6 Ocular Accommodation

In the previous sections we saw that the power of a In dealing with distance vision corrections, we con-
distance spectacle correction can differ from that of a sider only unaccommodated eyes. Let us now take
contact correction because of the change in vergence accommodation into account.
of the wavefronts as they move from the spectacle The crystalline lens in the eye adds converging
plane to the eye. Historically, this phenomenon has power when the eye accommodates. In our thin lens
been described under the name of lens effectivity. and screen model, we can simulate this change by
Instead of directly mentioning the change in vergence, adding power to the thin lens. In the model, let Pu be
people would refer to the effect of the spectacle lens at the power of the unaccommodated eye, and Pa be the
the cornea. power of the accommodated eye. The amount of
Consider a vertex distance of 12 mm. Then, the ocular accommodation A 0 is the difference in the two
effect of a +10.00 D spectacle lens at the eye is values, or
+ 11.36 D. This statement means that for plane waves
incident on the +10.00 D lens, the vergence of the A0 = P a - P u . (6.2)
light leaving the lens is +10.00 D, but the vergence of Let ve be the distance between the thin lens and the
the light at the eye is +11.36 D. screen. Since the eye does not change length when it
Plus lenses are said to gain plus effectivity as they accommodates, ve is a constant. In order to have a
move away from the eye. This statement assumes that clear retinal image, the vergence Ve of the light leav-
plane waves are incident on the lens, so this is simply ing the thin lens representing the eye must equal the
another way of stating that converging wavefronts gain reciprocal of ve, or
curvature as they travel from the spectacle lens to the
eye. (6.3)
Minus lenses are said to lose minus effectivity, or to
gain plus effectivity, as they move away from the eye. When the object is at the far point of the eye and
Thin Lens Eye Models 89

the eye is unaccommodated, then the retinal image is A0 = U f p - U x ,


clear and
A 0 = - 2 . 5 0 D - ( - 4 . 0 0 D),
ve = P + ufp (6.4) Ao=+1.50D.
When the object is moved inside the far point to a When uncorrected, Larry has an ocular accommoda-
point x, i.e., closer to the eye, then the eye accommo- tive demand of +1.50D when viewing the book at
dates to maintain a clear retinal image, and 25 cm.
V e = P a + Ux, (6.5) Suppose Larry holds the book only 10 cm in front
of his cornea. How much accommodation is needed
where Ux is the vergence incident on the cornea when then? In this case,
the object is at point x. Figure 6.9 shows the case for a
myope. 100
U = = -10.00 D.
Equation 6.5 can be subtracted from Eq. 6.4 to (-10)
give From Eq. 6.7,
o = Pu + u f p - p a - u x , A0 = U f p - U x ,
or
Pu = U f p - U x . (6.6) and

Then, from Eqs. 6.2 and 6.6, A o = - 2 . 5 0 D - (-10.00 D),


Ao=+7.50D.
A0 = U f p - U x . (6.7)
Larry's ocular accommodative demand for the object
Equation 6.7 states that the amount A 0 of ocular at 10 cm is +7.50 D.
accommodation needed to maintain a clear retinal Suppose Larry looks at an object 1 m from his eye.
image must just offset (or neutralize) the increase in In this case, the image inside Larry's eye would be
vergence incident on the eye when the object is moved formed in front of his retina resulting in a blurred
in from the far point to the point x. In real eyes, the retinal image. If Larry accommodates, the image is
amount of accommodation that actually occurs (the pulled even further in front of the retina, thus increas-
accommodative response) may not match the accom- ing the amount of retinal blur. We say that the object
modation needed. To distinguish the two, the amount appears blurred because it is outside Larry's far point.
of accommodation needed is called the ocular accom- Suppose we blindly went ahead and calculated the
modative demand. Equation 6.7 is the fundamental ocular accommodative demand anyway. For an object
working equation for computing the ocular accom- at 1 m,
modative demand.
Larry is an uncorrected myope with a far point U =-1.00D.
40 cm in front of his eye. So, Then,
100 A0 = U f p - U x ,
Ufp
(-40) = -2.50D.
A0=-2.50-(-1.00),
Suppose Larry looks at a book 25 cm in front of his Ao=-1.50D.
eye. How much accommodation is needed for Larry to
have a clear retinal image? The negative value indicates a need for less plus power
When the book is 25 cm in front of Larry's eye, the (in this case, 1.50 D less plus). Since we assumed that
vergence incident on Larry's eye when he looks at the an unaccommodated eye has the least plus power and
book is that accommodation adds plus power, then a negative
value for A 0 indicates that the object is outside the far
U point. (There is physiological evidence for a positive
=F!)=-4-OOD resting level of accommodation, but that does not
From Eq. 6.7, the ocular accommodative demand is, change the basic considerations here.)

fp

FIGURE 6.9. a. Wavefronts from the far


point incident on a myope's eye. b. Wavefronts
b) from a point x inside the far point.
90 Geometrie, Physical, and Visual Optics

6.7 Near Point and Range of Clear Now consider Marcie who wears a -5.00 D dis-
Vision: Myope tance vision contact correction and has a 3.00 D am-
plitude of accommodation. What is Marcie's uncorrec-
As a person moves an object closer and closer to his ted range of clear vision?
eye, more and more accommodation is used. Eventu- From Eq. 6.1,
ally, the maximum amount is reached. The maximum Ufp = - 5 . 0 0 D ,
amount of accommodation is called the amplitude of
accommodation (A m ). and
The near point of an eye is the point conjugate to
u{p= -20.0 cm.
the axial retinal point of the maximally accommodated
eye. The near point is also known by its Latin name of From Eq. 6.7,
punctum proximum. The near point can be found as a A0 = U f p - U x ,
special case of Eq. 6.7 by setting A 0 equal to the
amplitude of accommodation A m . The points that or
successively become conjugate to the retina as a per-
Ux = U f p - A 0
son increases his or her accommodation constitute the
person's range of clear vision. Ux=-5.00-3.00,
Suppose Larry, the myope in the preceeding sec- U =-8.00D.
tion, has an amplitude of accommodation of 9.00 D.
Where is Larry's near point? Finally,
We can solve Eq. 6.7 for U x : 100
M
np - "x - (_ 8 oo) '
A 0 = Ufp Uv
unp= -12.5 cm.
Therefore,
Ux = U f p - A 0 . Marcie's uncorrected range of clear vision extends
from her far point at 20 cm to her near point at
Larry's far point was 40 cm, and 12.5 cm.
Ufp = -2.50D. The amplitude of accommodation gradually di-
minishes with age. At the age of 8 years, the am-
When A is set equal to the 9.00 D amplitude of plitude of accommodation is approximately 14.00 D,
accommodation, the point x is the near point. at 20 years it has fallen to 11.00 D, at 30 years it is
Ux = - 2 . 5 0 - 9 . 0 0 , approximately 9.00 D, and at the age of 50 years it is
less than 2.00 D. Because of the decline of the am-
or plitude of accommodation with age, a person's near
U =-11.50 D. point steadily recedes from the eye and approaches
the far point.
Then
" v =

u. 6.8 Range of Clear Vision: Emmtrope


100
Accommodation is particularly easy to calculate for
" v =
(-11.50)'
M
np="x = -8.70cm. the emmtrope. The general equation is again Eq. 6.7:
A 0 = Ufp LL
Larry's range of clear vision extends from his far point
at 40 cm to his near point at 8.7 cm (Figure 6.10). However, for an emmtrope, Ufp is zero.

range of clear vision

FIGURE 6.10. Range of clear vision for a myope.


Thin Lens Eye Models 91

Norman is an emmtrope who looks at a poster adds plus converging power, it can be used to over-
50 cm from his eye. What is Norman's ocular accom- come some or all of the hyperopia. Since the un-
modative demand? corrected hyperope uses accommodation to see distant
Here, objects, he or she may not have enough left for near
objects. Furthermore, the ciliary muscle, like any
Uv = J ^ = -2.00D, muscle in the body, is subject to fatigue under condi-
(-50) tions of exertion for long periods of time. Therefore,
and from Eq. 6.7, while a high hyperope may be able to see a distant
Ao = 0 - ( - 2 . 0 0 ) . object clearly by accommodating, he may not be able
to sustain the accommodation.
or An uncorrected hyperope has a virtual far point.
A=+2.00D. As the hyperope accommodates, the virtual point
conjugate to the retina moves away from the eye and
If Norman moves the poster in to a point 25 cm in toward optical infinity. If a hyperope has enough
front of his eye, then accommodation, an object at optical infinity can be
100 imaged clearly on the retina. Whatever accommoda-
U = = -4.00D. tion is left can then be used to see real objects that are
(-25)
closer than optical infinity.
and Consider Paul, a hyperope who needs a +8.00 D
A o = 0 - (-4.00) = +4.00D. contact lens. From Eq. 6.1,
Ufp = Pc = +8.00D,
For an object 20 cm in front of his eye, Norman has an
ocular accommodative demand of +5.00 D, and for an and
object 10 cm in front of his eye he has an ocular 100
accommodative demand of +10.00 D. ufnfp = (+8.00) = +12.5 cm.
Octavia is an emmtrope with an amplitude of
accommodation of 7.00 D. What is Octavia's range of Paul's far point is virtual and located 12.5 cm behind
clear vision? his cornea.
To find Octavia's near point, use Eq. 6.7, set A 0 When Paul accommodates +3.00 D, what point is
equal to the amplitude of accommodation, and solve conjugate to his retina? From Eq. 6.7,
for U. A0 = U f p - U x ,
Ux = U f p - A 0 .
or
But Ufp is zero, therefore, 3 = 8-Ux,
Ux = 0 - (7.00) = - 7 . 0 0 D ,
and
or Uv = 8 - 3 = + 5 . 0 0 D .
100
= -14.3 cm. Then
(-7.00)
Octavia's near point is 14.3 cm in front of her eye. 100
+20.0 cm.
Since Octavia is an emmtrope, her far point is at (+5.00)
optical infinity. Octavia's range of clear vision extends When Paul accommodates 3.00 D, the point conjugate
from optical infinity to 14.3 cm in front of her eye. to his retina remains virtual but moves from 12.5 cm to
As Octavia ages, her near point will move out away 20.0 cm. An uncorrected 5.00 D hyperope has a far
from her eye toward her far point at optical infinity. point at 20.0 cm, so in a sense, by using 3.00 D of
Eventually, she will have difficulty seeing near objects, accommodation Paul can now function as a 5.00 D
a signal that presbyopia has arrived. hyperope instead of an 8.00 D hyperope (Figure 6.11).
What is Paul's ocular accommodative demand to
see an object at optical infinity? From Eq. 6.7,

6.9 Range of Clear Vision: Hyperope A0 = U f p - U x .


For the object at optical infinity
A hyperope does not have enough converging power
in the unaccommodated eye. Since accommodation u x = o,
92 Geometrie, Physical, and Visual Optics

fp
toOCZX- - to O O
12.5 cm| FIGURE 6.11. Point x is conjugate to
the retina for an 8.00 D hyperope who is
20.0 cm accommodating 3.00 D.

and since 6.10 Range of Clear Vision: Summary


Ufp=+8.00D,
The circular representation can be used to represent
then the range of clear vision for the myope and emm-
trope as well as for the hyperope. Figure 6.13 shows
A 0 = + 8 - 0 = + 8 . 0 0 D.
four cases. In each case, the eye is looking left. The
When uncorrected, Paul needs to accommodate points on the left half of the circle represent optically
+8.00 D to clearly see an object at optical infinity. In real points, while the points on the right half of the
effect, Paul neutralizes the 8.00 D of hyperopia by circle represent virtual points. The far point of a
accommodating 8.00 D. myope is a real point located a finite distance in front
What is Paul's ocular accommodative demand to of the eye, while the far point of a hyperope is a
read a book held 25 cm in front of his eye? Now virtual point located a finite distance behind the eye.
The far point of an emmtrope is at optical infinity.
n = ^ 4 . 0 0 D . The first case is that of a myope. The far point is
(-25)
real and a finite distance in front of the eye. The near
From Eq. 6.7, point is also real and closer to the eye than the far
A0 = U f p - U x , point. Distance objects are not included in the range
of clear vision.
A0=+8-(-4),
The second case is that of an emmtrope. The far
or point is at optical infinity, and the near point is real
and closer to the eye than the far point.
A =+12.00 D.
The third case is that of a hyperope with enough
Paul's ocular accommodative demand to read the accommodation to overcome the hyperopia and see
book 25 cm in front of his eye is +12.00 D. some near objects. The far point is virtual and a finite
Note that Paul needed 8.00 D to see an object at distance behind the eye, while the near point is real
optical infinity. In order to get from optical infinity to and a finite distance in front of the eye. The range of
a book that is 25 cm in front of his eye, Paul requires clear vision includes optical infinity.
an additional 4.00 D of accommodation. The 8.00 D The fourth case is that of a hyperope who does not
plus 4.00 D provides the total ocular accommodative have enough accommodation to overcome the hy-
demand of 12.00 D that Paul needs to read the book. peropia. The far point is virtual and located a finite
Figure 6.12 provides a circular representation of Paul's distance behind the eye, and the near point is also
accommodation. virtual and located a finite distance behind the eye.

* X * CDO X *-
< 1 <

-x*-
FIGURE 6.12. Range of clear vision as an
25.0 cm 12.5 cm 8.00 D hyperope accommodates 12.00 D.
Thin Lens Eye Models 93

-CCD

FIGURE 6.13. Ranges of clear vision.


a. Myope, b. Emmtrope, c. Low hy-
perope. d. High hyperope.

The near point is closer to optical infinity than the far trope puts on a -8.00 D contact lens and looks at a
point. No real object points are imaged on the retina. distant object, the object will appear blurred (and
In each of the cases, the range of clear vision maybe not even visible). However, if the emmtrope
extends counterclockwise from the far point to the accommodates 8.00 D, the distant object will appear
near point. The region inside, or counterclockwise clear. In other words, in the thin lens eye model the
from, the near point is the region where objects are 8.00 D of accommodation neutralizes the -8.00 D
not clearly imaged on the retina because the person contact lens.
does not have enough accommodation available. The Suppose the emmtrope looks through a -8.00 D
region outside, or clockwise from, the far point is the spectacle lens at the distant object. Again the emm-
region where objects are not clearly imaged because trope can accommodate to neutralize the spectacle
the unaccommodated eye already has too much con- lens and see the distant object clearly. In this case, we
verging power, and the clear image is formed in front tend to say the -8.00 D spectacle lens causes 8.00 D
of the retina, i.e., in the vitreous humor. of accommodation, but, in fact, the actual ocular
Suppose each of the eyes represented in Figure accommodative demand is not 8.00 D. For the distant
6.13 has a 4.00 D amplitude of accommodation and object, the vergence of the light leaving the spectacle
the respective refractive states are: 3.00 D myopia, lens is 8.00D, but 14 mm downstream the vergence
emmetropia, 3.00 D hyperopia, and 6.00 D hyperopia. incident on the eye is 7.19D. The vergence Ufp is
The respective far points are then: real and 33.3 cm in zero for the emmtrope, so the emmetrope's ocular
front of the eye; at optical infinity; virtual and 33.3 cm accommodative demand is +7.19 D.
behind the eye; and virtual and 16.67 cm behind the In the above case, we label the 8.00 D as the
eye. You should now be able to show that the respec- spectacle accommodative demand. The spectacle ac-
tive near points are: real and 14.3 cm in front of the commodative demand is related (but not equal) to the
eye; real and 25.0 cm in front of the eye; real and ocular accommodative demand. Clinically, it is expedi-
100.0 cm in front of the eye; and virtual and 50.0 cm ent to think in terms of the spectacle accommodative
behind the eye. Note the near point variation even demand even though it can differ from the ocular
through each has a 4.00 D amplitude of accommo- accommodative demand.
dation. Another example of spectacle accommodative de-
mand is that of a spectacle corrected ametrope looking
through the spectacle lenses at a near object. Assume
6.11 Spectacle Accommodative Demand that the spectacle corrected ametrope can see a distant
object clearly without accommodating. For the distant
Consider an unaccommodated emmtrope with a object, the vergence incident on the spectacle lens is
10.00 D amplitude of accommodation. When this emm- zero. When the object is 25 cm from the spectacle
94 Geometrie, Physical, and Visual Optics

lens, the vergence incident on the spectacle lens is Ps

-4.0 D. To see this object clearly, the person has to


A

accommodate enough to neutralize the 4.00 D change.


The 4.00 D is the spectacle accommodative demand.
?
We can find the actual amount of accommodation
needed by the eye (the ocular accommodative de-
mand) by calculating Ufp and Ux (the vergence inci- *
u^J
a)
dent on the eye) and using Eq. 6.7. This is shown in
the next section.
We can formalize the above argument by defining
the spectacle accommodative demand A s as the differ-
ence in the vergences incident on the spectacle lens.
For the distance object (no accommodation) the inci-
dent vergence is zero. For the near object, the ver-
gence incident on the spectacle lens is U s . The specta-
cle accommodative demand is the difference between
the two vergences incident on the spectacle lens, or
FIGURE 6.14. a. A spectacle-corrected hyperope viewing a
A = 0-U. distant object, b. The spectacle-corrected hyperope viewing a
near object.
Then
A =-U.. (6.8)

EXAMPLE 6.1 Because of the vergence changes across the gap be-
Bobby Brickbat is wearing a distance vision specta- tween the lens and the eye, the ocular accommodative
cle correction. What is the spectacle accommoda- demand and the spectacle accommodative demand
tive demand when Bobby holds his bat 20 cm from may differ.
his spectacle lens and examines it for a suspected
crack? EXAMPLE 6.2a
Hilda is hyperope with a +12.00 D spectacle cor-
For the bat at 20 cm (the proximal object), rection at a 14 mm vertex distance. What is Hilda's
US = -5.00D, spectacle and ocular accommodative demands when
and looking at a tapestry located 25 cm in front of her
As = - ( - 5 . 0 0 D) =+5.00 D. spectacle lens?
For the tapestry (the proximal object), the ver-
Note that the power of the spectacle lens does not gence incident on the spectacle lens is
matter for the spectacle accommodative demand.
However, the spectacle lens power does matter for US = -4.00D.
the ocular accommodative demand. From Eq. 6.8, Hilda's spectacle accommodative
demand is
As = -(-4.00D) = +4.00D.
6.12 Ocular Accommodation Through For the tapestry, the light leaving the spectacle
Spectacles lens has a vergence
Figure 6.14 shows a spectacle corrected hyperope
looking first at a distant object (Figure 6.14a), and Vs = +12.00 D + (-4.00 D) = +8.00 D.
then at a near object (Figure 6.14b). For the distant For a +8.00 D vergence, the vergence Ux at a
object, plane waves are incident on the spectacle lens, distance 14 mm downstream is +9.01 D. When
and the vergence leaving the spectacle lens is equal to plane waves are incident on the +12.00 D lens, the
the power Ps of the spectacle lens. Converging light vergence leaving is +12.00 D, and the vergence Ufp
leaves the spectacle lens, and the vergence Ufp inci- 14 mm downstream is +14.42 D. From Eq. 6.7, the
dent on the cornea is greater than Ps. For the near ocular accommodative demand is
(proximal) object, the vergence Vs leaving the specta- A0 = U f p - U x ,
cle lens is less than P s , and again a vergence change A 0 =+14.42 D - ( + 9 . 0 1 D),
occurs across the gap between the lens and the eye. A 0 =+5.41D.
The vergence incident on the eye is U x . From Eq. 6.7,
the ocular accommodative demand is So while Hilda's spectacle accommodative demand
is +4.00 D, her ocular accommodative demand is
A = U Uv +5.41 D.
Thin Lens Eye Models 95

EXAMPLE 6.2b While the previous examples clearly show the ef-
Myra is myope with a -12.00 D spectacle correc- fect, they do not provide an intuitive feeling for the
tion at a 14 mm vertex distance. What is Myra's result. The following argument provides some intui-
spectacle and ocular accommodative demands tion even though the argument is flawed. The image
when looking at the tapestry located 25 cm in front formed by the spectacle lens is the object for the eye.
of her spectacle lens? For the distant object, the spectacle lens image is at
As in Example 6.1a, the spectacle accommoda-
tive demand is the eye's far point. Figure 6.15a shows the spectacle
lens images for the hyperope. The hyperope's far
A S =+4.00D. point is virtual. As one starts to move the object for
For the tapestry, the light leaving the spectacle lens the spectacle lens in from optical infinity, the spectacle
has a vergence lens' image moves from the far point out towards
V S = PS + U S , optical infinity.
Vs = -12.00 D + (-4.00 D) = -16.00 D. Figure 6.15b shows the corresponding case for the
myope. Here the eye's far point is real. As one starts
For a -16.00 D vergence, the vergence Ux at a
to move the object for the spectacle lens in from
distance 14 mm downstream is -13.07 D.
When plane waves are incident on the -12.00 D optical infinity, the spectacle lens's image moves from
lens, the vergence leaving is 12.00 D, and the the far point in towards the lens. When Figure 6.15a
vergence Ufp 14 mm downstream is -10.27 D. and Figure 6.15b are compared, we see that the
From Eq. 6.7, the ocular accommodative demand myope's eye has only a small distance change to
is accommodate for, whereas as the hyperope's eye has a
A0 = U f p - U x , much larger distance change to accommodate for.
The above argument is flawed because it does not
A 0 = -10.27 D - (-13.07 D),
show the effect of the spectacle lens's vertex distance
A o =+2.80D. on the ocular accommodative demand. We can use an
So while Myra's spectacle accommodative demand algebraic argument to build some legitimate intuition
is +4.00 D, her ocular accommodative demand is about the vertex distance dependence. First, we need
only +2.80 D. to consider an algebraic equation for downstream
vergence.
Note that the spectacle accommodative demand
was the same for Hilda and Myra. However, the
ocular accommodative demands are very different.
Hilda, the hyperope, has an ocular accommodative 6.13 The Downstream Vergence Equation
demand that is higher than the spectacle accommoda-
tive demand; while Myra, the myope, has an ocular We have been numerically solving downstream ver-
accommodative demand that is lower than the specta- gence problems since Chapter 2. Here we use the
cle accommodative demand. This difference is due to same technique to derive an algebraic equation for the
the vergence effectivity changes between the spectacle downstream vergence problem.
lens and the eye. In effect, for an object at the same Figure 6.16a shows converging wavefronts in the
distance, spectacle corrected hyperopes have to ac- same bundle. The wavefront at position 2 is a distance
commodate more than spectacle corrected myopes. d downstream from position 1. The wavefront at

spectacle lens
real images
a)

M
spectacle lens
virtual images FIGURE 6.15. Spectacle lens image move-
A ment as object is moved in from optical
b) infinity, a. Hyperope. b. Myope.
96 Geometric, Physical, and Visual Optics

FIGURE 6.16. Downstream vergence


geometry, a. Converging wavefronts. b.
Diverging wavefronts.

position 1 has a radius of curvature ql and the wave- EXAMPLE 6.3


front at position 2 has a radius of curvature of q2. For converging wavefronts (Figure 6.16a), when
From the figure, it is clear that the vergence at position 1 is +4.00 D, what is the
vergence at position 2 that is a distance of 15 cm
q2 = ql- d.
downstream?
The vergence of the wavefront at position 2 is given by Note that you should also be able to work this
particular problem numerically in your head by
using the techniques of Chapter 2. From Eq. 6.9,

where q2 is in meters. Then V = +4.00 D


2
~ [l-(0.15m)(4.00D)]'
v = i
2
(9,-d)' _ +4.00 D
2
provided ql and d are in meters. The numerator and " [1 - 0.60] '
the denominator of the above equation can each be +4.00 D
divided by qx with the result that V2 = = +10.00 D.
0.40
y = "ft
2
(l-d/9l) EXAMPLE 6.4
For diverging wavefronts (Figure 6.16b), when the
Since the vergence at position 1 is given by
vergence at position 1 is -20.00 D, what is the
vergence at position 2 which is a distance of 45 cm
downstream?
Again from Chapter 2, you should be able to
the expression for V2 becomes work this problem numerically in your head with-
V out using Eq. 6.9. From Eq. 6.9,
= m meters
^2 i -HV ' ^ - (6-9) -20.00 D
V9 =
[l-(0.45m)(-20.00D)]'
Equation 6.9 relates the vergence at position 1 to
the vergence at position 2 that is a distance d down- -20.00 D
stream. For this reason, I like to call Eq. 6.9 the V9 =
[1 - (-9.00)] '
downstream vergence equation. However, one can also
say that Eq. 6.9 gives the effect at a distance d -20.00 D
2
downstream of the vergence at position 1. Therefore, [l-(-9.00)]'
it can also be referred to as a vergence effectivity
-20.00 D
equation. V,= = -2.00D.
+ 10.00
Equation 6.9 is a general equation, and can be
derived from diverging wavefronts as well as from We can also solve Eq. 6.9 for V, in terms of V 2 ,
converging wavefronts (Figure 6.16b). Note the diop- in which case the result is the upstream vergence
ter is the reciprocal of a meter, and the distance d in equation:
Eq. 6.9 must be in meters so that the product dV, is
dimensionless. V,= d in meters. (6.10)
1 + dV,
Thin Lens Eye Models 97

6.14 Algebraic Approach: Accommodation approximation, Eq. 6.13, in Eq. 6.11 to obtain
Through Spectacle Lenses
A 0 ~ P s ( l + d P s ) - V s ( l + dV s ),
Figure 6.14 shows the situation for the ocular accom- or
modative demand through spectacle lenses. For ver- A 0 ~ P s - V s + d(Pf-Vf).
gence incident on the eye, the fundamental equation is
Now
Ac = U f p - U x .
P s - V f = (P s -V s )(P s + V s ),
When the person is looking at the distant object, plane
waves are incident on the spectacle lens, and, con-
sequently, the vergence of the light leaving the specta- and
cle lens is equal to the dioptric power of the lens. - U S = P S -V S .
Then from the downstream vergence Eq. 6.9, the
vergence Ufp incident on the eye is We can combine the above four equations to obtain
A 0 ~ - U s + d(-U s )(P s + V s ),
U
Uf = A 0 ~ - U , [ l + d(P, + V,)],
P 1-dP
When the person looks at the near (proximal) object and
the vergence incident on the spectacle lens is U s and A 0 ~ - U , [ l + d ( 2 P , + U,)].
the vergence leaving the spectacle lens is Vs. Then the
vergence Ux incident on the eye is Let us additionally assume that the object is just
starting to move in from optical infinity so that U s is
v small relative to 2PS. Then
U =
1 - dV/ A0~-Us[l+2dPs] (6.14)
When the expressions for Ufp and U x are substituted
From Eq. 6.8, the spectacle accommodative demand
into the accommodation equation, the result is
A s is equal to - U s , so
A = (6.11) A0~A,[l+2dP,]. (6.15)
1 - dR 1 - dVc
Equation 6.15 is an approximation to Eq. 6.11.
Equation 6.11 gives the ocular accommodative de-
According to Eq. 6.15, the difference between the
mand in terms of the dioptric power of the spectacle
ocular accommodative demand and the spectacle ac-
lens, the vergence Vs of the light leaving the spectacle
commodative demand is due to the 2dPs term. How-
lens, and the vertex distance d. It is clear from Eq.
ever, this term has opposite effects for myopes and
6.11 that the vertex distance has an effect, but it is not
hyperopes. For hyperopes, the spectacle lens power Ps
clear what the pattern of that effect is. We can make
is positive and the term adds so that A 0 is greater than
the pattern emerge by considering some approxima-
A s . For myopes, Ps is negative and the term subtracts
tions to Eq. 6.11. so that A 0 is less than A s .
Consider an equation of the form
Ginny is a myope who wears a 5.00D spectacle
y=h- < 612 > lens at a vertex distance of 12 mm. From Eq. 6.15,
when Ginny looks at an object 100 cm in front of her
When x is small relative to 1, the equation can be spectacle lens, the ocular accommodative demand is
approximated by
A 0 ~ - ( - 1 . 0 D)[l + 2(0.012 m)(-5.00 D)],
y ~ l + x. (6.13)
A0~+1.00D[l-0.12],
For example,
A 0 ~ +1.00 D[0.88] = +0.88 D.
1 The exact ocular accommodative demand for Ginny is
J = 1.11,
(1-0.10) also +0.88 D, so the approximation is very good.
while the approximation gives Gerry is a hyperope who wears a +5.00 D spectacle
lens at a vertex distance of 12 mm. From Eq. 6.15,
y ~ l + 0.10 = 1.10. when Gerry looks at an object 100 cm in front of his
The approximation gets better for smaller x values and spectacle lens he needs to accommodate
worse for larger x values. A 0 ~ - ( - 1 . 0 0 D)[l + 2(0.012 m)(+5.00 D)],
The vergence effectivity terms in Eq. 6.11 have the
form of Eq. 6.14 with x equal to dPs or dVs. When A 0 ~ +1.00 D[l +0.12],
these terms are small relative to 1, we can use the A o ~+1.00D[1.12] = +1.12D.
98 Geometrie, Physical, and Visual Optics

The exact ocular accommodative demand for Gerry is where the vergences Ufp and Ux are incident on the
also +1.12 D so the approximation is very good. cornea. From Eq. 6.1,
In the above cases, the d term added 0.12 D of
u f p = pc.
accommodation for the hyperope (Gerry) and sub-
tracted 0.12 D of accommodation for the myope The light leaving the contact lens is immediately inci-
(Ginny). The vertex distance and the resulting ver- dent on the eye with no vergence change. Therefore,
gence effectivity increases the ocular accommodative ux=vc.
demand for a hyperope and decreases the ocular
accommodative demand for a myope. By combining the above three equations, we get
For Gerry and Ginny the approximation gives the
A = P C -V C .
same result (to two digits) as the exact procedures.
However, the approximation starts to break down for Since
Hilda and Myra in Examples 6.2a and 6.2b. Both
- U C = P C -V C ,
Hilda and Myra had a +4.00 D spectacle accommoda-
tive demand. For Hilda, the hyperope, the approxima- the equation for the ocular accommodative demand
tion gives +4.67 D, which is greater than +4.00 D, simplifies to
while the exact ocular accommodative demand is
A0 = - U c . (6.16)
+5.41 D. For Myra, the myope, the approximation
gives +3.33 D, which is less than +4.00 D, while the For the contact corrected ametropes, the vergence
exact ocular accommodative demand is +2.80 D. of the light incident on the contact lens is equal to the
vergence of the light incident on the cornea of an
emmtrope who is standing at the same distance from
the object. Thus, in the thin lens and screen model, all
contact corrected ametropes have the same ocular
6.15 Accommodation Through Contact accommodative demand as an emmtrope.
Lenses For example, a myope wearing a -4.00 D contact
lens and viewing an object 50 cm in front of the lens
The differences in accommodative demand in the has a U c of -2.00 D, and from Eq. 6.16 an ocular
spectacle corrected cases were due to the different accommodative demand of +2.00 D. Similarly, a hy-
vergence changes that occurred between the spectacle perope wearing a contact correction of +4.00 D and
lens and the eye. However, contact lens corrections sit viewing an object 50 cm in front of the lens has an
on the eye; therefore, no vergence changes occur ocular accommodative demand of +2.00 D. An
between the contact lens and the eye. emmtrope viewing the same object 50 cm in front of
We can consider a contact lens as a spectacle lens his or her eye also has an ocular accommodative
with a zero vertex distance. In this case, Eq. 6.15 demand of +2.00 D.
becomes exact, and shows that For Myra and Hilda in the previous section, the
A 0 = AS. contact corrections are -10.27 D and +14.43 D, re-
spectively. The ocular accommodative demand for the
So, for a contact corrected ametrope, the ocular ac- contact corrected girls to see the tapestry 25 cm from
commodative demand equals the spectacle accom- the spectacle plane, or 26.4 cm from the eye is found
modative demand (spectacle at zero vertex distance). as follows:
In turn, Eq. 6.9 gives
A0 = - U c , (6.16)
where the subscript s (for spectacle) has been changed and
to c (for contact).
We can also derive Eq. 6.16 from first principles. A 0 = - U C = +3.79D.
Let Pc be the power of the contact lens, U c be the The accommodative demand of 3.79 D is the same as
vergence incident on the contact lens, and Vc be the that of an emmtrope viewing the tapestry held
vergence of the light exiting the contact lens. Then 26.4 cm in front of the cornea.
VC = PC + UC. Myra, the myope, needs 2.80 D of accommodation
to see the tapestry when spectacle corrected and
From Eq. 6.7, the ocular accommodative demand is 3.79 D when contact corrected. Hilda, the hyperope,
given by requires 5.42 D of accommodation to see the tapestry
A0 = U f p - U x ) when spectacle corrected but only 3.79 D when con-
Thin Lens Eye Models 99

E

o
.>
<D

c3 vertex distance 14 mm
E object 20 cm in

II front of spectacle plane

8.00D4

7.00D-4-

e
*pv
X*f>{
6.00D 1

contact corrected or emmtrope


4.67D

4.00D

nyope
3.00D

-h +
-+- FIGURE 6.17. Ocular accom-
2D 4D 6D 8D 10D 12D 14D 16D modative demand vs spectacle and
spectacle lens powers contact lens powers.

tact corrected. In general, when a hyperope changes is only a small difference in ocular accommodative
from a contact correction to a spectacle correction, demand for the various individuals. However, the
more ocular accommodation is needed. When a difference steadily increases as the correcting lens
myope changes from a contact correction to a specta- power increases.
cle correction, less ocular accommodation is required.
Keep in mind that any difference has to be 0.25 D or
greater to be clinically significant and that low hy- 6.16 Corrected Near Points
peropes and myopes may not show any clinically
significant differences. The point conjugate to the retina for the correcting
The graph in Figure 6.17 shows the ocular accom- lens and maximally accommodated eye is the correc-
modative demand as a function of correcting lens ted near point. For contact corrected individuals, the
powers. The vertex distance of the spectacle correc- corrected near point is very easy to find. In fact, it is
tions is 14 mm. For low correcting lens powers, there equivalent to finding the near point of an emmtrope.
100 Geometrie, Physical, and Visual Optics

Myra, the myope in Section 6.12, has a -10.27 D Myra Hilda


contact lens correction and an amplitude of accommo-
dation of +6.00 D. What is Myra's contact corrected "x -6.15 cm + 11.86 cm
near point?
-(6.15-1.4) cm (11.86+ 1.4) cm
Use Eq. 6.16, and set A 0 equal to the amplitude: *>s

+6.00D=-UC "s -4.75 cm + 13.26 cm

or vs -21.05 D +7.54 D
UC=-6.00D. us VT - P V S -P S
s * s

When maximally accommodated, Myra can handle an us (-21.05 +12.00) D (+7.54-12.00)0


incident vergence of -6.00 D. Myra's corrected near
point is then real and 16.67cm (i.e., 100/6D) in front us -9.05 D -4.46 D
of her contact lens. -11.05 cm -22.43 cm
Hilda, the hyperope in Section 6.12, also had an "s

amplitude of accommodaton of +6.00 D. Thus, Hil-


da's contact corrected near point is also 16.67 cm in
front of the contact lens plane.
The spectacle corrected near point is not as easy to The magnitude of the corrected near points meas-
find. For a given object, spectacle corrected hyperopes ured from the spectacle plane for Myra, the myope,
have to accommodate more than spectacle corrected and Hilda, the hyperope, are 11.05 cm and 22.43 cm,
myopes. Therefore, if they have the same amplitude respectively. So indeed, the spectacle corrected hy-
of accommodation, the spectacle corrected hyperope perope has a near point that is farther away than that
will have a corrected near point that is farther from of the myope (and of an emmtrope). This is apparent
the eye than the near point of the spectacle corrected clinically in that spectacle corrected hyperopes tend to
myope. start showing presbyopic symptoms at a younger age
Again consider Myra and Hilda of Section 6.12. than spectacle corrected myopes.
Hilda's spectacle correction at a vertex distance of
14 mm is +12.00 D, while Myra's spectacle correction
at 14 mm is -12.00 D. When the object is at the
corrected near point, the vergence incident on the
spectacle lens is U s , the vergence leaving the spectacle 6.17 Improper Corrections
lens is Vs, and the vergence incident on the eye is Ux
(Figure 6.14b). The ocular accommodative demand Suppose that a lens in front of an eye is not the
equation is optically proper correcting lens. Then the "corrected
far point" is not at optical infinity and the person will
A0 = U f p - U x .
not function like an emmtrope.
In each case, we know the amplitude and the Ufp Quinn is a myope who needs a -6.00 D distance
values. We can then solve for U x , vision contact lens but instead is given a 8.00D
contact lens. Quinn is over-minused by 2.00 D. With
Ux = U f p - A o ;
the 8.00D contact lens on his eye, Quinn is func-
and work backwards to successively find Vs, then U s , tioning like a 2.00 D hyperope. Quinn can overcome
and then the object distance us from the spectacle lens the 2.00 D of excess minus by using 2.00 D of ocular
to the corrected near point. The numbers for Myra accommodation. If Quinn does not accommodate,
and Hilda are as follows: then the corrected far point, i.e., the point conjugate
to the retina for the 8.00D lens-eye system, is
virtual and 50 cm from the lens. In other words, an
Myra Hilda incident vergence of +2.00 D is required to neutralize
the 2.00 D of over-minus.
U fp -10.27 D +14.43 D
Rita is a hyperope who needs a +5.50 D spectacle
Amp 6.00 D 6.00 D lens but is given a +3.50 D spectacle lens. Rita is
under-plussed by 2.00 D, which is equivalent to being
Ux U fp - Amp U fp - Amp over-minused by 2.00 D. Rita then functions like a
Ux -10.27 D - 6 . 0 0 D +14.43 D - 6.00 D
2.00 D spectacle hyperope and can compensate for the
2.00 D shortage with 2.00 D of spectacle accommoda-
Ux -16.27 D +8.43 D tion. When Rita is not accommodating, the point
Thin Lens Eye Models 101

conjugate to her retina for her +3.50 D lens-eye sys- 6.18 Range of Clear Vision Through
tem is virtual and located 50 cm from the spectacle Various Lenses
lens. In other words, a vergence of +2.00 D incident
on the +3.50 D spectacle lens compensates for the Suppose an ametrope, Randy, looking through a
2.00 D shortage. +4.00 D lens in the spectacle plane can clearly see real
Sam is a hyperope who requires a +3.00 D contact objects ranging from 100 cm to 20 cm in front of the
lens but is given a +4.00 D contact lens. Sam is lens. What is Randy's range of clear vision when the
over-plussed by 1.00 D. For incident plane waves, the +4.00 D lens is replaced by a +6.00 D lens?
clear image formed by the +4.00 D lens-eye system One way to proceed is to figure the bounds needed
will be in front of Sam's retina (in the vitreous for the vergence leaving the spectacle plane. For the
humor), and accommodation pulls it even further in object at 100 cm in front of the +4.00 D lens, the
front of the retina. With the +4.00 D contact lens on vergence leaving the lens is +3.00 D. For the object at
his eye, Sam will function like a 1.00 D myope. The 20 cm in front of the +4.00 D lens, the vergence
corrected far point for the +4.00 D lens-eye system is leaving the lens is -1.00D. Randy can handle ver-
real and 100 cm in front of the +4.00 D lens. In other gences leaving the spectacle plane ranging from
words, an incident vergence of 1.00D neutralizes +3.00 D down to -1.00 D.
the 1.00 D of excess plus. For the +6.00 D lens, the range of incident ver-
Tina is a myope who needs a 8.50D spectacle gences are:
lens but instead is given a 7.00 D spectacle lens. Tina
is under-minused by 1.50 D, i.e., 7.00 D of her myopia U=V-P,
has been corrected leaving 1.50 D uncorrected. When U = +3.00 D - (+6.00 D) = -3.00 D,
Tina views a distant object through the -7.00 D spec-
tacle lenses, the clear image formed by the 7.00D and
spectacles and her eye will be located in front of her U = -1.00 D - (+6.00 D) = -7.00 D.
retina. Accommodation will pull the image even fur-
ther in front of her retina. With the -7.00 D specta- So the range of incident vergences is from -3.00 D to
cles, Tina is functioning like a 1.50 D spectacle myope. -7.00 D, or Randy's range of clear vision through the
Her corrected far point is real and 66.67 cm in front of +6.00 D lens is 33.3 cm in to 14.3 cm.
her spectacles. In other words, an object placed Incidentally, for distance vision Randy needs a
66.67 cm in front of her spectacles results in a ver- +3.00 D lens. Randy's amplitude of accommodation
gence incident on her spectacles of -1.50 D, which in terms of the spectacle accommodative demand is
compensates for the - 1 . 5 0 D shortage. + 3 . 0 0 D - ( - 1 . 0 0 D ) = +4.00D. Through the +3.00
Any person who is over-plussed or under-minused D lens, Randy's range of clear vision is from optical
functions like a myope. When these people view a infinity in to 25 cm.
distant object without accommodating, their eye forms A bifocal lens is a lens with a distance vision area
the clear image in the vitreous humor and the retinal of one power and a near vision area of a higher plus
image is blurred. When they accommodate, the clear (or less minus) power. The difference between the
image is pulled even further away from the retina and near vision power and the distance vision power is
the blur of the retinal image increases. These people called the add. An unsophisticated bifocal can be
are sometimes said to be fogged because their distance achieved by simply gluing a small plus lens to the
vision is blurred and accommodation blurs it even distance vision lens (Figure 6.18).
more.
Any person who is under-plussed or over-minused
functions like a hyperope. Accommodation can be
used to compensate for the under-plussed or over-
minused amount.
Ursula is wearing a +3.50 D spectacle lens. While
wearing these spectacles, she has a corrected far point
that is real and 40 cm in front of her spectacle plane.
What is Ursula's proper correction?
The corrected real far point 40 cm in front of the
spectacle plane indicates that Ursula needs an addi-
tional -2.50 D in her spectacle correction. Her proper
spectacle correction is:
add
+3.50 D + (-2.50 D) = +1.00 D. FIGURE 6.18. A bifocal lens.
102 Geometrie, Physical, and Visual Optics

We could consider the +4.00 D lens of the above The letters on a 20/20 visual acuity line subtend an
example as the near vision region of a +3.00 D dis- angle of 5' with the fine detail subtending an angle of
tance vision lens with a +1.00D add. Similarly, the . The retinal image size of such a letter is,
+6.00 D lens can be considered as a +3.00 D distance
vision lens with a +3.00 D add for near vision. Note ^ = t a n ,
that increasing the add makes the range of clear vision 16.67 mm
closer and smaller. This is an important feature in the |I| = 16.67 mm tan(5').
clinical selection of an add.
Since there are 60' per degree, then 5' equates to
0.083. Therefore,
|I| = 16.67 mm tan(0.083),
6.19 Retinal Image Sizes
|I| = 0.0241 mm,
Retinal images are small and inverted relative to the
real world object. Consider a 2-m tall basketball or
player standing 10 m from a typical emmtrope. In the
|I|=24.1 Mm,
thin lens and screen model, the typical lens-screen
distance was 16.67 mm. From the nodal ray, where \ stands for micrometer. (1 = 10~ m.
That's small!)
tan6 = An alternate method for dealing with the tangents
of small angles is to use the small angle approximation

and in which the tangent is replaced by the angle itself
expressed in radians. For the above case 5' equals
|i| 0.083. Let us convert the angle to radians:
tan6 =
16.67 mm '
9 = 0.083/(57.3/radian),
Therefore,
= 1.45 10"3 radians.
|i| |o|
16.67 mm 10 m ' Then
2m |I| = 16.67 mm tan ,
1III
1 = 16.67 mm i r . ,
10 m
| I | ~ 16.67 mm ,
|l| = 3.33 mm.
| I ] ~ (16.67 mm)(1.45xl0" 3 ),
The retinal image of the 2-m tall basketball player is
3.33 mm in size. |I| ~ 24.1 x 10"3 mm = 24.1 .

FIGURE 6.19. a. Image size for a contact-


corrected myope, b. Deviation of ray by the
spectacle correction, c. Superposition to com-
pare spectacle and contact-corrected retinal
image sizes.
Thin Lens Eye Models 103

far point FIGURE 6.20. Spectacle vs contact lens image


plane sizes in a myope's far point plane.

6.20 Retinal Image Sizes: Spectacle vs parallel in object space because they come from the
Contact Lenses same distant object point. For the myope, the path of
the ray traveling through the spectacle lens shows that
Retinal image sizes of corrected ametropes differ de- the spectacle corrected retinal image is smaller than
pending on whether a contact correction or a spectacle the contact corrected retinal image. Many myopes
correction is used. Figure 6.19a shows the nodal ray prefer contact lenses to spectacle lenses because of the
for a contact corrected myope viewing a distant off- larger retinal image size they get with the contact
axis object point. The nodal ray specifies the retinal lenses.
image size I c . When the myope is corrected with a Another way to reach the same conclusion is to
spectacle lens, the same incident ray is bent up by the consider the image size of the two different correcting
spectacle lens and is no longer the nodal ray of the eye lenses. The image of these correcting lenses is the
(Figure 6.19b). Figure 6.19c shows a ray (solid lines), object for the eye, and this image is formed in the far
which after being bent by the spectacle lens becomes point plane of the eye. The image size in the far point
the nodal ray for the eye and specifies the retinal plane is easily determined by considering the nodal ray
image size I s . For comparison purposes, the nodal ray for the correcting lens. Figure 6.20 shows a superposi-
(dashed line) for the contact lens is superimposed on tion of the spectacle lens case (solid lines) and the
Figure 6.19c. The two rays, solid and dashed, are contact lens case (dashed lines). The two incident

^ ^ - ^ t
4 h
Y
a)

FIGURE 6.21. a. Image size for a contact


corrected hyperope. b. Deviation of ray by
the spectacle lens. c. Superposition to com-
pare spectacle and contact corrected retinal
image sizes.
104 Geometrie, Physical, and Visual Optics

FIGURE 6.22. Spectacle vs contact correc-


far point ted image sizes in a hyperope's far point
plane plane.

nodal rays are parallel to each other because they this section. Note that we are talking about the retinal
come from the same distant object point. Clearly, the image size of the same ametrope when corrected by
contact lens image in the far point plane is larger than spectacle lenses vs contact lenses. We are not compar-
the spectacle lens image in the far point plane. Since ing different ametropes. A more quantitative treat-
the correcting lens images are both formed in the far ment is given in Chapter 14.
point plane and both serve as the object for the eye,
then the contact corrected retinal image will be larger
than the spectacle corrected retinal image. For high
myopes who switch back and forth between contacts Problems
and spectacles, this effect is clearly noticeable.
Similar diagrams for a hyperope are shown in For each of the next four problems, represent the
Figure 6.21a-c. Figure 6.21a shows the nodal ray for unaccommodated eye by a -1-60.00 D thin lens and a
the contact corrected eye. Figure 6.21b shows that this screen (retina) located a distance x behind the lens.
ray is not the nodal ray for the eye when a spectacle The amplitude of accommodation is listed as A. For
correction is worn. Figure 6.21c shows the nodal ray the uncorrected eye, find the far point and near point,
(solid) for the eye when a spectacle lens is worn sketch the range of clear vision, classify the refractive
together with a superposition of the nodal ray (dash- error (myopia, hyperopia, emmetropia), and give the
ed) for the contact corrected eye. The hyperope's distance vision correction at the cornea and the dis-
retinal image is larger when spectacle corrected. This tance vision correction at a spectacle plane located
is exactly opposite to that of the myope. 15 mm from the cornea.
Figure 6.22 shows the correcting lens image sizes in
the hyperope's far point plane. The dashed ray is the 1. Sally Jones, age 30, x = 19.9 mm, A = 7.00 D.
nodal ray for the contact lens, while the solid ray is the 2. Richard Nixit, age 55, x = 13.9 mm, A = 1.25D.
nodal ray for the spectacle lens. Note that in the far 3. Roger Stub, age 25, x = 16.0 mm, A = 8.50 D.
point plane the spectacle lens image is larger than the 4. Thelma Elf, age 43, x = 17.2 mm, A = 2.50 D.
contact lens image. Consequently, since these images 5. Molly Finnigan is a 4.50 D ocular myope with a
serve as the object for the eye, then the spectacle 6.00 D amplitude of accommodation. When un-
corrected retinal image for the hyperope will be larger corrected, can Molly clearly see the fine print in a
than the contact corrected retinal image. book held 40 cm from her cornea? If not, why
Table 6.1 summarizes the qualitative findings of not? If so, what is the ocular accommodative
demand?
6. Mike Finnigan is a 3.50 D ocular hyperope with a
5.00 D amplitude of accommodation. When un-
TABLE 6.1. corrected, can Mike clearly see the fine print in a
Retinal Image Size Comparison: Same Ametrope book held 40 cm from his cornea? If not, why not?
If so, what is the ocular accommodative demand?
Contact Spectacle 7. Jim Finnigan is a 1.50 D ocular myope with a
Corrected Corrected
2.00 D amplitude of accommodation. When un-
corrected, can Jim clearly see the fine print in a
Myope Larger Smaller
Hyperope Smaller Larger book held 40 cm from his cornea? If not, why not?
If so, what is the ocular accommodative demand?
Thin Lens Eye Models 105

8. Katie Finnigan is a 2.00 D ocular hyperope with a objects at a distance of 37 cm to 75 cm from the
5.50 D amplitude of accommodation. When un- lens. When the +5.00 D lens is replaced by a
corrected, can Katie clearly see the fine print in a +4.00 D lens (same vertex distance), what is
book held 40 cm from her cornea. If not, why not? Wilhelmina's range of clear vision? What is
If so, what is the ocular accommodative demand? Wilhelmina's distance vision correction (same ver-
9. Rose Opal has a +13.50 D distance vision specta- tex distance)? What is Wilhelmina's range of clear
cle correction at a 12 mm vertex distance. Rose vision through her spectacle correction?
looks through her spectacles at a statue located 12. Elroy Frost puts on a - 1 . 5 0 D spectacle lens at a
19 cm in front of the spectacles. What is the 13 mm vertex distance. With the -1.50 D lens,
spectacle accommodative demand? What is Rose's Elroy can clearly see real objects from 26 cm to
ocular accommodative demand? 50 cm from the lens. What is Elroy's range of clear
10. Zeke Feldstein has a -9.50 D distance vision vision through a +1.00D spectacle lens (same
spectacle correction at a 13 mm vertex distance. vertex distance)? What is Elroy's distance vision
Zeke looks through his spectacles at a small auto- spectacle correction? What is his range of clear
motive part that he holds 14 cm in front of his vision through his correction?
spectacles. What is the spectacle accommodative 13. Jill Harkins is an emmtrope with an unaccommo-
demand? What is Zeke's ocular accommodative dated eye represented by a +64.00 D thin lens and
demand? screen (retina). What retinal image size (in mi-
11. Wilhelmina Roads puts on a +5.00 D spectacle crons) does Jill have when viewing a distant object
lens that has an 11 mm vertex distance. With the that subtends an angle of 16'?
+5.00 D lens, Wilhelmina can clearly see real
CHAPTER SEVEN

Single Spherical
Refracting
Interfaces

7.1 Convex/Concave Terminology surfaces. The lens itself is thus a converging lens. As
viewed from the outside, which is where a human
observer would be, each surface is convex. This lens is
Consider the spherical glass (n = 1.50) surface shown conventionally called a biconvex lens.
in Figure 7.1. In the middle, the glass bulges out into However, a discrepancy in terminology has de-
the air. In Section 3.1, we saw qualitatively that such a veloped in the description of each surface. Consider
surface converges plane waves no matter which way Figure 7.4b, which shows light incident from the left
the waves propagate through the surface. Consequent- on the biconvex lens. Initially the light in air ap-
ly, a spherical glass surface with the glass in the middle proaches the left surface of the lens, and this surface
bulging into the air is always a converging surface. viewed from the left is convex. Then the light enters
The surface is actually an interface between two the glass and approaches the right surface of the lens.
different optical media, air and glass. Some other But the right surface of the lens, as viewed from inside
examples of optical interfaces are water and glass, the glass, is curved inward like the inside of a bowl.
plastic and water, oil and glass, air and the cornea of Consequently, some authors call the right surface
the eye, the aqueous humor and the front surface of concave, but according to such terminology, the bicon-
the crystalline lens, the back surface of the crystalline vex lens, in which biconvex literally implies two con-
lens and the vitreous humor. In general, when the vex sides, has a convex left side and a concave right
interface is spherical and the middle of the higher side.
index media bulges out into the lower index media, In order to avoid the above illogical terminology,
then the interface is converging no matter which way let us label a surface or interface as being convex or
the light propagates through the interface. Figure 7.2 concave from the perspective of the lower index
shows examples of converging interfaces. Conversely, medium, regardless of which way the light is traveling.
when the middle of a lower index media bulges into a Thus, let us call the right surface of the lens in Figure
higher index media, then a spherical interface diverges 7.4b convex. With this convention, a biconvex lens has
incident light no matter which way the waves prop- a convex left side and a convex right side, as one
agate through the interface. Figure 7.3 shows some might expect.
diverging spherical interfaces. With the assignment of convex and concave being
Historically, people have sought to identify surfaces made from the perspective of the lower index
by their geometric characteristics rather than their medium, then a convex refracting interface always
optical characteristics. The word convex is defined as turns out to be a converging interface (Figures 7.1 and
being curved outward like the exterior of a sphere. 7.2), and a concave refracting interface always turns
The word concave is defined as being curved inward out to be a diverging interface (Figure 7.3). (When
like the interior of a bowl or the interior of a cave. referring to other books, be sure to check what ter-
Figure 7.4a shows a glass lens with two converging minology they are using.)
107
)8 Geometric, Physical, and Visual Optics

7.2 Sagittas and the Sagittal Approximation


The next two sections deal with the derivation of the
air f glass
vergence equation for a single spherical refracting
interface. If you wish to study the equation and re-
lated examples before going through the derivation,
then skip to Section 7.4.
FIGURE 7.1. A spherical interface. The sagittal approximation is frequently used in

air / plastic water f plastic glass ( oil


= 1.00)1 (n = 144) 1.33)l
(n = 1.33)1 (n = 1.44)
1.44) (n = 1.52)1
1.52)1 (n = 1.70)

glass air glass plastic ] water


(n = 1.50) f (n
(n == 1.00)
1.00) (n = 1.50) (n = 1.44)/(n = 1.33)

FIGURE 7.2. Converging interfaces, all con-


vex to the lower index medium.

glass plastic water diamond I glass


= 1.50) (n = 1.44)1
1.44)l (n = 1.50) (n = 2.40)\
2.40) (n = 1.50)

glass glass
(n = 1.50) /(n = 1.50)

FIGURE 7.3. Diverging interfaces, all con-


cave to the lower index medium.

o -o
geometric optics to quantitatively treat the relation-
ship between the curvature of a spherical surface and
the dioptric power of that surface, as well as the
relationship between the curvature of a wavefront and
the vergence of that wavefront. Before treating the
sagittal approximation, an exact relationship for the
sagitta of a spherical surface is defined and discussed.
Figure 7.5a shows a cross-section of a long horizon-
tal glass rod with a spherical front surface. The center
of curvature of the surface is labeled C. The horizontal
a) b) line through C is normal (i.e., perpendicular to) the
FIGURE 7.4. Biconvex lens. spherical surface at point A. In Figure 7.5b, a line
Single Spherical Refracting Interfaces 109

FIGURE 7.5. Spherical interface and sagitta.

called a chord is perpendicular to the normal line. The The half chord length h is 0.9 cm. Then
chord intersects the surface of the sphere at points B s = 10 cm - V(10cm) 2 -(0.9cm) 2 ,
and D, and intersects the normal at point E. The line
s = 10 cm - V(100-0.81) cm,
segment AE, which is normal to the surface, is called
the sagitta. The point A is called the vertex or pole of s = 10 cm - 9.96 cm,
the sphere. s = 0.04 cm,
Let r equal the radius of curvature of the sphere, s = 0.40 mm.
and s equal the length of the sagitta. In Figure 7.5c, r
equals both the length AC and the length BC. Let h As illustrated in the above example, h2 is frequent-
equal half the chord length, which is the length BE. ly much smaller than r2. As a result, Eq. 7.3 can be
Let g equal the length EC. Then from the Pythag- accurately approximated by a simpler equation. First,
orean theorem applied to the triangle BEC, rewrite Eq. 7.3 as
r 2 = g 2 + h 2, s = r - rVl - (h 2 /r 2 ). (7.4)
or From calculus, the square root can be approximated
(7.1) by a Taylor series expansion:
From Figure 7.5c, h4
AC = AE + EC, (7.5)
or 2 2
When h is much smaller than r , only the first two
r = s + g, terms of the expression are significant, or
and
s= r-g. (7.2)
By substituting Eq. 7.1 into Eq. 7.2, we obtain The square root is equivalent to an exponential of 1/2,
s = r-Vr2-h2. (7.3) which is the source of the 2 in the denominator of the
second term.
Equation 7.3 is an exact relationship between the When we substitute the above equation into Eq.
sagittal length s, the half chord length h, and the 7.4, we obtain
radius of curvature of the spherical interface or wave-
front.
- r - r ( l - ^ ) ,
EXAMPLE 7.1
A spherical convex glass surface, as in Figure 7.5, or
has a radius of curvature of 10.00 cm. What is the
sagittal length for a chord length of 1.80 cm? s= r-r+2?.
110 Geometric, Physical, and Visual Optics

Finally, while
h^ h2 = (0.9 cm)2 = 0.81 cm2.
provided h2 <^ r2. (7.6)
2r' When s2 is much less than h2, s2 can be neglected in
Equation 7.6 is called the sagittal approximation. It the numerator of Eq. 7.8, and then
provides a very simple relationship between s, h, and
r. Furthermore, since the curvature R of a sphere is r =
equal to the reciprocal of the radius r, it follows from 2^'
or
the sagittal approximation that
s = 2r'
s = y R, provided h2 < 1/R2. (7.7) which is the sagittal approximation.
Equation 7.7 shows that the sagittal length is directly As shown in the next section, the sagittal approxi-
related to the curvature R of the sphere for those sit- mation is immensely useful. However, we need to use
uations is which the approximation is accurate. the exact relationship, Eq. 7.3, when h2 is not small
relative to r2, in which case, s2 is not small relative to
EXAMPLE 7.2 h2.
Use the sagittal approximation to calculate s for the
surface used in Example 7.1. There, the radius r
was 10 cm and the half chord length h was 0.9 cm.
From Eq. 7.6, 7.3 Derivation of the Vergence Equation
2 2
(0.9 cm) 0.81cm Figure 7.6 shows a long glass rod with a convex
s = -^^-r = -^ = 0.04 cm,
2(10 cm) 20 cm spherical surface. The spherical surface is optically a
or single spherical refracting interface (SSRI) between
s = 0.4 mm, air and glass.
The medium that the light is initially in has a
which is the same value obtained with the exact refractive index n^ The medium that the light is in
equation. when leaving the interface has a refractive index n 2 .
The incident wavefront (in medium n t ) is diverging,
For those who are not familiar with the Taylor
while the exiting wavefront (in medium n 2 ) is converg-
series expansion, there is a more intuitive way to ob-
ing. The incident wavefront is just touching the inter-
tain the sagittal approximation. First, take the exact sag-
face at point B, which is the vertex of the interface.
ittal equation, Eq. 7.3, and solve for r. The steps are:
The exiting wavefront has moved to point C on the
normal to the interface.
s = r-Vr2-h2,
The time t for the wavefront to move from B to C,
s-r^-VT^h2". in terms of the directed distance BC and the velocity
w2 of the wavefront in the second medium, is
Square both sides to obtain
(s-r)2 = r2-h2. BC
t=
w. '
Then
During this time, the part of the incident wavefront at
s2 - 2rs + r = r - h2, point E has moved from E to H. In the paraxial
s2 - 2rs = -h 2 , approximation, the path of the wavefront in going
from E to H is well approximated by the horizontal
2rs = s2 + h2,
line defined by the points E, F, G, and H. In other
and finally words, any vertical movement is negligible relative to
the horizontal movement. In moving from E to H, the
s2 + h2 wavefront enters the second medium at G. The time t
(7.8)
2s. that it takes for the wavefront to get from E to F to G
Equation 7.8 is still an exact relationship between s, h, to H, in terms of the directed distances EF, FG, and
and r. However, when h2 is much less than r2, then s2 GH, and the velocities w, and w2 of the wavefront in
is also much less than h2. In the example previously the first and second mediums, respectively, is
given, EF FG GH
=
s2 = (0.04 cm)2 = 0.0016 cm2, w, w, w^
Single Spherical Refracting Interfaces 111

FIGURE 7.6. Wavefronts passing through the spheri-


cal interface.

Since the two previous equations are for the same or


period of time, they can be set equal to each other
n 2 CD = (n2 - n J B C + njBA. (7.11)
resulting in
The distances CD, BC, and BA are the respective
EF FG GH
- C7.9) sagittas of the exiting wavefront, the single spherical
W, W, Wo
refracting interface, and the incident wavefront. Un-
The respective velocities Wj and w2 of the light in the der par axial conditions, the sagittal approximation is
first and second media are related to the speed of light accurate. Then
c in a vacuum, and the respective indices of refraction
nl and n2 by the equations
-.
w, = - where r is the radius of curvature of the interface and
h is the half-chord length,
w7 = A h2
BA=
2-u>
When the velocity relationships are substituted into
where u is the radius of curvature of the incident
Eq. 7.9 and the common factor c is cancelled on both
wavefront (u is then, by definition, the object dis-
sides of the equation, the result is
tance), and
^ E F + n ^ G + n 2 GH = n 2 BC. (7.10) i 2

In general, the index of refraction n times a distance CD =


2v9
along a wavefront direction (or ray) is called the
where v is the radius of curvature of the exiting
optical path length. Equation 7.10 is an example of a
wavefront (v is then, by definition, the image
general principle that states: When a wavefront prop-
distance).
agates from one place to another, the optical path
lengths for different parts of the wavefront are equal. By substituting the sagitta relationships into Eq.
Note that it is the optical path lengths that are equal, 7.11 and cancelling the common factor h 2 /2 on both
as opposed to the actual path lengths. sides, we obtain
From Figure 7.6, the following distance equalities 2 _ ( n
2~ n
i) + i
hold:
EF = AB, But n2/v is the vergence V of the exiting wavefront,
FG = BC, and nju is the vergence U of the incident wavefront;
i.e.,
GH = CD.
n
Furthermore, the directed distance AB is opposite in
v= ' V
sign to the directed distance BA, or u=Hi.
AB = - B A . u
So
By using the preceding distance equalities in Eq. 7.10,
+ U. (7.13)
we obtain v=
- i ^ B A + n x BC + n 2 CD = n 2 BC, Equation 7.13 is a vergence equation for the single
112 Geometrie, Physical, and Visual Optics

spherical refracting interface. Evidently, the dioptric written as


power P is given by
P = (n2-ni)R. (7.20)
P = ^ i , (7.14) Equation 7.20 says that the dioptric power of the
interface increases when the curvature of the interface
in which case, Eq. 7.13 can be written as increases (i.e., the surface gets steeper), and decreases
V = P + U. (7.15) when the curvature of the interface decreases (i.e., the
surface gets flatter). The equation also says that the
dioptric power of the interface increases when the
difference in the refractive indices increases, and de-
7.4 Dioptric Power creases when the difference in the refractive indices
decreases.
The derivation in the preceding section shows that the The dioptric power P is expressed in diopters with
vergence equation 1 D being dimensionally equal to the reciprocal of 1 m.
In Eq. 7.19, either r must be expressed in meters, or a
V = P + U, (7.16) conversion factor must be used. Just as with ver-
describes the paraxial imaging by a single spherical gences, the conversion factor can be explicitly written
refracting interface. The (generalized) vergence U of in the numerator.
the incident wavefront is given by
EXAMPLE 7.3a
V = nl/u, (7.17) A spherical convex interface between water of
where is the index of the medium that the light is index 1.33 and glass of index 1.53 has a radius of
initially in and u is the radius of curvature of the curvature of 10 cm (Figure 7.7). What is the diop-
incident wavefront. The radius of curvature u of the tric power of the interface when light is incident
from the water side?
incident wavefront is, by definition, the object dis- Before calculating, you should be able to tell
tance for the interface. The (generalized) vergence V whether the interface is converging or diverging.
of the wavefront exiting the interface is given by Here the interface is converging so P must be
V=n 2 /i;, (7.18) positive.
The parameters for the calculations are:
where n2 is the index of the medium that the light is in n, = 1.33, n2 = 1.53, and r=+10.00 cm.
after it passes through the interface and v is the radius Then from Eq. 7.19,
of curvature of the wavefront leaving the interface.
n7 - n,L
The radius of curvature v of the wavefront leaving the P=- ,
interface is, by definition, the image distance. The sign
convention is the same as for thin lenses. In fact, the (1.53- 1.33)100 cm/m
only difference between the basic SSRI equations and + 10 cm
the thin lens in air equations is the appearance of the +20 cm/m
refractive indices. P= = +2.00D.
+ 10 cm
The derivation of Eq. 7.16 also provided the rela-
tionship between the dioptric power P of the spherical EXAMPLE 7.3b
interface, the radius of curvature r, and the indices of What is the dioptric power of the above interface
when light is incident from the glass side?
refraction nl and n 2 : We already know that the interface is a converg-
ing interface for light going either way. Therefore,
P = n? - n, (7.19) P must be positive for light going either way.
The same sign convention is used for u, v, and r. The
directed distance r is positive if going from the inter-
face to the center of curvature is in the same direction
that the light is traveling when it leaves the interface.
The directed distance r is negative if going from the
interface to the center of curvature is opposite to the water
direction that the light is traveling when it leaves the
interface.
Since the radius r is reciprocally related to the
curvature R of the interface, then Eq. (7.19) can be FIGURE 7.7. Convex interface.
Single Spherical Refracting Interfaces 113

For the reversed direction,


nl = 1.53, n 2 = 1.33, and r=-10cm. 11, = 1.00 1.336
Then
,-,
P=

( 1 . 3 3 - 1.53)100 cm/m
P= cornea
- 1 0 cm
- 2 0 cm/m FIGURE 7.9. Cornea modelled as an SSRI.
P= = +2.00D.
- 1 0 cm
Note that the dioptric power of the interface is
+2.00 D no matter which way the light is propagat- D_ (1.00-1.50)100 cm/m
ing through the interface. The sign of the radius of +5 cm
curvature r does change since it depends on the -50 cm/m
sign convention for directed distances. However, P= = -10.00 D.
+5 cm
the sign of the difference in indices also changes
with the result that the sign of the dioptric power P Again note that the dioptric power of the interface
is invariant with respect to the direction that the is -10.00 D, irrespective of which way the light
light is propagating. travels through the interface.

EXAMPLE 7A EXAMPLE 7.5


A spherical concave interface between air (n = As shown in Figure 7.9, the cornea and aqueous
1.00) and glass (n = 1.50) has a radius of curvature humor of the eye can be approximated as an SSRI
of magnitude 5 cm. What is the dioptric power of between air and the ocular contents (n = 1.336). A
the interface when light is incident from the air typical radius of curvature of the cornea is 7.5 mm.
side? What is the dioptric power of the cornea?
Note that before calculating, since the interface The cornea is convex, so P must be positive. For
is concave, it is diverging and the dioptric power light going into the eye,
should be negative. From Figure 7.8, nl = 1.00, n 2 = 1.336, and r=+7.5mm.
= 1.00, 2 = 1.50, and r = - 5 cm. Then
Then, P=
P = n9 - n , (1.336-1.000)1000 mm/m
(1.50-1.00)100 cm/m +7.5 mm
P= 336 mm/m
-5 cm P= =+44.80 D.
50 cm/m +7.5 mm
P= = -10.00 D. A typical cornea has a dioptric power of almost
- 5 cm
+45 D, while a typical human eye has a dioptric
What is the dioptric power of the above inter- power of +60 D. Clearly, the cornea is the
face when the light is incident from the glass side? strongest converging element of the eye.
The interface is still divergent, so P must be
negative. Here n1 = 1.50, n 2 = 1.00, and r = +5 cm.
Since
P= 7.5 Focal Points

The secondary focal point (F 2 ) is the axial image point


that results when plane waves are incident on the
interface. Alternatively, the secondary focal point is
the image point conjugate to optical infinity. The
paraxial rays associated with the secondary focal point
are shown in Figure 7.10a.
Frequently, we use a larger scale vertically so that
we can easily see the rays (Section 4.12). The larger
vertical scale stretches out the spherical surface so that
FIGURE 7.8. Concave interface. it appears flat on the ray diagram (Figure 7.10b). This
114 Geometrie, Physical, and Visual Optics

1 / n2

a)

ni / n2

5^^--^^^

FIGURE 7.10. Paraxial rays. a. Actual, b.


Scaled.

is equivalent to saying that the paraxial rays are The secondary focal length f2 is the image distance
refracted at or near the vertex point of the interface. when the object is at optical infinity, or alternatively it
As an aid in remembering which way the spherical is the directed distance from the interface to the
surface is actually curving, dashed curves at the top secondary focal point F 2 . To find the secondary focal
and bottom of the vertical line represent the interface. length, we use
The center of curvature C provides the same informa-
V = P + U,
tion, but not as quickly.
The secondary focal point is a real image point for and set U equal to zero (incident plane waves). Then
a converging SSRI and a virtual image point for a
diverging SSRI (Figure 7.11). Once the location of the V=R
secondary focal point is known, the incident parallel Since
rays become predictable, and can be used in a ray
diagram. V

ni n2

>
> ~ ~~.;
C ^ F 2

y
\^s '

i
b) ni > n2

FIGURE 7.11. Predictable rays as-


d) n2 > n, sociated with the secondary focal point.
Single Spherical Refracting Interfaces 115

and for this case


?=-^ (7.22)
f2 = v, fi "
one can combine the above three equations to obtain Equations 7.21 and 7.22 are similar to Eqs. 5.4 and
5.5 for the focal lengths of a thin lens in air except that
(7.21) now the indices of the different media appear. The
primary and secondary focal points are not equidistant
The primary focal point ( F J is the axial object from the interface since nx and n2 are not equal. By
point that results in plane waves leaving the interface. equating 7.21 and 7.22, we obtain
Alternatively, the primary focal point is the object
point conjugate to optical infinity. The primary focal (7.23)
point is a real object point for a converging interface n2
and a virtual object point for a diverging interface or
(Figure 7.12).
The rays that originate at the primary focal point (7.24)
fj n/
all leave the interface parallel to the axis. Once the
Equation 7.24 shows that in magnitude, the ratio of
location of is known, any incident ray associated
the focal lengths equals the ratio of the indices of
with Fj becomes predictable, and can be used in a ray
diagram. refraction. The larger the difference in the indices of
refraction, the larger the difference in the focal
The primary focal length fr is the object distance lengths. It also follows that the longer focal length is
for which the conjugate image point is at optical always associated with the higher index medium and
infinity, or alternatively it is the directed distance from vice versa.
the interface to F x .
The secondary focal length is the image distance for
To find the primary focal length, use
incident plane waves. As such, the secondary focal
V = P + U, length is really the radius of curvature of the wave-
front that leaves the interface when plane waves are
and set V equal to zero (exiting plane waves). Then incident. The exiting wavefront is physically in the
P=-U. medium of index n 2 , and the secondary focal length is
Since then associated with n2 as is indicated mathematically
in Eq. 7.21. This is true even when the secondary focal

and for this case


"- point is virtual, in which case the exiting wavefront is
diverging (Figures 7.11c and 7.lid).
Similarly, the primary focal point is the object
fi = U, point conjugate to optical infinity. As such, the pri-
mary focal length is really the radius of curvature of
one can combine the above three equations to obtain the incident wavefront that results in plane waves

ni
f "2^

r^^
Fl
X^J c
>v

a) Ik n2 >
n2 < n ,

^ n2

~0-^F, --cir^F,
^
/ -
*
FIGURE 7.12. Predictable rays as-
n 2 < : n-, d) n2 > n! sociated with the primary focal point.
116 Geometrie, Physical, and Visual Optics

leaving the interface. The incident wavefront is phys- 7.13c in which C and F! are on the same side. From
ically in the medium of index n 1} and the primary focal Eq. 7.25,
length is then associated with n l 5 as is evident mathe- f2 = r - f 1 ,
matically in Eq. 7.22. This is true even when the
primary focal point is virtual, in which case the inci- or in absolute values
dent wavefront is converging (Figures 7.12c and
7.12d). IfJHr-fJHCFj,
Besides Eqs. 7.23 and 7.24, another relationship where CT\ is the distance from C to Fl. Consequently,
exists between the focal lengths. By adding Eqs. 7.21 for the cases where C and F1 are on the same side, the
and 7.22 together, we obtain distance f2 is equal to the distance CT\.
f + f = ^_i
L
2 ~ M D D ' EXAMPLE 7.6
In Example 7.3, there was a +2.00 D SSRI be-
or tween water (n = 1.33) and glass (n = 1.53). What
= 2- are the focal lengths for the +2.00 D interface?
L
2 ^ Ll p For light incident from the water side,
But from Eq. 7.19, nj = 1.33 and n2 = 1.53.
n, - n , From Eq. 7.21,
P ' n2 _ (1.53)100 cm/m
so L
2 -
P +2.00 D '
f2 + f ! = r. (7.25)
or
Equation 7.25 says that the radius of curvature of 153 cm/m
the interface is equal to the sum of the primary and 2 ~~ ^ n n n = + 7 6 5 cm.
secondary focal lengths of the interface. This relation- The secondary focal point is 76.5 cm behind the
ship is a useful self-consistency check when doing interface (Figure 7.14).
SSRI calculations. It is also useful in setting up "quick From Eq. 7.22,
and dirty" ray diagrams.
In Figure 7.13a, F 2 and C are both on the right side f =-iii
of the interface. From Eq. 7.25,
or
f,=r-f2; 133 cm/m
f, = - -66.5 cm.
or in absolute values +2.00 D
The primary focal point is 66.5 cm in front of the
|f,l = |r- f2l = |cF 2 |, interface.
Let us use Eq. 7.25 as a self-consistency check.
where CF 2 is the distance from the center of curvature From Example 7.3, the radius of curvature of this
C to F 2 . Consequently, when making a quick ray interface is +10 cm. From Eq. 7.25,
diagram for this case, one can select the horizontal
scale, plot C and F 2 , and then simply insert on the r = f!+f 2 ,
opposite side at a distance equal to CF 2 . r = -66.5 cm + 76.5 cm = +10.0 cm,
A similar relationship exists for Figures 7.13b and which checks.

FIGURE 7.13. Center of curvature and


focal points.
Single Spherical Refracting Interfaces 117

-*- +

FIGURE 7.14. A +2.00 D SSRI.

EXAMPLE 7.7 2.40). Thus, the incident light is in the diamond,


In Example 7.4, a -10.00 D SSRI between air and and
glass (n = 1.50) had a radius of curvature of - 5 cm
for light incident from the air side. What are the =2.40, while n2 = 1.40.
focal lengths? Here, nl = 1.00 and n2 = 1.50. From Then either Eq. 7.21 or Eq. 7.22 can be used to
Eq. 7.21, calculate P. From Eq. 7.22,
(1.50)100 cm/m _ n 1 = -(2.40)100cm/m
hf =5*
p = -10.00 D f, ~ +40.00 cm
150 cm/m
f2 = = -15.0 cm. ^ -240 cm/m , ^
-10.00 D
+40.00 cm
The secondary focal point is virtual and 15 cm in
front of the interface (Figure 7.15). From Eq. 7.22, You should check that Eq. 7.21 gives the same
value for P.
-(1.00)100 cm/m
fl
P -10.00 D
or
-100 cm/m 7.6 The Nodal Point
f1 = = +10.0 cm.
-10.00 D
The primary focal point is virtual and located 10 cm Consider a point source located at the center of
behind the interface. curvature of an SSRI (Figure 7.16). The wavefronts
The self-consistency check is: leaving the point source are spherical and centered on
r = f 1 +f 2 , the point source. Since the point source is located at
r = (10.0 cm) + (-15.0 cm) = -5.0 cm, the center of curvature of the interface, the wavefront
reaching the interface has the same curvature as the
which checks. interface, and all parts of the wavefront enter the
second medium at the same time. Consequently, the
EXAMPLE 7.8
A single spherical refracting interface between plas- whole wavefront changes speed at the same instant,
tic (n = 1.40) and diamond (n = 2.40) has a primary and no curvature change occurs as the wavefront
focal length of +40.00 cm and a secondary focal passes through the interface. The wavefronts leaving
length of -23.33 cm. What is the dioptric power of the interface are spherical and still centered on the
the interface? center of curvature of the interface.
The secondary focal length is negative so the Because there is no wavefront curvature change,
interface is a diverging interface and P must be the rays associated with the wavefronts are straight
negative. lines that pass through the interface without bending
Since the longer focal length is associated with (Figure 7.16b). This is consistent with Snell's Law,
the higher index medium, the primary focal length
(+40.00 cm) is associated with the diamond (n = sinGj = n2 sin6 r ,

-** +
\ Fi

W\ FIGURE 7.15. An SSRI with a 15 cm secondary


-15 cm- -+10cm
focal length.
118 Geometrie, Physical, and Visual Optics

n2

) )

a) FIGURE 7.16. Point source at the center of


curvature, a. Wavefronts. b. Wavefronts and
rays.

FIGURE 7.17. Nodal rays. a. b. Actual, c. d. Scaled.

since a ray from the center of curvature is normal to image. The image size is I and the object size is O. As
the interface so that the incident angle 4 is equal to before, the sizes are considered positive for objects or
zero, and thus the angle of refraction 9r is also zero. images above the axis and negative for objects and
The rays that pass straight through an interface are images below the axis. The lateral magnification m is
the nodal rays, and hence the nodal point of an SSRI defined as the ratio of the image size I to the object
is at the center of curvature of the interface (Figure size O, or
7.17). For a thin lens in air, the nodal point is at the m = I/0. (7.26)
lens itself. But for the SSRI, the nodal point is not at
the interface, but rather at the center of curvature of The image subtends the angle w' at the nodal point C,
the interface. and from Figure 7.18
The displacement of the nodal point away from the -I
interface is a result of the asymmetry of the SSRI. A tan w = .
v- r
thin lens in air is symmetric in the sense that the same
media is on both sides of the lens. An SSRI has a The object subtends the angle w at the nodal point C
different medium on each side, and thus is asym- and from Figure 7.18
metric. O
In general, any ray associated with the nodal point tan w = +
(either passing through it or pointing toward or away
Note that M is a negative quantity in Figure 7.18, so
from it) is a predictable ray, and passes straight
that u is a positive quantity.
through the interface without bending. The paraxial
Since the nodal ray is a straight line, the angles w
ray diagrams with the expanded vertical scale are
and w' are equal and from the above two equations
shown in Figures 7.17c and 7.17d.
-I O
v- r -w + r '

7.7 Lateral Magnification


m = (7.27)
Figure 7.18 shows a real object, its conjugate real x u
image, and the nodal ray connecting the object and Equation 7.27 can be used to calculate the lateral
Single Spherical Refracting Interfaces 119

FIGURE 7.18. Nodal ray and image size.

magnification; however, an easier expression can be for a thin lens in air by the presence of the indices of
derived as follows. refraction.
The basic vergence Eq. 7.16 is For an SSRI,
V = P + U, and U = ^ .
v u
which can be written as
n0 Mn, n, When the vergence equations are substituted into Eq.
7.29, the result is
V X + U
U
A common denominator can be found: m= (7.30)
n2rw _ n2vu nxvu \
vru VTU vru vru Equation 7.30 is a straightforward relationship be-
tween the lateral magnification and the vergences U
The common denominator can be eliminated to give and V. Furthermore, Eq. 7.30 is identical in form to
n2rw = n2vu nxvu + n^v, Eq. 5.8 for thin lenses in air. This identity occurs
because the vergences U and V implicitly contain the
or indices or refraction. Equation 7.30 makes lateral
n2rw - n2vu = r\xvu + n^y. magnification calculations extremely easy.
Then
n2u(r - v) = nji;(r - w),
7.8 Imaging Examples
or
rv (7.28) EXAMPLE 7.9
n9w T u' A goldfish in water (n = 1.33) is 40 cm from a
Equations 7.27 and 7.28 can be combined to give +5.00 D water-high index glass (n = 1.83) interface
(Figure 7.19). First, use a ray diagram to find the
nji;
(7.29) image, and determine whether it is virtual or real,
m = n0u ' erect or inverted, larger or smaller. Then answer
the same questions by calculation.
Equation 7.29 is an equation for the lateral magnifica- In order to draw the ray diagram, we first need
tion m in terms of the image distance v, the object to find the focal points and the radius of curvature
distance w, and the indices of refraction nl and n 2 . of the interface. Here n1 = 1.33, n2 = 1.83, u =
Equation 7.29 differs from the corresponding Eq. 5.7 -40 cm.

glass (n=i.83)

to>
water

v FIGURE 7.19. Goldfish example.


40 cm
120 Geometrie, Physical, and Visual Optics

From Eq. 7.19, The image distance of +109 cm agrees with the ray
diagram. The lateral magnification is
U -3.32 D
m = - - = -1.98.
(1.83-1.33)100 cm/m + 1.68 D
+5.00 D The image is inverted and 1.98 times larger than
the object, and that agrees with the ray diagram.
50 cm/m For instructional purposes, let us also use Eq.
= +10.0 cm.
+5.00 D 7.29 to calculate m:
From Eq. 7.21, riji;
_ n 2 _ (1.83)100 cm/m m = n.w '
2
~T~ +5.00 D (1.33)( + 109cm)
183 cm/m m = (1.83)(-40cm) = -1.98.
U = +5.00 D = +36.6 cm.
Note that the indices are needed in the above
From Eq. 7.22, equation, and that we cannot obtain m by merely
n - 1 3 3 cm/m _, , taking the image distance over the object distance.
f11 = = ^ ^ = -26.6 cm.
P +5.00 D
EXAMPLE 7.10
Equation (7.25) can be used as a self-consistency
The goldfish in the previous example swims in to a
check: distance of 14 cm in front of the interface. Now
r = f,+f 2 , where is the conjugate image? Is it real or virtual,
r = (-26.6 cm) + (+36.6 cm) = +10.0 cm. erect or inverted, larger or smaller?
The focal points and nodal points are un-
Figure 7.20 shows F n F 2 , C, and the predictable
changed. The primary focal length is -26.6 cm, so
rays. the fish is inside F j . Therefore, we would expect a
From the ray diagram, the image is real, inver- virtual image that is erect and larger. A quick ray
ted, larger, and approximately 100 cm from the diagram (Figure 7.21) confirms these expectations.
interface. Note that the nodal ray goes through C The calculations are:
and not through the pole of the interface. The _ n^ _ (1.33)100 cm/m
calculations are:
u - 1 4 cm '
V = P + U,
u = 1 3 3 c m A n = _ 9 5 0 D
_ ni _ (1.33)100 cm/m
- 1 4 cm
u - 4 0 cm
Then
133 cm/m
U = -3.32D. V = P + U,
-40 cm
Then V = +5.00 D + (-9.50 D),
V = +5.00 D + (-3.32 D), V=-4.50D.
V=+1.68D. The light leaving the interface is in glass, and is
The light in the glass leaving the interface is con- diverging.
verging, which agrees with the ray diagram. The image distance is
n2 183 cm/m
y= = = 40 67cm
"=V' v ^n5ir - - '
183 cm/m which agrees with the ray diagram. Note that the
= +109 cm. image position is at the center of curvature of the
+ 1.68 D

FIGURE 7.20. Ray diagram for Figure


7.19.
Single Spherical Refracting Interfaces 121

FIGURE 7.21. Goldfish inside the


primary focal point.

wavefronts leaving the interface and these wave- 40.67 cm+ 5 cm). The magnitude of the vergence
fronts are physically in the glass. at B is
The lateral magnification is . , (1.83)(100cm/m)
U _ -9.50 D 1 BI
45.67 cm
V " -4.50 D = +2.11.
m = 77 =

So the image is erect and 2.11 times as large as the |VB| = H = 4 . 0 1 D .


object, which also agrees with the ray diagram. Since the light is diverging at B, the vergence is
negative and
EXAMPLE 7.11
For the previous example compute the vergence of VB = -4.01 D.
the light at 5 cm in front of the interface (point A Thus, the light lost about a half diopter of minus
in Figure 7.22), and at 5 cm behind the interface vergence in traveling from the interface to B.
(point B in Figure 7.22).
The light incident on the interface had a ver- EXAMPLE 7.12
gence of -9.50 D. At 5 cm in front of the interface, A concave interface between air and plastic (n =
the light is in water and is diverging. The wavefront 1.44) has a butterfly imbedded in the plastic at a
there is curved more than the wavefront hitting the distance of 10.0 cm from the interface. The conju-
interface so that we expect the vergence to be more gate image is virtual and located a distance of
minus than -9.50 D. Position A is 5 cm closer to -6.02 cm from the interface. What is the dioptric
the object so the radius of curvature of a wavefront power of the interface? Also compute the lateral
at A is 9 cm (i.e., 14 cm - 5 cm). The magnitude of magnification and draw a ray diagram.
the vergence is The interface is concave from the perspective of
the lower index medium. Figure 7.23 shows a
(1.33)100 cm/m sketch of the given information. The light incident
|u A | = 9 cm on the interface is in the plastic and is diverging.
Here n, = 1.44, n9 = 1.00, and u = -10 cm. Then
133 cm/m
|u A | = 14.78 D. U=
9 cm
u
Since the wavefront is diverging, 144 cm/m
U= = -14.4D.
U A =-14.78 D. -10 cm
The light leaving the interface has a vergence of The light leaving the interface is in air and is
-4.50 D. At 5 cm behind the interface, the light is diverging away from the virtual image position. In
in the glass and is diverging. The center of curva- other words, a wavefront leaving the interface is in
ture of the diverging wavefront at B is at the image air and is centered on the virtual image. Here
position that is 45.67cm away from B (i.e., v = -6.02 cm.

^ +

FIGURE 7.22. Wavefronts for vergence


values.
122 Geometrie, Physical, and Visual Optics

The focal lengths are:


100 cm/m
f,= = -45.5 cm,
-2.2 D
nt -144 cm/m
f = =
' " 7 -2.2 D =+655cm-
The self-consistency check is
r = f 1 +f 2 ,
FIGURE 7.23. Butterfly example. r = +65.5 cm + (-45.5 cm) = +20.0 cm.
Figure 7.24 shows two of the three predictable rays.
You should draw in the predictable ray associated
Then with F j .
n9
EXAMPLE 7.13
100c^m=_
v = 1 6 6 D For the situation in the preceding example, calcu-
-6.02 cm late the vergence of the light at a position 4 cm in
front of the interface (A in Figure 7.24), and at a
From the vergence equation,
position 5 cm behind the interface (B in Figure
P=V-U, 7.24).
P = - 1 6 . 6 D - ( - 1 4 . 4 D), The wavefront that is physically at A is diverg-
ing, in the plastic, and centered on the object. The
P=-2.2D.
object is 10 cm from the interface, so position A is
The lateral magnification is 6 cm from the object, and the radius of curvature of
the diverging wavefront is then 6 cm in magnitude.
U
m= The magnitude of the vergence is
V
-14.4 D (1.44)(100cm/m)

m = - 1 6 . 6 D = +0.87. I^AI 6 cm
Alternatively, 144
|u A | = = 24.00 D,
,;
m= , or
n2w U A = - 2 4 . 0 0 D.
(1.44)(-6.02cm)
m
-(1.00)(-10.0cm)-+087 The vergence of the wavefront incident on the
interface was - 1 4 . 4 D , which is consistent with the
We need to find the center of curvature and the fact that the diverging wavefronts lose curvature as
focal points of the interface in order to draw the they propagate from A to the interface.
ray diagram. The radius of curvature is At B, the wavefront is diverging, in air, and is
n,-n, centered on the virtual image. The image is 6.02 cm
from the interface. The radius of curvature of the
wavefront at B has a magnitude of 11.02cm (i.e.,
(1.00-1.44)100 cm/m 6.02 cm + 5 cm). The magnitude of the vergence at
-2.2 D B is
-44 cm/m (1.00)100 cm/m
-2.2 D
+20.0 cm. |v B | = 11.02 cm
= 9.07D,

^+

FIGURE 7.24. Ray diagram for Figure


7.23.
Single Spherical Refracting Interfaces 123

or since the wavefront is diverging For instructional purposes, calculate the ver-
VB = -9.07D. gence of the wavefront at C, the center of curva-
ture of the interface. The wavefront at C is con-
The wavefront leaving the interface has a vergence verging, in air, and centered on the real image that
of -16.6D, and 5 cm downstream (at B) its ver- is 18.5 cm away from C (i.e., 38.5 cm - 20.0 cm).
gence is -9.07 D. The vergence at C is:
EXAMPLE 7.14 100
Converging light in plastic is incident on the same + 18.5 = +5.40D.
plastic-air interface discussed in the previous two Note that the wavefront leaving the interface has a
examples. The resulting virtual object is located vergence of +2.60 D, and its vergence has in-
30.0 cm from the interface. Find the conjugate creased to +5.40 D at C.
image, and specify whether it is real or virtual,
larger or smaller, erect or inverted.
From Example 7.12, r = +20 cm, f2 = 45.5 cm,
fj = +65.5 cm, and P = -2.2 D. 7.9 The Symmetry Points
Figure 7.25 shows two predictable rays. The
exiting rays show that the image is real, erect, and
larger. You should draw in the predictable ray For thin lenses in air, the plane 2F2 was conjugate to
associated with F 2 . the plane 2F 2 , and the lateral magnification was - 1 .
The calculations are: Let us now consider what conjugate plane gives a
u = +30 cm, lateral magnification of - 1 for an SSRI. Since
n, m - - - - l ,
U=,
u
144 U=-V.
U= +4.80 D.
+30 Then
Note that the incident converging wavefront is in V = P + U,
the plastic. Then
V = P + U, becomes
V = - 2 . 2 D + (+4.80D), V=P-V,
V=+2.60D.
or
The light leaving the interface is in air and converg- 2V=P,
ing in agreement with the ray diagram. The image
distance is: and
n0
v-i-
=
V'
100
v = +2.60 D = +38.5 cm. Then the expression for V and P in Eqs. 7.18 and 7.21
can be inserted to obtain
The lateral magnification is: n
n2 2
U +4.80 D
=

m = = +2.60 D = +1.85.
i; 2f2'

So the image is erect and 1.85 times larger than the or


object. v = 2f7

plastic - +
f
/

'~s6
L t r :

F2 Sc cL^) F 7
^
S^ object
FIGURE 7.25. Incident converging light in
the plastic.
124 Geometrie, Physical, and Visual Optics

The image plane is again at 2F 2 . r = f 1+ f 2 ,


Similarly, it is easy to show that r = +18.75 cm + (-30.0 cm) = -11.25 cm.
u = ~2fl, Figure 7.26 gives the ray diagram. You could also
so that the object plane is again at 2F t . draw in the nodal ray. For instructional purposes,
let us also consider the vergences U and V.
Therefore, just as for a thin lens in air, 2F t is
conjugate to 2F2 and the lateral magnification is - 1 . u = n 1 = _J50_
= +4.0D,
The axial points 2Fj and 2F2 are called the symmetry u +37.5 cm
points, and the planes containing the symmetry points 240
are called the symmetry planes. Since the primary and = -4.0D,
V -60.0 cm
secondary focal points of an SSRI are not equidistant and then
from the interface, the symmetry points are also not
equidistant. When the interface is a converging inter- U +4.0 D
m =
face, the symmetry points are both real, and when the V=^4^D=" 1 0

interface is a diverging interface the symmetry points
are both virtual.
7.10 Unit Magnification
EXAMPLE 7.15 For thin lenses in air, an object distance of zero
Find the symmetry points for a -8.00 D single resulted in an image distance of zero and a lateral
spherical refracting interface between glass (n =
1.50) and diamond (n = 2.4) with the light incident magnification of +1. Let us check if this also holds for
from the glass side. Here nl = 1.50, and n2 = 2.40. an SSRI.
Then From
240 V = P + U,
U= =-30.0 cm,
-8.00 n,== P + ^
and V
u
-150 Then
11
P = +18.75 cm.
-8.00 n2 wP + nj
For the symmetry points: V u
u = 2f1= 2( + 18.75 cm) = +37.5 cm, or
v = 2f2 = 2(-30.0 cm) = -60.0 cm. V u
n, + ^
The lateral magnification is:
n^ As the object is moved closer and closer to the
m= , interface, u gets closer and closer to zero, and the wP
n2w term in the denominator becomes negligible compared
(1.50)(-60.0cm) _ to nj. So in the limit of u going to zero.
(2.40)(+37.5cm) i; u
The radius of curvature of the interface is

-*+
glass (n = 1.50) diamond (n == 2.40)

/1
virtual object
^ - - ' F C
/
2
virtual image

s >
FIGURE 7.26. Symmetry points for a diverging
interface.
Single Spherical Refracting Interfaces 125

Since image distance goes to zero and the lateral magnifica-


n^ tion goes to + 1 . This means that when an object is
m= nu' placed directly in contact with the SSRI, the interface
0
essentially has no imaging effect.
we can combine the above two equations to obtain
m = +1.
Therefore, for an SSRI, an object distance of zero
7.11 The Nodal Plane
gives an image distance of zero and a lateral magnifi-
cation of -Hi.
In addition to a zero object distance, there is another
case in which the object and image are both the same
EXAMPLE 7.16
Consider a +6.00 D convex air-glass (n = 1.50) in- distance away from the SSRI. This case involves the
terface. Find the image distance and lateral magni- nodal point that is at the center of curvature C of the
fication for a piece of paper in air at distance of SSRI. Figure 7.27 shows an object in the nodal plane
1 cm, 0.1 cm, and 0.01 cm. Does the lateral magni- of a converging SSRI. Note that the focal points in
fication go to +1 as u goes to zero? Here n1 = 1.00 Figure 7.27 are placed so as to satisfy Eq. 7.25. The
and n2 = 1.50. The calculations for 1 cm are: two predictable rays show that the image is virtual,
u = - 1 cm, located in the same plane as the object, and larger
than the object.
u-*-^IOOD, The ray diagram can be confirmed analytically as
u -1 follows:
V = P + U, V = P + U,
V=+6.00D + (-100D) = -94.0D,
n2^ n2-ni i
n2 r
150 = -1.60 cm, v u
-94.0 Let u = r. Then
-100D
m = = -94 D = +1.06.
n
2 = 2~ni + 1
v r r '
The results for the other object distances are com-
puted similarly, and shown in Table 7.1. or

v r
TABLE 7.1 and
Lateral Magnification as Object Distance Goes to Zero
v = r.
u (cm) v (cm) The lateral magnification is
nji;
-1.00 -1.60 + 1.06 m n2w '
-0.10 -0.15 + 1.006
-0.01 -0.015 + 1.0006 n
i r

m = n7r (7.31)

The SSRI shown in Figure 7.27 is converging, and the


From the numerical results presented in Table 7.1, higher index medium must be on the left. Therefore,
it is clear that as the object distance goes to zero, the for this case, nt must be greater than n 2 . Consequent-

cjlass

image < .
^ -- [object\
Fi c FT-

FIGURE 7.27. Object in the center of curva-


ture plane.
126 Geometrie, Physical, and Visual Optics

ly, m is greater than + 1 , and the image must be erect j\ student's eye
and larger than the object, which agrees with the ray
diagram.

EXAMPLE 7.17 glass paperweight


A glass (n = 1.50) hemisphere of radius 5 cm is
used as a paper weight. A 2-mm tall letter is
centered under the hemisphere (Figure 7.28). The letter
light diverging away from the letter is in the glass
so the letter effectively acts like an object in the FIGURE 7.28. Paperweight example.
glass (i.e., refer back to the zero object distance
discussion). Since the letter is centered, it is at the
center of curvature of the interface. Where is the The image size is
conjugate image? What is its size? Here ni = 1.50 I = mO,
and n 2 = 1.00 Then I = ( + 1.50)(2mm) = 3mm.
u = r = - 5 cm, An observer looking at the letter through the
and the ray diagram in Figure 7.27 describes the paperweight would see a virtual image at the same
situation. The dioptric power is given by position as the letter but 1.5 times larger than the
letter.
P= The adjective apparent is used to describe vari-
ous aspects of the image formed by an optical
(1.00-1.50)100 cm/m system placed between the object and an observer's
P=
- 5 cm eye. The apparent position of the letter is 5 cm
from the interface, and its apparent size is 3 mm.
P=^ = +10.0D. Note that the word apparent refers to the charac-
teristics of the interface's image (virtual in this
The incident vergence is given by case), and not necessarily to the visually perceived
image of the observer.
150
U = -30.0 D.
-5 In the preceding discussion, the image is larger
Then than the object by the ratio of the indices of refrac-
V = P + U, tion. In general, an image at the center of curvature of
V = +10.0 D + (-30.0 D) = -20.0 D. the interface can be either larger or smaller than its
conjugate image by the ratio of the indices of refrac-
The light leaving the interface is in air and diverg- tion. Figure 7.29 shows a ray diagram for a converging
ing. The image is virtual, and the image distance is interface with a virtual object and a smaller real image
100 at the center of curvature.
v = V -20.0 D '
i; = 5 cm,
which is equal to r.
7.12 Strange Cases
The lateral magnification is
Chapters 4, 5, and 6 developed optics intuition by
U = -30.0 D considering thin lenses in air. Much of this intuition
+ 1.50.
V " -20.0 D has carried over to the discussion of SSRIs. In particu-

glass
^XsT* real
virtual

TF^J * object
F
<^* C '^\.

FIGURE 7.29. Virtual object at the center of cur-


vature.
Single Spherical Refracting Interfaces 127

FIGURE 7.30. Real object inside the center of curva-


ture for a converging interface.

lar, the role of the focal points Fl and F 2 is the same,


the symmetry points are again at 2F1 and 2F 2 , and an P=
object distance of zero gives an image distance of zero , _ ( 1 . 0 0 0 - 1.336)1000 mm/m
with a lateral magnification of + 1 .
- 7 . 5 mm
However, there is no thin lens in air analog for an
object at the center of curvature of an SSRI. In that -336.0
P= = +44.8D,
sense, an object at the center of curvature of an SSRI -7.5
constitutes a strange case. Actually, all SSRI cases for and
which an object or an image falls between the inter- TJ_i _ (1.336)1000mm/m
face and the center of curvature are strange in that u - 3 . 6 mm
they have no thin lens in air analog, and thus our thin 1336.0
lens in air intuition breaks down. U= -371.1 D.
-3.6
Consider the ray diagram in Figure 7.30. The inter-
From
face is converging. The object is inside Fl and inside
C. Our thin lens in air intuition would tell us to expect V = P + U,
the image to be virtual, larger, and further from the V = + 4 4 . 8 D + ( - 3 7 1 . 1 D),
interface than the object. In the ray diagram, how- V = -326.3 D.
ever, the two predictable rays show that while the
The light leaving the cornea is in air and the image
image is indeed virtual and larger, it is closer instead is virtual. The image distance is
of being further away.
_ 2 _ 1000 mm/m
V
EXAMPLE 7.18 ~ V ~ -326.3 D '
Consider a simplified model of the anterior cham- v = -3.06 mm.
ber of the eye (Figure 7.31). Assume that the
So the virtual image is at 3.06 mm which is closer to
cornea and the anterior chamber both have an
the cornea than the iris is.
index of 1.336, and that the cornea is spherical with
The lateral magnification is
a radius of curvature of 7.5 mm. The iris of the eye
sits in the aqueous humor about 3.6 mm from the U -371.1 D
cornea. When you look at the iris you do not see m = - = -326.3 D = +1.14,
the iris itself, but rather an image of the iris formed
so the virtual image is larger than the iris.
by the cornea. Find the image position and the
Thus, the apparent iris that you see is both
lateral magnification.
closer and larger than the actual iris. You might
The light leaving the iris is initially in the aque- keep this in mind the next time you are gazing into
ous humor. Therefore, n1 = 1.336, n 2 = 1.000, and someone's beautiful eyes.
r = - 7 . 5 mm. Then

EXAMPLE 7.19
An object in air is located 12.5 cm from a -2.00 D
air-glass (n=1.50) interface. Find the conjugate
image and specify whether it is real or virtual, erect
or inverted, larger or smaller, and closer or further
from the interface. Here n1 = 1.00 and n 2 = 1.50.
The focal lengths and the radius of curvature
can be determined from the usual equations. The
results are as follows r = -25.0 cm, fj = +50.0 cm,
FIGURE 7.31. Iris as object for the cornea. and f, = -75.0 cm.
128 Geometrie, Physical, and Visual Optics

air ^S glass

image _ ^ j r "" """*"


^2 y^^ object-^ Fi

FIGURE 7.32. Real object inside


the center of curvature for a diverg-
ing interface.

Figure 7.32 shows two of the predictable rays. 7.13 Object Space/Image Space
(You should supply the third predictable ray.) Ac-
cording to our thin lens in air intuition, an object in The concepts of object space and image space were
front of a minus system would result in a virtual introduced in Section 3.9. The set of all points, rays,
image that is erect, smaller, and closer. The ray or ray extensions associated with the light incident on
diagram shows that this is a strange case because an SSRI constitutes the object space for the interface.
the virtual image is erect, smaller, and further from
the interface than the object is. The calculations All object space entities are associated with the index
are as follows: nl. The set of all points, rays, or ray extensions
associated with the light leaving the interface consti-
u = -12.5 cm, tutes image space. All image space entities are as-
U=-8.00D, sociated with the index n 2 .
V = P + U = -2.00 D + (-8.00 D), Remember that object space and image space are
mathematical sets. An object space point might be in
V= -10.00 D, front of the interface (as for a real object) or behind
n0 150 the interface (as for a virtual object). Similarly, an
= -15 cm, image space point might be behind the interface (as
V -10
U -8.00D for a real image) or in front of the interface (as for a
m virtual image).
" V " -10.00 D - + 8
Consider a glass-air interface (Figure 7.33). A real
image in air immediately indicates that the light leav-
EXAMPLE 7.20
An ophthalmic crown glass (n = 1.523) lens has a ing the interface is in air and is converging. Similarly,
front surface power of +14.00 D, a back surface a virtual image in air indicates that the light leaving
power of -3.00 D, and a central thickness of 8 mm. the interface is in air and diverging. Pay special atten-
What is the apparent thickness of this lens as tion to the fact that the virtual image in Figure 7.33b
viewed from the front? has no association with the incident light that is phys-
The question is essentially asking for the image ically in the glass. Based on the information given so
position of the back surface of the lens as seen far, we do not know whether the incident light is
through the front. The object distance is converging or diverging. Remember that the virtual
u = - 8 mm, image point is the center of curvature of the exiting
and diverging wavefront that is physically in the air.
Figure 7.34 shows a water-plastic interface. A real
11! = 1.523 while n 2 = 1.000. object in water indicates immediately that the light
Then incident on the interface is in the water and is diverg-
u=n1=1523mm/m=1904 ing. Similarly, a virtual object in water indicates that
u - 8 mm the light incident on the interface is in the water and is
V = P + U, converging. Pay special attention to the fact that the
virtual object in Figure 7.34b has no association with
V = +14.00 D + (-190.4 D) = -176.4 D, the exiting light that is physically in the plastic. Based
1000 on the information given so far, we do not know
= v = - -190.4 = -5.7 mm. whether the exiting light is converging or diverging.
The image of the back surface is virtual and located Remember that the virtual object point in Figure
5.7 mm from the front surface. Thus, the apparent 7.34b is simply the center of curvature of the incident
thickness, as seen from the front, is 5.7 mm. converging wavefront that is physically in the water.
Single Spherical Refracting Interfaces 129

real image- FIGURE 7.33. a. Real image in air.


a) b. Virtual image in air.

water water plastic

virtual object
real object

FIGURE 7.34. a. Real object in


water, b. Virtual object in water.

Imagine an interface between glass and plastic with front of an SSRI, but now the nodal point is at the
a virtual object in plastic. What does this mean in center of curvature of the interface and not at the
terms of the incident light? The virtual object indicates interface itself.
that the incident light is converging. In plastic indi- Figure 7.35a shows a converging interface and the
cates that the incident light is physically in the plastic. nodal ray for a distant object that subtends an angle w
So converging light in the plastic is incident on the at the nodal point. The image is in the secondary focal
interface. The virtual object location is at the center of plane, which is a distance of f2 minus r from the nodal
curvature of the incident converging wavefront. You point. The angle w' that the image subtends at the
should draw a sketch similar to Figure 7.34b for this nodal point is equal to w. From the triangle involving
situation. the C, the optical axis, and the image:
Imagine an interface between beer and glass with a |l| = ( f 2 - r ) t a n w . (7.32)
virtual image in the beer. What does this mean in
terms of the exiting light? The virtual image indicates Note that Eq. 7.32 has a different form than the
that the exiting light is diverging. In beer indicates that analogous Eq. 5.11 for a thin lens in air.
the exiting light is physically in the beer. So diverging
light in the beer is leaving the interface. The virtual EXAMPLE 7.21
image location is at the center of curvature of the A distant object in air is in front of a long glass
diverging wavefront that is leaving the interface. You (n=1.50) rod with a +1.00 D front surface. The
should draw a sketch similar to Figure 7.33b for this object subtends an angle of 2 at the interface.
situation. Where is the conjugate image? What is its size?
As suggested in Section 3.9, it sometimes helps to
color code the incident light and everything associated f2 = y = ^ y = +150cm,
with it (i.e., everything in object space), and use a
different color for the exiting light and everything n9 n, 50
= +50 cm,
2
associated with it (i.e., everything in image space). P +1
I = 100 cm tan w = (100 cm)(0.0349) = 3.49 cm.
It is interesting to note that
7.14 Imaging a Distant Object
The calculation of the image size for a distant object in so that Eq. 7.32 can be written
front of a thin lens in air was discussed in Section 5.13. |l| = |fi|tanw. (7.33)
There, the angle subtended at the nodal point of the
thin lens in air was used to calculate the image size. Figure 7.35b shows the geometry involved. The paral-
The same technique can be used for a distant object in lel ray in image space shows that the distance AB is
130 Geometrie, Physical, and Visual Optics

\ n1 ' n2

n2
f 2 -r
W| ^ ^ F2
F
'iSsS. A C\.
I

"V" B FIGURE 7.35. Image size for a dis-


f2 tant object, a. Nodal ray. b. Ray as-
a) b)
sociated with the primary focal point.

equal to the image size I, and then the above equation eliminated and an equation derived that relates the old
follows directly from the triangle ABFj. dioptric power to the new dioptric power.
In the original situation, we first found the radius of
curvature from
(n2~ni)old

7.15 Changing Media Without P


Changing Curvature
Then we used r to get the new dioptric power from
It is instructive to consider the dioptric power changes (n2-ni)ne
P =
that occur for an SSRI when a media change occurs x
new
without a change in curvature. We can then combine the above two equations to
obtain
(n2~ni)new
EXAMPLE 7.22 P = (7.34)
Suppose we have a +2.00 D SSRI between water (n2_ni)oK
(n = 1.33) and glass (n = 1.53), and the water is Equation 7.34 says that to get the new dioptric power,
drained off. What is the dioptric power of the one can take the old dioptric power, divide out the old
resulting air-glass interface? difference in indices (which we do not want anymore),
Let us assume that the light is initially incident and multiply in the new difference in indices. When
in the water, although this assumption is not crucial
since the dioptric power is invariant relative to the the new difference in indices is greater than the old
direction that the light is traveling. Here n1 = 1.33 difference in indices, the new dioptric power will be
and n2 = 1.53. greater in magnitude than the old dioptric power.
From Eq. 7.19 When the new difference in indices is less than the old
difference in indices, then the new dioptric power will
be less in magnitude than the old dioptric power.
(1.53-1.33)100 cm/m EXAMPLE 7.23
r= +2.00 D Let us reconsider the case in Example 7.22. There,
20 a +2.00 D water (n = 1.33)-glass (n = 1.53) inter-
r = = +10 cm. face had the water drained off. Since the index
For the air-glass interface, difference increases, the new dioptric power is
greater than the old dioptric power.
n^ - n , The old indices are:
P=
n2 = 1.53 and nx = 1.33.
(1.53-1.00)100 cm/m The new indices are:
P=
10 cm
n2 = 1.53 and 1^ = 1.00.
53
P= +5.30 D. Then from Eq. 7.34
10
The dioptric power of the air-glass interface is (1.53- 2 00
+5.30 D, which is considerably higher than the ^Ni--S^ ^
+2.00 D of the water-glass interface. Since the 0.53
curvature is unchanged, the increase is due to the (+2.00 D),
change in the index difference. 0.20
Pnew = | ( + 2 . 0 0 D ) ,
Actually, in the case of changing media while the
curvature is unchanged, the radius can be algebraically Pnew = (2.65)(+2.00D)=+5.30D.
Single Spherical Refracting Interfaces 131

The +5.3 D is the same result as that obtained in


the previous example, and the increase in dioptric
power agrees with our expectations. Note that with
a little practice, the next to the last step (the one
containing 53/20) can be written down im-
mediately.

EXAMPLE 7.24
Consider a -13.00 D air-plastic (n = 1.44) inter-
face. The room is suddenly flooded with water
(n = 1.33). What is the dioptric power of the result-
ing water-plastic interface? FIGURE 7.36. Lens clock.
Stop! Before calculating, do you expect a diop-
tric power that is larger or smaller in magnitude?
The difference in indices decreases, so the diop-
tric power should decrease in magnitude. For the
calculations, let us assume that the light is initially
in air, although the assumption is not crucial. lated directly from the lens clock reading and the
The old indices are: differences in indices.
^ = 1.00 and n2 = 1.44. The calculation is based on the fact that the lens
clock is actually measuring a sagitta, and the sagitta is
The new indices are: directly related to the curvature. When two SSRIs
nl = 1.33 and n2 = 1.44. have the same curvature, the lens clock reading is the
Then from Eq. 7.34 same on both surfaces. In effect, when going from one
surface to the other, the media is being changed
_ (1.44-1.33) without changing the curvature. Hence, Eq. 7.34 can
(-13.00 D),
(1.44-1.00) be used with the old dioptric power as the lens clock
11 reading, and the new dioptric power as the true
Pe= 44 (-13.00 D), dioptric power of the surface, i.e.,
Pnew = -3.25D.
p _ \n2 ~ n i ) t r u e p , <-\
The -3.25 D agrees with the expectations of a
r r
true (r\ n \ clock * \I.JJ)
Vn2
1 J clock
dioptric power smaller in magnitude.
The true dioptric power is larger in magnitude than
the lens clock reading when the true index difference
is larger than the assumed index difference, and vice
versa.
7.16 Lens Clock Readings

A lens clock (also known as a lens measure) is a EXAMPLE 7.25


three-legged device that actually measures the sagitta A lens clock calibrated for an index of 1.53 reads
-4.50 D on a spherical interface with a 1.71 index.
of a spherical surface (Figure 7.36). However, the What is the true dioptric power to the nearest
sagitta is directly related to the curvature, and for a quarter diopter?
given set of indices the curvature is directly related to Stop! Before calculating, do you expect the true
the dioptric power. Therefore, the lens clock can be dioptric power to be larger or smaller in magnitude
calibrated directly in terms of the dioptric power of an than the lens clock reading? From Eq. 7.35
assumed index of refraction. _ (1.71-1.00)
The tools for making glass spectacle lenses are true (-4.50 D),
(1.53-1.00)
usually calibrated for a 1.53 index of refraction. Typi-
cally, lens clocks are also calibrated for the 1.53 index, Ptrue = ^ ( - 4 . 5 0 D ) ,
although a few are calibrated for a 1.49 index of
refraction. A lens clock, calibrated for index 1.53, Ptrue = -6.03 D ~ -6.00 D.
reads the true dioptric power on an air-glass SSRI As expected, the true dioptric power is larger in
provided 1.53 is the index of the glass. magnitude that the lens clock reading since the 71
When the lens clock is used on an SSRI with an was multiplied in and the 53 was divided out. Note
index different from the calibrated index, then it gives that with a little practice, the line containing 71/53
an incorrect reading. The true power may be calcu- can be written down immediately.
132 Geometrie, Physical, and Visual Optics

EXAMPLE 7.26 a. a real object in air 50 cm from the interface.


A lens clock, calibrated for a 1.53 index, reads b. a real object in air 10 cm from the interface.
+ 14.00 D on a hard resin plastic (n = 1.49) spheri- c. a virtual object in air 20 cm from the interface.
cal surface. To the nearest quarter diopter, what is d. a virtual object in air 50 cm from the interface.
the true dioptric power of the interface? Before e. a real object in glass 25 cm from the interface.
calculating, do you expect the true dioptric power
to be larger or smaller? f. a virtual object in glass 10 cm from the in-
The difference in indices is smaller than the terface.
difference assumed for the lens clock, so the true g. a virtual object in glass 45 cm from the in-
dioptric power is also smaller. From Eq. 7.35 terface.
3. An air bubble is 5 cm from the center of a glass
(n = 1.61) sphere of radius 12 cm. Where does the
bubble appear to be to an observer, (a) when
looking in at the side nearest the bubble? (b)
Ptrue= 53 ( +14.00 D), when looking in at the side farthest from the
p
true = +12.94 D ~ +13.00 D. bubble?
4. Given a plastic (n = 1.49) sphere of radius 10 cm,
When properly used, lens clocks supply the needed calculate where parallel incident rays come to a
information very quickly. However, lens clocks are focus after passing through the sphere.
not smart. The people who use the lens clocks must 5. Given a paper weight of glass, index 1.60, radius
supply the intelligence. One can put a lens clock on a 15 cm convex, 18 cm thick; a printed letter is
rock, and it might read +4.00 D. This does not mean under the paper weight, flush against its flat face.
the rock has dioptric power. The index assumptions The letter is 6 mm high. Locate the position and
size of the conjugate image.
and the fact that a lens clock is actually measuring a
sagitta need to be kept in mind. 6. A thick lens is made of glass, index 1.50. The
front surface has a radius 25 cm convex. The back
surface has a radius 50 cm concave. The lens is
8.5 cm thick. Considering the lens as two single
spherical refracting surfaces find the primary and
secondary focal points of the lens.
Problems 7. A real object in water (n = 1.33) is located 25 cm
from a water-glass (n = 1.63) SSRI. The conjugate
1. Given a convex air-glass (n = 1.56) SSRI with a image is real and located in the glass 60.82 cm
14-cm radius of curvature, calculate conjugate from the interface. What is the lateral magnifica-
image positions and magnifications for the follow- tion? What is the dioptric power of the interface?
ing object points. Include a ray diagram for each 8. A 10-cm long glass (n= 1.53) rod has a +6.00 D
part. Also, calculate the vergence of the light 5 cm spherical front surface and a -2.00 D spherical
in front of the interface, and 5 cm behind the back surface. For plane waves incident on the
interface. front, what is the vergence of the wavefront leav-
a. a real object in air 1 m from the interface. ing the back surface?
b. a real object in air 50 cm from the interface. 9. Represent an aphakic eye with a +44.00 D cor-
c. a real object in air 35 cm from the interface. nea. If the cornea to retina length is 22.27 mm and
d. a real object in air 20 cm from the interface. the ocular index is 1.336, what is the aphake's
e. a virtual object in air 14 cm from the interface. spectacle correction at a 10 mm vertex distance?
f. a virtual object in air 40 cm from the interface. 10. Represent an eye by a +63.00 SSRI between air
g. a real object in glass 150 cm from the interface. and the ocular media (n = 1.336). If the length of
h. a real object in glass 40 cm from the interface. the eye is 19.0 mm, what contact correction (if
i. a real object in glass 15 cm from the interface. any) is needed?
j . a virtual object in glass 48 cm from the in- 11. A long plastic rod (n = 1.44) has a +3.80 D
terface. spherical front surface. What is the image size in
2. Given a concave air-glass (n = 1.52) SSRI with a the plastic for a distant object that subtends an
13-cm radius of curvature, calculate conjugate angle of 3?
image positions and magnifications for the follow- 12. Represent an emmetropic eye by a +56.00 D
ing object points. Include a ray diagram for each SSRI between air and the ocular media (n =
part, and calculate the vergence of the light 5 cm 1.336). What is the retinal image size (in microns)
in front of and 5 cm behind the interface. for a distant letter that subtends an angle of 5'?
Single Spherical Refracting Interfaces 133

13. The anterior surface of the crystalline lens is an 14. An air-glass (n = 1.53) interface has a power of
SSRI between the aqueous humor (n = 1.336) and -10.00 D. If the room is filled with water (n =
the crystalline lens cortex (n = 1.386). If the radius 1.33), what is the power of the water-glass inter-
of curvature is 10.0 mm, what is the dioptric face? If the room is filled with oil (n = 1.73), what
power of this surface in the eye? If the crystalline is the power of the oil-glass interface?
lens is removed from the eye, and the anterior 15. A lens clock calibrated for n = 1.53 reads +6.00 D
surface remains spherical with the same radius, on a spherical highlite (n = 1.71) glass surface.
what is the dioptric power of the air-crystalline What is the true dioptric power of the interface?
lens interface?
CHAPTER EIGHT

Plane Refracting
Interfaces and
Reduced Systems

8.1 Plane Interfaces as a Special Case direction of travel of the incident plane waves by
of Spherical Interfaces Snell's Law (Figure 8.1c and 8.Id),
nx sinOj = n2 sin9 r . (8.1)
Consider a series of single spherical refracting inter-
faces (SSRIs) where each successive interface is flatter When paraxial angles are involved, the small angle
than the preceding one. The SSRI equations in Chap- approximation to Snell's law is valid, i.e.,
ter 7 hold for each of these interfaces. The limiting
| = 2 (8.2)
case of this series is a flat or plane interface, and
hence, in the limit, the SSRI equations also hold for When n2 is greater than n1? the rays bend toward the
the flat interface. normal (Figure 8.Id). For example, the ray bends
As shown in Eq. 7.20, the dioptric power P of an towards the normal when light in air is incident on a
SSRI is directly proportional to the curvature R of the plane air-water interface. When nx is greater than n 2 ,
interface. As spherical interfaces are made flatter, the rays bend away from the normal (Figure 8.2). For
their curvature decreases and their dioptric power gets example, the ray bends away from the normal when
closer to zero. A flat or plane interface has zero light in glass is incident on a plane glass-air interface.
curvature and zero dioptric power. Note that the
radius r is reciprocally related to the curvature R, and
hence r goes to infinity for a plane interface.
Consider plane waves in a medium of index nt 8.2 Diverging and Converging Wavefronts
incident on a plane refracting interface between media
nl and n 2 . When the plane waves are incident normal- When diverging or converging wavefronts are incident
ly, all parts of the wavefront reach the interface at the on a plane interface, part of the wavefront reaches the
same time and all parts change speed at the same interface first, and this part changes speed before the
time. The wavefronts leaving the interface are still remaining parts. The result is that diverging or con-
plane waves and still normal to the interface (Figure verging wavefronts change their curvature as they pass
8.1a). through plane interfaces. Figure 8.3a shows the wave-
When the plane waves are incident on a plane fronts for a case in which nl is greater than n 2 . Figure
interface at an angle to the normal, they are neither 8.3b shows the rays associated with the wavefronts.
converged nor diverged so that the waves leaving the Since nt is greater than n 2 , each ray in Figure 8.3
interface are still plane waves, as expected for a zero bends away from the normal. The object point at A is
dioptric power interface. However, the direction of real, and the image point at B is virtual and closer
travel of the plane waves is deviated (Figure 8.1b). As than the object to the interface. The image location
discussed in Section 1.12, the direction of travel of the can be determined from the vergence Eqs. 7.16-7.18.
plane waves leaving the interface is related to the Since P is zero,
135
136 Geometrie, Physical, and Visual Optics

-** +
n2

a)

FIGURE 8.1. Plane waves incident on a flat inter-


face. a. Incident normal, b. Incident at an angle to
the normal, c. Wavefronts and rays. d. One ray.

V = P + U, Equation 8.3 relates the object and image distance for


the plane interface. It is also possible to derive Eq. 8.3
results in directly from the small angle version of Snell's law
v=u, without using the SSRI equations.
where
EXAMPLE 8.1
A small fish is swimming in water (n = 1.33) 100 cm
and below the (plane) surface. To an observer in air,
what is the apparent depth of the fish?
An observer in air looking at the fish underwa-
u ter sees the virtual image of the fish. In Figure 8.3,
For a plane interface, it follows that A represents the position of the fish and B repre-
v _ u sents the position of the virtual image. The appar-
(8.3) ent depth of the fish is just the location of the
n, n, ' virtual image. The fact that the rays bend away
from the normal indicate that the virtual image of
the fish is closer than 100 cm. In this case, nl =
1.33, n2 = 1.00, and u = -100 cm. Then

f ^ ^ * ^ l normal N n,
u
and
133
U= = -1.33D.
FIGURE 8.2. Ray incident in the higher index medium. -100

ni |/n 2

FIGURE 8.3. a. Diverging wave-


fronts incident on a flat interface, b.
Wavefronts and rays. c. Rays only.
Plane Refracting Interfaces and Reduced Systems 137

From EXAMPLE 8.3


What is the apparent thickness of a 9-cm thick flat
V=U, glass (n = 1.5) slab?
V=-1.33D; An observer in air looking through the slab does
not see the actual backside of the slab but instead
and from sees the virtual image of the backside. By drawing
a figure, such as Figure 8.3c, it is easy to predict
V= that the apparent thickness will be less than the
actual thickness.
100 The object distance u is - 9 cm and from Eq.
= -75 cm.
-1.33 D 8.3,
Alternatively, we can use Eq. 8.3 directly. - 9 cm
u 1.50 '
or
-100 cm v = - 6 cm.
v
1.33 The apparent thickness is 6 cm as compared to the
i; = - 7 5 cm.
actual thickness of 9 cm. (Note the same principle
applies to aquariums, which is why they look nar-
The apparent depth of the fish is 75 cm below the rower than they actually are.)
water's surface.
The (reduced) vergence of a wavefront is equal to
EXAMPLE 8.2 the index of the medium that the wavefront is in times
A mayfly is flying 50 cm above a pond. A hungry the curvature of the wavefront. Even though the
trout swimming in the water below the pond's curvature of a converging or diverging wavefront
surface sees the mayfly. To the trout, what is the changes at a plane interface, the (reduced) vergence of
apparent height of the mayfly? the wavefront does not change. This means that the
Figure 8.4 shows the situation for the trout. The curvature change is exactly offset by the index change.
light leaving the mayfly is initially in air (nx = 1.00),
and then enters the water (n2 = 1.33). In this case,
n2 is greater than nl9 and the rays bend toward the
normal. The trout sees a virtual image of the
mayfly, and the virtual image is farther than the 8.3 Lateral Magnification
mayfly from the surface. From Eq. 8.3,
-50 cm The lateral magnification of an SSRI is equal to U/V.
1.00 : Since the vergences U and V are equal for a plane
33
interface, the lateral magnification is + 1 . This is easily
or confirmed by using rays incident along the normal.
v = -66.5 cm. Figure 8.5 shows such a case for the fish underwa-
The apparent height of the mayfly is 66.5 cm above ter. The two rays incident normally are the vertical
the surface. lines from the front and back of the fish. The respec-

f>
FIGURE 8.4. The virtual image of the mayfly serves as the FIGURE 8.5. The virtual image is the same size as the object
object for the trout's eye. (the fish), so the lateral magnification is + 1 .
138 Geometric, Physical, and Visual Optics

v = - 1 5 cm.
The object distance for the emmetrope's eye is
ux = - ( 1 5 + 5) cm = - 2 0 cm.
For the emmtrope, U fp is zero and the accom-
d
> modative demand is
A0 = U f p - U x ,
or
100
A =- = +5D.
- 2 0 cm
The image distance for the emmetrope's eye is
FIGURE 8.6. Comparison for the observer viewing a fish that 16.67 mm, and the lateral magnification is
is an actual distance, d, away. When the water is drained, the
observer views the actual fish at the distance d. When the water is + 16.67mm /,
m = =
present, the observer views the virtual image at a distance less 9nn
2UU mm -0834
than d. The retinal image size I is
I = mO,
I = (-0.0834)(30 mm) = -2.50 mm.
tive virtual image points also lie along these two
vertical lines, so the virtual image of the fish is the
EXAMPLE 8.4b
same size as the fish itself.
The ice in Example 8.4a melts, and the emmtrope
Whenever any object is brought closer to an obser- can now hold the arrowhead in air 25 cm from his
ver, the observer's retinal image of the object gets eye and look directly at it. Calculate the accom-
larger even though the object's size is unchanged. This modative demand and the retinal image size of the
same principle holds when viewing virtual images. emmtrope, and compare it with the results in
Figure 8.6 shows an observer looking at the sub- Example 8.4a. Now
merged fish and then looking at the fish in air. The - 2 5 cm, and U x = - 4 D.
actual distance between the fish and the observer's eye
is the same in each case. The observer looking at the From
submerged fish does not see the actual fish, but rather A0 = U f p - U x ,
sees the virtual image of the fish at the apparent A = +4D.
depth. Since the apparent depth is less than the actual The accommodative demand is less when the em-
depth, the observer's retinal image of the submerged mtrope is viewing the arrowhead in air as com-
fish is larger than when he or she looks at the fish in pared to viewing it in ice.
air. The lateral magnification of the plane interface is The image distance remains at 16.67 mm, so
+ 1, and so, it does not cause the increase. Instead, the
v + 16.67 mm
shortened object distance for the observer when view- = -0.0667,
m= - -250.0 mm
ing the virtual image causes the increase.
and the retinalu image size is
EXAMPLE 8.4a I = mO,
An arrowhead, 3 cm in size, is imbedded 20 cm I = (-0.0667)(30 mm) = -2.00 mm.
deep in a block of ice (n = 1.33) with flat sides. An
emmtrope stands with his eye 5 cm from the block The emmetrope's retinal image size when the ar-
of ice (i.e., 25cm from the arrowhead). Use the rowhead is imbedded in the ice is 25% larger than
typical +60.00 D thin lens and screen model for the when the ice is melted.
emmetrope's eye to calculate the accommodative
demand and the retinal image size that the emmet-
rope has when observing the arrowhead.
Since nl is greater than n 2 , Figure 8.3 applies to 8.4 Ray Tracing Through a Flat Slab
the situation. The emmetropic observer sees the in Equi-index Media
virtual image of the arrowhead. The virtual image
of the arrowhead is still 3 cm in size but is at an Figure 8.7 shows a ray passing through a flat slab of
apparent depth given by Eq. 8.3. index n 2 . The indices of the media in front of and
-20 cm behind the slab are and n 3 , respectively. Assume
1.33 ' that nx equals n 3 .
Plane Refracting Interfaces and Reduced Systems 139

FIGURE 8.7. Ray through a flat slab.

/about} /abouti (about) *


The angles of incidence and refraction at the first
surface are Wj and wj, respectively. The angles of FIGURE 8.8. Flat slab in front of a lens.
incidence and refraction at the rear surface are w2 and
w2, respectively. From Snell's law,
nl sinwj = n 2 sin wj, ing rays are parallel to the incident rays but displaced
in such a manner that the virtual image for the back
n2 sin w2 = n 3 sin w2.
surface of the slab is closer than the actual point
The two normals (at the front and back surfaces of the source to the lens.
slab) are parallel. From geometry, the alternate inter- Assume that the distances d1? d2, and d3 are all
ior angles to parallel lines are equal, and so wj and w2 positive. The object distance for the lens can be found
are equal. Then from the above two equations, by stepping through the slab. From Eq. 8.3, the image
n3 . . distance for the first surface is
sin w, = sin w,,
n2 n
2 "i = ~di

or
n3 sin w2 = n! sinwj. or
But since n3 equals n1? it follows that *>i = - n g d i .
In this case, v1 is negative, and
sin w2 = sin Wj,
or u2 = vl d2.
Equation 8.3 applied to the second surface is
The equality of w2 and v/x means that the outgoing ray
in n3 is parallel to the incident ray in nl. 1 V
The fact that the outgoing ray is parallel to the or
incident ray makes it easy to quickly draw qualitative
ray sketches for flat slabs in equi-index media. How- v2 = -di .
ever, when is not equal to n 3 , the outgoing ray is n
g
not parallel to the incident ray. The object distance for the lens is
u3 = v2-d3,
or
8.5 Flat Slabs
3=-[d1 + ^ + 4
Figure 8.8 shows a point source in air at a distance dj (8.4)
from the front surface of a glass slab. The slab itself
has a thickness d2 and an index n g . A thin lens in air is In Eq. 8.4, dj and d3 are actual distances in air,
located a distance d3 behind the back surface of the while d2 is a thickness of glass and is divided by the
slab. The object for the lens is the image formed by index n g . Evidently, the thickness d2 in glass of index
the back surface of the slab. ng is equivalent to an air thickness of d 2 /n g .
Each of the diverging rays leaving the point source By the same reasoning, it is easy to show that any
bends toward the normal on entering the slab and thickness d of a material of index n is equivalent to an
away from the normal on leaving the slab. The outgo- air thickness of d/n. Let us temporarily label the
140 Geometrie, Physical, and Visual Optics

equivalent air thickness as d, or oil (Figure 8.9). What distance must the camera be
focused for?
(8.5) The object for the camera is the virtual image
n that the top oil surface forms. From the equivalent
Then we can write Eq. 8.4 as air system, the object distance for the camera lens
is
i*3 = - [ d 1 + d 2 + d 3 ]. (8.6)
7 16 5 1
M
EXAMPLE 8.5 ^ = -Ll44 + r33 + r72 + 1 J C m '
A 4-mm tall object in air is located 17 cm from the ".ens = -[^.86 + 12.03 + 2.91 + 10] cm,
front surface of a 33-cm thick glass (n = 1.50) slab.
A thin lens in air is located 8 cm behind the glass "ien. = -29.80 cm.
slab. What is the object distance for the lens? What Note that the 7 cm of plastic is equivalent to
is the size of the (apparent) object for the lens? 4.86 cm of air, the 16 cm of water is equivalent to
The object for the lens is the virtual image 12.03 cm of air, and the 5 cm of oil is equivalent to
formed by the back surface of the slab, and Figure 2.91 cm of air. Note also that the apparent depth of
8.8 shows the situation. the butterfly is 19.80 cm from the top oil surface,
Equation 8.6 gives the object distance for the and finally the virtual image of the butterfly at the
lens: apparent depth is the same size as the imbedded
butterfly.
33 cm
3 = - [ l 7 cm + + 8 cm
] Whenever the refractive index n is greater than 1,
u3 = -[17 + 22 + 8] cm = -47 cm. the equivalent air distance d is less than the actual
Note that the 33 cm of glass of index 1.5 is equival- distance d. For this reason, the equivalent air distance
ent to 22 cm worth of air. is usually called the reduced distance. In the preceding
The lateral magnification for each plane surface example, the reduced distance for the 7 cm of plastic
is +1, so the size of the object for the lens is still was 4.86 cm, the reduced distance for the 16 cm of
4 mm. water was 12.03 cm, and the reduced distance for the
5 cm of oil was 2.91 cm. In this context, reduced is a
Equation 8.6 is easily extended to systems with synonym for equivalent air.
multiple plane interfaces. Figure 8.9a shows three flat
slabs in front of a thin lens in air. The object distance EXAMPLE 8.7
for the lens can be found by stepping through the A 3-cm thick oil layer (n = 1.72) floats on top of
plane interfaces, a somewhat slow process, or by 26.6 cm of water (n = 1.33), which in turn is on a
setting up the equivalent air distances as shown in 27-cm thick flat glass slab (n = 1.50). A camera is
Figure 8.9b and writing directly that held so that its lens is 4 cm above the oil's top
surface. A beetle is clearly imaged in the camera
. = - [ d , + d2 + d3 + d 4 ]. (8.7) when it is focused for an object distance of 21.5 cm.
Where is the actual beetle?
EXAMPLE 8.6 One method of solving this problem is to step
A butterfly is imbedded 7 cm deep in a plastic back through the surfaces, again a slow process. An
(n = 1.44) slab. The plastic is covered by 16 cm of alternate method is to use the equivalent air (or
water (n = 1.33), and a 5 cm layer of oil (n = 1.72) reduced) system to determine the apparent location
is floating on the water. A camera is held in the air of the beetle relative to the apparent boundaries of
so that its lens is 10 cm above the top surface of the the different media.

I1
I
air air | air
I
I
I
I
**'v - ^ , * * v
d2 d3 d4 T di d4 FIGURE 8.9. a, A series of plane interfaces in
di front of a lens. b. The equivalent air (or reduced)
a) b) system.
Plane Refracting Interfaces and Reduced Systems 141

4 ,
4cm {
U
water

53 cm
i-7cm czr"_izr.L 21
20cm \ J virtual image
. ^ of beetle
18 cm .

FIGURE 8.10. Reduced (equivalent air) system for beetle ex- FIGURE 8.11. Converging light incident on a flat interface.
ample.

face has a virtual object (shown at A in Figure


Figure 8.10 shows the reduced system. The re-
8.11). Since the rays are initially in the lower index
duced thicknesses are:
medium, they bend towards the normal. The result
3 cm is that the image, shown at B, is real and farther
d
oil = = 1.7 cm,
1.72 than the (virtual) object from the interface. For this
26.6 cm situation,
= 20.0 cm,
1.33 ^ = 1.00 and n 2 = 1.33.
27 cm P is zero for the plane interface, so
a
glass
= 18.0 cm.
1.50 V=U=+4.00D.
The positions corresponding to the media boun- Now
daries are marked by the dashed lines in the figure.
The object distance for the camera is 21.5 cm. In
the reduced (or equivalent air) system, the beetle is V
at distance x below the marker corresponding to 133
the oil-water interface, and = +33.25 cm,
+4D
x = 21.5 cm - 4 cm - 1.7 cm = 15.8 cm. which confirms that the image is real and farther
The 15.8 cm is the reduced (or equivalent air) from the interface than the virtual object. The
distance corresponding to an actual distance in image would be visible on a screen placed in the
water. We can then find the actual water distance water 33.25 cm from the interface.
by unreducing, or Alternatively, Eq. 8.3 can be used. Here
100
water water water' = +25 cm.
+4D
d W ater = (1.33)(15.8 cm) = 21.0 cm.
The actual beetle is located in the water 21 cm Then
below the oil-water interface.
u
n?
gives
v +25 cm
8.6 Converging Light and Reduced Systems
L33 1.00 '
The techniques for plane interfaces work for converg-
ing light as well as for diverging light. The problems v = (1.33)(+25 cm) = +33.25 cm.
may be solved either by stepping through the inter- The lateral magnification for the interface is still
faces or by using a reduced system. +1, so the real image is the same size as the virtual
object.
EXAMPLE 8.8
Converging light in air is incident on an air-water EXAMPLE 8.9
(n=1.33) interface. When the vergence of the An 8-mm spider is suspended on a web 40 cm in
incident wavefront is +4.00 D, where is the image front of a +8.50 D lens. The lens has 4 cm of air
for the interface located? behind it followed by a 12-cm thick glass (n = 1.5)
Before calculating, you should draw a ray sketch slab, a 13.3-cm thickness of water, and a 25-cm
to see what characteristics you expect the answer to thick plastic (n = 1.44) slab. Where is the physically
have. The incident light is converging, so the inter- real image of the spider? What is the image size?
142 Geometrie, Physical, and Visual Optics

16.67 cm 8.7 Single Spherical Refracting Interfaces


and Reduced Systems
An equivalent air (reduced) system can also be set up
for a single spherical refracting interface (SSRI). In
the reduced system, a thin lens in air supplies the
convergence or divergence that the SSRI supplies in
the actual system. As shown below, all corresponding
4 cm 8 cm 10 cm 17.4 cm vergences in the reduced and the actual systems are
equal, and, consequently, the dioptric power of the
FIGURE 8.12. Reduced system for spider example. thin lens in the reduced system must be equal to the
dioptric power of the actual SSRI.
Consider an object in a medium of refractive index
Figure 8.12 shows the reduced system. The re- nl at a distance u in front of an SSRI of dioptric power
duced distances are: P. The image is in a medium of index n 2 at a distance i;
12 cm away from the interface. Figure 8.13 shows an exam-
glass
~ 1.50 = 8 cm, ple of such a system. In the actual SSRI system, the
13.3 cm vergence equations are:
d water 1.33 10 cm,
u a = ^,
25 cm U
= 17.4 cm.
4
plastic 1.44 and
For the lens.
a
V
u = - 4 0 cm,
The dioptric power of the SSRI is P a and
= -2.5D,
- 4 0 cm Va = P a + U a .
V = P + U,
The distance u in the medium of index n x between
V = +8.50 D + (-2.50 D) = +6.00 D,
the interface and the object is equivalent to an air
100 distance of where
v = +6.00 D + 16.67 cm,
u
u = .
and Hi
U -2.50 D
m= = = -0.42. The distance i; in the medium of index n 2 is equivalent
V +6.00 D
to an air distance v where
In the actual system, the lens' image is optically
i;
real but not physically real. In the reduced system, v = .
the image is 16.67 cm from the lens. From Figure
8.12, the image is located in the region correspond- In an equivalent air (or reduced) system correspond-
ing to the water at a distance of x past the marker ing to the SSRI, the incident vergence is given by
corresponding to the glass-water boundary, and
x = 16.67 - 4 - 8 = 4.67 cm. u.-i.u
r
The 4.67 cm is the equivalent air or reduced dis-
tance corresponding to the actual distance in the and the exiting vergence is given by
water. Therefore, in the actual system, the phys-
ically real image is in the water at a distance of v.-4.r
V
(1.33)(4.67cm) = 6.21cm,
from the glass-water interface. The image would be
visible on a screen placed in the water at this
position. n2
The lateral magnification for each plane inter-
face is + 1 , so the total lateral magnification is +1
times the lateral magnification m of the lens. The
image size is u
I = mO = (-0.42)(8 mm) = - 3 . 3 3 mm. FIGURE 8.13. Actual system.
Plane Refracting Interfaces and Reduced Systems 143

The dioptric power of the thin lens in the reduced The reduced object distance is
system is Pr and
-66.5 cm
Vr = Pr + U r . u = 1.33 =-50.0 cm,

By substituting the equation for the reduced distance and the reduced image distance is
into the above equation for U r , we see that +37.5 cm
v = 1.50 = +25.0 cm.
Ur=
(^S) = "^ = U a ' Thus, the reduced system consists of a +6.00 D
thin lens in air. The (real) object is located 50 cm in
where Ua is the vergence in the actual system. By front of the lens and the (real) image is located
substituting the expression for the reduced distance v, 25 cm behind the lens.
we find that In the reduced system:
10
r a
u= -50.0 cm
= -2.00D.
(v/n2) v '
and
where Va is the exiting vergence in the actual system. 100
Since the corresponding vergences are equal, the sub- V= = +4.00D.
scripts can be dropped: +25.00 cm
From U and V, it is clear that the dioptric power of
u = ua = ur, the thin lens in the reduced system must be
+6.00 D.
and In the actual system:
v=v =v. U= M
133
=-2.00D,
It then follows that the dioptric power of the thin lens -66.5 cm
in the reduced system must be equal to the dioptric 150
power of the actual SSRI, or V= = +4.00D.
+37.5 cm
P = P M. = PX .. The respective vergences are equal, and in both
M. a r
systems the lateral magnification is U/V or -0.5.
The lateral magnificaton of an SSRI is equal to the
vergence ratio U/V. Since the respective vergences are EXAMPLE 8.11
the same in the actual system and in the reduced A real object in plastic (n = 1.44) is located 18 cm
system, the lateral magnification is the same in the two in front of a +3.00 D plastic-water (n = 1.33) inter-
systems. Thus, given the same object size, the image face. The conjugate image is virtual and located
size I is the same in the actual and in the reduced -26.6 cm from the interface. Set up the reduced
system. system, and verify that the vergences are equal.
Figure 8.14a shows the actual system and Figure
EXAMPLE 8.10 8.14b shows the reduced system. The thin lens in
A real object in water (n = 1.33) is 66.5 cm in front the reduced system has a power of +3.00 D. The
of a -1-6.00 D water-glass (n = 1.50) interface. The reduced object distance is given by
conjugate image is real and located in the glass
u= -18 cm = -12.5 cm.
u
37.5 cm from the interface (Figure 8.13). Set up the
equivalent air system and verify that the vergences n, 1.44
and the lateral magnification are equal in the two Now look at Figure 8.14a again. What index should
systems. we use to reduce the image distance? The correct

-> + + 3.00D
+ 3.00D
A n = 1.44 \ n = 1.33 air ' air

12.5 cm

- 2 0 . 0 cm y i FIGURE 8.14. a. Actual system, b. Re-
b) duced system.
144 Geometrie, Physical, and Visual Optics

answer is the index of water. Remember that a and


virtual image point is simply the center of curvature V = P + U,
of a diverging wavefront that leaves the interface.
The exiting diverging wavefront is physically in the gives
water. Actually, the equation for the reduced V = - 2 . 8 0 D + (-2.50 D) = -5.30 D.
image distance indicates the index automatically,
and hopefully it now makes sense. The reduced image distance is
100
i; 26.6 cm = - 1 8 . 9 cm.
v= -20.0 cm. -5.30 D
n2 1.33
The light leaving the actual interface is diverging
The vergences in the reduced system are: and in the glass. Therefore, the actual image dis-
100 tance is found by unreducing.
U = -8.00 D,
-12.5 cm v = n2v = (1.50)(-18.9 cm) = -28.3 cm.
and Again, pay special attention to the fact that the
100 image space index is n 2 (or 1.50 in this case)
v = -20.0 cm
-5.00 D. regardless of whether the image is real or virtual.
In the actual system, the exiting wavefront is phys-
In the actual system: ically in the glass. Virtual simply means the wave-
144 front is diverging instead of converging.
u = -18.0 cm
-8.00 D, The lateral magnification is
133
v = -26.6 cm
-5.00 D. m = ^ = +0.47.

The vergences in the two systems are indeed equal.

EXAMPLE 8.12 8.8 Systems with Multiple Spherical Surfaces


A real object is in air 68 cm from a +7.00 D
air-glass (n = 1.50) interface. Suddenly, the room Reduced systems greatly simplify the calculations of a
floods with oil (n = 1.70). Now the object is in oil system of multiple plane interfaces. Reduced systems
68 cm from an oil-glass interface. Set up the re- can also be used for systems of multiple spherical
duced system, and use it to find the image and the surfaces. In the reduced system, the multiple spherical
lateral magnification. interface system is equivalent to a series of thin lenses
Be careful when setting the dioptric power for in air. The reduced system can aid one's intuitive
the interface in the reduced system. The power of
understanding of the optics. In fact, several different
the lens is equal to the power of the interface in the
actual system. The actual interface is now an oil- actual systems can have the same reduced system, and
glass interface. From Eq. 7.34, the new dioptric the reduced system analysis can provide information
power is about each of the different actual systems.
In the actual system, the refractive indices are used
_(n2-n1)new at each step, and a common mistake is to use the
new old
(,-^, ' wrong index. A main advantage of reducing a multiple
spherical system is that the refractive indices are used
(1.50-1.70)
P ( + 7 0 0 D ) only at the beginning and at the end, and there is less
~ (1.50-1.00) ' of a chance of making a mistake. Another advantage
occurs when the final image space is air because then
Pnew = ^ f j (+7.00 D) = -2.80 D. the reduced system provides the actual answer.
In the next series of examples, both the actual
Note that with the flood of high index oil, the system calculations and the reduced system calcula-
interface changes from converging to diverging. tions are done. You should compare the two systems.
The dioptric power of the thin lens in the reduced
system is -2.80 D.
EXAMPLE 8.13a
The reduced object distance is
As shown in Figure 8.15a, a real object is in water
u - 6 8 cm 44.33 cm from a +8.00 D water-glass interface. The
= - 4 0 cm. glass is 12 cm thick and has a -2.00 D back inter-
n, 1.70
face with water. The refractive indices of the water
Then and glass are 1.33 and 1.50, respectively. Where is
the final image? What is the total lateral magnifi-
U = - ^ = -2.50D,
- 4 0 cm cation?
Plane Refracting Interfaces and Reduced Systems 145

+ 8.00D -2.00D
+ 8.00D 2O0D
water /^glass f water

y
44.33 cm 12 cm
a) b) 33.33 cm 8 cm FIGURE 8.15. a. Actual, b. Reduced.

The calculations for the actual system are: See Figure 8.15b. Then
10
ux = -44.33 cm,
u,=
-33.33 cm
- 3 . 0 0 D,
133
Ux = = -3.00D, ^, + ,,
-44.33 cm
^, + ,, V1 = +8.00 D + (-3.00 D) = +5.00 D,
V1 = +8.00 D + (-3.00 D) = +5.00 D. 100
= +20 cm.
+5.00 D
l >1 =

The light leaving the interface is in the glass, and


The reduced object distance for the -2.00 D inter-
150 face is then
u, = = +30.00 cm.
+5.00 D 2 = +20 cm - 8 cm = 12 cm,
The object distance for the -2.00 D interface is
100
then U,= - = +8.33D,
+ 12 cm
u2 = +30.00 cm - 12 cm = +18 cm.
V2 = P2 + U 2 ,
The light incident on the interface is still in the
glass so V2 = -2.00 D + 8.33 D = +6.33 D,
100
U 22 = *g = +8.33D, 2 = +15.8 cm.
+18 cm +6.33 D

V2 = P2 + U 2 , In the actual system, the light leaving the -2.00 D


interface is in water, and so the actual image
V2 = -2.00 D + 8.33 D = +6.33 D. distance is
The light leaving the interface is in water so v2 = 1.33i72,
133 v2 = (1.33)( + 15.8 cm) = +21.0 cm,
^ = T633D=+21cm which agrees with the results in Example 8.13a.
Since the vergences in the reduced system are
The final image is real and located 21 cm behind
identical to the corresponding vergences in the
the glass-water interface.
actual system, the computations for the total lateral
The total lateral magnification is
magnification are exactly the same as in Example
m
tot = m 1 m 2 , 8.13a.
EXAMPLE 8.14a
mtot
" vx v 2 ' Figure 8.16a shows a system consisting of a
+5.00 D oil-plastic interface followed by a -9.00 D
m tot = (-0.60X + 1.32) = - 0 . 7 9 . plastic-water interface. The plastic is 28.8 cm thick,
and the refractive indices of the oil, plastic, and
EXAMPLE 8.13b water are 1.70, 1.44, and 1.33, respectively. A real
Use the reduced system and rework Example object is in the oil 85 cm in front of the +5.00 D
8.13a. The reduced system consists of a +8.00 D interface. Where is the final image? What is the
thin lens in air separated by a distance d from a total lateral magnification?
-2.00 D thin lens in air. The distance d is given by The calculations for the actual system are:

- 12 cm ux = - 8 5 cm,
d= =8cm
T5- = -2.00D,
U
The reduced object distance for the first lens is ' - 8 5 cm
-44.33 cm
= -33.33 cm.
^, + ,,
" l =
1.33 V, = +5.00 D + (-2.00 D) = +3.00 D.
146 Geometrie, Physical, and Visual Optics

-9.00D + 5.00D -9.00D

water
t Y
plastic
air air air

y y i A
85 cm 28.8 cm
b) 50 cm 20 cm FIGURE 8.16. a. Actual, b. Reduced.

The light leaving the interface is in the plastic, and Vj = +5.00 D + (-2.00 D) = +3.00 D,
144 100
+48.00 cm. v1 = +33.33 cm,
+3.00 D +3.00 D
The object distance for the -9.00 D interface is 2 = +33.33 cm - 20.0 cm = +13.33 cm,
then
100
u2 = +48.00 cm - 28.8 cm = +19.20 cm. U2 = +7.50D,
+13.33 cm
The light incident on the interface is still in the
plastic so V2 = P2 + U 2 ,
144 V2 = -9.00 D + 7.50 D = -1.50 D,
U9 = +7.50 D,
+ 19.20 cm
100
V2 = P 2 + U 2 , = -66.67 cm.
-1.50D
V2 = -9.00 D + 7.50 D = -1.50 D.
The final image is virtual and located 66.67 cm to
The light leaving the interface is in the water so the left of the -9.00 D lens. In the actual system,
133 the light leaving the -9.00 D interface is in the
= -88.67 cm. water. So the actual image distance is
-1.50D
The final image is virtual and located 3.67 cm to 2 = 1.332,
the left of the plastic-water interface.
The total lateral magnification is 2 = (1.33)(-66.67 cm) = -86.67 cm.
Note that the vergences at each step in Examples
8.14a and 8.14b are identical. Therefore, the total
U, U , lateral magnification is the same in both the re-
mtrtt = duced and in the actual system. Consequently, the
(-0.67)(-5.00) = +3.33. image size I is the same in both the reduced and in
mt, the actual system.
The final image is erect and 3.33 times larger than
the object.
The above examples show the equivalency of the
actual system and the reduced system. However, in
EXAMPLE 8.14b general, the numerical roundoff errors are different in
Solve Example 8.14a by using the reduced system. the actual system vs the reduced system. So when
Figure 8.16b shows the reduced system. The diop-
comparing the two systems, one needs to keep in mind
tric powers in the reduced system are the same as
those in the actual system, namely, +5.00 D and the number of significant figures.
-9.00 D. The lenses are separated by the reduced
distance d where
EXAMPLE 8.15
28.8 cm
d= = 20.0 cm. A real object is in water 133 cm from -3.00 D
1.44 water-plastic interface. The refractive indices of the
water and plastic are 1.33 and 1.44, respectively.
The calculations for the reduced system are: The plastic is 14.4 cm thick and followed by a
- 8 5 cm 4.02-cm thick slab of ice (n = 1.34). The ice is
-50 cm, followed by 13.5 cm of glass (n = 1.50). The back
1.70
surface of the glass is a +10.00 D glass-air inter-
100 face. The plastic-ice and ice-glass interfaces are
U -2.00 D,
- 5 0 cm plane interfaces (Figure 8.17a). Where is the final
^,+,, image? What is the lateral magnification?
Plane Refracting Interfaces and Reduced Systems 147

-3.00D +10.00D
-3.00D 1
waterNplastid

133 cm 14.44 cm 13.5 cm


100 cm
a) 4.02 cm b)

water ^Nplasticl ice

F,

11.7cm 12.2 cm
FIGURE 8.17. a. Actual, b. Reduced, c. Focal
points in actual system.

Since the system has some plane interfaces, it is In the actual system, the final medium is air (n =
easier to work this problem with the reduced sys- 1.00), so it is trivial to find the actual distance.
tem. The reduced system consists of a -3.00 D thin u2 = 1.00 i72,
lens in air at a distance of d in front of a +10.00 D
thin lens in air (Figure 8.17b). The reduced dis- i;2 = (1.00)( + 12.7cm) = +12.7cm.
tance d is given by The lateral magnification is the same in both the
reduced and the actual system. From the reduced
- 14.4cm 4.02cm 13.5cm .. _ system,
d= + +
-T4^ -T3r ^^-=22-0cm
m
tot m j m 2 ,
The reduced object distance for the - 3 . 0 0 D lens is
u, u2
=
v, v2 >
m
- 1 3 3 cm tot
= -100 cm.
1.33 - I D -2.13 D
m
tot = - 4 D '+7.87 D '
Then
mtot = (+0.25)(-0.27) = -0.068.
100
u, - cm = - 1 . 0 0 D ,
= -100
^, + ,,
8.9 Front and Back Vertex Powers
W, = -3.00 D + (-1.00 D) = -4.00 D,
Consider a centered multiple refracting interface sys-
100
= - 2 5 cm, tem with an image space index of n{ (the refractive
-4.00 D
index of the medium behind the system), and an
u2 = - ( 2 5 cm + 22 cm) = - 4 7 cm, object space index of n 0 (the refractive index of the
medium in front of the system) (Figure 8.18a). Three
100 dioptric values are used to characterize the paraxial
U,= = -2.13D,
- 4 7 cm imaging properties of such refractive systems. These
values are the equivalent dioptric power, the back
V2 = P2 + U 2 , vertex dioptric power, and the front vertex or neut-
V2 = +10.00 D + (-2.13 D) = +7.87 D, ralizing dioptric power.
In the basic vergence equation, V = P + U, the
100 dioptric power P is operative for any incident vergence
i>, + 12.7 cm. U. Of the three dioptric values for refractive systems,
+7.87 D
148 Geometrie, Physical, and Visual Optics

* + the optical axis, the final image point is called the


secondary focal point F 2 of the system. The back focal
system
length fb is the distance from the back vertex of the
many different system to F 2 , and as such is just the radius of curva-
indices
ture of the wavefront leaving the back of the system.
Figure 8.18b shows a converging wavefront leaving the
back resulting in a real F 2 . Figure 18.8c shows a
a)
diverging wavefront leaving the back resulting in a
>+ virtual F 2 . Since Pv is equal to the vergence Vb, which
in turn equals the image space index divided by the
plane
waves (real F2) radius of curvature of the wavefront, it follows that
incident
P = (8.8)
V
Let Uj be the vergence of the light incident on the
front of the system. When plane waves leave the back
plane of the system (Vb = 0), we say the system neutralizes
waves (virtual F2) the vergence U 1? or that the system has a neutralizing
incident dioptric power Pn of Ul5 i.e., for Vb = 0, Pn = - U , .
For example, when light of vergence 3.00 D is
incident on a system with a neutralizing power of
+3.00 D, then plane waves leave the back of the
system. Pn is also called the front vertex power.
When plane waves leave a system along the optical
plane waves axis, the initial object point is the primary focal point
emerging x of the system. The front focal length ff, is the
distance from the front vertex of the system to F 1? and
thus, is the radius of curvature of the incident wave-
front. Thus,

P = -- (8.9)

plane waves Figure 8.18d shows a system with a positive Pn. For
emerging emerging plane waves, the incident light is diverging
and F 1 is real. Figure 8.18e shows a system with a
negative P n . For emerging plane waves, the incident
light is converging and Fj is virtual.
Consider a system with a back vertex power of
FIGURE 8.18. a. Paraxial representation of multi-interface +7.00 D. When plane waves are incident on the sys-
system, b. Positive back vertex power, c. Negative back vertex tem, the vergence of the converging light leaving the
power, d. Positive (or plus) neutralizing power, e. Negative (or back vertex is +7.00 D. Now reverse the light. When
minus) neutralizing power. the light is reversed, diverging light of vergence
-7.00 D is incident on the back surface of the system,
and according to the principle of reversibility, plane
only the equivalent dioptric power is operative for any waves leave the front vertex (Figure 8.19a). For the
incident vergence. The front and back vertex powers reversed situation, the system has a neutralizing diop-
are operative only for situations involving plane tric power of +7.00D, i.e., it neutralizes light of
waves, and thus are not real dioptric powers in the vergence -7.00 D. In general, when the light through
V= P + U sense. Here only the back vertex and neut- a system is reversed, the original back vertex power
ralizing powers are discussed. The equivalent dioptric becomes the new neutralizing power.
power is discussed in Chapter 11. Similarly, consider light of vergence -5.00 D inci-
The back vertex power of a system is the vergence dent on a system with a neutralizing power of
Vb of the light leaving the back of the system when +5.00 D. Then plane waves leave the back of the
plane waves are incident on the front (Ul = 0), i.e., system. Now reverse the light so that the plane waves
for U, = 0 , P v =V b . For plane waves incident along are incident on the back of the system. From the
Plane Refracting Interfaces and Reduced Systems 149

original Then
Ux=0,
Vj = +10.00 D,
^ = +10.00 cm,
2 = +10.00 cm - 22.00 cm = -12.00 cm,
reversed U2=-8.33D,
a)
V2 = -3.00 D + (-8.33 D) = -11.33 D.
original So in the reversed system,
(Pv)rev= - 1 1 . 3 3 D ,
or in the original system
Pn = -11.33 D.

b) reversed EXAMPLE 8.17


Find Fx and F 2 for the system in the previous two
FIGURE 8.19. Principal of reversibility and vertex powers. examples.
In the actual system, the image space index ni is
1.00 and the object space index n 0 is 1.33.
From Eq. 8.8,
principle of reversibility, the light emerging from the 100
= +12.2 cm.
front vertex of the system now has a vergence of
IK
+8.19D
+5.00 D (Figure 8.19b). Thus, in the reversed situa-
F 2 is real and located +12.2 cm behind the back
tion, the original neutralizing power of +5.00 D plays vertex of the system.
the role of back vertex power. In general, when the From Eq. 8.9,
light through a system is reversed, the original neut-
133
ralizing power becomes the new back vertex power. ff'=- = +11.7 cm.
-11.33D

EXAMPLE 8.16
F 1 is virtual and located 11.7 cm behind the front
Find the neutralizing and back vertex powers of the vertex of the system (Figure 8.17c).
system in Example 8.15 (shown in Figure 8.17a).
Since the vergences are the same in the actual Note that the back vertex power provides informa-
and in the reduced system, we can use the reduced tion only when plane waves are incident on a system,
system. The reduced system consists of a -3.00 D and the neutralizing power provides information only
thin lens 22 cm in front of a +10.00 D thin lens when plane waves are leaving a system. For the system
(Figure 8.17b). To find the back vertex power, in Example 8.15, the back vertex power is +8.19 D,
consider plane waves incident on the system front, and the neutralizing or front vertex power is
which is the -3.00 D lens, then 11.33 D. Clearly, the two vertex powers can be very
different, but this is not always the case.
Consider a thin lens of dioptric power P. When
V^-3-OOD, plane waves are incident on the lens:
vl = 33.33 cm,
V = P + U = P,
2 = -(33.33 cm + 22.00 cm) = -55.33 cm,
so the back vertex power of the thin lens is equal to P.
U2=-1.81D, When plane waves leave the lens:
V2 = +10.00 D + (-1.81 D) = +8.19 D. V = 0 = P + U,
So or
P=+8.19D. P=-U,
The easiest way to find the neutralizing power is to so the neutralizing power of the thin lens is equal to P.
reverse the light and let plane waves be incident on Thus, for a thin lens of power P, the back vertex
the back lens. For the reversed system, relabel the
power and the neutralizing power are both equal to P.
lens as
As shown in Chapter 11, the equivalent dioptric power
Pj = +10.00 D and P2 = -3.00 D. also equals P.
150 Geometrie, Physical, and Visual Optics

8.10 Vertex Neutralization 8.10 gives the same result. Note that the informa-
tion given does not specify the neutralizing power
Assume that optical systems A and B are in contact of A. Also, the back vertex power of B is not
with each other, with A in front of B. Assume that useful here since plane waves are not incident on
whenever plane waves are incident on A, plane waves B.
leave B. Then we say that the systems have neutral-
EXAMPLE 8.19
ized each other. What condition is needed for this to
A system has a neutralizing power of -3.75 D and
happen? a back vertex power of -6.25 D. What thin lens
Since plane waves are incident on A, the vergence placed against the back of the system will neutralize
leaving it is equal to its back vertex power (P V ) A . it?
Since the systems are in contact, this vergence is When plane waves are incident on the front of
incident on B without change. For plane waves to the system, the vergence of the light leaving the
leave B, the vergence incident on it must equal back is -6.25 D. Therefore, a +6.25 D thin lens is
" ( P J B where (P n ) B is the neutralizing power of needed to give plane waves leaving.
system B. Therefore, the condition is that the back What thin lens placed against the front of the
vertex power of system A is equal in magnitude but system will neutralize it?
opposite in sign to the neutralizing power of system B, To get plane waves leaving the back of the
system, the vergence of the light incident on the
or front must be +3.75 D. The thin lens in front of the
(PV)A = - ( P ) E (8.10) system must take incident plane waves and convert
them to the needed vergence. A +3.75 D thin lens
EXAMPLE 8.18
will do the job.
System B has a neutralizing power of +4.00 D and
a back vertex power of +7.00 D. System B is
placed in contact with the back of system A. What
parameters must A have to neutralize B? 8.11 Reduced Angles
To get plane waves out of B, the vergence
incident on it must equal -4.00 D. Therefore, the Figure 8.20a shows a paraxial ray incident on a spheri-
vergence leaving A must be -4.00 D, and A must cal surface. The object space index is n. The axial
have a back vertex power of -4.00 D. Equation point is a distance d from the surface. The ray makes

actual refractive index n > 1 | air

"M
a)
I =n0

i
equivalent air (reduced) air air

y
^
-I -T

b)

1f
FIGURE 8.20. a. Actual ray angle, b. Reduced ray angle.
Plane Refracting Interfaces and Reduced Systems 151

an angle with the axis and hits the surface at a the water (n = 1.33) outside the window. What is
distance y above the axis. Therefore, the ocular accommodative demand when the fish
y is 21 cm from the outside of the submarine's win-
tan ~ = ^ . dow and when the lens representing the hy-
perope's eye is 15 cm from the inside of the
Figure 8.20b shows the same ray in a reduced system. window?
Now the axial point is a (reduced) distance d from the
5. A lens in air has exiting light of vergence
surface, where
+3.50 D. Seven centimeters behind the lens is a
flat air-glass (n = 1.53) interface. The glass is
a-4. 11cm thick followed by a flat glass-water (n =
n 1.33) interface. The water is 9 cm thick followed
Since image sizes are the same in an actual system and by a flat water-plastic (n = 1.44) interface. The
in a reduced system, it follows that the lateral distance plastic is 5 cm thick followed by a flat plastic-air
y is the same in the two systems. The (reduced) angle interface. Where and in what medium is the phys-
that the ray makes with the axis is given by ically real image formed?
~ tan = =.
y 6. Find the reduced (equivalent air) system for a
d 7-cm thick glass (n = 1.61) lens with a +7.00 D
We can then substitute for the reduced distance to front surface and a -3.00 D back surface. Specify
obtain the plastic (n=1.44) lens (surface powers and
y central thickness) that has the same reduced
A - y
- -ny system.
d d/n d 7. A system consists of an ophthalmic crown (n =
or 1.523) glass component in front of a barium crown
= . (8.11) (n = 1.617) glass component. The ophthalmic
crown glass is 4 cm thick and has a +10.00 D front
Equation 8.11 relates the reduced angle to the angle surface. The barium crown component is 3 cm
in the actual system. Remember that reduced is a thick and has a -4.00 D back surface. The inter-
synonym for equivalent air. While a reduced distance d face between the two glass types is flat. Find the
is smaller than the actual distance d, the reduced angle reduced system, the back vertex power, and the
is the larger than the actual angle . neutralizing power.
8. A +2.00 D thin lens is located 6 cm in front of a
-6.00 D thin lens, which in turn is 3 cm in front of
a +7.00 D thin lens. What is the back vertex
Problems power? What is the neutralizing power?
9. A +8.00 D thin lens is located 15 cm in front of a
1. A butterfly is 18.5 cm above the surface on a pond -10.00 D thin lens. Find the back vertex power
(n = 1.33). What is the apparent height of the and the neutralizing power of the system.
butterfly as seen by a fish under water in the 10. A three lens system has a +3.00 D neutralizing
pond? If the fish is 250 cm below the pond's power, and a +7.00 D back vertex power. Rela-
surface, what is the apparent depth of the fish as tive to the front surface of the system, where can
seen by the butterfly? an object be placed in order for plane waves to
2. While tubing down the Muskegon River (n = leave the back of the system?
1.33), a student notices a rock that appears to be 11. A system has a distant object, a +4.00 D neut-
80 cm vertically below the surface. What is the ralizing power, and a +6.50 D back vertex power.
actual depth of the rock? What thin lens placed against the front of the
3. A real object is located 12.5 cm in front of a 15-cm system results in plane waves leaving the back of
thick glass (n = 1.50) slab. A lens is located 7 cm the system?
from the back surface of the slab. What is the 12. A system has a distant object, a 5.00D neut-
vergence of the wavefront incident on the lens? ralizing power, and a +8.00 D back vertex power.
What is the equivalent air thickness (reduced What thin lens placed against the back of the
thickness) of the slab? system results in plane waves leaving the thin
4. A submarine has a 12-cm thick plastic (n = 1.58) lens?
window with flat sides. A 3.25 D uncorrected hy- 13. What is the neutralizing power of a +4.50 thin
perope in the submarine views a tropical fish in glass (n = 1.53) lens?
152 Geometrie, Physical, and Visual Optics

14. A system has a distant object, a +5.00 D back d. neutralizing power 17.00 D, back vertex pow-
vertex power, and a +12.00 D neutralizing power. er anything.
Which of the following systems placed against the e. back vertex power -5.00 D, neutralizing power
back of the first system will result in plane waves anything.
leaving the back of the second system. f. back vertex power -12.00 D, neutralizing pow-
a. neutralizing power -5.00 D, back vertex power er anything.
anything. g. back vertex power -7.00 D, neutralizing pow-
b. neutralizing power -12.00 D, back vertex pow- er anything.
er anything. h. back vertex power -17.00 D, neutralizing pow-
c. neutralizing power -7.00 D, back vertex pow- er anything.
er anything.
CHAPTER NINE

Lenses Revisited

9.1 Lens Shapes


A lens consists of two refractive surfaces separated by
a certain thickness. The words convex and concave are
frequently used to classify lenses as well as to classify
the individual surfaces. A convex lens is a converging a) b) c) d)
lens, and a concave lens is a diverging lens. FIGURE 9.2. Diverging lenses, a. Biconcave, b. Equiconcave.
Converging lenses in air tend to be thicker in the c. Planoconcave, d. Meniscus-concave.
middle than at the edge. A converging lens with two
converging surfaces is called a biconvex lens (Figure
9.1a). When the two converging surfaces have the 9.2a). When the two diverging surfaces have the same
same curvature, the lens is called equiconvex (Figure curvature, the lens is called equiconcave (Figure
9.1b). When a converging lens has one plane surface 9.2b). When a diverging lens has one plane surface
and one converging surface, the lens is called plano- and one diverging surface, the lens is called planocon-
convex (Figure 9.1c). When a converging lens has a cave (Figure 9.2c). When a diverging lens has a con-
converging surface and a diverging surface, the lens is verging surface and a diverging surface, the lens is
called meniscus-convex (Figure 9.Id). (The word called meniscus-concave (Figure 9.2d).
meniscus comes from the Greek word meniskos, which While the shape of a thin lens is not important, the
means moon. A meniscus convex lens is crescent- shape of a thick lens is. Ophthalmic lenses used to
shaped like a new moon.) correct the eye typically have a meniscus shape with
Diverging lenses in air tend to be thinner in the the convex surface away from the eye and the concave
middle than at the edge. A diverging lens with two surface toward the eye (Figures 9.Id and 9.2d).
diverging surfaces is called a biconcave lens (Figure

9.2 Thin Lens Power


In general, the optics of a lens depends on the dioptric
powers of the front and back surface of the lens, and
on the reduced thickness of the lens. Figure 9.3 shows
light incident on the convex surface of a meniscus-
a) ~ b) ^ c) d) convex lens. Pj and P2 are the dioptric powers of the
FIGURE 9.1. Converging lenses, a. Biconvex, b. Equiconvex. front and back surfaces, respectively. The surface
c. Planoconvex, d. Meniscus-convex. powers are related to the refractive indices and radii of

153
154 Geometrie, Physical, and Visual Optics

EXAMPLE 9.1
A thin glass (n = 1.60) biconvex lens has surface
radii of magnitude 12 cm and 20 cm, respectively
ni ln2 n3 (Figure 9.1a). What is the dioptric power of the
lens?
Unless specified otherwise, we assume the lens
is in air. Since the lens is biconvex, the surface
powers are both positive. From Eq. 9.1,
FIGURE 9.3. Lens surface powers. (1.60-1.00)(100cm/m)
1_
+12 cm
curvature by or
Pl =
p.=
l
and p2 =
n,-n, (9.1) ^ = + 5 0 0 D
l
Similarly,
where xx and r 2 are the respective radii of curvature of
the surfaces, n 2 is the refractive index of the lens P2 = ^ = +3.00D.
material, and where nx and n 3 are the respective
refractive indices for the mediums in front of and From Eq. 9.3,
behind the lens. Pt = ( + 5.00 D) + (+3.00 D) = +8.00 D.
Let Uj and Vj be the respective incident and
exiting vergence of the light at the first surface. Let U 2
EXAMPLE 9.2
and V2 be the respective incident and exiting vergence
A thin plastic (n = 1.50) meniscus-convex lens has a
of the light at the second surface. The vergence equa- convex side with a radius of curvature 5 cm and a
tions at each surface are: concave side with a radius of curvature of 25 cm
(Figure 9.Id). What is the dioptric power of the
^ , + ,, lens?
and For the convex side:
V2 = P2 + U 2 .
P, = y = +10.00 D.
When the central thickness of the lens is small enough,
no significant vergence changes occur as the light For the concave side:

travels across the lens, and
U2=V,.
; = ! = -2000
Then
Then
V2 = P 2 + V Pt = ( + 10.00 D) + (-2.00 D) = +8.00 D.

or The lenses in the above two examples are made out of


V2=P2 + P , + U different materials and have different shapes, but both
are +8.00 D thin lenses.
which can be written as
V2 = P t + U (9.2) EXAMPLE 9.3
A -4.50 D thin meniscus-concave lens has a con-
with
vex surface of radius 25 cm and a concave surface
Pt = P2 + P,. (9.3) of radius 10 cm (Figure 9.2d). What is the refrac-
tive index of the lens material?
Equation 9.2 directly relates any vergence Ux inci- From Eqs. 9.1 and 9.3
dent on the lens to the vergence V2 leaving it. Pt is the
( n 2 - 1 ) 1 0 0 cm/m (1 - n 2 )100cm/m
dioptric power of the thin lens. According to Eq. 9.3, t _ +
Pt is the sum of the surface powers Px and P 2 . The 25 cm 10 cm '
refractive index and the surface curvature information Pt = [ ( n 2 - l ) 4 ] + [ ( l - n 2 ) 1 0 ] ,
is hidden in Pt and does not explicitly occur in Eq. 9.2. Pt = ( n 2 - l ) ( - 6 ) ,
Since addition is commutative, we can interchange
Pj and P2 in Eq. 9.3. Thus, within the accuracy of the
paraxial approximation, the shape of a thin lens is not
important. Remember that the front and back vertex
power of a thin lens are both equal to P t . n, = 1.75.
Lenses Revisited 155

9.3 Thick Lenses: Front and Back V ^ P j + 1 ^ =+12.00 D,


Vertex Powers
100
^ = TlzD = + 8 ' 3 3 c m '
When a lens is thick enough there is a significant
vergence change across the interior of the lens, so the 2 = +8.33
100cm - 1.31 cm = 7.02 cm,
vergence U 2 incident on the back surface of the lens is 2
+7.02 cm
not equal to the vergence V t leaving the front surface, = +14.25 D.
i.e., Note the light waves gain +2.25 D of vergence
while crossing the interior of the lens, i.e., 14.25 -
12.00. Then
As a result, Eqs. 9.2 and 9.3 do not hold for thick V2 = P2 + U 2 = -4.00 D + ( + 14.25 D),
lenses.
PV = V2 =+10.25 D.
When plane waves are incident on the front of a
lens, the vergence of the light leaving the back surface The answer agrees with our expectations and is
is called the back vertex power Pv (Section 8.9). In greater than the thin lens value by the +2.25 D of
vergence gained by the converging light as it cros-
other words, when \Jl = 0 , Pv = V 2 . sed the interior of the lens.
The back vertex power is a characteristic of a lens.
The back vertex power is the parameter specified in a EXAMPLE 9.4b
prescription for a distance vision ophthalmic correc- What is the neutralizing power of the lens in Exam-
tion. ple 9.4a?
When a thick lens neutralizes incident vergence of The easiest way to find the neutralizing power is
Uj (i.e., gives plane waves leaving), the lens' neut- to use the reduced system, reverse the light, and
ralizing dioptric power P n is equal to minus U l 5 i.e., consider plane waves incident on the back of the
when V2 = 0, P n = \J1. The neutralizing power is a lens. Before calculating be sure to make your own
characteristic of the lens, and in general is not equal to estimates for what the answer should be.
the back vertex power. The neutralizing power is also In the reversed system, the calculations are:
called the front vertex power. Remember that both the ?, = -4.00 D and P2 = +12.00 D,
front and back vertex powers are not real dioptric u f = o,
powers in the V = P + U sense, because both are V ^ P ! + U 1 = -4.00D.
operative only in their respective plane wave situa-
Then
tions.
100
v, = -4.00 D = - 2 5 . 0 cm,
EXAMPLE 9.4a
What is the back vertex power of a glass (n = u2 = -(25.0 cm + 1.31 cm) = -26.31 cm,
1.523) lens with a central thickness of 2 cm and
with front and back surface powers of +12.00 D 100
U,= = -3.80D,
and -4.00 D, respectively? -26.31 cm
First what would we expect the answer to be if V2 = P2 + U 2 = +12.00 D + (-3.80 D ) ,
the lens were thin? We would expect the answer to
be the thin lens power that is the sum of the surface V2=+8.20D.
powers or +8.00 D in this case. However, for the So in the forward system, when light of vergence
back vertex power, plane waves are incident on the -8.20 D is incident on the front surface, plane
+12.00 D surface and converging light crosses the waves leave the back surface. The neutralizing
interior of the lens. The converging light gains dioptric power of the lens is then +8.20 D.
vergence across the interior, so we expect that the The lens in this example has a back vertex
back vertex power will be greater than +8.00 D. power of +10.25 D and a neutralizing (or front
This example can be worked either with the vertex) power of +8.20 D. The difference of the
reduced system or with the actual system. Let us values from each other and from the thin lens value
use the reduced system since that minimizes the use of +8.00 D comes from the different vergence
of the refractive index. The reduced central thick- changes across the interior of the lens.
ness d is
2 cm EXAMPLE 9.5
d = 1.31cm. A high index glass (n = 1.70) lens has a central
1.523
thickness of 4 cm, a front surface power of
The reduced system then consists of a +12.00 D +6.00 D, and a back surface power of -15.00 D.
thin lens in air 1.31cm in front of a - 4 . 0 0 D thin What are the front and back vertex powers of the
lens. The calculations are: lens?
U,=0, What are your expectations? The thin lens re-
156 Geometrie, Physical, and Visual Optics

suits would be -9.00 D. For the back vertex Consider a meniscus-convex polymethylmethacry-
power, the converging light across the interior adds late (PMMA) contact lens correction with a central
plus vergence to the -9.00 D (or reduces the thickness of 0.36 mm, a front surface radius of
minus). For the neutralizing power, (in the re- 6.30 mm, and a back surface radius of 7.80 mm. The
versed system) the diverging light loses minus ver- refractive index of P M M A is 1.49. The front and back
gence across the interior. surface powers are +77.78 D and - 6 2 . 8 2 D , respec-
Again let us use the reduced system
tively. The sum of the surface powers is +14.96 D.
4 cm The back vertex power is +16.45 D , and the neutraliz-
P T Q =2.35 cm. ing power is +15.90 D. The contact lens has a small
For the back vertex power: central thickness, but the surface powers are high.
^ = +6.00 0 and P2 = -15.00 D. Therefore, a clinically significant vergence change oc-
curs across the interior of the lens, and this contact
For lens must be treated as a thick lens. Clearly, the
Uj=0, ^ = +6.00 0 , surface powers as well as the central thickness must be
taken into account in deciding whether or not a lens
100
l>i = = +16.67 cm, can be treated as thin.
+6.00 D
2= + 1 6 . 6 7 c m - 2 . 3 5 c m = +14.31 cm,
100
U,= = +6.98D,
+ 14.31 cm 9.4 Thin Lenses in Different Media
V2 = P2 + U 2 = -15.00 D + 6.98 D,
P V =V 2 = - 8 . 0 2 D , Typically, lenses are used in air and so the thin lens
power Pt is the power of the lens in air. Suppose the
which agrees with our expectations that the result medium in front of or behind the lens is changed.
would be less minus than the thin lens value of What happens to the dioptric power? To answer this
-9.00 D. question, we need to know the refractive index n 2 of
For the reversed system:
the lens material.
P, = -15.00 D and P2 = +6.00D, Let n1 be the refractive index of the medium in
1^ = 0 gives V! = - 1 5 . 0 0 D , front of the lens (the object space index), n 2 be the
and then refractive index of the lens material, and n 3 be the
- 100 m refractive index of the medium behind the lens (the
l =
^i^D="6-67cm' image space index). When the object space index nl is
changed, the lens' front interface power F{ changes
2 = -(6.67 cm + 2.35 cm) = -9.02 cm,
according to
100
U2 = -11.09 D, (n2 n
-9.02 cm /D\ _ i)new / p x
(9.4)
VMJnew in n \ Vrl/old
V2 = P2 + U 2 = +6.00 D + (-11.09 D). U
\ 2 n
lJold

Since this is the reversed system When the image space index n 3 is changed, the lens'
Pn=V2 = -5.09D, back interface power P2 changes according to
which also agress with our expectations that the /p \ ( n 3 n2)new / p v
result would be less minus than the thin lens value V r 2^new /_ _ _ \ V r 2;old>
(n3 - n2)olc
of -9.00 D.
This lens has a back vertex power of -8.02 D or
and a front vertex (or neutralizing) power of /D\ _ (n2 n
3)new / p x
-5.09 D. Again, the difference is due to the differ- V r 2^new ( n - nr> \)
in
2
V^r 2o /lodl d
3 old
( 9
5 )

ent vergence changes across the interior of the lens.


In general, the new thin lens dioptric power can be
Consider a polycarbonate (n = 1.58) spectacle cor- calculated by adding the new interface powers to-
rection with a central thickness of 3 mm, a +4.00 D gether.
front surface, and a - 2 . 0 0 D back surface. The back A simplification occurs when the object and image
vertex power is +2.03 D, and the neutralizing power is space indices remain equal during the change. Then,
+2.01 D. The sum of the surface powers is +2.00 D. since nl is equal to n 3 , Eqs. 9.4 and 9.5 can be added
Variations much smaller than 0.25 D are negligible in to give
ophthalmic correcting lenses, so the polycarbonate p _ (n2 n
i)new p
(9.6)
lens can be treated as a thin lens. new
(n2-n,)old r
'
Lenses Revisited 157

where Pnew is the new thin lens power and Pold is the TT 1000 mm/m ^
old thin lens power, i.e., U
' = -15.72 mm = ~^>
*new
VM/new ' V*2>Jnew5 Vl = P, + Uj = +19.00 D + (-63.61 D)
"old =
( * l ) o I d "*" ( " 2 ) o l d = -44.61 D,
Note that in Eq. 9.6 the interface powers have _ 1000 mm/m
dropped out and only the thin lens powers remain. " 1 = -44.61 D = - 2 2 - 4 2 m m >
2 = -(22.42 mm + 2.70) = -25.12 mm,
EXAMPLE 9.6 1000
A +13.00 D thin ophthalmic crown lens is dropped U2 = = -39.81 D,
into water (n = 1.333). What is the power of the 25.12
lens while under water? V2 = P2 + U 2 = +47.00 D + (-39.81 D),
The refractive index of ophthalmic crown is
1.523. Then from Eq. 9.6, V2=+7.19D,
_ _ (1.523-1.333) 100
( + 13.00 D), v? = = +13.91 cm.
(1.523-1.000) +7.19 D
190 The light coming out of the eye is converging and has
Pnew= 523 ( +13.00 D)=+4.72 D. a vergence of magnitude 7.19 D. This eye is a 7.19 D
ocular myope. The far point is real and located a
distance of 13.91 cm in front of the eye.
EXAMPLE 9.7 What (thick) spectacle lenses worn at a vertex
Assume the crystalline lens of the human eye is
thin and has a refractive index of 1.400. The power distance of 12 mm corrects this eye for distance vision?
of the lens in the eye is about +19D. What is the For this model, assume the vertex distance is from the
power of the crystalline lens in air? back vertex of the lens to the cornea.
In the eye, the crystalline lens is bounded on the To get a clear image on the retina, the vergence
front by the aqueous humor, and on the back by incident on the cornea must be -7.19 D. The far point
the vitreous humor, both of which have a refractive is 13.91 cm minus 1.2 cm or 12.71 cm from the specta-
index of 1.336. From Eq. 9.6, cle plane. The vergence Vb leaving the back of the
= (1.400-1.000)
spectacle lens must be
new
(1.400-1.336) ^Lyu> ^11^ 100
Actually, the crystalline lens should be treated as a
Vb
12.71 cm = -7.87D.
thick lens. At this vertex distance, any spectacle lens with a back
vertex power of -7.87 D corrects the eye for distance
vision.
9.5 Another Eye Model

Previously, we dealt with thin lens in air and single 9.6 Exploding a Single Spherical
spherical refracting interface eye models. Now con- Refracting Interface
sider a slightly more sophisticated eye model. The
cornea is represented as a +47.00 D SSRI between air Figure 9.4a shows a system with water (n = 1.33) in
and the aqueous humor (n = 1.336), while the crystal- front, air (n = 1.00) in the middle, and glass (n = 1.53)
line lens is represented as a +19.00 D thin lens with behind. Both the concave water-air interface and the
the aqueous humor in front of it and the vitreous convex air-glass interface have a radius of curvature of
humor (n = 1.336) behind it. The distance from the 4 cm. From
cornea to the crystalline lens is 3.6 mm, and the p_ n
2-"i
distance from the crystalline lens to the retina is r '
21 mm. What is the ocular refractive status of this eye? the dioptric powers of the two interfaces are, respec-
The easiest way to answer this question is to reduce tively,
the system and reverse the light. The reduced system -33
consists of a +47.00D lens located 2.70mm (i.e., Pi = +4 = -8.25D,
3.6/1.336) in front of a +19.00 D lens, which in turn is and
located 15.72mm (i.e., 21/1.336) in front of the re- +53
p2 = = +13.25 D
tina. Then in the reversed system: +4
158 Geometrie, Physical, and Visual Optics

water glass water

FIGURE 9.4. a. Two interfaces each with a


4 cm radius of curvature, b. Single interface with
b) 4 cm radius of curvature.

Let Uj be the vergence of the light in water incident that, for analysis purposes, we can conceptually split
on the first interface and Vt be the vergence of the the water and glass apart, and consider a thin air layer
light in air leaving the first interface. Let U 2 be the in between them.
vergence of the light in air incident on the second The above argument can be repeated for an SSRI
interface and V2 be the vergence of the light in glass between media of refractive index n0 and n-. Thus, the
leaving the second interface. The vergence equations interface can be split in two with a thin air gap in
at the two interfaces are between. (Conceptually, you might think of one
molecular layer of air.) The curvature of each inter-
vr = p, + u,, face remains the same as the original interface. A
and system with each interface split in this manner is
V 2 = P2 + U 2 . sometimes called an exploded system. Exploded sys-
tems are useful in the analysis of contact lens correc-
In general, there is a vergence change across the air tions.
gap so that U 2 is not equal to \ x .
Suppose the air gap is so thin that no significant EXAMPLE 9.8
vergence change occurs across it. Then Find the exploded system for a -2.00 D spherical
v, = u2, refracting interface between plastic (n = 1.44) and
high index glass (n = 1.74).
and similar to the thin lens situation, the above three Figures 9.5a and 9.5b show the actual and the
equations can be combined to give exploded system, respectively. One way to proceed
is to find the radius of curvature r, and then
V2 = P, + U calculate the powers in the exploded system. How-
where ever, since the curvature is unchanged, we can also
use the Pnew-Pold equations.
PS = PI + P 2 The old interface is between plastic and the high
Then index glass. In the exploded system, the first new
interface is between plastic and air, so
Ps = +13.25 D + (-8.25 D) = +5.00 D.
Consider a single spherical water-glass interface P, = P... = 1.74-1.44
^ ^ ( - ^ D ) ,
with a radius of curvature of 4 cm (Figure 9.4b). The -44
dioptric power of the water-glass interface is Pi +30
(-2.00 D) =+2.93 D.
(1.53-1.33)100 _ 20 In the exploded system, the second new interface is
P= = +5.00D. between air and high index glass. So,
+4 cm +4
Essentially, the +5.00 D water-glass interface acts = P...= i ^ ? ? ( - 2 . 0 0 D ) ,
paraxially just like the system consisting of the 1.74-1.44
-8.25 D water-air interface separated by a thin air gap +74
from the +13.25 D air-glass interface. This means (-2.00 D) =-4.93 D.
+30

plastic glass

FIGURE 9.5. a. Plastic-glass spherical inter-


a) face. b. Exploded system.
Lenses Revisited 159

The exploded system consists of a +2.93 D plastic- surface of the contact lens is a diverging surface in
air interface separated by a thin air layer from a air. In the exploded system, the contact lens has a
-4.93 D air-glass interface. As a check, the sum back surface power of
should equal the original dioptric power.
(1.00-1.49)1000 mm/m
-4.93 D +2.93 D = - 2 . 0 0 D. +7.15 mm '
-490
P= = - 6 8 . 5 4 D.
+7.15
9.7 Hard Contact Lens Corrections The rest of the problem is easy. For plane waves
incident on the front of the contact lens, we want
light of vergence -7.19 D coming out the back (in
A knowledge of the eye's far point enables us to
the exploded system). If we reverse the light, we
calculate the vergence U f p of light in air incident on then have light of vergence +7.19 D incident on
the cornea that results in a clear retinal image. A the back surface of the contact lens and
spectacle lens correction sits in the air in front of the
eye and has a back vertex power that gives the proper
U fp value. However, a contact lens correction sits on V, = -68.54 D + 7.19 D = -61.35 D,
the cornea and wipes out the air-corneal interface. 1000
Nevertheless, we can use the exploded system concept
> = -61.35 D = _ 1 6 - 3 0 m m
and consider a thin layer of air (perhaps one molecu-
lar layer thick) between the contact lens and the The reduced central thickness of the contact lens is
cornea. This enables us to continue to use the U fp - 0.20 mm
d= , =0.13 mm.
information in the same manner as it is used for 1.49
spectacle lenses. Then
2 = -(16.30 + 0.13) mm = -16.43 mm,
EXAMPLE 9.9
The eye considered in Section 9.5 was a 7.19 D 1000
U2 = -60.86 D.
ocular myope, so U fp equals -7.19 D. The cornea -16.43 mm
was a +47.00 D spherical refracting interface be- Note that the vergence changed by 0.49 D (a clini-
tween air and the aqueous humor (n = 1.336). Con- cally significant amount) across the interior of the
sider a PMMA (n = 1.49) hard contact lens that has contact lens.
the same back surface radius of curvature as the Light of vergence -60.86 D is incident on the
cornea and a central thickness of 0.2 mm (Figure front surface of the contact lens, and plane waves
9.6). What must the front surface power of the must leave. Therefore, the front surface power
contact lens be? must be +60.86 D. The radius of curvature of the
The radius of curvature of the cornea is: front surface of the contact lens is
n7 - n , (1.49-1.00)1000
r= r= +60.86 D
(1.336-1.000)1000 mm/m 490
= +8.05 mm.
+47.00 D +60.86 D
336
r = +47.00 D = +7.15 mm. EXAMPLE 9.10
A +9.25 D ocular hyperope has a cornea with a
Thus, the back surface of the contact lens has a radius of curvature of 7.8 mm. A PMMA (n = 1.49)
radius of curvature of +7.15 mm. Clearly, the back contact lens has the same back surface radius of
curvature as the cornea and a central thickness of
0.4 mm. What is the front surface radius of the
distance vision contact correction?
First, explode the system. The back surface
(diverging) power in the exploded system is
(1.00-1.49)1000
P=
+7.8 mm
-490
P= = - 6 2 . 8 2 D.
+7.8
The light emerging from the back surface in the
exploded system should have a vergence of
FIGURE 9.6. Contact lens correction. +9.25 D. Now reverse the light. Then
160 Geometrie, Physical, and Visual Optics

^, + ,, tears must now be taken into account in determining


the correction. In an exploded system, the tears act
V, = -62.82 D + (-9.25 D) = -72.07 D, like a converging lens (Figure 9.7b). When a hard
1000 contact lens has a back surface that is flatter than the
cornea, the tears act like a minus lens in the exploded
The reduced central thickness of the contact lens is system (Figure 9.7c and 9.7d). For a contact lens fit
0.40 mm ^ with the same curvature as the cornea, the tears act
d = rrr = 0.27 mm. like a thin piano lens (which is why we did not
1.49
consider the tears in Section 9.7).
Then
= -(13.88 + 0.27) mm = -14.15 mm, EXAMPLE 9.11
1000 The patient in Example 9.10 was a 9.25 D ocular
U2 = = -70.67 D. hyperope and had a cornea with a radius of curva-
-14.15 mm ture of 7.8 mm. A PMMA (n = 1.49) contact lens is
Note that the vergence changed by 1.40 D across fit with a central thickness of 0.4 mm, and a back
the interior of the contact lens, which is a clinically surface radius of 7.4 mm (Figure 9.8a). First, what
significant amount. must the back vertex power of the contact lens be?
Light of vergence -70.67 D is incident on the Second, what is the front surface power and radius?
front surface of the contact lens, and plane waves Since the back surface of the contact lens does
must leave. Therefore, the front surface power not have the same curvature as the cornea, we need
must be +70.67 D. The radius of curvature of the to take the tear layer into account. Assume the
front surface of the contact lens is then tears have a refractive index of 1.333. From Figure
(1.49-1.00)1000 9.8b, the tears act like a converging lens in the
exploded system. The tear's front surface radius of
+70.67 D curvature is 7.4 mm and the back surface radius is
490 7.8 mm. The surface powers of the tear lens in the
r = +70.67 D = +6.93 mm. exploded system are
_ (1.333-1.000)1000
A
~ +7.4 mm
333
9.8 The Tear Lens P
A=T7^ = + 4 5 0 0 D'
Figure 9.7a shows a hard contact lens that has a back (1.000-1.333)1000
PB =
surface steeper (curved more) than the cornea. This fit +7.8
creates a gap between the back of the contact lens and -333
the cornea, and the gap fills up with tear fluid. The PB = -42.69 D.
+7.8

FIGURE 9.7. a. Tears (black) for minus contact


lens that is fit steep, b. Exploded system, c. Tears
(black) for a plus contact lens fit flat. d. Exploded
system.
Lenses Revisited 161

FIGURE 9.8. a. Tears (black) for a plus contact


lens fit steep, b. Exploded system.

The thin lens power of the tears is must be +71.74 D. The radius of curvature of the
Pt = +45.00 D + (-42.69 D) = +2.31 D. front surface of the contact lens is then
(1.49-1.00)1000
In the exploded system, the light leaving the tears r=
must have a vergence of +9.25 D. The tears have a +71.74 D
thin lens power of +2.31 D. So from the thin lens 490
vergence equation, the light incident on the tears = +6.83 mm.
+71.74 D
must have a vergence of +6.94D (i.e., 9.25
2.31). Thus, the needed back vertex power of the EXAMPLE 9.12
contact lens is +6.94 D. Suppose the same 9.25 D ocular hyperope is fit with
The effect of the tear layer is to change the a contact lens that has a back surface radius of
needed back vertex power of the contact lens in air. 8 mm. Now the back surface is flatter than the
The patient is a 9.25 D ocular hyperope, but the cornea (radius 7.8 mm), so in the exploded system,
tear lens corrects 2.31 D of the hyperopia. So the the tear lens acts like a diverging lens. What must
back vertex power of the contact lens in air needs the back vertex power of the contact lens be?
only to be +6.94 D. In the exploded system, the front surface power
Given the back vertex power of +6.94 D, what of the tear lens is
is the front surface power of the contact lens? (1.333-1.000)1000
The back surface (diverging) power in the ex- =
+8.0 mm
ploded system is
333
(1.00-1.49)1000 PA = +8.0 = + 4 1 . 6 3 D.
P=
+7.4 mm
The back surface power is the same as the previous
-490 example, or -42.69 D. The thin lens power of the
P= -66.22 D.
+7.4 tears is
The light emerging from the back surface in the Pt = +41.63 D + (-42.69 D) = -1.07 D.
exploded system should have a vergence of
+6.94 D. Now reverse the light. Then In the exploded system, the light leaving the tears
must have a vergence of +9.25 D. The tears have a
^^+^, thin lens power of -1.07 D. So from the thin lens
% = -66.22 D + (-6.94 D) = -73.16 D, vergence equation, the light incident on the tears
must have a vergence of +10.32D (i.e., 9.25 +
_ 1000 1.07). Thus, the back vertex power of the contact
Vl = 13.67 mm.
~ -73.16 D lens must now be +10.32 D.
The reduced central thickness of the contact lens is The effect of the tear layer is to change the
0.40 mm needed back vertex power of the contact lens in air.
d = ,_ = 0.27 mm. The patient is a 9.25 D ocular hyperope, but the
1.49
tear lens adds another 1.07 D of divergence. So the
Then needed back vertex power of the contact lens in air
is +10.32 D.
u = -(13.67 + 0.27) mm = -13.94 mm,
_1000_
2 = - 7 1 . 7 4 D.
-13.94 mm
Note that the vergence changed by 1.42 D, a clini-
cally significant amount, across the interior of the
9.9 Newton's Equation
contact lens.
Light of vergence -71.74 D is incident on the The single spherical refracting interface is the fun-
front surface of the contact lens, and plane waves damental building block for the paraxial analysis of
must leave. Therefore, the front surface power spherical refracting systems. The (reduced) vergence
162 Geometrie, Physical, and Visual Optics

->- +

FIGURE 9.9. Ray giving x' dependence.

equation for a spherical system was heavily used by is labeled x. The ray leaves the surface a distance h
Allvar Gullstrand (1862-1930). The corresponding from the axis, but since this ray is parallel to the axis,
image distance-object distance equation was first intro- h is equal to the image size I. Since the ray angles w
duced by Carl Gauss (1777-1855). There is an alter- and w' at the primary focal point are equal, it follows
nate formulation of the optics of an SSRI that is due from the tangents that
to Issac Newton (1642-1727). The alternate analysis is
I _ -O
as follows.
fi * '
Figure 9.9 shows a paraxial ray parallel to the or
optical axis incident on a spherical interface. At the
1
interface, the ray is bent down and passes through the
m = (9.8)
secondary focal point and then on to the image posi- =" X '
tion. The directed distance from the secondary focal We can set Eqs. 9.7 and 9.8 equal to each other and
point to the image is called the secondary extra-focal obtain
length and is labeled x'. The incident ray hits the
surface at a distance h from the axis, but since the f2 '
incident ray is parallel to the axis, h equals the object or
size O. Since the angles labeled w and w' are equal, it
is easily shown from the tangents of the angles that '=^2. (9.9)

-I O Equation 9.9 is called Newton's equation.


f '
EXAMPLE 9.13
or A real object in air is 45 cm in front of an air-glass
(n = 1.50) interface. The primary focal length of
m = - = (9.7) the system is -20 cm, and the secondary focal
f," length is +30 cm (Figure 9.11). Use Newton's
Figure 9.10 shows a similar drawing for the object equation to find the image, and then find the lateral
space ray that passes through the primary focal point. magnification m.
The directed distance from the primary focal point to The object is 25 cm in front of F1? so the
the object is called the primary extra-focal length and extra-focal distance x is -25 cm. Then,
(-25 cm)x' = (-20 cm)( + 30 cm),
x' = +24 cm.
-* +
- +

t F.

1
FIGURE 9.11. Newton's parametersreal object and real
FIGURE 9.10. Ray giving x dependence. image.
Lenses Revisited 163

The image is 24 cm past F 2 . (An image at that * +


position must be real.)
From Eq. 9.8,
__f.__Zgcm__o.80.
x - 2 5 cm
t F2
. Fi

You should check the above results with the V = x'


P + U equations.
V
y
\
x J
FIGURE 9.13. Newton's parameters for a diverging interface
EXAMPLE 9.14 with a real object and a virtual image.
The real object in the previous example is moved in
to 15 cm in front of the interface. Figure 9.12 shows
the situation for light traveling to the right. Use
Newton's equation to find the image, and then find EXAMPLE 9.15
the lateral magnification. A real object in air is located 18 cm in front of a
The object is now 5 cm past F 1 5 so the extra- diverging thin lens with a secondary focal length of
focal distance x is +5 cm. Then - 1 2 cm. Use Newton's equation to find the image,
(+5cm)x' = (-20cm)(+30cm), and then find the lateral magnification.
Figure 9.13 shows the diagram for light traveling
x' = - 1 2 0 cm. to the right. The object is 30 cm to the left of F , , so
The image is 120 cm to the left of F 2 , or 90 cm to x is - 3 0 cm. From Eq. 9.11,
the left of the interface. (An image at that position
must be virtual.) (-30cm)x' = - ( - 1 2 c m ) 2 ,
From Eq. 9.8, 144
x = + -r~- = +4.8 cm.
f, - 2 0 cm
m= - - = - = +4.0.
x +5 cm The image is located 4.8 cm to the right of F 2 , or
Again, you should check the above results with the 7.2 cm to the left of the lens. (An image at this
V = P + U equations. location must be virtual.)
x' +4.8 cm
The primary focal length can be eliminated from m=- - = - = +.40.
f2 - 1 2 cm
Newton's equation. From Eq. 7.23,
Newton's equation can also be written in terms of
(reduced) vergences. For an SSRI, the vergence X of
the incident light at F! is given by
We can substitute the above result into Eq. 9.9 to
obtain =, (9.12)

xx' = - ^ (f 2 ) 2 . (9.10) while the vergence X' of the outgoing light at F 2 is


given by
Newton's equation can also be derived for a thin
X'=2?. (9.13)
lens in air. There, ni and n 2 are equal, and Newton's
equation simplifies to Then from Eqs. 9.9, 7.21, and 7.22,
xx' = - ( f 2 ) 2 . (9.11)
XX' = - P 2 . (9.14)

EXAMPLE 9.16
Rework Example 9.13 using Eq. 9.14. In Example
* - -- 9.13, n, = 1.00 and n 2 = 1.50. The secondary focal
/ length of the interface is +30 cm, and the dioptric
power of the interface is
,. P = ^ _ 150_ = +5.00D.
X F2 f2 +30 cm
Y
> From Example 9.13, x is - 2 5 cm. The object is
'
^ then 25 cm in front of F j , and X is

FIGURE 9.12. Newton's parametersreal object and virtual = (i.oo)ioo c m / m = _4Q0D


image. - 2 5 cm
164 Geometrie, Physical, and Visual Optics

From Eq. 9.14, 5.0 cm. Find the back vertex power and the neut-
2 ralizing power of the lens.
-P2 -(+5.00 D)
5. A thick plastic (n = 1.49) lens has a +5.00 D front
X' =
X -4.00 D surface and a -12.00 D back surface. The central
-25.00 thickness is 6.0 cm. Find the back vertex power
X' = = +6.25D.
-4.00 and the neutralizing power.
The light at F2 is converging, and a real image is 6. A thick lens has a +4.00 D back surface power
rom F22 where
located a distance x' from and a +7.00 D back vertex power. Find the back
(1.50)100 cm/m focal length for the lens.
n, 7. A thick lens has a +5.00 D front surface power
X' +6.25 D
and a -3.00 D neutralizing power. What vergence
150 = +24 cm. must the incident wavefront have in order to get
+6.25 plane waves leaving the back of the lens? Find the
primary focal point of the lens.
EXAMPLE 9.17
Rework Example 9.14 using Eq. 9.14. In Example 8. A thick lens has a +8.00 D neutralizing power and
9.14 the interface is the same as in Examples 9.13 a +10.00 D back vertex power. What thin lens
and 9.16. From Example 9.16, = 1.00, n2 = 1.50, held against the back of the thick lens will neutral-
and P=+5.00D. ize it? (Plane waves out for plane waves in.)
From Example 9.14, x is +5 cm. The object is 9. Dr Pepper is a 7.00 D ocular myope with a spheri-
then 5 cm past F\. The light at Fx must be converg- cal cornea of radius of curvature 8.20 mm. Doc is
ing toward the object position and X is fit with a contact lens that has a back surface
(1.00)100 cm/m radius of 8.20 mm. What is the front surface
X = =+20.00 D. power and radius if the contact lens central thick-
+5 cm
ness is 0.30 mm and it has a 1.49 index?
From Eq. 9.14,
10. Maribell Maple is a 4.50 D ocular myope.
- P 2 = -(+5.00 D)2 Maribell's cornea has a 7.00 mm radius of curva-
X'
X ~ +20.00 D ture. Maribell is fit with a contact lens (n = 1.49)
-25.00 of back surface radius 7.20 mm. What back vertex
power must the contact lens have in air? (Assume
The light at F2 is diverging, and a virtual image is 1.333 is the refractive index of Maribell's tears.)
located a distance x' from F2 where 11. Baron Smith is a 3.50 D ocular hyperope. The
baron's cornea has a 7.60 mm radius of curvature
, _ n2 _ (1.50)100 cm/m and the baron is fit with a contact lens (n = 1.49)
X' -1.25 D of back surface radius of 7.90 mm. What back
150 vertex power must the contact lens have in air?
x = =-120 cm. (Assume 1.333 is the refractive index of the
-1.25
baron's tears.)
12. Sugar Brains is a 4.25 D ocular hyperope. Sugar's
cornea has a 7.50 mm radius of curvature. Sugar is
Problems fit with a contact lens (n = 1.49) with a back
surface radius of 7.35 mm. What back vertex
1. A thin meniscus-convex polycarbonate (n = 1.58) power must the contact lens have in air? (Assume
lens has surfaces with radii of curvature of 9.0 cm 1.333 is the refractive index of Sugar's tears.)
and 6.0 cm, respectively. What is the dioptric 13. Represent the cornea of an eye by a +45.00 D
power of the lens? SSRI between air and the ocular media (n =
2. A thin meniscus-concave highlite (n=1.71) lens 1.336). Represent the crystalline lens as a
has surfaces with 12.0 cm and 8.0 cm, respective +20.00 D thin lens between the aqueous and vit-
radii of curvature. What is the dioptric power of reous humor (n = 1.336 each). The distance from
the lens? the cornea to the crystalline lens is 4.0 mm, and
3. A thin biconvex ophthalmic crown (n = 1.523) the distance from the cornea to the retina is
lens has radii of curvature of 16 cm and 25 cm, 20.0 mm. Where is the far point for this eye? Is it
respectively. What is the dioptric power of the real or virtual? What correction is needed at the
lens? cornea?
4. A thick ophthalmic crown (n = 1.523) lens has a 14. A +14.00 D thin glass (n = 1.523) trial lens is
front surface power of +12.00 D and a back sur- placed underwater (n = 1.33). What is the power
face power of -4.00 D. The central thickness is of the lens while underwater?
Lenses Revisited 165

15. An SSRI between water (n = 1.33) and glass (n = 17. A spherical concave ophthalmic crown (n = 1.523)
1.73) has a real object in water and a real image in surface is held against a convex spherical barium
glass at a distance of 20.0 cm from the secondary crown (n = 1.617) surface. The dioptric power of
focal point. If the image is two times larger than the barium crown-ophthalmic crown interface is
the object, what is the dioptric power of the +2.00 D. If the surfaces are slightly separated so
interface? that a thin air gap exists between them, what is
16. An SSRI between glass (n = 1.61) and air has a the dioptric power of the air-ophthalmic crown
real object in glass at a distance of 13 cm from the interface, and what is the dioptric power of the
primary focal point and a real image in air at a air-barium crown interface? What is the sum of
distance of 35 cm from the secondary focal point. these two powers?
What is the dioptric power of the interface?
CHAPTER TEN

Reflection

10.1 Mirrors and Other Reflecting Interfaces ence, whereas for lenses (refraction) the word convex
is associated with convergence. A concave metallic
Chapters 4 through 9 treated image formation by re- surface converges light, so that a converging mirror is
fraction. Images are also formed by reflection. The also referred to as a concave mirror (Figure 10.2b).
reflection may be from metallic surfaces that reflect Thus, for reflection the word concave is associated
most of the incident light, or the reflection may be with convergence, whereas for lenses (refraction) the
from dielectric surfaces, such as glass, which reflect word concave is associated with divergence.
only a small percentage of the incident light. An air-glass single spherical refracting interface
The law of reflection states that (SSRI) convex with respect to the lower index
(10.1) medium converges the transmitted light no matter
fl = A, which way the light travels through the interface. The
where { and 0S are the angles of incidence and reflec- same interface can either diverge or converge the
tion, respectively (Figure 10.1). The media involved reflected light depending on which way the incident
influence the percentage of light reflected, but the light is traveling. The interface diverges the reflected
geometry of Eq. 10.1 is independent of the media. light incident from the air side, but converges the
This chapter covers paraxial image formation by reflected light incident from the glass side (Figures
spherical or plane mirrors. A convex metallic surface 10.2c and 10.2d). Thus, we need to take the direction
diverges light, so that a diverging mirror is also re- of propagation into account in deciding whether a
ferred to as a convex mirror (Figure 10.2a). Thus, for particular interface acts like a converging or a diverg-
reflection the word convex is associated with diverg- ing mirror.
When light in a transparent medium of refractive
index nl is incident on a transparent medium of
refractive index n 2 , the percentage of light R % that is
spherical surface
paraxially reflected is given by Fresnel's law:

Re 100%. (10.2)

The refractive indices n1 and n2 can be exchanged in


Eq. 10.2 without changing the value of R % . This
means that the percentage of light reflected at an
interface is the same no matter which medium the
light is initially in.
Fresnel's law for paraxial light reflected from an
FIGURE 10.1. Ray reflection at a spherical surface. interface between two transparent media can be de-

167
168 Geometrie, Physical, and Visual Optics

a)

FIGURE 10.2. a. Divergence by reflection from a


convex silvered mirror, b. Convergence by reflec-
tion from a concave silvered mirror, c. Divergence
by reflection of light incident in air. d. Convergence
d) by reflection of light incident in glass.

rived from Maxwell's classical electromagnetic equa- From Fresnel's law, the percentage of light re-
tions. As such, R % is a fundamental electromagnetic flected at an interface between two transparent media
property of the materials involved. This chapter treats depends on the difference in refractive indices. The
only the paraxial situation, but you should be aware higher the difference, the larger the percent reflection,
that the percent reflection increases outside the paraxi- and vice versa. The percentage reflected at an air-
al region. water (n = 1.33) interface is 2.0%, while the percen-
tage reflected at an air-diamond (n = 2.4) interface is
EXAMPLE 10.1 17.0%. For CR-39 (n = 1.498), a hard resin plastic
Light in air is paraxially incident on an air-ophthal- used for spectacle lens corrections, the percent reflec-
mic crown glass (n = 1.523) interface. What percen- tion in air is 4.0%. For highlite glass (n = 1.71), a high
tage of the incident light is reflected? index glass used for spectacle corrections, the percent
From Eq. 10.2, reflected in air is 6.9%.
1.523-1.000 A 2 % , 3 % , or 4 % reflection is a small percentage,
R<2 100%, but the luminance of the reflected light can be large if
1.523 + 1.000
the luminance of the incident light is high. As an
R. analogy, consider the fact that 4 % of a billion dollars
is 40 million dollars. It is only 4 % , but it is still a lot of
R % = (0.207) 2 100%, money.
R % = (0.043)100% = 4.3%; A lens has two interfaces, and some light is re-
flected at each interface. A first approximation to the
4.3% of the incident light is reflected at the air- total percent reflected is just twice that reflected at
glass interface. Therefore, 95.7% of the incident
light is transmitted through the interface and re- each interface. A more accurate calculation takes into
fracted. The percent reflected is independent of account the fact that not all the incident light gets to
which medium the light is initially in. So if the light the second interface, and a still more accurate calcula-
is reversed and is incident on the air-glass interface tion would take into account multiple reflections back
from the glass side, then again 4.3% of the incident and forth across the interior of the lens.
light is reflected.
EXAMPLE 10.3
EXAMPLE 10.2 Taking into account only single reflections, what
Light is paraxially incident on a water (n = percent of the incident light is transmitted by an
1.33)-plastic (n = 1.49) interface. What percent of ophthalmic crown (n = 1.523) lens in air?
the incident light is reflected? From Example 10.1, 4.3% is reflected at each
surface. Therefore, 95.7% of the incident light is
=/1.49 - 1.33x2
2
transmitted through each surface (Figure 10.3). As
%
\ 1.49+ 1.33:) 100%, a first guess, since the lens has two surfaces, we
R % = (0.057) 2 100% = 0.3%. would expect the lens to lose 8.6% due to reflection
(i.e., 2 times 4.3%). For a more accurate calcula-
Only 0.3% of the incident light is reflected at the tion, let I 0 , I 1? and I 2 be the intensity of the light,
water-glass interface, so 99.7% is transmitted. respectively, incident on the first surface, trans-
Reflection 169

Consider parallel light incident from the right on a


spherical concave mirror. The mirror converges the
light, and a real image point is formed at the sec-
ondary focal point F 2 (Figure 10.4a). The reflected
light is traveling to the left, which is then the positive
direction. Similar to lenses and refracting interfaces,
FIGURE 10.3. Reflection losses for a ray passing through a the directed distances (focal lengths, radii of curva-
lens. ture, image and object distances) are distances from
the reflecting interface or mirror to the position in-
volved.
mitted through the first surface and incident on the Figure 10.4b isolates some incident paraxial rays.
second surface, and transmitted through the second In the paraxial diagram, the spherical mirror appears
surface (Figure 10.3). The transmittance factor at flat because the vertical scale is expanded. The ray
the first surface is 0.957, or that travels along the optical axis (through the center
Ij= 0.95710. of curvature C) is incident normally on the mirror and
The transmittance factor at the second surface is has a zero angle of incidence. From Eq. 10.1 the angle
again 0.957, or of reflection is also zero, and the reflected ray travels
in the opposite direction along the optical axis.
I2 = 0.957 I i ;
The off-axis ray hits the mirror at point E, a
so that vertical distance h from the axis, and makes the angle
I2 = (0.957)(0.957)I0, { with the normal (which passes through C). The
reflected ray makes angle fts with the normal, and
or passes through F 2 . Vertical lines have been drawn
I2 = 0.91610. though C and F 2 , and these lines intersect the off-axis
Thus, the transmittance factor for both surfaces is ray at the points A and B, respectively.
0.916, or 91.6% of the light incident on the lens is For paraxial rays, the small angle approximations
transmitted through it. The other 8.4% is lost by are valid. Then from the triangle AEC,
reflection. The 8.4% differs only slightly from the
first guess of 8.6%. For interfaces with a higher
percent reflection the first guess is not as close to ft ~ tan ft = (10.3)
the answer, whereas for interfaces with a lower
percent reflection the first guess is better. where r is the radius of curvature of the mirror. From
the triangle BEF 2 ,

(ft, +ft,)-1311(4 + 0,) = (10.4)


10.2 Focal Points
V
where f2 is the secondary focal length of the mirror.
We have been using a sign convention in which the From Eq. 10.1, and fts are equal, so Eq. 10.5
positive direction is the direction that the light is becomes
traveling when it leaves the system. For refraction
systems the outgoing direction is the same as the 2,= r . (10.5)
l
2
incident direction, but for single reflection systems the
outgoing direction is opposite to the incident direc- Then from Eqs. 10.3 and 10.5,
tion. Let us continue to take the positive direction as
(10.6)
the direction that the outgoing light is traveling.

+ *-
N

+ -*- A B E

h

1 ^ ^ 1 /
C F2
a) b) ) FIGURE 10.4. a. Actual rays associated with the (sec-
ondary) focal point, b. Scaled ray.
170 Geometrie, Physical, and Visual Optics

Equation 10.6 shows that the secondary focal length is


one half the radius of curvature, or F 2 is halfway
between the center of curvature C and the mirror.
Equation 10.6 follows directly from the law of reflec-
tion, and is independent of the medium in front of the
mirror. Even if the medium in front of the mirror is
changed, f2 remains the same. Equation 10.6 also
holds for diverging mirrors, in which case F 2 is virtual.
FIGURE 10.5. Paraxial rays associated with the (secondary)
EXAMPLE 10.4a focal point of a diverging (convex) mirror.
A concave mirror in air has a radius of curvature of
20 cm. Where is the image conjugate to a distant
object? The convex mirror is a diverging mirror. In
The concave mirror is a converging mirror. The Figure 10.5, the reflected light is traveling left, so
image conjugate to a distant object is at the sec- the positive direction is to the left. Then
ondary focal plane (by definition). From Eq. 10.6, r = - 1 0 0 cm,
the secondary focal plane is 10 cm in front of the
mirror. In Figure 10.4, the positive direction is to and from Eq. 10.6,
the left, so f2 = "-50 cm.
r = + 2 0 c m , and f2 = +10cm.
For incoming plane waves, a virtual image would
The image would be visible on a small screen held be formed 50 cm behind the mirror. An observer
10 cm in front of the mirror. looking at the mirror would see the virtual image.

EXAMPLE 10.4b Let us apply the principle of reversibility to the


Unfortunately, the converging mirror in Example case where plane waves are incident on the mirror
10.4a was on the Titanic. As the ship sank, and the (Figure 10.4a). Then the rays leaving the point mark-
room flooded with water (n = 1.33). What hap- ed F 2 would reflect off the lens and leave parallel to
pened to the secondary focal point of the mirror?
each other. This means that the point marked F 2 is
The law of reflection, Eq. 10.1, is independent
of the media involved. Because of this, Eq. 10.6 is also the primary focal point of the mirror. Refract-
also independent of the media involved. As the ing interfaces and lenses always had their primary and
room filled with water, the radius of curvature r secondary focal points on opposite sides, but for a
remained +20 cm, the secondary focal length f2 mirror the primary and secondary focal points coincide
remained +10 cm, and F 2 remained in the same at the same point. For this reason, the subscripts on
location relative to the mirror. the focal points for a mirror are usually dropped:
F = F2 = F 1 ;
EXAMPLE 10.5
A convex mirror has a radius of curvature of and
100 cm. What is the secondary focal length of the (10.7)
mirror (Figure 10.5)? f = f2 = f, = 2

>
FIGURE 10.6. Predictable rays. a. b. Con-
verging (concave) mirror, c. d. Diverging (con-
d) vex) mirror.
Reflection 171

EXAMPLE 10.6 change needed is to reverse the direction of the arrow.


Where must an object be placed in order for a Since this reflected ray is not bent up or down at the
converging mirror of radius 20 cm to form an image reflection, it is considered the nodal ray, and the nodal
at optical infinity? point for the mirror is the center of curvature C.
The object must be placed at the primary focal (Remember that the nodal point for an SSRI was also
point, which like the secondary focal point, is at its center of curvature.)
located halfway between the mirror and the radius
of curvature. Here Consider a converging (concave) mirror with a real
object located outside C. The focal point is located
r=+20cm, and f = f1 = +10cm. halfway between C and the mirror. The incident ray
You should compare this answer to that in Example traveling parallel to the axis reflects at the mirror and
10.4a. then travels through F (Figure 10.8a). The incident ray
that passes through F reflects at the mirror and comes
Once the focal point is known, then any ray inci- out parallel to the axis. The two reflected rays are
dent parallel to the axis is predictable, and passes out converging and indicate that a small, inverted real
through F for converging mirrors (Figure 10.6a), or image is formed between C and F. The third predict-
appears to be coming from F for diverging mirrors able ray is drawn from the object through C to the
(Figure 10.6c). Any incident ray incident either com- mirror (Figure 10.8b). This ray also passes through the
ing from F for converging mirrors or pointing toward conjugate point because when it reflects at the mirror
F for diverging mirrors is predictable, and leaves the it comes straight back along itself. An observer look-
mirror parallel to the axis (Figure 10.6b and 10.6d). ing in the mirror would see the small, inverted image.
In fact, because the object is a person, the person sees
his or her own reflected image as inverted and smaller.
10.3 Ray Diagrams and the Nodal Ray The real object is now placed between C and F.
Again, the incident parallel ray reflects and passes
The third predictable ray is the nodal ray, and it is through F, and the incident ray through F reflects and
associated with the center of curvature C of the mir- comes out parallel (Figure 10.9a). The two reflected
ror. Any ray through C for a converging mirror or rays indicate that the outgoing light is converging and
pointing toward C for a diverging mirror is normal to a large, inverted real image is formed outside of C.
the mirror's surface, and has an incident angle of zero. The nodal ray, drawn through C to the mirror and
From Eq. 10.1, the reflected angle is also zero, and then back along itself, also passes through the conju-
the reflected ray travels back along the path of the gate image point (Figure 10.9b).
incident ray (Figure 10.7). On a ray diagram, the only Next, the real object is moved inside of F. The

FIGURE 10.7. Object point at a mir-


ror's center of curvature, a. Converging
mirroractual, b. Converging mirror
scaled, c. Diverging mirroractual, d.
c) d) Diverging mirrorscaled.
172 Geometrie, Physical, and Visual Optics

*-

C "^F^O

a)
J
FIGURE 10.8. a. Two predictable rays
for object outside C. b. Nodal ray.

FIGURE 10.9. a. Two predictable rays for ob-


S ject between C and F. b. Nodal ray.

virtual
image virtual
image

FIGURE 10.10. a. Two predictable rays


for object inside F. b. Nodal ray.

incident parallel ray again reflects and passes through (Figure 10.10b). When the reflected ray is extended
F, and the incident ray that appears to be coming from backwards, it also passes through the virtual image
F reflects and goes out parallel to the axis (Figure point. The virtual image point is just the center of
10.10a). The two outgoing rays indicate that the re- curvature of the diverging wavefront that is physically
flected light is diverging and the image is virtual. The leaving the mirror, so the image space medium is the
virtual image is located by extending the outgoing rays medium in front of the mirror even though the virtual
backwards until they intersect. The virtual image is image location is behind the mirror.
erect and larger than the object. When the object is a Now, consider an object in the center of curvature
person, that person will see his/her own erect and plane, and remember that F is exactly halfway be-
larger image in the mirror. This is the principle of tween C and the mirror. The ray incident parallel to
cosmetic and shaving mirrors. The nodal ray is drawn the axis reflects and passes through F, while the ray
from the object to the mirror so that it appears to be incident through F reflects and comes out parallel to
coming from C, and it reflects directly back along itself the axis. The two outgoing rays intersect at the center
Reflection 173

the ray diagram is essentially the reverse of Figure


10.12. The incident ray parallel to the axis reflects and
object goes out through F (Figure 10.13a). The incident ray
that appears to be coming from C reflects and goes
back out through C. The two outgoing rays are paral-
lel to each other, so the conjugate image is at optical
infinity. In this case it is not possible to draw an
image incident ray through F.
Actually, a fourth ray is easy to draw for each of
the mirror cases. The fourth ray is a ray from the
object drawn to the place where the optical axis
FIGURE 10.11. Object in the center of curvature plane.
intersects the mirror (i.e., the vertex or pole of the
mirror). The outgoing ray is drawn by making the
reflected angle 0S equal to the incident angle { (Figure
of the curvature plane, and an inverted real image is 10.13b). Even on a quick freehand sketch, one can
formed there (Figure 10.11). The inverted real image draw this ray fairly accurately.
is the same size as the object (m = - 1 ) . Note C is at Finally, consider a virtual object. In this case,
2F, and is the symmetry point for the mirror. It is not converging light is incident on the converging mirror
possible to draw the nodal ray for this case. so the reflected light is even more convergent. The
Next, consider an object at optical infinity. The incident rays all point towards the virtual object point
off-axis incident rays are parallel to each other, but (Figure 10.14a). The ray parallel to the axis reflects
not to the axis (Figure 10.12). Two of the rays are and goes through F (Figure 10.14b). The incident ray
predictable: the one through F, which comes out that passes through F reflects and comes out parallel
parallel to the axis; and the one through C, which to the axis. The two reflected rays show that the
comes directly back along itself. The two reflected rays outgoing light is converging, and a smaller, erect, real
intersect in the (secondary) focal plane forming a image is formed between F and the mirror. The third
small, inverted image there. Note from Figure 10.12 predictable ray is incident through C, and the reflected
that the size I of the image can be found from the ray passes through the image point and then back
subtended angle w and the incident ray through F through C (Figure 10.14c).
(i.e., |I|=ftanw). Now consider a diverging (convex) mirror. The
When the real object is placed in the focal plane, focal point F is a virtual point, and is still halfway
between the center of curvature C and the mirror. The
predictable rays associated with F are shown in Fig-
ures 10.6c and 10.6d. The nodal ray is shown in Figure
10.7d.
Figure 10.15a shows a real object in front of the
diverging mirror. The incident ray parallel to the axis
reflects at the mirror, and the outgoing ray appears to
be coming from F. The incident ray that points toward
F reflects at the mirror, and the outgoing ray is
parallel to the axis. The two outgoing rays are diverg-
FIGURE 10.12. Predictable rays for extended object at optical ing, and their extensions backwards intersect at the
infinity. virtual image location. The virtual image is smaller

FIGURE 10.13. a. Two predictable rays for extended


object in the focal plane, b. Another predictable ray.
174 Geometrie, Physical, and Visual Optics

and erect. The nodal ray is drawn from the object extension backwards passes through the virtual image
towards C, and its extension also passes through the point (Figure 10.16c).
virtual image point (Figure 10.15b). (This type of In Figure 10.17a, the virtual image is between F
mirror gives a large field of view, and is used as a and the mirror. Again, all the incident rays point
truck and car side mirror or as a security mirror in toward the object. The incident ray parallel to the axis
stores.) reflects, and the outgoing ray appears to be coming
Figure 10.16a shows a virtual object located farther from F. The incident ray pointing toward F reflects,
away than C. The incident rays all point toward the and the outgoing ray is parallel to the axis (Figure
virtual object point. The incident ray parallel to the 10.17b). The two outgoing rays indicate that the re-
axis reflects at the mirror, and the outgoing ray ap- flected light is converging, and a larger, erect real
pears to be coming from F (Figure 10.16b). The image is formed. You should check that the nodal ray
incident ray that points toward C reflects, and the also passes through the image point.
outgoing ray goes straight back along the path of the Now consider a virtual object in the focal plane.
incident ray. The two outgoing rays are diverging, and All the incident rays point toward the virtual object.
their extensions give a virtual image that is smaller and The nodal ray points toward C, and the outgoing ray
inverted relative to the virtual object. The third pre- goes straight back along the path of the incident ray
dictable ray points towards F, as well as towards the (Figure 10.18a). The incident ray parallel to the axis
object. The outgoing ray is parallel to the axis, and its reflects, and the outgoing ray appears to be coming

FIGURE 10.14. a. Converging light inci-


dent on a converging mirror, b. Two pre-
dictable rays. c. The nodal ray.

FIGURE 10.15. a. Two predictable rays for a


real object in front of a diverging mirror, b. The
nodal ray.
Reflection 175

virtual virtual
object r
x object
v
/ 1
pN /c
i
/ virtual
image
/
/
b) y f

FIGURE 10.16. a. Incident converging


light giving virtual object outside C. b. Two
predictable rays. c. Third predictable ray.

FIGURE 10.17. a. Incident converging


light giving a virtual object inside F. b.
Image location.
176 Geometrie, Physical, and Visual Optics

FIGURE 10.18. a. Incident converging


virtual light giving a virtual object in the focal
object plane, b. Predictable rays.

from F. The two outgoing rays are parallel to each


other, indicating that the reflected light consists of v = nv2 ' (10.10)
plane waves. In this case, it is not possible to draw an n2-
incident ray pointing toward the off-axis object point P= 1
(10.11)
r
and also toward F. However, the fourth predictable n2
ray, the one incident at the vertex of the mirror, can P= (10.12)
be drawn, and the outgoing ray is parallel to the other
and
outgoing rays (Figure 10.18b).
Finally, Figure 10.19 shows parallel light incident P= (10.13)
on the diverging mirror. The ray pointing toward C fi'
reflects and goes back out along the incident path. The The lateral magnification equations are
incident ray pointing toward F reflects, and the outgo-
ing ray is parallel to the axis. The two outgoing rays U
m= V (10.14)
are diverging, and their backward extensions locate
the virtual image, which is in the secondary focal or
plane. nji;
m = n w" (10.15)
7

These equations can be modified to describe image


formation by reflection by making specific selections
for the indices nj and n 2 . Assume that the refractive
index in front of the reflecting surface is n. In the


reflection process, the light is always in this medium,
but the reflected light and the incident light travel in
opposite directions. Therefore, let
F C
virtual n2 = +n, and ! = - n , (10.16)
image
where the minus sign describes the change in direction
FIGURE 10.19. Ray diagram for incident parallel rays from a that occurs at the reflection. (Other choices are pos-
distant off-axis object. sible, but no matter what choice is made a minus sign
needs to be introduced to describe the reflection.)
Equation 10.8 remains the same, while Eq. 10.11
becomes
10.4 Vergence Equations +n-(-n)
P=
The vergence equations for reflection from a spherical
surface can be obtained from the sagittal approxima- 2n
P= (10.17)
tion, from the law of reflection, or as a special modifi-
cation of the single spherical refracting interface equa- Equation 10.17 describes the converging or diverging
tions. The latter method is used here. The vergence power of a spherical interface as a result of reflection.
equations for a SSRI are: The power P in Eq. 10.17 is referred to as the
(10.8) reflecting power or as the catoptric power. (The word
V = P + U,
catoptric comes from katoptrikos, the Greek word for
mirror.) The catoptric power acts exactly like dioptric
where
power in that it is measured in diopters, and in that it
(10.9) is positive for converging power and negative for
U= diverging power.
Reflection 177

When Eqs. 10.12, 10.16, and 10.17 are combined, When a real object is 50 cm in front of any system, the
the result is incident light is diverging and the incident vergence U is
2n __ +n -2.00D. i.e.,
r f ' 100 cm/m
luh 50 cm
= 2.00D,
or
r and since the light is diverging,
f2 =
2' U=-2.00D.
which is just Eq. 10.6. Note that while Eq. 10.17 for P Then, from Eq. 10.8,
depends on n, the relationship between the focal
length and the radius of curvature of the mirror is V= P + U = +10.00 D + (-2.00 D),
independent of n, and the focal point for reflection is V=+8.00D.
always halfway between the center of curvature and The outgoing light is converging, and the image is real
the mirror. We can also derive Eq. 10.6 by combining and located 12.5 cm in front of the mirror (i.e., 100/8 D).
the primary focal length Eq. 10.13 with Eqs. 10.16 and From Eq. 10.14, the lateral magnification is
10.17. U -2D
m =
V = T8D=-25'
so the image is inverted and considerably smaller than the
object (in agreement with the ray diagram).
10.5 Imaging Examples in Air
The above example shows that if you think in terms of
EXAMPLE 10.7
vergences, you really do not need to worry about the
A real object is 50 cm in front of a concave mirror with a signs of the object distance u and the image distance
radius of curvature of 20 cm. Where is the conjugate
image? Is it real or virtual, erect or inverted, larger or i;.
smaller? For completeness, let us consider the object dis-
Assume the object is in air unless otherwise specified. tance. The positive direction is to the left, so
A concave mirror is a converging mirror, and P must be u = +50 cm,
positive. In Figure 10.20, the outgoing light travels left,
so r is positive. From Eq. 10.15, which feels wrong based on our refraction intuition.
2(1.00)(100cm/m) However, Eqs. 10.9 and 10.16 still give
P=
+20 cm (1.00)100 cm/m
200 u u +50 cm
+20 = +10.00 D. or
Alternatively, we could use Eq. 10.6 to find f, -100
U= = -2.00D.
_ r +20 cm +50
f= = 2 = + 1 0 c m > Our intuition in terms of the vergences is unchanged.
and then use Eq. 10.12 to find P: Diverging light still has a minus vergence, and con-
100cm/m verging light a positive vergence. Therefore, you
P= : =+10.00 D. should check the vergences instead of the object and
+ 10 cm image distances to see whether your answers make
Draw a quick ray diagram, and make your guesses at the sense sign-wise.
answers. Your ray diagram should resemble Figure 10.8. The image distance is
v = +12.5 cm,
which means that in Figure 10.8 the image is 12.5 cm
glass
to the left of the mirror.

EXAMPLE 10.8
A real object is 16.66 cm in front of the same
mirror used in Example 10.7. Where is the conju-
20 cm
gate image? Is it real or virtual, erect or inverted,
larger or smaller?
FIGURE 10.20. Outgoing direction for light in air reflecting off Be sure to draw your own quick ray diagram
a spherical air-glass interface. and guess the answers before calculating. Figure
178 Geometrie, Physical, and Visual Optics

10.9 is the ray diagram that roughly fits this situ- From your quick ray diagram, what are your
ation. guesses? Figure 10.14 is the corresponding dia-
The light incident on the mirror is still diverging, gram.
and the incident vergence U is - 6 . 0 0 D . i.e., The incident light is converging and the mag-
nitude of the incident vergence is 5 D (i.e., 100/
100 20). Then
U = 6.00D,
16.67 I
U = +5.00D,
and since the light is diverging,
V = P + U = +10.00 D + 5.00 D,
U = -6.00D.
and
Then
V = +15.00 D.
y = p + u = +10.00 D + (-6.00 D),
The outgoing light is converging, so the image is
V=+4.00D.
real and located 6.67cm (i.e., 100/15) in front of
The light leaving the mirror is converging so the the mirror. The lateral magnification is
image is real and the image 25 cm in front of the
U + 5 D
mirror (i.e., 100/4D). The lateral magification is ^n r i
U -6
m = v = _ = -1.5, so the image is erect and smaller than the object.
For completeness, the object distance is nega-
so the image is inverted and larger than the object. tive in this case, or
Again, the sign of U is chosen because the incident
light is diverging, as opposed to using the sign of -20 cm.
the object distance u. For completeness, Then from
+ 16.67 cm,
and u
n, n -100
U = = = + 16.67 = - 6 . 0 0 D . = (-i.ooKioo cm/m) = z ioo +500D
u u - 2 0 cm -20
The image distance v is +25 cm.

EXAMPLE 10.9 EXAMPLE 10.11


The real object is placed 5 cm in front of the same A real object is 40 cm in front of a convex mirror
mirror. Where is the conjugate image? Is it real or with a radius of curvature of 25 cm. Where is the
virtual, erect or inverted, larger or smaller? conjugate image? Is it real or virtual, erect or
From your quick ray diagram, what are your inverted, larger or smaller?
guesses? Figure 10.10 is the corresponding dia- Be sure to draw a quick ray diagram, and make
gram. your estimate for the answer. The ray diagram in
The magnitude of the incident vergence is 20 D Figure 10.15 roughly matches this situation.
(i.e., 100/5). Since the incident light is diverging, A convex mirror is a diverging mirror so P must
the vergence is be negative. In Figure 10.15, the outgoing light
travels left, so r is negative. From Eq. 10.15,
U = -20.00 D
Then 2(1.00)(100cm/m)
P=
V = P + U = +10.00 D + (-20.00 D), - 2 5 cm
200
V = - 1 0 . 0 0 D. P= = -8.00D.
-25
The outgoing light is diverging, so the image is
virtual and located 10cm (i.e., 100/10D) behind Alternatively, we could use Eq. 10.6 to find f,
the mirror. The lateral magnification is r _ - 2 5 cm
f= = -12.5 cm,
U -20 D 2 ~
m =
V=^D=+2 and then use Eq. 10.12 to find P,
The image is erect and twice as large as the object. 100cm/m o
(This is the cosmetic or shaving mirror example.) P=pr^ = -8.00D.
-12.5 cm
EXAMPLE 10.10 When a real object is 40 cm in front of any system,
Consider the same mirror with a virtual object the incident light is diverging and the incident
located 20 cm behind the mirror. Where is the vergence U is - 2 . 5 0 D , i.e.,
conjugate image? Is it real or virtual, erect or 100 cm/m
inverted, larger or smaller? |U|: = 2.50D,
40 cm
Reflection 179

and since the light is diverging U + 5 D


m = =
U=-2.50D. V =3D=~166
Then The image is inverted and larger than the object.
For completeness, the object distance is nega-
V = P + U = -8.00 D + (-2.50 D), tive in this case, or
and u = - 2 0 cm.
V=-10.5D. Then from
The outgoing light is diverging, and the image is Hi
virtual and located 9.52cm behind the mirror (i.e., u
100/10.5 D). From Eq. 10.14, the lateral magnifica- ( _ 1 .00)(100cnl/m) - 100
tion is u\J -
- 2 0 cm -20
= +5.00D.
U _ -2.50 D The image distance is
= +0.24,
V " -10.5 D
v -33.33
- cm1.
so the image is erect and considerably smaller than
the object (in agreement with the ray diagram).
EXAMPLE 10.13
Again, the above example shows that if you
think in terms of vergences, you really do not need A real object in air is 50 cm to the left of a -2.50 D
to worry about the signs of the object distance u single spherical refracting interface between air and
and the image distance v. glass (n = 1.50); 4% of the incident light reflects at
For completeness, let us consider the object the interface. Where is the reflected image? Is it
distance. The positive direction is to the left, so real or virtual, erect or inverted, larger or smaller?
From the refraction information, the radius of
u = +40 cm, curvature r of the interface is
which feels wrong based on our refraction intuition. (1.50-1.00)100
However, Eqs. 10.9 and 10.16 still give -2.50 D
-n -(1.1O0)l()0cm/m 50
u=u u +50 cm -2.50 D
-20 cm.

or While the interface diverges the refracted light, it


converges the reflected light so the catoptric (re-
u = -100
+40
= .-2.50 D.
flecting) power of the interface must be positive.
For the reflected light, the outgoing direction is
Remember that you should check the vergences opposite to that of the refracted light (Figure
instead of the object and image distances to de- 10.20), and the sign of r needs to be changed:
termine whether your answers make sense sign-
wise. r = +20 cm.
The image distance is The catoptric power of the interface is then
v = -9.52 cm. 2n 2(1.00)100 _ _
P= = - ^ r = +10.00 D.
r +20 cm
EXAMPLE 10.12 The remainder of the problem is identical to Exam-
Consider the same diverging mirror with a virtual ple 10.6, and the reflected image is real and located
object located 20 cm behind the mirror. Where is 12.5 cm to the left of the interface. The lateral
the conjugate image? Is it real or virtual, erect or magnification is - 0 . 2 5 . Note that the reflected light
inverted, larger or smaller? never enters the glass and the refractive index of
From your quick ray diagram, what are your the glass is not used in the reflection calculation.
guesses?
The incident light is converging and the mag-
nitude of the incident vergence is 5.00D (i.e.,
100/20). Then 10.6 Imaging Examples for n ^ 1
U=+5.00D,
EXAMPLE 10.14
V = P + U = - 8 . 0 0 D + (+5.00 D), Consider the preceding example (10.12) again. The
and real object is 50 cm in front of the air-glass inter-
face, which has a radius of curvature of magnitude
V=-3.00D. 20 cm. Suddenly, a flood fills the room with water
The outgoing light is diverging, so the image is (n = 1.333). Now where is the reflected image?
virtual and located 33.33cm (i.e., 100/3) behind What is the lateral magnification? There are at least
the mirror. The lateral magnification is three ways to solve the problem.
180 Geometrie, Physical, and Visual Optics

Method 1 Method 3
Since the law of reflection and the interface's center The third method is to reduce the system. There is
of curvature do not depend on the index, the ray no advantage to reducing the system for a single
diagram is unchanged by the water, and one can mirror, but in mixed lens-mirror systems there is an
solve the problem as if air was still present. This is advantage. The dioptric or catoptric power in the
exactly the solution of the previous example. The reduced system has to be the same as in the actual
image is still real and formed 12.5 cm in front of the system, so from method 2, the reflecting power is
mirror. The lateral magnification is still -0.25. + 13.33 D. The reduced object distance is 50/1.333
or 37.51 cm. A real object in air 37.51 cm from a
Method 2 surface gives an incident vergence on the surface of
The solution can also be worked out in the actual -2.67 D. Then
system. Here, for the reflected light V = P + U = +13.33 D + (-2.67 D),
n = 1.333, and r=+20cm. V = +10.66 D.
Then Note that the vergences in the reduced system are
2n = 266.6 the same as those in the actual system (except for
P= = +13.33 D. possible numerical round-off differences). Then
r +20
The light incident on the mirror is diverging, and 100
= +9.38 cm.
the incident vergence U is negative. Then + 10.66 D
u = +50 cm, The image space index is water, so the actual image
nt -n distance is
U = ^ = , i; = (1.333)(9.38 cm) = +12.5 cm.
u u
133 The lateral magnification is still given by U/V, and
u = +50 = - 2 . 6 6 D , equals - 0 . 2 5 . Again, these answers are the same as
those in methods 1 and 2.
V = P + U = +13.33 D + (-2.66 D),
and EXAMPLE 10.15
V = +10.67 D. A thick polycarbonate lens (n = 1.58) has a front
The light leaving the mirror is converging so the surface dioptric power of +10.00 D, a back surface
image is real and in front of the mirror, and dioptric power of -3.00 D, and a central thickness
of 3.16cm (Figure 10.21a). You stand 50 cm from
n
+n
2 the front surface of the lens, and can see your own
V ' reflected images from the front and back surfaces
+ 133 of the lens. Describe the location and size of these
= +12.5 cm. images.
+ 10.67 D
The lateral magnification is
For the Image Formed by Reflection from the
U -2.66 D Front Surface
m = -0.25.
V +10.67 D From the refraction information, the radius of the
These answers agree with those of method 1. front surface is

3.00D
50 cm 3.16cm

a)

+ 10.00D -3.00D
50 cm

FIGURE 10.21. a. Geometry for an observer to view his or her


2 cm own reflected images off the front and back surfaces of the lens. b.
b) Reduced system.
Reflection 181

(1.58-1.00)100 cm/m = +12.50 cm,


r= +10.00 D 2 = +12.50 cm - 2.00 cm = +10.50 cm,
+58
r = ^ 1 0 = +5.8 cm. 100
U, = +9.52D.
+ 10.50 cm
The front surface of the lens is convex and acts like
a diverging mirror. Therefore, the catoptric power
is negative. With the sign convention for the re- So far, the sign convention for the light that
flected light, traveled through the +10.00 D lens has been used.
The converging light gained vergence (from
r = - 5 . 8 cm, +8.00 D to +9.52 D) as it traveled from the lens to
and the mirror. Then
2n 200 V2 = -16.34 D + 9.52 D = - 6 . 8 2 D.
P= -34.48 D.
r - 5 . 8 cm Now use the plus direction for the reflected light,
Then and
U=-2.00D, 100
Vl = -14.66 cm,
-6.82 D
V = P + U = -34.48 D + (-2.00 D),
w3 = - ( 1 4 . 6 6 + 2) cm = -16.66 cm,
V = -36.48 D,
100
U3 = = -6.00D.
= -2.74 cm, -16.66 cm
-36.48 D
The diverging light lost vergence (from -6.82 D to
and -6.00 D) as it crossed back from the mirror to the
U -2.00 D lens. At the lens,
m= + = +0.0548.
V -36.48 D V3 = +10.00 D + (-6.00 D) = +4.00 D,
The reflected image is virtual, erect, much smaller, and
and located 2.74 cm behind the front surface.
100
V = +25 cm.
For the Image Formed by Reflection from the " * +4.00 D
Back Surface The total lateral magnification is
Here the light has to pass through the front surface, m = m1m2m3,
reflect off the back surface, and pass back out
through the front surface. _ -2.00 D +9.52 D -6.00 D
m
For the positive direction specified by the direc- ~ +8.00 D -6.82 D +4.00 D '
tion of the refracted light, the back surface radius m = (-0.25)(-1.40)(-1.50) = -0.52.
of curvature is
So the image is real, inverted, about one half the
-58
= +19.33 cm. size, and located 25 cm in front of the +10.00 D
surface.
The back surface of the lens acts like a diverging
mirror for the light incident in the glass, so the
catoptric power must be negative. With the positive EXAMPLE 10.16
direction specified by the reflected light, A thin -4.00 D CR-39 lens (n = 1.50) has a
+4.00 D front surface and a -8.00 D back surface.
r = - 1 9 . 3 3 cm, A distant object is in front of the lens. Find the
2n _ 2(1.58)100 cm/m image formed by the light that reflects from the
r -19.33 cm back surface of the lens and comes back out the
front.
and First, we need to find the catoptric power of the
316
P= = - 1 6 . 3 4 D. back surface. The light incident on the back surface
19.33 cm is in the CR-39, and the back surface diverges the
The reduced thickness of the lens is 2.00cm (i.e., reflected light. For the positive direction specified
3.16 cm/1.58). The reduced system consists of a by the transmitted light, the radius of curvature is
real object 50 cm in front of a +10.00 D lens that is given by
2 cm in front of a -16.34 D mirror (Figure 10.21b).
Then starting with the +10.00 D lens,
P
U! = - 2 . 0 0 D , -50
r= + 6.25 cm.
V, = +8.00D,
182 Geometrie, Physical, and Visual Optics

The plus direction is opposite for the reflected pletely fogged like an uncorrected high myope viewing
light, so a distant object. As the reflected image approaches
r = -6.25 cm. optical infinity, the wavefronts entering Emmy's eye
become flatter and the fog decreases. Eventually,
Then
Emmy begins to see a blurred, erect, and large re-
2n _ 2(1.50)100 cm/m flected image.
P=
(-6.25 cm) Why erect? Emmy is a real object for the mirror.
300 The mirror's image is real and inverted. The inverted
P= =-48.00 D. image is a virtual object for Emmy's eye. The real
-6.25
Since the lens is thin, no significant vergence image formed by Emmy's eye is erect relative to her
changes occur across the interior of the lens, and eye's virtual object. Thus, Emmy's blurred retinal
we can simply add the dioptric powers to get the image is erect relative to the virtual object, or inverted
total dioptric power. The light goes through the relative to the mirror's real objectwhich is herself.
front surface first, is reflected, and then goes back Emmy perceives inverted retinal images as erect.
out through the front surface. The total dioptric (Note that your retinal image of these letters is inver-
power is ted, but you perceive the letters as erect.)
Pt = (+4.00 D) + (-48.00 D) + (+4.00 D), When Emmy reaches F, plane waves leave the
Pt = -40.00 D. mirror. Then, without accommodating, Emmy can see
her erect and magnified reflected image clearly. As
The irriage is at the secondary focal point for the Emmy starts to move from F toward the mirror, the
catadioptric system or
light leaving the mirror starts to diverge. Emmy now
100 needs to accommodate to see her erect (and large)
f = -40.00 D = -2.5 cm.
reflected image clearly.
The image for the light that passes through the At some point Emmy uses all of her accommoda-
front surface, reflects off the back surface, and tion, and her reflected image starts to blur again. As
comes back through the front surface is virtual and Emmy's nose swoops in toward the mirror, her virtual
located 2.5 cm behind the lens. reflected nose also swoops in toward the mirror. As
Emmy's nose meets the mirror, her virtual nose meets
her real nose (i.e., an object distance of zero gives an
image distance of zero). However, unless Emmy's
10.7 Through a Glass Darkly nose is very long, she cannot clearly see the meeting.
(I suggest that you draw ray diagrams for each of
Consider Emmy, an emmtrope, standing in front of the above situations.)
her cosmetic (converging) mirror. When Emmy is far
away from the mirror, the reflected light forms a real EXAMPLE 10.16a
image at F. The light then diverges away from F and Emmy's cosmetic mirror is a +4.00 D mirror. What
returns to Emmy. The diverging wavefronts become is Emmy's accommodative demand when she
plane waves by the time they reach her. Emmy can stands 100 cm from the mirror?
then clearly see her reflected image, which is inverted For the mirror, f=+25cm, and r=+50cm.
and smaller, without accommodating. When Emmy is 100 cm in front of the mirror, the
As Emmy walks toward the mirror's center of vergence U incident on the mirror is - I D . Then
curvature C, her inverted reflected image moves out V= P + U = +4.00 D + (-1.00 D) = +3.00 D,
toward C and gets larger. To see it clearly, Emmy and
needs to accommodate. At some point prior to Emmy
reaching C, she will run out of accommodation. v = +33.33 cm.
Emmy's view of her large, inverted reflected image The reflected image is 66.67 cm from Emmy, and
then starts blurring. As Emmy moves from the first the diverging light incident on Emmy's eye has a
blur position to C, her view of the reflected image vergence of -1.50 D (Figure 10.22a). Since Emmy
blurs out completely. (To simulate this condition, hold is emmetropic, her accommodative demand is
this book against your face. Are the letters clear or + 1.50D.
blurred?)
As Emmy moves from C toward F, the reflected EXAMPLE 10.16b
image moves from C toward infinity. The reflected Besides the 100 cm of the previous example, at
light going into Emmy's eye is now converging. The what other distance would Emmy have an accom-
object for Emmy's eye is virtual, and Emmy is com- modative demand of 1.50 D?
Reflection 183

(not to scale) (not to scale)


Emmy Emmy

FIGURE 10.22. a. Emmy outside C. b.


b) 66.67 cm Emmy inside F.

Let us think a little before calculating. Figure The two solutions are:
10.22a shows the situation of the previous example.
The reflected image is 66.67 cm from Emmy's eye. U = - 1 D and U = - 6 D .
The other possible case occurs when Emmy is The - I D solution gives the 100 cm distance of the
closer than F and the image is virtual (Figure previous example. The - 6 D solution specifies that
10.22b). A guess might be 20 cm. when Emmy is 16.67 cm in front of the mirror, her
We could check the 20 cm guess by calculating accommodative demand is +1.50D. (Note the
the accommodative demand, and then modify the closeness of the initial guess made before cal-
guess if needed. This is essentially an iteration culating.)
procedure, which with a calculator can be fairly
fast. Alternatively, from Figure 10.22b,
-66.67 cm = v - w,
10.8 Unit Lateral Magnification and
the Radiuscope
v = -66.67 cm - u.
Let us work in terms of vergences, and for a mirror Cases in which the lateral magification is equal to plus
nn -1 or minus one are of special interest. For a mirror (just
U= as for a thin lens or an SSRI), when the object
u distance goes to zero, the image distance goes to zero,
Diopters are the reciprocal of meters, so convert
66.67 cm to 2/3 m. Then and the lateral magnification is +1.
For a mirror, 2F coincides with the center of
2 1 curvature C. An object at C results in an image at C,
and a lateral magnification of - 1 . Figure 10.11 shows
2U + 3 the ray diagram for converging mirrors and Figure
10.23 shows the ray diagram for a diverging mirror.
-3U For the diverging mirror, an object and image at C are
V V 2U + 3 both virtual. Thus, the center of curvature C of a
mirror serves both as the nodal point and as the
From
symmetry points.
V = P + U, The radiuscope (optical microspherometer) is an
ophthalmic instrument that uses both the -1-1 and - 1
^ L = 4 + TT
2U + 3 ' cases. The radiuscope measures the radius of curva-
-3U = (2U + 3)(4 + U), ture of the central part of a back contact lens surface.
The radiuscope works by first forming a target image
-3U = 8U + 12 + 2U2 + 3U, at point A as shown in Figure 10.24a. The light that
and formed the target image is traveling away from the
radiuscope and does not return unless reflected off
2U2 + 14U+12 = 0
. something.
The standard quadratic equation solution gives The contact lens is oriented so that its back surface
-14Vl4 2 -(4)(2)(12) faces the radiuscope. Some of the light traveling away
U= from the radiuscope reflects at the back surface of the
2(2)
contact lens and returns to the radiuscope. An obser-
-14 10 ver looking in the radiuscope can clearly see the
U=
reflected image of the target when the reflected image
184 Geometrie, Physical, and Visual Optics

FIGURE 10.23. A virtual object in the center of curvature plane.

light light
source source

FIGURE 10.24. Radiuscope. a. Illumi-


nation system, b. Setting for a clear image.
c. The other setting for a clear image.

is also formed at A; otherwise, the observer sees a contact lens until the observer's image clears (Figure
blurred image of the target. The reflected image is 10.24b). The distance measuring dial is then set to
clear at two different locations of the radiuscope from zero. Now the radiuscope is moved closer to the
the contact lens. The first (Figure 10.24b) is when the contact lens. The observed image first blurs and then
point A is at the center of curvature of the back comes clear again (Figure 10.24c). The distance on the
surface of the contact lens. Then the reflected image is measuring dial (i.e., the distance between the two
also at A and is the same size as the target image (but clear image positions) is the radius of curvature of the
inverted). The second is when the point A is at the contact lens' back surface.
back surface of the contact lens (Figure 10.24c). Then
the target image constitutes an object at zero distance
for the back surface of the lens. The reflected image is
also at A (zero image distance) and is the same size as 10.9 Keratometry
the target image (erect this time). Dust or scratches on
the contact lens surface are also clear in position two. Another ophthalmic instrument that uses reflected
The distance the radiuscope moves between the light is the keratometer (ophthalmometer). We use the
two clear images is just the radius of curvature of the keratometer to determine the radius of curvature r of
back surface of the contact lens. For example, suppose the anterior surface of the cornea. The keratometer
one starts with the radiuscope relatively far away from works by using a doubling procedure to measure the
the contact lens. The image that the observer sees is size I of the reflected image. The reflected image size I
blurred. The radiuscope is then moved toward the is then used to determine r. The anterior surface of
Reflection 185

the cornea acts like a diverging mirror, so the reflected instrument tolerances, is then read directly off the
image is virtual and erect (Figure 10.17). keratometer scale.
The keratometer dioptric readings are referred
EXAMPLE 10.17 to as the K readings. When a contact lens, as in
The keratometer target (called a mire) is 7 cm from Sections 9.7 and 9.8, is fit with the same back
the cornea. The mire is 5 cm in size, and the surface radius as the cornea, the fit is described as
measured size of the reflected image is 2.45 mm. on K. When the contact lens is fit with a smaller
What is the corneal radius of curvature? back surface radius of curvature (higher curvature)
The lateral magnification is than the cornea, the fit is described as steeper than
K. When the contact lens is fit with a larger radius
I +0.245 cm of curvature (smaller curvature) than the cornea,
m = - = = +0.0490.
O +5 cm the fit is described as flatter than K.
The target is real and 7 cm from the cornea, so the In the older keratometer, the mires were located
light incident on the cornea has a vergence U of farther away from the patient's cornea. Then, due
-14.29 D. Then from Eq. 10.14, to the high catoptric power of the cornea, the
reflected image was formed essentially at the focal
-14.29 D point (r/2), and the calculations shown in Example
= - 2 9 1 . 6 3 D.
m +0.0490 10.15 could be simplified.
From Eq. 10.8,
P = V - U = (-291.63 D) - (-14.29 D)
= -277.34 D;
10.10 Plane Mirrors
and from Eq. 10.17,
= 2n = 2000 mm/m Plane or flat mirrors are special cases of spherical
= -7.21mm.
~ P ~ -277.34 D mirrors. As the radius of curvature r of a spherical
The radius r was read off the scale on the mirror gets longer and longer, the mirror gets flatter
original keratometers. The newer keratometers and flatter. A plane mirror has a radius of curvature r
have been modified to give a refractive power of infinity and a curvature R of zero. The plane mirror
reading for the cornea. The cornea has a refractive has zero catoptric power. Therefore,
index of about 1.376 and about a - 5 D back inter-
face with the aqueous humor. The negative back V = P + U,
surface is compensated for by assuming a lower gives
refractive index for the cornea. The assumed index
is 1.3375. The plus direction for the refracted light V=U.
is opposite that of the reflected light so that r is
now +7.21 mm. Then Then from Eqs. 10.9, 10.10, and 10.16,
(1.3375 - 1.0000)1000 mm/m
P=
7.21 mm
From Eq. 10.14 the lateral magnification is + 1 . So the
+ 337.5 image is located the same distance away from the
P= = +46.81 D.
+7.21 plane mirror as the object, but on the opposite side.
The +46.80 D, or its approximation due to the The image is erect and the same size as the object. We

FIGURE 10.25. a. Plane mirror image, b.


a) plane b) Minimum size to see all of self in a plane
mirror mirror.
186 Geometrie, Physical, and Visual Optics

can construct ray diagrams for plane mirrors by using Historically, the mirror's image was said to be perver-
the law of reflection, Eq. 10.1. ted. A contemporary description is that the mirror's
image is reversed in handedness.
EXAMPLE 10.18 Spherical mirrors with real objects also give re-
An uncorrected 3.00 D ocular hyperope stands versed virtual images. In contrast, when a thin lens
25 cm in front of her bathroom (flat) mirror. What forms a virtual image of a real object, the image is
is the hyperope's accommodative demand when
looking at her reflected image? erect and no reversal in handedness occurs. Catadiop-
The ray diagram is shown in Figure 10.25a. One tric systems, including prisms, can be designed to
ray is incident normal to the surface and reflects maintain the same handness or to reverse it.
straight back along the path of the incident ray. A
second ray is incident at a particular incident angle
0, and the outgoing ray also makes an angle of
with the normal. The reflected image is virtual and 10.12 Ghost Images and Partially
located 25 cm behind the mirror. Silvered Mirrors
The reflected image is 50 cm from the hy-
perope's eye, so the light incident on the eye has a
vergence of -2.00 D. The accommodative demand The light that is reflected at lens surfaces can form
is then the sum of 2.00 D (for the incident ver- images that were not intended to be there. These
gence) plus 3.00 D (for the uncorrected hyperopia), images are called ghost images. Figure 10.26a shows
or 5.00 D. the reflections that give one of the ghost images for
people wearing spectacle corrections. This ghost image
What is the smallest plane mirror in which you can involves a reflection from the back surface of the lens
see your whole body? The answer is one half of your followed by a reflection from the front surface. This
height. The reasoning is shown in Figure 10.25b. For particular ghost image sometimes bothers speakers
simplicity, the eye is shown on top of the head. The who are trying to look at their audience, and instead
top ray is normal to the surface and reflects directly they see the ghost image of the overhead lights.
back. The bottom ray originates at the foot and obeys Figure 10.26b shows another ghost image that oc-
the law of reflection. From the law of reflection and a curs for spectacle wearers. This ghost image is of an
little trigonometry, one can show that the normal for object behind and to the side of the person and
the bottom ray bisects the person, so the mirror's size involves a single reflection at the back surface. For
is equal to one half the person's height. The argument distant objects, I can see this ghost image clearly. It
is easily adjusted for an eye that is not on the top of enables me to spy on people or objects behind me
the head, and the answer remains one half the per- while I am looking the other way.
son's height. There are a number of other possible ghost images
for spectacle lenses (see the problems). Ghost images
are sometimes visible on developed camera pictures
and in ophthalmic instruments.
10.11 Left-Right Inversion The appearance of the ghost images together with
the desired image was due to the fact that a certain
When you stand in front of your bathroom mirror, percentage of the light is reflected at an interface and
your reflected image is virtual and erect. However, a certain percent is transmitted. Partially silvered mir-
when you hold your right hand up, the reflected image rors are frequently used to superimpose two images,
holds its left hand up. In short, your bathroom mirror as shown in Figure 10.27a. The appearance to an
images your right hand as a left hand and vice versa. observer is shown in Figure 10.27b. Since the two

light bulb

object of
regard

FIGURE 10.26. a. Ray path for re-


flected (ghost) image of light bulb
superimposed on object of regard, b. Ray
path for reflected (ghost) image of tree
b) superimposed on object of regard.
Reflection 187

partially silvered
mirror

FIGURE 10.27. a. Ray paths for a partially


silvered mirror, b. Appearance of superim-
a) b) posed images.

objects are in different places they can be manipulated mirror parameters can also be found in a reduced
separately. This concept is very useful for gathering system, and then
information about the human visual system.
Pe=i. (10.19)

10.13 Equivalent Mirrors EXAMPLE 10.19


A +6.00 D thin lens is located 5 cm to the left of a
The reflected image from the anterior corneal surface -4.00 D mirror. What is the position and power of
the equivalent mirror?
is used in keratometry. The reflected images from the Let us first find the focal point F of the system
front and back surface of the crystalline lens provide since F is also the focal point of the equivalent
information about the accommodation mechanism of mirror. Consider plane waves incident from the left
the lens in the living human eye. These reflected on the +6.00 D lens (Figure 10.28a). Then the
images from the eye are called the Purkinje-Sanson vergence leaving the lens is +6.00 D, and the image
images. distance is 16.67 cm. The converging wavefronts
The light involved in the reflected images from the gain curvature as they travel the 5 cm to the mirror,
crystalline lens must pass through the cornea, reflect, and the vergence incident on the mirror is +8.57 D
and then pass back out through the cornea. Therefore, (i.e., 100/11.67cm). The mirror has a power of
-4.00 D, so the vergence leaving the mirror is
both refraction and reflection are involved, and the +4.57 D. The converging wavefronts gain curva-
system is a catadioptric system. ture again as they travel back toward the lens, and
When there is a single reflecting surface behind a the vergence incident on the lens from the right is
number of refracting surfaces, we can find a single +5.93 D (i.e., 100/16.88 cm). The vergence leaving
mirror that is equivalent to the catadioptric system. the +6.00 D lens is then +11.93 D, and the focal
The equivalent mirror concept simplifies the repeated point is 8.38cm to the left of the lens (i.e., 100/
analysis of data from such systems including those that 11.93 D).
form the Purkinje-Sanson images. The position H of the equivalent mirror is the
Two parameters that specify the equivalent mirror apparent position of the actual mirror as seen
are its position H and its power P e . The position H of through the lens, so we consider the mirror as a
the equivalent mirror is the apparent position of the real object for the lens. In Figure 10.28b the plus
direction is to the left. For the lens,
reflecting surface as seen through the refracting sur-
faces. Since the equivalent mirror must give the same w = -5cm and U = -20.00 D.
images as the actual system, the focal point of the Then
equivalent mirror must be at the focal point F of the V = P + U,
actual system. The focal length fe of the equivalent
V = +6.00 D + (-20.00 D) = -14.00 D,
mirror is then the directed distance HF where the plus
direction is the direction of the reflected light. Then and
i; = -7.14 cm.
Pe = ^ , (10-18) The image of the mirror is virtual and is located
7.14 cm to the right of the lens. This is the position
where nj is the image space index. The equivalent H of the equivalent mirror.
188 Geometrie, Physical, and Visual Optics

4.00D
+ 6.00D

>
^

: >
Y L

-4.00D "N equivalent


+ 6.00D
mirror
+ 6.44D
I

FIGURE 10.28. A +6.00 D lens in


front of a -4.00 D mirror, a. Incident
parallel rays and focal point for the
system, b. The mirror as an object for
the lens gives the equivalent mirror
location H. c. d. Two incident rays
converging toward the axial point H
give the conjugate image at H. e. Ray
paths for object at the center of the
curvature of the equivalent mirror.

The focal length fe of the equivalent mirror is thus the image distance should be zero and the
the distance HF, with the plus direction being that lateral magnification +1. Figure 10.28c shows two
of the reflected light or left in Figure 10.28b. So incident rays converging toward the axial point on
fe = +(8.38 cm + 7.14 cm) = +15.52 cm, H of the actual system in the above example. One
Then of the rays is incident along the axis, and travels
back along the axis when exiting. Since for the lens
p
e= T T ^ = +6.44D. H is conjugate to the mirror, the incident off-axis
+ 15.52 cm ray bends at the lens and goes to the vertex of the
The single mirror equivalent to the lens-mirror mirror. There, it reflects according to Eq. 10.1 and
system has a power of +6.44 D and is located is incident on the lens again. When this ray leaves
7.14 cm to the right of the +6.00 D lens position. the lens, it again has to be pointing away from H
If an object is placed at the postion H, it has a (just as in Figure 10.28b). So the conjugate image
zero object distance for the equivalent mirror, and for the actual system is indeed at H.
Reflection 189

Figure 10.28d shows two off-axis rays incident giving


towards H. It is easy to prove that for the actual
^ = +55 cm and w2 = +50cm.
system, the total lateral magnification is + 1 . Let Uj
be the vergence initially incident on the lens and Vj Then for the -4.00 D mirror,
be the vergence initially leaving the lens. Then the U 2 = +2.00D and V2 = - 2 . 0 0 D .
lateral magnification m, of the lens is
Since the light has reflected, the plus direction is
now to the left. Then
y2 = 50 cm and w3 = - 5 5 c m ,
The image of the lens is at the mirror, so the giving
mirror's object and image distances are zero, and
the lateral magnification of the mirror m 2 is + 1 . U3 = - 1 . 8 2 D and V3 = +4.18D.
Let U 3 be the vergence of the light incident on the The light leaving the lens is converging and
lens from the back side and V3 be the light leaving v3 = +23.90 cm.
the lens. This time the lateral magnification m 3 for
the lens is So the final image is at the same position as the
object. The total lateral magnification is given by
mt = m 1 m 2 m 3 ,
The total lateral magnification m t is = -4.18 D +2.00 D -1.82 D
mt
~ +1.82 D -2.00 D +4.18 D ~ '
mt = m 1 m 2 m 3 .
which checks. Note from u19 v19 and u2 (or from
According to Figure 10.28d, the initial image dis- Figure 10.28e) that C e is conjugate to the center of
tance for the lens 1 is equal in magnitude to the curvature C of the actual mirror. This relationship
object distance w3. This, together with the fact that can be used as a quick check on the accuracy of
the light initially leaving the lens is converging equivalent mirror calculations.
while the light incident on the lens from the back
side is diverging, means that U 3 is equal to - V j . It
EXAMPLE 10.21
follows from the vergence equation for the lens that
V3 is equal to Ul. Then Find the equivalent mirror for a -5.00 D thin lens
located 8 cm to the left of a +10.00 D mirror.
First, find the focal point of the system by
m
t = ^ + ^2' considering plane waves incident from the left (Fi-
gure 10.29a). Then the vergence leaving the lens is
m - U l
"Vl -+1 -5.00 D. The diverging wavefronts lose curvature
as they cross the 8 cm to the mirror, and the
vergence incident on the mirror is -3.57 D. The
Another characteristic of an equivalent mirror, mirror has a power of +10.00 D, so the vergence
as with any mirror, is that the symmetry points leaving it is +6.43 D. The converging wavefronts
coincide at its center of curvature C e . The radius of gain vergence as they travel back to the lens, and
curvature r e of the equivalent mirror is the distance the light incident on the lens from the back has
HC e , and is equal to twice the equivalent focal a vergence of +13.24 D. The light leaving the
length fe. An object in the actual system at a -5.00 D lens has a vergence of +8.24 D, and the
position coinciding with C e must be conjugate to image (which is at F) is real and located 12.14 cm to
itself and have a lateral magnification of 1. the left of the lens.
To find the location H of the equivalent mirror,
EXAMPLE 10.20 take the mirror as a real object for the -5.00 D
For the lens-mirror system in the previous example, lens. The mirror is 8 cm from the lens, so the light
verify that an object at C e results in a conjugate incident on the lens has a vergence of -12.5 D, and
image at C e . the light leaving the lens has a vergence of
In the previous example fe was +15.52 cm. Then -17.5 D. The image is virtual and located 5.71 cm
to the right of the lens (Figure 10.29b).
r e = 2fe = + 3 1 . 0 4 cm.
The distance HF is the equivalent focal length.
So C e is 31.04 cm to the left of H, or 23.90 cm to So from Figure 10.29b,
the left of the +6.00 D lens (Figure 10.28e). An
fe = +(12.14 cm + 5.71 cm) = +17.85 cm,
object at this position is a real object for the actual
system, and and
ux = -23.90 cm.
Then The equivalent mirror is a +5.60 D mirror located
U! = - 4 . 1 8 D and V^+1.820, 5.71 cm to the right of the lens.
190 Geometrie, Physical, and Visual Optics

FIGURE 10.29. A -5.00 D lens in front of


a +10.00 D mirror, a. Incident parallel rays
showing focal point for the system, b. Loca-
tion H of the equivalent mirror, c. Ray paths
for an object at the center of curvature of the
P e =+5.60D equivalent mirror.

As a check, let us see whether C e is conjugate to the rays from C e are drawn all the way through the
the center of curvature C of the actual mirror. The actual system, they come right back to C (Figure
actual mirror had a power of +10.00 D, and from 10.29c).
Eq. 10.17, r is +20 cm. So C is 20 cm to the left of
the mirror or 12 cm to the left of the lens. For the EXAMPLE 10.21
equivalent mirror,
A 15-cm thick glass (n = 1.50) rod has a front
r e = 2fe =+35.70 cm, surface of refracting power +20.00 D, and a plane
back surface. What is the equivalent mirror for
and C e is located 35.70 cm left of H, or 29.99 cm light incident from the front and reflecting from the
left of the -5.00 D lens. back?
For the actual system an object point 29.99 cm This example can be worked in either the actual
left of the -5.00 D lens gives an incident vergence system or in the reduced system. Let us work in the
of -3.33 D, and the vergence leaving is - 8 . 3 3 D. actual system.
The image is virtual, located 12 cm to the left of the To find F, take plane waves in air incident on
lens, and thus is at C. So everything checks. When the +20.00 D surface. Then
Reflection 191

V1 = + 2 0 . 0 0 D, 150
U = = -10.00 D,
150 -15.00 cm
= +7.50 cm, V = +20.00 D + (-10.00 D) = +10.00 D,
+20.00 D
u2 = -(15.00 cm - 7.5 cm) = - 7 . 5 cm, and
where the plus direction is still that of the light 100
= +10.00 cm.
before it reflects. The object for the plane surface is +10.00 D
7.5 cm in front of it, so the reflected image will be H is located 10 cm in front of the +20.00 D surface
7.5 cm behind it. For the light traveling back left, (Figure 10.30a).
the reflected image is the object for the +20.00 D The distance HF is - 2 . 5 cm and the image space
surface, which is 15 cm from the plane surface or index is air (Figure 10.30b). Therefore
22.5 cm from the reflected image. Therefore,
100
u3 = -22.5 cm = -40D.
-2.50 cm
U 3=
15Q
-6.67 D, So the equivalent mirror is a - 4 0 D mirror located
-22.5 cm 10 cm in front of the +20.00 D surface.
and To check, the actual reflecting surface is flat, so
V3 = +20.00 D + (-6.67 D) = +13.33 D, its center of curvature C is at optical infinity. For
the equivalent mirror,
100 r e = 2(-2.50 cm) = - 5 . 0 0 cm,
y
3=Ti333D=+7'50cm
so C e is 5 cm behind H or 5 cm in front of the plus
So F is 7.50 cm in front of the +20.00 D surface.
+20.00 D surface. An object placed at C e is a real
To find the location H of the equivalent mirror,
object for the actual system and the incident ver-
take the plane surface as a real object for the
gence on the +20.00 D surface is -20.00 D. So the
+20.00 D surface. Then
vergence leaving the +20.00 D surface is zero, and
u= -15.00 cm, the conjugate image is at optical infinity and coin-

v
15 cm
a)

I 5.0 cm 5.0 cm

b)

+ 20D FIGURE 10.30. a. The long glass rod of the


example. F is the focal point for light passing
through the spherical surface, reflecting at the flat
surface, and passing back through the spherical
surface. H is the equivalent mirror location, b.
Pe=-40D
Representation of the equivalent mirror, c. Ray
paths for incident light converging to the equivalent
mirror's center of curvature.
192 Geometrie, Physical, and Visual Optics

cides with C. The reflected rays then pass back tion calculation). Let us use the equivalent mirror.
through C e (Figure 10.30c). The penlight is 5 cm in front of the cornea, or
5.6874 cm in front of the equivalent mirror. Both
EXAMPLE 10.22 the object and image space are air. The vergence
Consider the following eye model (Figure 10.31a). that would be incident on the equivalent mirror is
The cornea is a +43.00 D SSRI between air and the then -17.58 D, and the vergence leaving it is
aqueous (n = 1.336). The crystalline lens (n = V = +346.15 D + (-17.58 mm) = +328.57 D.
1.416) has a central thickness of 3.6 mm, a front
The image distance is
surface that is a +8.00 D SSRI with the aqueous
and a back surface that is a +13.33 D SSRI with 1000 mm/m
= +3.044 mm,
the vitreous (n = 1.336). The front surface of the +328.57 D
crystalline lens is located 3.6 mm behind the cor-
nea. The fourth Purkinje-Sanson image is the and the lateral magnification is
image formed by the light that reflects off the back -17.58 D
surface of the crystalline lens and comes back out m= = -0.054.
+328.57 D
of the eye. This light first passes through the cor-
nea, and then the front surface of the crystalline The fourth Purki ije-Sanson image is inverted and
lens, then reflects, and passes back through the two considerably smaller than the penlight. The image
refracting surfaces again. By using the techniques is formed 3.044 mm in front of the equivalent
of this section, we can show that the equivalent mirror position, or 3.803 mm behind the cornea
mirror for the catadioptric system that forms the (Figure 10.31b). Think of the light leaving the
fourth Purkinje-Sanson image is a +346.15 D mir- equivalent mirror as converging to the real image
ror located 6.837 mm behind the cornea. Where is position and then diverging away. The light at the
the fourth Purkinje-Sanson image for a penlight position corresponding to the cornea is then diverg-
held 5 cm in front of the cornea? What is the lateral ing. In the actual system the final image formed by
magnification? the last corneal refraction is virtual and identical in
This question can be answered either from the position and size to the equivalent mirror image.
actual catadioptric system (which involves four re- If you wish, you are welcome to check the
fractions and a reflection calculation), or from the results by doing the four refraction and one reflec-
equivalent mirror (which involves only one reflec- tion calculation of the actual system.

+ 43.00D

n = 1.336

a)

FIGURE 10.31. a. Eye model, b. Equivalent mirror for the


fourth Purkinje-Sanson image.
Reflection 193

Problems 6. A thin biconvex glass (n = 1.50) lens has a front


surface refracting power of +3.00 D and a back
1. A concave (converging) spherical mirror has a surface refracting power of +2.00 D. When plane
25 cm radius of curvature. Locate the conjugate waves are incident on the front, find the resulting
image, specify whether it is real or virtual, and image for the light that passes through the front
give the lateral magnification for the following surface, reflects internally at the back surface, and
objects: passes back through the front. What would be the
a. real object, 50 cm from the mirror. dioptric power of a single mirror that would give
b. real object, 25 cm from the mirror. the same image position?
c. real object, 18 cm from the mirror. 7. What percent of the incident light is reflected at
d. real object, 10 cm from the mirror. an air-highlite (n = 1.70) glass interface? Taking
e. virtual object, 50 cm from the mirror. surface reflections into account, what percent of
2. A convex (diverging) spherical mirror has a 20 cm the incident light is transmitted through a highlite
radius of curvature. Find the conjugate image, lens in air?
specify whether it is real or virtual, and give the 8. Plane waves in air are incident on a -5.50 D
lateral magnification for the following objects: refracting surface (n = 1.50). Where is the image
a. real object, 50 cm from the mirror. for the light that is reflected?
b. real object, 15 cm from the mirror. 9. A real object in air is located 20 cm in front of a
c. virtual object, 5 cm from the mirror. +45.00 D cornea (n = 1.3375). What percent of
d. virtual object, 15 cm from the mirror. the incident light is reflected? Where is the image
e. virtual object, 20 cm from the mirror. for the reflected light? Is the image real or virtual,
f. virtual object, 50 cm from the mirror. erect or inverted? What is the lateral magnifi-
3. A person stands 50 cm in front of the bathroom cation?
mirror. Where is the conjugate image? What is 10. A real object 12.0 cm in size is located 20 cm from
the lateral magnification? How much must an a cornea (n = 1.3375). The reflected image has a
emmtrope accommodate to clearly see his or her size of 2.10 mm. What is the corneal radius of
image in the mirror? curvature? What is the refracting power of the
4. A +9.00 D lens is 10 cm in front of a diverging cornea? What is the keratometer reading for this
mirror of unknown power. When a real object is cornea?
placed 20 cm in front of the +9.00 D lens, the 11. An emmtrope stands 14 cm from a concave mir-
light travels through the lens, reflects from the ror with a 40 cm radius of curvature. What ocular
mirror, travels back through the lens, and forms accommodative demand does the emmtrope have
an image in the same plane as the original real when viewing his or her own image?
object. The image is inverted and the same size as 12. An uncorrected 6.00 D hyperope stands 25 cm
the object. What is the radius of curvature and the from a concave mirror with a 16 cm focal length.
dioptric power of the mirror? What is the ocular accommodative demand for the
5. A convex mirror has a 32 cm radius of curvature. hyperope to see his or her own image?
What is the dioptric power and focal length of the 13. A keratometer reads +44.50 D on the front sur-
mirror in air? What is the dioptric power and focal face of a contact lens (n = 1.49). What is the true
length of the mirror in water (n = 1.33)? Compare power of the contact lens surface? What is its
your results. radius of curvature?
CHAPTER ELEVEN

The Gauss
System

11.1 The Secondary Principal Plane In general, H 2 might be in front of, inside, or
behind the system (Figures 11.1c, 11.2, and 11.3). In
Consider plane waves traveling along the optical axis each case, H 2 is the intersection plane for the incom-
incident on a system of centered refracting interfaces. ing rays parallel to the axis and the outgoing rays that
The image point is the secondary focal point F 2 of the pass through F 2 . The point A on H 2 is the intersection
system. The incident rays are parallel to the axis. The point for the top incident ray. Note that H 2 is a
rays can be traced through each interface, and the geometric concept (the set of intersection points) and
result is a zigzag path through the system. The exiting that, in general, nothing physically real exists at the
rays either point toward F 2 (if F 2 is real) or point away H 2 position.
from F 2 (if F 2 is virtual). Figure 11.4 shows the zigzag paraxial ray trace
Figure 11.1a shows the zigzag paraxial ray trace for (solid lines) for two rays parallel to the axis incident
a system with a real F 2 . A ray bends at each internal on a system with a virtual F 2 . The image space rays
interface in the system. (The internal interfaces are appear to be coming from F 2 . For each incident ray,
not shown, but their locations can be determined from there is an intersection point between the object space
the ray trace.) In object space, the incident rays are ray (or its extension) and the corresponding image
parallel to the axis. In image space, the outgoing rays space ray (or its extension). Point A is the intersection
leave the back of the system and pass through F 2 . The point for the top ray. Again, all of the intersection
object and image space rays can be extended forward points lie in the secondary principal plane H 2 .
and backward. At some point, the object space ray or How is the secondary principal plane H 2 useful?
its extension intersects the image space ray or its When we know the locations of H 2 and F 2 , then we
extension (e.g., point A, Figure 11.1b). can take an incident ray parallel to the axis and
Figure 11.1c shows two paraxial rays incident paral- immediately draw the exiting ray without having to
lel to the axis. The image space rays both pass through follow all the internal zigzags throught the system. For
F 2 . There is an intersection point for each of the example, Figure 11.5a shows H 2 and F 2 for a system.
incident rays. The two intersection points lie in a The individual interfaces, which determine the loca-
plane, marked H 2 , which is perpendicular to the opti- tions, are not shown. The incident ray parallel to the
cal axis. It turns out that, for a given system, all of the axis is extended, and the extension reaches H 2 at the
intersection points for incident paraxial rays parallel to point labeled A. Since H 2 is the intersection point
the axis lie in the same plane. The plane H 2 containing between the object and image space ray, the exiting
these intersection points is called the secondary princi- (image space) ray must lie along the line from A to F 2
pal plane. The point at which the optical axis passes (Figure 11.5b). The line from A to F 2 hits the back of
through H 2 is called the secondary principal point. the system at the point B. We do not know the actual
The position of the secondary focal point F 2 is zigzag path of the ray through the system, but we do
determined by the zigzag ray paths through a system, know that the actual image space ray travels from B to
and F 2 is thus a characteristic of the system. The F 2 . All of the image space rays can be constructed in
zigzag ray paths through a system also determine the the same manner (Figure 11.5c).
location of the secondary principal plane H 2 , and H 2 is Figure 11.6 shows a system with H 2 in front of it
also a characteristic of the system. and F 2 behind it. The object space rays are incident

195
196 Geometrie, Physical, and Visual Optics

FIGURE 11.1. a. Zigzag paraxial ray path


through a multi-interface system, b. A is the
intersection point between the object and image
space rays. c. The intersection plane H2 for a
series of incident rays parallel to the axis.

parallel to the axis. The top incident ray intersects H 2 travels, and this line hits the back surface of the
at point A. The image space ray again lies on the line system at point B. We do not know the actual zigzag
from A to F 2 , and this line hits the back of the system path of the ray through the system, but we do know
at point B. We do not know the actual zigzag path of that the exiting ray travels from B to F 2 and then on
the ray through the system, but we do know that the out through A. The path of the bottom ray is con-
actual exiting ray travels from B to F 2 . The path of the structed similarly (Figure 11.7b).
bottom incident ray is constructed similarly. Figure 11.8 shows the construction for a case in
Figure 11.7a shows the construction for a case which F 2 is virtual and H 2 is somewhere inside the
where F 2 is real and H 2 is behind F 2 . Again, the system. The top incident ray (parallel to the axis) is
incident ray is extended and intersects H 2 at point A. extended, and the extension intersects H 2 at point A.
F2 and A specify the line along which the exiting ray The exiting ray must lie along the line that contains
The Gauss System 197

FIGURE 11.2. Secondary principal plane (in-


tersection plane) in front of a system.

FIGURE 11.3. Secondary principal plane be-


hind a system.

FIGURE 11.4. Secondary principal plane and


virtual secondary focal point.

both A and F 2 , and this line hits the back of the Figure 11.9 shows a system in which F 2 is virtual
system at B. We do not know the actual zigzag path of and H 2 is in front of F 2 . The top incident ray intersects
the ray through the system, but we do know that the H 2 at point A. The line passing through A and F 2 hits
actual exiting ray leaves B and appears to be coming the back of the system at point B. The actual outgoing
from F 2 . The path of the bottom ray is constructed ray leaves point B and appears to be coming from F 2 .
similarly. You may draw the path for the bottom incident ray.
198 Geometrie, Physical, and Visual Optics

I*

a)

A
B

. '
FIGURE 11.5. a. Object space ray extension for a
system with a known H2 and F 2 . b. Image space ray
must lie on line connecting A and F 2 . Ray actually
c) emerges from system at B. c. Multiple image and
object space rays.

-> r

FIGURE 11.6. Object and image rays for sec-


ondary principal plane in front of the system.
The Gauss System 199

FIGURE 11.7. a. Object and image space ray


for secondary principal plane behind the system.
b) A is the intersection point. B is where the ray
actually leaves the system, b. Second ray.

-31--
F2

FIGURE 11.8. Rays for a virtual secondary


focal point.

FIGURE 11.9. Prediction of image space ray


path.
200 Geometrie, Physical, and Visual Optics

11.2 The Primary Principal Plane intersection point between the extension of the top
object space ray and the extension of the correspond-
ing image space ray. All of the intersection points
Consider a point source at the primary focal point between the object space rays associated with F\ and
for a system of centered refracting interfaces. The the corresponding image space rays parallel to the axis
waves leaving the system are plane waves traveling lie in the primary principal plane H P
along the optical axis. Figure 11.10a shows the paraxi- In general, the primary principal plane Hl can be in
al ray diagram for a system with a real F P The rays front of, coincident with, or behind H 2 . The individual
that originate at zigzag through the system and interfaces in the system determine the location of F l 5
leave parallel to the optical axis. and Fj is thus a characteristic of the system. The
Both the incident and the exiting rays can be individual interfaces in the system also determine the
extended forward or backward. At some point, the location of H l 5 and thus is a characteristic of the
incident ray or its extension intersects the exiting ray system. When the location of Hl and Fx are known,
or its extension. Point E in Figure 11.10b is the then for an object space ray associated with x we can
intersection point for the top incident ray. The bottom immediately determine the actual path of the image
incident ray also has an intersection point, and the two space ray that is parallel to the axis without having to
intersection points lie in the plane labeled H l 5 which is follow the zigzag path of the ray through the system.
perpendicular to the axis. It turns out that, for all of Figure 11.12 shows a system with a real and an
the incident paraxial rays associated with F 1? the Hj located inside the system. The top incident ray
intersection points lie in the same plane. The plane Hl originates at F 1? and its extension intersects Hj at
containing these intersection points is called the pri- point E. The outgoing ray must lie along the line
mary principal plane. The point at which the optical parallel to the optical axis and passing through E. This
axis passes through is called the primary principal line hits the back surface of the system at G. We do
point. not known the actual zigzag path of the ray through
Figure 11.11 shows the paraxial ray diagram for a the system, but we do know that the actual image
system with a virtual . The incident rays that point space ray leaves G parallel to the axis.
toward zigzag through the system and leave parallel Figure 11.13a shows a system with in front of a
to the optical axis. In Figure 11.11b, the point E is the real . The incident ray from x can be extended

I ^
a)

>G -

F ^ \ ^ FIGURE 11.10. a. Zigzag paraxial ray path for


an object point at the primary focal point of a
-<r*^ multisurface system, b. For the top ray, E is the
intersection point between the object space and
image space ray, and G is where the ray actually
b) leaves the system. The primary principal plane H,
is the plane containing the respective intersection
Hi points.
The Gauss System 201

\ A 1^Lr^-^F
FIGURE 11.11. Primary principal plane for a
Hi virtual primary focal point.

Hi FIGURE 11.12. Object and image space rays.

A
Fi

a)
Hi

v\
\
\ >

/ \
/
E
/ G
FIGURE 11.13. a. E is the intersection point
for the object and image space rays. G is where
b)
Hi the ray actually leaves the system, b. Second ray.
202 Geometrie, Physical, and Visual Optics

FIGURE 11.14. Ray prediction for virtual


primary focal point.

backwards until it intersects Hx at E. The image space not know the actual zigzag path of the ray through the
ray must lie on a line passing through E and parallel to system, but we do know that the actual outgoing ray
the axis. This line hits the back of the system at G. We leaves G parallel to the axis.
do not know the actual zigzag path of the ray through Figure 11.15a shows a system with Fl (virtual)
the system, but we do know that the actual image behind the system and behind F^ The incident ray
space ray leaves G parallel to the axis. The path of the points towards F{ and its extension through Fx inter-
second ray is found similarly (Figure 11.13b). sects Hj at E. The outgoing ray must lie along the line
Figure 11.14 shows a system with inside the that is parallel to the axis and passes through E. This
system and F{ (virtual) behind the system. The top line hits the back surface of the system at G. We do
incident ray points toward Fx and intersects Hx at not know the actual zigzag path of the ray through the
point E. The outgoing ray must lie along the line that system, but we do know that the actual ray leaves G
passes through point E and is parallel to the axis. This parallel to the axis. The path of the bottom incident
line hits the back surface of the system at G. We do ray is constructed similarly (Figure 11.15b).

b) FIGURE 11.15. a. Image space ray pre-


Hi diction, b. Second ray.
The Gauss System 203

11.3 Ray Diagrams incident parallel to the axis intersects H 2 at A. The


outgoing ray must lie along the line AF 2 , and this line
For a given system, the principal planes together with hits the back surface of the system at B. We do not
the focal points provide us with two predictable tays. know the actual zigzag path of the ray through the
Consider a system with real focal points and the system, but we do know that the actual outgoing ray
principal planes located inside the system (Figure leaves the system at B and passes through F 2 .
11.16). Remember that the positions of the focal The other incident predictable ray passes through
points and the principal planes are determined by the F 1? and the outgoing ray must be parallel to the axis.
interfaces in the system. The intersection plane for these rays or their exten-
A real object of size O is located outside Fx at a sions is the primary principal plane Hj. The incident
distance of Uj from the front surface of the system. ray passes through Fl and its extension intersects at
One of the predictable rays is incident parallel to the point E. The outgoing ray must lie along a line parallel
axis, and the outgoing ray must pass through F 2 . The to the axis and passing through E. This line hits the
intersection plane for this ray or its extension is the back surface of the system at G. We do not know the
secondary principal plane H 2 . The extension of the ray actual zigzag path of the rays through the system, but

FIGURE 11.16. Real image by use of principal


Hi H2 planes.

A*
virtual
image

/
/
real
object

Ui

H2 HT

vb FIGURE 11.17. Virtual image location.


204 Geometrie, Physical, and Visual Optics

virtual
image

FIGURE 11.18. Image location for virtu-


al object.

we do know that the actual outgoing ray leaves G F 2 . The incident ray that is pointing toward Fx inter-
parallel to the axis. sects Hj at E. The exiting ray must lie along the line
The two actual exiting rays tell us that the image parallel to the axis and passing through E, and the
space light is converging and a real image of size I is actual ray leaves the system at G. The two exiting rays
formed at a distance of vh from the back surface of the indicate that the actual light leaving the system is
system. We do not know what the light inside the diverging. The image is virtual and located by extend-
system is doing, but we do know that the light leaving ing the two exiting rays backward until they intersect
the system is converging. each other. The image is inverted and the same size as
Figure 11.17 shows another system with real focal the object.
points. H 1 and H 2 are inside the system with H 2 in
front of Hj. A real object is inside F\ at a distance ux
from the front surface of the system. Again, the
extension of the ray incident parallel to the axis 11.4 Conjugacy of the Principal Planes
intersects H 2 at A. The exiting ray must lie along the
line AF 2 , and the actual exiting ray leaves the system Figure 11.19 shows a system with real focal points.
at B. When extended, the incident ray that appears to The primary principal plane H 1 is in front of the
be coming from Fx intersects at E. The exiting ray system (but closer than Fx), and the secondary princi-
lies on the line parallel to the axis that passes through pal plane H 2 is inside the system. A real object
E, and the actual exiting ray leaves the system at G. happens to be at the same position as H^ The exten-
The two actual exiting rays indicate that the image sion of the ray incident on the system parallel to the
space light is diverging. The virtual image is located by axis intersects H 2 at A. The outgoing ray lies along the
extending the actual outgoing rays back until they line AF 2 and leaves the system at B.
intersect each other. The virtual image is erect, larger The ray incident on the system that appears to be
(size I), and located a distance vb from the back coming from Fx intersects Hx at E, which is the same
surface of the system. point as the top of the object. The outgoing ray lies on
Figure 11.18 shows a system with virtual focal the line through E that is parallel to the axis, and the
points. Hl and H 2 are located in front of the system. actual ray leaves the system at G. Since E coincides
A virtual object of size O is located behind the system with the top of the object and the line through E and
at a distance of ux from the front surface. The two G must pass through the conjugate image, the image is
predictable incident rays are converging and both erect and the same size as the object.
point toward the virtual object point. The incident ray The two outgoing rays indicate that the light leav-
parallel to the axis intersects H 2 at A. The exiting ray ing the system is diverging. The image is virtual and
must lie along the line AF 2 , and the actual exiting ray located by extending the outgoing rays backward until
leaves the system at B and appears to be coming from they intersect each other. In this case, the intersection
The Gauss System 205

E, /
>
A G ,
>-
virtual B
y ' real image
s> object

Fi \

FIGURE 11.19. Object at H, gives erect image of


Hi the same size at H,.

occurs at the point A on H 2 (Figure 11.19). So for an points are real, and the principal planes are inside the
object at H 1? the conjugate image is at H 2 . The image system. A real object is located outside Fj and the
is erect and the same size as the object, so the lateral conjugate real image, found by tracing the two pre-
magnification m is +1. dictable rays, is real and located outside F 2 . A third
We can construct similar ray diagrams for other incident ray, when extended, hits the point K o n H j .
systems, and an object at Hl is always conjugate to an We know that the corresponding image space ray must
image at H 2 with a lateral magnification of + 1 . There- pass through the bottom image point. We also know
fore, we say that Hj and H 2 are unit conjugate planes. that an object point at K on H, is conjugate to an
Note that for light going in the same direction, an image point at L on H 2 with a lateral magnification of
object at H 2 does not give an image at Hl (Figure + 1. Therefore, L and K are the same perpendicular
11.20). distance from the axis, and can be connected by a line
In the previous section, we learned how to predict parallel to the axis (Figure 11.21b). Since L is conju-
the path of two special rays. In general, when the gate to K, the image space ray must lie on the line that
conjugate object and image positions are known, we connects L and the bottom of the image, and this line
can use the conjugacy of the principal planes to trace hits the back surface of the system at M. We do not
the path of any image space ray corresponding to an know the ray path inside the system, but we do know
object space ray. that the actual image space ray leaves the system at M
Consider the system in Figure 11.21a. The focal and travels to the image point.

virtual
object
A
B

E G* '

-~^r, (real
FT^
image

FIGURE 11.20. Object at same loca-


H, tion as the secondary principal plane.
206 Geometrie, Physical, and Visual Optics

Hi H2

FIGURE 11.21. a. Object space ray


intersecting primary principal plane at
K. Image space ray must leave the sec-
ondary principal plane at the same height
(L). b. Other rays.

The same procedure can be used for any other ray. 11.5 Vergence Equations
The bottom ray in Figure 11.21b is an example. The
incident ray is extended until it hits point Q on Hj. Figure 11.21b is a ray diagram in which all of the
Then we go across parallel to the axis to point S on object space rays and their extensions are straight until
H 2 . The exiting ray must lie on the line from S to the they hit the primary principal plane H 1? and all image
bottom image point, and the actual ray leaves the back space rays and their extensions are straight once they
of the system at T. leave the secondary principal plane H 2 . Figure 11.21b
The general rule is that once an object space ray or resembles a single spherical refracting interface
its extension hits a point on the primary principal (SSRI) ray diagram except that the interface is peeled
plane H 1? a line parallel to the optical axis is drawn into two layers, and H 2 . Hx interacts with the
from that point to a point on the secondary principal incident rays and is thus the object space layer, while
plane H 2 , and the corresponding image space ray lies H 2 interacts with the exiting rays and is thus the image
on the line that includes the point on H 2 and the space layer.
image point. Note that in the previous figures that the We can imagine the incident (object space) wave-
two predictable rays also follow this rule. fronts propagating all the way to H and we can
The Gauss System 207

compute the (reduced) vergence U H that they would 5.11 cm in front of the system's front surface, and
have there: the secondary focal point is located 12.78 cm be-
hind the system's back surface. The primary princi-
U = - ^ (11.1) pal plane is located 6 cm behind the front surface,
*H, and the secondary principal plane is located 8 cm
where uH is the object distance measured from H{ behind the front surface (or 2 cm in front of the
and n 0 is the object space index. back surface). A real object is located 19 cm from
the front surface of the system. Where is the
Similarly, we can imagine the exiting (image space)
conjugate image? What is the lateral magnification?
wavefront propagating away from H 2 , and we can Figure 11.22a shows the ray diagram for the
compute the (reduced) vergence V H that they would situation. For the calculations, n 0 = 1.00 and n{ =
have there: 1.33. The real object is 19 cm from the front sur-
face, or 25 cm from H1. Therefore,
VH = (11.2)
uH = - 2 5 cm.
where i;H is the image distance measured from H 2 and The incident light is diverging and in air. We are
n4 is the image space index. pretending that this incident light goes all the way
to H , . From Eq. 11.1,
As suggested by the ray diagrams, the incident and
exiting vergences are related by an equation that has 100
UH = = -4.00D.
the same form as the SSRI vergence equation, -25 cm
namely, Then
V H , = P. + U H I , (11.3) VH = P e + U p
where P e is called the equivalent dioptric power of the gives
system. (Equation 11.3 is derived in the next chapter.) VH2 = +9.00 D + (-4.00 D) = +5.00 D.
The name equivalent dioptric power refers to the We are pretending the light leaving H 2 is converg-
fact that the system is acting like an SSRI of dioptric ing and in the image space medium, water. From
power Pe between mediums of refractive indices n 0 Eq. 11.2,
and nj. Just as for an SSRI, Eq. 11.3 enables us to get 133
from the object space to the image space in one step =
without having to do the calculations for all the inter- ^ T^D=+26-6cm-
nal interfaces. In fact, the refractive indices for the The image is located +26.6cm from H 2 , or
mediums inside the system do not appear in Eq. 11.3. +24.6 cm from the back surface of the system.
From Eq. 11.4,
In Section 8.9, we discussed the back vertex power
Pv and the neutralizing power P n for a system. For -4.00 D
1 = -0.8.
plane waves incident on the system, Pv is the vergence +5.00 D
of the wavefront leaving the system's back vertex. For The calculated results agree with the ray diagram.
plane waves leaving the system, P n is equal in mag-
nitude but opposite in sign to the vergence incident on EXAMPLE 11.2
the system's front vertex. Pv and Pn provide informa- The object in the previous example is moved
tion only in their respective plane wave situations, and against the front surface of the system. Where is
hence are not true dioptric powers in a V = P + U the conjugate image? What is the lateral magnifi-
sense. On the other hand, Eq. 11.3 is good for any cation?
incident or exiting vergence, and thus P e functions as a Figure 11.22b shows the ray diagram. The real
true dioptric power. object is 6cm in front of H j . Therefore,
The total lateral magnification m t , defined as the uH = 6 cm.
ratio of the system's image size I to the system's object As far as Hl is concerned, the incident light is
size O, is given by diverging and in air. Here we are pretending that
U
this light goes all the way to H r From Eq. 11.1,
H,
mt (11.4) 100
U, -16.67 D.
-6 cm
EXAMPLE 11.1 Then
A multiple interface system with air (n = 1.00) in :
P. + U t
front and water (n = 1.33) behind has a central
thickness of 10 cm and an equivalent dioptric power gives
of +9.00 D. The primary focal point is located VH = +9.00 D + (-16.67 D) = -7.67 D.
208 Geometrie, Physical, and Visual Optics

6cm. 2cm

a) 19.11cm

n0 = 1.00 n, = 1.33

virtual
image
/
Fi

b)
H,H 2

v FIGURE 11.22. a. Predictable rays to locate a real


image, b. Predictable rays to locate a virtual image.
19.35cm

We are pretending the light leaving H 2 is diverging It follows from Eqs. 11.2 and 11.3 that
and in the image space medium, water. From Eq.
11.2, Pe = r> (5>
133
H = -17.35 cm. where f2 is the distance from H 2 to F 2 and is called the
2 -7.67 D
secondary equivalent focal length.
The image is virtual and located -17.35 cm from
Remember that the distance from the back vertex
H 2 , or -19.35 cm from the back surface of the
system (Figure 11.22b). From Eq. 11.4, L b of the system to F 2 is called the back focal length
fb, and Eq. 8.8 was
16.67 D
mt = = +2.17.
-7.67 D (11.6)
The image is erect and a little over two times larger
than the object. For incident plane waves, we can consider a hypo-
thetical image space wavefront that is either propagat-
ing from H 2 to L b or from L b to H 2 depending on
which one is downstream. The vergence of the image
11.6 Power and Focal Length Relations space wavefront at H 2 is P e , and at L b the vergence of
the image space wavefront P v . So P e and Pv are related
Consider plane waves incident on the system along the by vergence effectivity in the image space.
optical axis. The image is located at the system's When the object is placed at F l of the system,
secondary focal point. Then plane waves leave the back of the system. Then
uHl=o, v H2 = o,
and and
^H2 = f2 " H l = fi
The Gauss System 209

It follows from Eqs. 11.1 and 11.3 that Here n0 = 1.33, and nj = 1.00. For incident
plane waves,
Pe = - ^ , (11.7)
u=o,
where fj is the distance from Hl to Fj and is called the and
primary equivalent focal length. v H = P e + U H = +5.00D.
Remember that the distance from the front vertex H2 is 7 cm in front of the system's back surface L b .
of the system to Fj is called the front focal length ff, To get the back vertex power, we can imagine a
and Eq. 8.9 was +5.00 D wavefront in the image space medium (air
in this case) propagating 7 cm downstream. From
P = --^ (11.8) downstream vergence effectivity relations, the ver-
ff gence of the wavefront at Lb is +7.69 D, and Pv is
For plane waves leaving the system, the vergence thus 7.69 D. On a more formal basis, the calcula-
of the wavefront incident on the system's front vertex tions are:
Lj is - P n , and the vergence of the wavefront hypo- = ni = _ L 0 0 _
= +20.00 cm,
thetically incident on Hl is P e . We can consider the 2
P. +5.00 D
incident object space wavefront as propagating either fb = +(20.00 - 7.00) cm = +13.00 cm,
from Lj to H1 or from to Lj depending on which
one is downstream. So Pn and Pe are related by n
i 100
vergence effectivity in the object space.
Pv=
= +13.00cm = +7.69D.
From Eqs. 11.5 and 11.7, the primary and sec- Now consider an object at F1 so that plane
ondary equivalent focal lengths are related by waves leave the back of the system. Then
f,
(11.9)
where Uj is the vergence incident on the system's
Note that when the object space index n 0 equals the front surface. For plane waves leaving,
image space index n i? then fj and f2 are equal in VM = 0,
magnitude but opposite in sign. and from Eq. 11.3,
U HH = - 5 . 0 0 D .
EXAMPLE 11.3 l
A multiple interface system with an equivalent Hj is 5 cm behind the system's front vertex Lj. So
dioptric power of +5.00 D has water (n = 1.33) in the wavefront incident on Lx is upstream from Hl9
front of it and air (n = 1.00) behind it. The system and if allowed to propagate in the object space
is 10 cm thick, its primary principal plane is located medium (water in this case), gets to H1 with a
5 cm behind the front surface of the system, and its vergence of -5.00 D. From upstream vergence ef-
secondary principal plane is located 7 cm in front of fectivity relations, the wavefront at Lj has a ver-
the back surface of the system (Figure 11.23). Find gence of -6.16 D. The system can neutralize a
the back vertex power and the neutralizing power. vergence of -6.16 D incident on its front surface,

7cm
5crrf~

n 0 = 1.33
rij = 1.33

j
Fi r
?
V ->
?

H 2 HT
FIGURE 11.23. System with +5.00 D equivalent
v dioptric power.
10cm
210 Geometrie, Physical, and Visual Optics

so Pn is +6.16 D. On a more formal basis, the Let us think a little before calculating. For
calculations are: incident plane waves the vergence leaving H 2 is
^ nQ = 133 -10.88 D, and these wavefronts lose vergence as
1 = -26.6 cm, they propagate downstream. However, the back
P -5.00 D vertex power is -11.83 D, so the back vertex L b
ff = -(26.6 - 5.00) cm = -21.6 cm, must be upstream from H 2 .
For Eq. 11.10 we need fb and f2:
133
P = - ^ = +6.16 D. 133
-21.6 cm
{l
Pe -10.88 D -12.22 cm,
As shown, we can find the vertex powers if we n, _ 133
know the equivalent power and the principal plane b -11.24 cm.
Pv -11.83 D
locations. Conversely, when we know P e , P v , and P n ,
we can find the principal plane locations. An easy way Then
to find H 2 is to picture a trip from the system's back L
bH2 =
fb _
f2>
vertex to H 2 by way of F 2 . In terms of directed dis-
tances, L b H 2 = - 1 1 . 2 4 c m - ( - 1 2 . 2 2 cm)
= +0.98 cm.
LbH2 = LbF2 + F2H2,
The secondary principal plane is 0.98 cm behind the
L
bH2 = L
bF2 ~ H 2 F 2 , back vertex of the system. This agrees with our
expectations that L b must be upstream from H 2 .
or Again, let us think about the primary principal
L
bH2_fb ~U' (11.10) plane location before calculating. For plane waves
leaving, the vergence of the object space wavefront
Equations 11.5 and 11.6 give the focal lengths needed at Hj must be +10.88 D, and the vergence of the
for Eq. 11.10. To find H 1 ? picture a trip from the wavefront incident on the system must be +6.97 D.
system's front surface Lj to H t via F 1# In terms of The converging wavefronts gain vergence, so Wx
directed distances, must be downstream from the front vertex L r
For Eq. 11.11, we need ff and f1#
I ^ H ^ L ^ + F ^ , ,
n
L
iHi =
L1F1 - H j F ^ f o = 100 + 14.35 cm,
f
P -6.97 D
or 100
(11.11) f, = - = +9.19 cm.
I^H^ff-f! -10.88 D
Equations 11.7 and 11.8 give the focal lengths needed Then
in Eq. 11.11.
L . H ^ f V f,
EXAMPLE 11.4 L^H, = 14.35 cm - 9.19 cm = +5.16 cm.
A multi-interface system has air in front (n = 1.00) The primary principal plane is +5.16 cm behind the
and water (n = 1.33) behind. The central thickness front vertex of the system, which agrees with our
of the system is 4 cm. P e , P v , and Pn are -10.88 D, expectations that H, must be downstream from LY.
-11.83 D, and -6.97 D, respectively. Where are In this case, the principal plane locations are quite
the principal planes? close together (Figure 11.24).

* +-
= 1.33
n 0 =1.00

F2 A F,

(not to scale)

/W w
H2 HT FIGURE 11.24. System with -10.88 D equival-
ent dioptric power.
The Gauss System 211

11.7 Systems in Air fb = L b H 2 + f2,


fb = - 2 3 cm + 22.22 cm = - 0 . 7 8 cm.
When the system is in air, both the object space index
So F 2 is 0.78 cm in front of the system's back
n 0 and the image space index ni are equal to 1.00. The vertex. Figure 11.25 shows the ray diagram.
indices of 1 simplify the calculations, and from Eq. The real object is 49.44 cm from the front ver-
11.9, the primary and secondary equivalent focal tex, or 44.44cm in front of H j . So
lengths become equal in magnitude but opposite in
u = -44.44 cm,
sign. In this case, we can think of the system as being
equivalent to a single thin lens, except the thin lens and
has been peeled into two layers. The object space 100
UH . = = -2.25D.
layer is at H 1 ? while the image space layer is at H 2 . -44.44 cm
The object space layer Hl interacts with the incident Then
light, while the image space layer H 2 interacts with the
V = P + u,
exiting light. In practice, may be in front of,
coincident with, or behind H 2 , but in any event, the gives
incident light is referenced to Hj and the exiting light +4.50 D + (-2.25 D) = +2.25 D,
to H 9 . and
100
= +44.44 cm.
EXAMPLE 11.5 +2.25 D
A 15-cm thick system with air on both sides has an The image is located 44.44 cm behind H 2 . Since H 2
equivalent dioptric power of +4.50 D. The primary is 23 cm in front of L b , the image is real and
principal plane is located 5 cm in front of the front located 21.44cm behind L b . From Eq. 11.4, the
vertex of the system, and the secondary principal lateral magnification is
plane is located 8 cm in front of the front vertex. A
real object is 49.44 cm from the front vertex of the -2.25 D
system. Where is the conjugate image? What is the =-1.
+2.25 D
lateral magnification?
To draw a ray diagram, we need to locate the
focal points. From Eq. 11.5,
100 11.8 The Symmetry Points
f = +22.22 cm,
-pl +4.50 D
Previously, we saw that the symmetry points 2Fj and
and from Eq. 11.7, 2F 2 gave a lateral magnification of 1. In the above
:
example, the lateral magnification is 1, the object
1 -22.22 cm. distance as measured from Hj (-44.44 cm) is twice the
Pp +4.50 D
primary equivalent focal length (-22.22 cm), and the
From Eq. 11.11,
image distance as measured from H 2 is twice the
ff = L 1 H 1 + f 1 , secondary equivalent focal length. Hence, we can
ff = - 5 cm + (-22.22 cm) = -27.22 cm. identify 2 and 2F 2 as the symmetry points for the
Gauss system, provided that by 2 and 2F 2 we mean
So Fj is 27.22 cm from the system's front vertex. H 2 that
is 8 cm in front of Ll, or 23 cm in front of the back
vertex L b . From Eq. 11.10, = 2f, and = 2L (11.12)

FIGURE 11.25. Inverted image is the


H 2 Hi same size as the object.
212 Geometrie, Physical, and Visual Optics

The above equations give the symmetry points (m = or


- 1 ) even when the object and image space indices ^,
differ from 1. Thus, 2 and 2F2 are the symmetry LjH^ (11.17)
points in all cases. Figure 11.18 is a ray diagram for a
symmetry point case in which the equivalent dioptric Note that the distance hlHl acts like an object space
power is negative. distance.

EXAMPLE 11.6
Where are the symmetry points for the system in Image Space Distances
Example 11.1?
From Example 11.1, n0 = 1.00, nj = 1.33, and 11 (11.18)
P = +9.00D. So v. = n
;

= ni = _ 1 3 3 _
2 = +14.78 cm, f = ^ (11.19)
P. +9.00 D n
i
and
-f = ^ (11.20)
100 b
fl = -11.11 cm. n,*
P +9.00 D Then from Eq. 11.10,
For the symmetry points
LbH2 = fb-f2> (11.21)
2f, = 2(11.11 cm) = -22.22 cm,
fu f.2
and LbH2 _ *b L

I>H2 = 2f2 = 2( +14.78 cm) = +29.56 cm.


n n
i i

You may verify that the lateral magnification is - 1 . fu -U


LbH2 _ _^b \2L

Note the small differences in object distance and Hi '


lateral magnification between this example and Ex-
ample 11.1. or
LuH.
LhH7 _ ^bAA2 (11.22)

Note that the distance L b H 2 acts like an image space


11.9 Reduced Systems distance. The vergence and dioptric power equations
are:
For a reduced system, the object and image space
indices are both 1. Furthermore, the dioptric powers V2 = l/ 2 , (11.23)
and vergences in the reduced system must equal those U H , = 1/"H,> (11.24)
in the actual system. Then from Eqs. 11.1 through (11.25)
P. = l/f2 = - - i / f
11.9, the reduced distances are as follows:
.= 1/
lb and Pn = -l/f f . (11.26)
Equations 11.13 through 11.26 enable us to work in a
Object Space Distances reduced system when advantageous to do so.

H
i
"1
n '
(11.13) EXAMPLE 11.7a
A 21-cm glass (n = 1.500) rod with spherical sur-
faces has ethyl alcohol (n = 1.362) in front and
(11.14) water (n = 1.333) behind. The respective dioptric
power of the front and back interfaces are -6.00 D
(11.15) and +6.00 D. The equivalent dioptric power of the
system is +5.040 D. When reduced, the system is
represented by lenses Lx and L b , representing the
Eq. 11.11, rod's front and back interfaces, separated by re-
(11.16) duced central thickness of 14 cm. In the reduced
M?>ff-fi, system, the primary principal plane is 2.667 cm
behind Lb and the secondary principal plane is
n0 n0 16.667 cm behind Lb (Figure 11.26a). Find the focal
points in the reduced system.
= ^ From Eq. 11.25, the equivalent focal lengths
The Gauss System 213

-6.00D 2.66cm
f + 6.00D J 1- (reduced)
3.17cm

Fi F2

t \ H,
_
H

JV
2
- .- . _y
a) 14cm 16.7cm 19.9
19.8cm

n, = 1.333 (actual)
6.00D
n 0 = 1.362
n = 1.50

Fi + 6.00D*^

4.3
Hi
Y y v FIGURE 11.26. a. Reduced system, b.
b) 21cm 22.2cm 26.4cm Actual system.

1 1QQ From Eq. 11.22,


?2 = = = +19.841 cm,
P. +5.040 D LbH2 - n j L b H 2 ,
L b H 2 = (1.333)( + 16.667cm) = +22.217 cm.
and
Figure 11.26b shows the actual system.
fi = -19.841 cm.
From Eq. 11.22,

11.10 The Nodal Points


= L f
^b bH2 + 2>
fb = +16.667 cm + 19.841 cm = +36.508 cm.
Hj is 2.667cm behind L b , or 16.667cm behind Lj. Consider Figure 11.21b, which is a diagram for an
From Eq. 11.17, actual system. A number of the top rays are bent
f ^ L J ^ + f,, down by the system, and a number of the bottom rays
are bent up by the system. Somewhere in the middle
ff == +16.667 cm + (-19.841 cm) there must be a ray that does not bend, although it
= - 3 . 1 7 4 cm. may be displaced due to the fact that H{ and H 2 are
not at the same position. The ray that does not bend is
the nodal ray.
EXAMPLE 11.7b The outgoing nodal ray is always parallel to the
For the previous example, find the principal planes incident nodal ray (Figure 11.27). The point at which
and the focal points in the actual system.
the incident nodal ray or its extension hits the optical
From Example 11.7a, f f = - 3 . 1 7 4 c m and fb =
+36.508 cm. From Eq. 11.15, axis is called the primary nodal point N{. The point at
which the exiting nodal ray or its extension hits the
ff = n0ff, axis is called the secondary nodal point N 2 . As shown,
ff = (1.362)(-3.174 cm) = -4.323 cm. the nodal points do not have to coincide with the
From Eq. 11.20, principal points.
Figure 11.28 shows an actual ray diagram for inci-
dent plane waves. The two incident rays are parallel to
fb = (1.333)(+36.508 cm) = +48.665 cm.
each other and both make an angle w with the axis.
From Eq. 11.17, One is the nodal ray and the other passes through F , .
I^H^nJ^H;, Since plane waves are incident on the system, the
LlHl =(1.362)( +16.667 cm) = +22.700 cm. image is in the secondary focal plane.
214 Geometrie, Physical, and Visual Optics

FIGURE 11.27. Primary and secondary nodal


points.

The exiting nodal ray, being parallel to the incident I = f, tan w. (11.29)
nodal ray, makes the same angle w with the axis. The Note that we can use either Eq. 11.28 or Eq. 11.29 to
incident nodal ray or its extension intersects H t at a calculate the image size for a distant object that
distance g above the axis, so the exiting nodal ray or subtends the angle w. In particular, fx is quickly
its extension has to leave H 2 at the same distance g determined from the equivalent dioptric power, and
above the axis. Then from the triangles involving H so, Eq. 11.29 is easy to use. From Eqs. 11.28 and
and N, 11.29,
g and tan w = (11.30)
tan w = H2N2' fi = - N 2 F 2
,,
Then but from the figure,
HjN^g/tanw and H 2 N 2 = g/tan w. N 2 F 2 = H 2 F 2 - HN,
Consequently, the displacement H J N J equals the dis- or
placement H 2 N 2 , and the subscripts are usually drop-
N 2 F 2 = f2 HN.
ped, i.e.,
HN = H 1 N 1 =H 2 N 2 . (11.27) It follows from Eq. 11.30 and the above equation that
From the triangle involving N2 and the image of HN = f 1 +f 2 . (11.31)
size I (which is negative), Equation 11.31 indicates that the displacement of the
I=-N2F2tanw. (11.28) nodal points from the principal planes is equal to the
sum of the primary and secondary equivalent focal
For the incident ray through F n the exiting ray is lengths.
parallel to the axis, leaves Hi at point A, and passes The nodal point of an SSRI is at the center of
through the image point. Consequently, the distance curvature of the SSRI. In other words, the nodal point
along Hj from the axis to A is equal to the image size for an SSRI is displaced from the interface by the
I. Then from the triangle involving H 1? A, and F 1? distance r where r is the interface's radius of curva-

"K
\

Hi
"K N1 H2 N^VVl

FIGURE 11.28. Predictable rays for incident


plane waves.
The Gauss System 215

ture. From Eq. 7.25 for an SSRI, EXAMPLE 11.8


When we look at an object, our eye is a refracting
r = f 1 +f 2 . system between air (the object space medium) and
Equations 11.31 and 7.25 are clearly analogous. the eye's vitreous humor (the image space
Equations 11.5 and 11.7 can be substituted into Eq. medium). In Gullstrand's #1 schematic eye, the
11.31 with the result that equivalent power is +58.64 D, the primary and
secondary principal planes are, respectively,
HN = ni - n n (11.32) 1.348 mm and 1.602 mm behind the front surface of
the cornea. Where are the nodal points of the eye
relative to the principal planes? (The vitreous index
Equation 11.32 is analogous to the familiar SSRI is 1.336).
equation First, let us think about which way we expect
r ^ n
2~"i the displacement to be. A positive SSRI between
P air and vitreous would have a center of curvature in
The analogous equations reemphasize the fact that the the vitreous, so the displacement should be toward
the vitreous (Figure 11.30).
system acts like an SSRI of power Pe between the From Eq. 11.32,
object space medium (index n 0 ) and the image space
medium (index nt). (1.336-1.000)1000 mm/m
HN:
Because of the SSRI analogy, it is easy to decide +58.64 D
which way the nodal points are displaced away from 336
the principal planes. We simply think of which way the HN = +5.730 mm.
+58.64 D
corresponding SSRI would be curved. Consider a The nodal points are displaced 5.73 mm from the
multi-interface system with a positive equivalent diop- principal planes towards the vitreous. Then Nx and
tric power separating water on the left from air on the N2 are, respectively, 7.078 mm and 7.332 mm be-
right. The corresponding SSRI between water and air hind the anterior cornea. In Gullstrand's #1 eye,
has a positive P, and is convex to the lower index the posterior surface of the crystalline lens is
medium, so C is to the left of the interface (Figure 7.2 mm behind the anterior cornea, so the nodal
11.29a). Consequently, the nodal points for the multi- points straddle this surface. It is interesting to note
interface system are displaced to the left (i.e., toward that a small posterior cataract in the nodal point
the water). When the equivalent dioptric power is region is much more detrimental to vision than a
large cataract elsewhere in the crystalline lens.
negative, the displacement is the other way (i.e.,
toward the air as shown in Figure 11.29b). Figure 11.31a shows two nodal rays, one traveling
When the object and image spaces have the same slightly downward and one traveling slightly upward.
refractive index, then HN is zero and the nodal points Consider a virtual point source located at N,. The
coincide with the principal points. In a reduced sys- incident rays in Figure 11.31a are two of the object
tem, the object and image space indices are both one. space rays for the virtual object point. The exiting rays
Therefore, in a reduced system, the nodal points are in Figure 11.31a are the corresponding image space
at the principal points. rays, and they show that the conjugate image is virtual

water air
water

Ni

a) : > 0
HT H2

water / air
water

Ni

FIGURE 11.29. SSRI analogy for nodal point


b) <0 displacement, a. Positive equivalent power, b.
HI H2 Negative equivalent power.
216 Geometrie, Physical, and Visual Optics

FIGURE 11.30. Principal planes and nodal points for the eye.

FIGURE 11.31. Object at primary nodal


l^ H2 plane, a. Axial rays. b. Off-axis rays.

and located at N 2 . We can repeat this type of diagram principal points (Hl and H 2 ), the primary and sec-
for any system, so that an object at Nx is always ondary nodal points (Nx and N 2 ), and the primary and
conjugate to an image at N 2 . For an extended object secondary symmetry points (2F t and 2F2where the
at N15 the lateral magnification is 0 /;. (You can doubling is done with the equivalent focal lengths).
verify the lateral magnification for the conjugate nodal When the direction of light through a system is re-
planes by using Eqs. 11.1, 11.2, 11.4, and 11.27.) The versed, each pair of cardinal points stay in the same
nodal point case is analogous to an SSRI when the place but exchange roles (i.e., Fj and F 2 exchange
object is at the center of curvature C. Figure 11.31b roles, Hj and H 2 exchange roles, Nx and N 2 exchange
shows an extended ray diagram for an object for a roles, and 2Fj and 2F2 exchange roles). Figure 11.32a
system with a positive P e , and ni greater than n 0 . shows the cardinal points for light traveling right, and
Figure 11.32b shows the same system for light travel-
ing left.
EXAMPLE 11.9
11.11 The Cardinal Points and Reversibility Gullstrand's #1 schematic eye (used in Example
11.8) is a model of the typical human eye. In
The following points are collectively called the cardi- Gullstrand's #1 eye, the retina is located 24.0 mm
nal points of a system: the primary and secondary from the anterior corneal surface. Where is the far
focal points (Fl and F 2 ), the primary and secondary point for Gullstrand's #1 schematic eye?
The Gauss System 217

-* +
light traveling
right

2F, Fi 2F 2

a)

Hi

+ < light traveling


left

2F 2 F2 N2 Ni Fi 2F,

b) FIGURE 11.32. a. Gauss parameters for light


traveling right, b. Same system for light traveling
H2 Hi left.

The far point is conjugate to the retina, and we From Eq. 11.2,
could find it by using the image distance and work- 100
ing backwards. Alternatively, we can consider light v = - 1 0 0 cm.
coming out of the eye (i.e., reversed), which is the -1.00 D
situation when doing ophthalmoscopy or retin- For the reversed light, H 2 is 1.348 mm behind the
oscopy. anterior cornea, so the far point is 100.1348 cm
For light going into the eye, the equivalent from the cornea. It is interesting to note that for
dioptric power, principal plane locations, and re- people aged 35 or younger, a 1 D hyperope has
fractive indices are given in Example 11.8 and accommodation to spare and essentially functions
shown in Figure 11.30. Figure 11.33 shows the the same as an emmtrope.
schematic eye for light coming out, in which case
Hj and H 2 have exchanged roles. Thus, the re-
versed H, is 1.602 mm behind the cornea, or EXAMPLE 11.10
22.398 mm from the retina. For the reversed light, Consider an axial myopic eye with refracting inter-
faces the same as those in Gullstrand's # 1 eye, but
n 0 = 1.336 and ^ = 1.000. with a lengthened vitreous chamber so that the
From Eq. 11.1, distance from the anterior cornea to the retina is
30 mm. Find the distance vision contact correction
1336 for the myopic eye.
-59.65 D.
-22.398 mm As compared to Example 11.9, the only change
The equivalent dioptric power does not change, is the axial length. So for the reversed situation
and from Eq. 11.3, (light coming out of the eye), the retina is
28.398cm from H, (i.e., 3 0 - 1 . 6 0 2 ) . Then
VH2 = +58.65 D + (-59.65 D) = - 1 . 0 0 D.
1336
The light leaving Gullstrand's # 1 eye is diverging, U = -47.05 D,
and the image is virtual. As referenced to the -28.398 mm
principal planes, the typical eye is 1 D hyperopic. VH = +58.64 D + (-47.05 D) = +11.59 D,

rl 2 H-
FIGURE 11.33. Retina as the object.
218 Geometrie, Physical, and Visual Optics

100 11.13 Newton's Equation


= +8.63 cm.
+ 11.59D
The eye is 11.59 D myopic as measured at the Newton's equation can be applied to a Gauss system.
principal planes. The far point is real and 8.63 cm Figure 11.36 shows a Gauss system ray diagram. Con-
in front of the reversed H 2 , or 8.49 cm in front of sider the incident ray through F,. The exiting ray
the cornea. The contact correction needed is then leaves parallel to the axis at a distance I from the axis.
-11.77D (i.e., 100 divided by -8.49). The triangle involving the incident ray, the object, and
the extrafocal distance x is similar to the triangle
involving the incident ray, the primary equivalent
11.12 SSRI and Thin Lens as focal length fl5 and Hj. Then
a Gauss System O -I
x
Consider an SSRI as the system. Where are the
principal planes? Figure 11.34 shows a standard parax- or
ial ray diagram for a converging SSRI. The rays bend I "fi
(are refracted) only at the interface, so the object and (11.33)
image space intersection points occur only at the Consider the incident ray parallel to the axis, which is,
interface. Hence, the principal points H{ and H 2
of course, a distance O from the axis. The triangle
coincide at the vertex of the interface. This is consis-
involving the outgoing ray, the secondary equivalent
tent with the fact that for an SSRI, an object distance
focal length f2, and H 2 is similar to the triangle
of zero gives an image distance of zero and a lateral
involving the outgoing ray, the extrafocal distance x',
magnification of +1. For the SSRI, the nodal points
and the image. Then
are displaced from the principal planes, and they are
both coincident with C, the center of curvature of the O -I
interface. U
For a thin lens in air, a standard ray diagram, such
as Figure 11.35, shows that the object and image space or
intersection points occur in the plane of the lens. l_
m (11.34)
Thus, Hj and H 2 are both in the plane of the lens. O f,
Since the object and image space indices are the same
(n = 1), the nodal points are not displaced and both N, Equations 11.33 and 11.34 can be combined to give
and N2 are located at the optical center of the lens. xx' =f1f2. (11.35)

real
image

FIGURE 11.34. Principal planes for


an SSRI.

FIGURE 11.35. Principal planes for a thin


lens.
The Gauss System 219

FIGURE 11.36. Newton's parameters.

In terms of the vergences at the focal points, Eq. 11.35 object placed 20 cm from the front of the system
becomes gives a conjugate real image 18.67 cm from the
XX' = - ( P e ) 2 (11.36) back of the system with a lateral magnification of
-0.667 (Figure 11.37a). What is the equivalent
Equations 11.33 or 11.34 can be used directly to dioptric power of the system? Where are the princi-
determine the equivalent focal length of a system even pal planes? The extrafocal distances are:
when the principal plane location is not yet known. x = - 1 5 c m and x' = +6.67cm.
From Eq. 11.34,
EXAMPLE 11.11
An 8-cm thick system in air has a front focal length x' -6.67 cm
of - 5 cm and a back focal length of +12 cm. A real U= m = +10.0 cm.
-0.667

real
A object

x'

J=i F2
X
real ^
'
image

a)

FIGURE 11.37. a. Newton's parame-


ters. b. Ray diagram.
220 Geometrie, Physical, and Visual Optics

Then to the object space nodal ray. The conjugate image


100 point lies on the nodal ray, so the conjugate image is
P = = +10.0D, displaced down relative to the unrotated lens in Figure
+ 10.0 cm
and from Eq. 11.7, fx equals -10.0 cm. From Eq.
11.38a. Figure 11.38c shows the results of a rotation
11.10, around the primary nodal point Nj. The nodal ray is
traced, and again the image is displaced down relative
L b H 2 = +12.0 cm - 10.0 cm = +2 cm. to Figure 11.38a. Figure 11.38d shows the results for a
From Eq. 11.11, rotation around the secondary nodal point N 2 . Here,
L ^ ! = - 5 . 0 cm - (-10.0 cm) = +5.0 cm. the image point has not moved relative to that of
Figure 11.38a, and thus, the image is stationary for
Figure 11.37b shows the results.
rotations around N 2 .
When the light through the system is reversed, then
F! and F 2 trade places. Hl and H 2 trade places, and Nj
11.14 The Nodal Slide and N2 trade places. Applying the principal of re-
versibility to Figure 11.38d shows that for an object at
The nodal slide is an apparatus that is used to measure F l 5 the image is at optical infinity and is stationary for
the location of a system's nodal points. For a system in rotations around Nj.
air, the nodal points are at the principal planes, which In summary, for an object at optical infinity, the
are then located by this method. stationary point is at N 2 ; for an object at F l 5 the
Assume the system has a horizontal optical axis. stationary point is Nj. As an object is moved from
The method consists of rotating the system back and optical infinity to F 1? the stationary point moves from
forth around a vertical axis and observing the effects N2 to .
on the image. Typically, the image moves back and The nodal slide is used to make the rotations. For a
forth on the image screen. However, the speed of distant object, N2 is determined by the stationary
movement of the image varies depending on the point point procedure. As opposed to using an object at Fx
of rotation. At some point, the image remains station- to find N 1? we usually turn the nodal slide around so
ary during the rotation. This is the end point of the that the light from the distant object is now passing
method. backwards through the system. We determine N2 for
Figure 11.38a shows a light from a distance axial the reversed situation by the stationary point proce-
object point, in which case the image is at F 2 . Figure dure, but the reversed N2 is just the N t of the original
11.38b shows the system rotated about some point R. system. So now both Nj and N2 of the original system
The image space nodal ray is displaced down relative are known.

F
2 after
rotation

light ray
in

FIGURE 11.38. a. Rays from a dis-


tant axial object point, b. System ro-
tated around point R. c. System ro-
out F2 after tated around primary nodal point, d.
rotation
Stationary image for system rotated
c) d) around the secondary nodal point.
The Gauss System 221

11.15 Thick Lenses effectivity Eq. 6.9 gives the vergence U 2 incident
on the back surface:
As shown, the equivalent power of a system can be Pi
found by using Eqs. 11.34 and 11.5. However, it U, = -
1-dP,
would be useful to be able to find the equivalent
power directly from the individual powers and (re- Then V = P + U for the second surface gives
duced) distances in the system. For systems with three
=p +
or more surfaces, the classical procedure for doing the
latter is very tedious and mistake-prone. The modern
Pv
* r -dP,'
procedure is a simple method that utilizes matrices,
and is covered in Chapter 12. For two surface systems, Pv P 2 ( i - dPJ + P,
there is a simple equation for P e . A geometric deriva- 1--d>,
tion is given below, and a derivation using matrices is and
given in Chapter 12. _ P, + P2 -dP,P 2
P (11.37)
A thick lens has two refracting surfaces. When 1-dR
reduced, the thick lens is represented by two sepa- We need Eq. 11.37 for the deviation of Eq. 11.41.
rated thin lenses. Therefore, the equations below Figure 11.39a shows the secondary principal
apply to either two separated thin lenses or to a thick plane and an incoming ray parallel to and at a
lens. distance a above the axis. The secondary principal
First let us review the procedure for finding the point is labeled H 2 . The point A on the secondary
back vertex power P v of a thick lens. The back vertex principal plane is the intersection point between the
power is just the vergence leaving the back surface extensions of the object space ray and the image
when plane waves are incident on the front surface. space ray. The actual image space ray leaves the
back surface at point B and travels to the system's
F 2 . The distance from the axis to A is a, and the
EXAMPLE 11.12 distance from the axis to B is b. The triangles
An 11.4-mm thick ophthalmic crown (n = 1.523) AH 2 F 2 and BL b F 2 are similar. It follows that
spectacle lens has a front surface power of
+16.00 D and a back surface power of -2.00 D. a b^
What is the back vertex power of the lens?
d 1.14 cm
= 0.75 cm.
n 1.523 aP. = bP
Let or
U,=0. a Pv (11.38)
Then b=
Vx = P j + U j = +16.00 D, Figure 11.39b shows the actual ray path. The points
Ll and L b are the vertex points of the first and
100 second surface, respectively. The incident ray is
v1 = +6.25 cm,
+16.00 D parallel to and at a distance a from the axis, and
2 = +6.25 cm - 0.75 cm = +5.50 cm, hits the first surface at point E. The ray leaving the
first surface points toward the secondary focal point
100 (F 2 )j of surface 1 and hits the second surface at B.
U = +18.18D,
+5.50 D The image space ray goes from B to the system's
F 2 . The triangles L 1 (F 2 ) 1 and BL b (F 2 ) 1 are simi-
V2 = P2 + U 2 = -2.00 D + ( + 18.18 D),
lar. It follows that
P V =V 2 = + 1 6 . 1 8 D.
(y> (f2),-d'
Consider the procedure algebraically (instead of
numerically). We want to determine the back ver- aP^
tex power of a lens with a refractive index n, a (1/PJ-d'
central thickness d, a front surface power P 1? and a
1
back surface power P 2 . The reduced system, shown (11.39)
in Figure 11.39a, consists of two thin lenses of b 1 - dPj '
dioptric powers Pj and P2 separated by the reduced By substituting Eq. 11.39 into 11.38, we obtain
thickness d (where d = d/n). For plane waves inci-
dent on the front, the vergence leaving the front P.
P = (11.40)
surface equals P19 and the downstream vergence O-dPi)
222 Geometrie, Physical, and Visual Optics

FIGURE 11.39. a. Secondary principal plane


for two lens system, b. Actual ray path for two
lens system.

Then by comparing Eqs. 11.40 and 11.37, we power comes from the downstream vergence equa-
obtain tion for the lens' first surface (power P J . For the
(11.41) reversed light case, the denominator of Eq. 11.42
P =R+P9-dRP9
comes from the downstream vergence equation for
Equation 11.41 gives the equivalent dioptric power the lens' back surface (power P 2 ).
of a two-surface system directly in terms of the When the reduced central thickness d goes to
powers and reduced central thickness of the sys- zero, the lens becomes thin. For a thin lens, Eqs.
tem. Eq. 11.41 is a linear equation as a function of 11.40 and 11.42 show that the back vertex power
the reduced thickness d. Therefore, a plot of Pe as and the neutralizing power are both equal to the
a function of d is a straight line with an intercept equivalent power. Equation 11.41 shows that the
equal to the sum of the surface powers and a slope thin lens equivalent power is equal to the sum of
equal to minus the product of the surface powers. the surface powers, or the equivalent power just
For a biconvex or biconcave lens, the slope is equals the thin lens power.
negative (Figure 11.40a). For a planoconvex or
planoconcave lens, there is a zero slope (Figure
11.40b); and for a meniscus lens the slope is posi- EXAMPLE 11.13
tive (Figure 11.40c). What is the equivalent power and neutralizing
power for the lens in Example 11.12? Also use Eq.
Equation 11.40 gives a connection between Pv
11.40 to verify the back vertex power obtained in
and P e . By reversibility, there must be a similar
Example 11.12.
connection between P and P
From Example 11.12,
P = (11.42) Px = +16.00 D, P2 = - 2 . 0 0 D ,
l-dP9 and
In Eqs. 11.40 through 11.42, the reduced dis- d = 7.5 mm = 0.0075 m.
tance d is always positive, and dimensionally has to From Eq. 11.41,
be in meters (since it is multiplying diopters).
Note from the derivation of Eq. 11.37, that the Pe = +16.00 D + (-2.00 D) - (0.0075 m)
denominator of Eq. 11.40 for the back vertex x ( + 16.00 D ) ( - 2 . 0 0 D ) ,
The Gauss System 223

+ b +5 + 5r
1
+ 4 biconvex +4 Piano- Convex
Pe(inD) Pe(inD) Pe(inD)
+ 3' +3 +3+

+ 2 +2

+1 ' +1

0 0
0.5 1.0 l \ 0.5 1.0 1.5 0.5 1.0 1.5
FIGURE 11.40. Equivalent
-1 -1
3(inm) d(inm)> d(inm) > dioptric power as a function of
-2- -2 thickness, a. Biconvex lens. b.
Planoconvex lens. c. Meniscus
b) c) lens.

Pe = +14.00 D - (-0.24 D) = +14.24 D. power, back vertex power, and neutralizing power as a
The sum of the surface powers is +14.00 D, and function of reduced thickness for a lens with a front
the thickness term contributed 0.24 D more. surface power of +10.00 D and a back surface power
From Eq. 11.42, of +5.00 D. The graph of the equivalent power is
linear with a negative slope and a power of +15.00 D
+ 14.24 D
P = for zero thickness. The reduced distance that gives
1-(0.0075 m ) ( - 2 . 0 0 D ) ' zero equivalent power is found from Eq. 11.41:
+14.24 D
= P e = 0 = P1 + P 2 - d P 1 P 2 ,
- = +14 030

0=+15.00 D-d(+50.00D2),
To verify the back vertex power, Eq. 11.40 gives
+14.24 D or
+ 15.00
v
" 1 - (0.0075 m)( +16.00 D) ' d = = 0.30 m = 30 cm.
+50.00
+ 14.24 D The back vertex power is equal to the vergence
+ 16.18D,
0.880 leaving the back surface when plane waves are inci-
which agrees. dent on the front surface. In Figure 11.41, the back
vertex power is +15.00 D at zero thickness and then
Figure 11.41 shows a plot of the equivalent dioptric starts to increase as the thickness increases. This

Pv/
PT ^r
+ 30
^
^y
+ 20"

+ 15*

^
+ 10 h-E:
+ 5

0
5 10 15 20 ^25-
-5
(in cm) ^
-10 ^

-20

-25 1
p
-30
\I 1i 1 / FIGURE 11.41. Graph of power parameters for
Pn a lens as a function of thickness.
224 Geometrie, Physical, and Visual Optics

increase occurs because of the vergence gain of the EXAMPLE 11.15


converging light across the interior of the lens. At a Small plastic (n = 1.49) spheres mounted on the
reduced thickness of 10 cm, the image from the end of a pencil are useful in visual testing of small
+10.00 D front surface is formed at the back surface children. For paraxial viewing through a sphere of
and the back vertex power goes to infinity. For a diameter 2 cm, what is Pe, Pv, and Pn? Where are
the principal points?
reduced thickness just greater than 10 cm, the image Let us first find the surface power of the sphere.
for the +10.00 D front surface is formed just in front The radius of curvature is 1 cm. Then for the front
of the back surface. The light diverges away from the surface,
internal image position, and highly divergent light is
incident on the back surface. The back vertex power is n.-n, +49
L
then a large negative number indicating that highly R = = r = +49.00 D.
1
divergent light leaves the back surface of the lens. As r +1
The sphere has the same surface power all around,
the thickness is increased further, the diverging wave- so the back surface power P2 is also +49.00 D.
fronts incident on the second surface become flatter, Then
and the light leaving becomes less divergent. Eventu-
d = TT77 = 134 cm = 0.0134 m.
ally, the vergence incident on the +5.00 D back sur- 1.49
face is -5.00 D, resulting in plane waves leaving the From Eq. 11.41,
lens and a back vertex power of zero. As is evident
Pe = +49.00 D + 49.00 D - (0.0134 m)(+49.00 D)
from Eq. 11.40, the zero back vertex power occurs at
the same distance (30 cm) that gives zero equivalent x (+49.00 D),
dioptric power. As the reduced thickness is increased Pe = +98.00 D - 32.23 D = +65.77 D.
past 30 cm, the light incident on the back surface has a From Eq. 11.40,
vergence that is negative but smaller in magnitude
than 5.00 D, so the light leaving the back of the lens +65.77 D
v
becomes convergent again. Eventually, the lens is so " 1-(0.0134)(+49.00D)'
thick that the wavefronts coming from the internal +65.77 D
P
image become plane waves before they are incident on -^34T- = +19216D

the +5.00 D back surface. In the limit of infinite Since the sphere is perfectly symmetrical, Pn should
thickness, the back vertex power goes to +5.00 D. equal Pv, and indeed Eq. 11.42 gives
By reversibility, a similar argument can be made Pn = +192.16D.
for the neutralizing power. Note that when the equi-
valent power is zero, the neutralizing power also
equals zero. A system with an equivalent power of For the Principal Points
zero is called an afocal system. Afocal systems neither
converge nor diverge light, but they do affect magnifi- From Eqs. 11.5 through 11.8,
cation. As discussed in Chapter 13, telescopic systems =
focused for emmtropes are afocal systems. f2 = .fie 7 7 D +1.52cm, ft = -1.52cm,

From Eqs. 11.10 and 11.11,


11.16 Special Thick Lens Examples LbH2 = +0.52 cm - 1.52 cm = -1.00 cm,
L1H1 = -0.52 cm - (-1.52 cm) = +1.00 cm.
EXAMPLE 11.14
Verify that the equivalent dioptric power given in Since the radius of the sphere is 1 cm, the principal
Example 11.7a is correct. points coincide at the center of the sphere (Figure
From Example 11.7a, ^ = -6.00 0 , P2 = 11.42a). Since the object and image space indices
+6.00 D, and d = 21 cm/1.5 = 14 cm = 0.14 m. are equal, the nodal points are at the principal
From Eq. 11.41, points and hence at the center of the sphere. We
can verify this by noting that any incident ray
Pe = -6.00 D + 6.00 D - (0.14 m)(-6.00 D) pointing toward the center of the sphere is incident
x(+6.00D), normally on the sphere and does not bend. These
rays travel straight through the center of the
Pe = 0 - (-5.040 D) = +5.040 D, sphere, are incident normally on the back side, and
which agrees. Note that here the entire contribu- again do not bend (Figure 11.42b).
tion to the equivalent power comes from the thick- Note that for paraxial imaging, we can computa-
ness term. tionally replace the sphere by a +65.77 D thin lens
The Gauss System 225

a) FIGURE 11.42. Sphere, a. Principal planes, b. Nodal


rays.

located at the position occupied by the sphere's For the Principal Plane Locations
center. This makes the analysis of the sphere's
optics particularly straightforward. From Eqs. 11.10, 11.7, and 11.8,

EXAMPLE 11.16
Consider a thick lens with a convex front surface LbH2=
and a plane back surface (Figure 11.43a). Where \;"v
are the principal planes located? 1-dP, 1
Here P2 = 0, and LbH2 = -d.
P P
r r
i i
^, + ,-,^,.
Then
From Eqs. 11.10, 11.5, and 11.6,
LiH^-f,,
p =
1 - dR ' L i H ^ - 1 + 1,
x
n e
becomes

Pv = L .
1 - dV,
For the given lens, H l is at the vertex of the front
For plane waves incident on the front, the vergence surface, which is the only surface with a non-zero
leaving the front is equal to P 1? and the above dioptric power (Figure 11.43b). H 2 is in front of the
equation is just the downstream vergence equation back surface by the reduced thickness of the lens.
for the reduced distance d. Since the back surface is In this example, the image space is air, so ni is 1.
plane, the exiting vergence (which is the back If n{ differs from 1, then the above derivation gives
vertex power) equals the vergence of the light LbH2 = -r^d.
incident on the back surface.
Since the back surface is plane, when we reverse
EXAMPLE 11.17
the light and consider plane waves incident on the
Given the results of the above example, where are
back surface, the plane waves proceed through and
the principal planes for a lens with a piano front
are incident on the front surface. The emerging
surface and a spherical back surface?
light then has a vergence equal to , and Pn equals
When the light through a lens is reversed, Hj
Pj. Indeed, Eq. 11.42 gives
and H 2 exchange places. By reversibility, Hl is a
Pe Pi distance d behind the piano surface, while H 2 is at
P =
=

1 - dP, 1-0 Pi the spherical back surface, which is the only surface
with a non-zero refracting power (Figure 11.44).

FIGURE 11.43. a. Planoconvex thick lens. b. Principal points. FIGURE 11.44. Principal points for a planoconcave lens.
226 Geometrie, Physical, and Visual Optics

11.17 Summary of the Important Equations Problems

General 1. An optical system in air has an 11-cm central


thickness. The primary principal plane is 3 cm
VH2 = Pc + U H l , (11.3) behind the front surface, and the secondary princi-
pal plane is 6 cm in front of the back surface. If
(11.1) the back focal length is 5 cm, what is the equival-
WHJ,
ent dioptric power? What is the neutralizing
power? What is the back vertex power? For a real
(11.2) object 20 cm from the front surface of the system,
where is the conjugate image relative to the back
U H" 1. surface? What is the lateral magnification?
mt = (11. 4)
V
H2
2. An optical system in air has a 7-cm central thick-
ness and a +6.50 D equivalent power. When a
For incident plane waves, Eqs. 11.3 and 11.2 give real object is placed 14 cm in front of the system,
n
i its conjugate image is erect, the same size as the
(11..5)
Pe
~2 object and located 3 cm in front of the back
From the back vertex power definition, surface. (It is a virtual image.) For a real object
located 40 cm in front of the system, where is the
(11 .6) conjugate image relative to the back surface?
Pv
What is the lateral magnification?
For plane waves leaving, Eqs. 11.3 and 11.1 give 3. An optical system with air in front and water
(n = 1.33) behind has 6-cm central thickness. The
n
P.= (11.7) equivalent power is +5.00 D, the primary princi-
f." pal plane is 8 cm behind the first surface, and the
From the neutralizing power definition secondary principal plane is 3 cm in front of the
back surface. For a real object in air 21 cm from
P =-'Js. (11.8) the front surface, where is the conjugate image
ff
relative to the back surface? What is the lateral
A trip from the back of the system to F 2 , and then to magnification? Where are the nodal points for this
H 2 results in system?
L b H 2 = f b -f 2 . (11.10) 4. An optical system in air has a 20-cm central
thickness. The neutralizing power is -10.36 D,
the equivalent power is +6.05 D, and the back
A trip from the front of the system to F 1? and then to
vertex power is +7.06 D. Find the primary and
H, results in
secondary focal points and the primary and sec-
L.H.-f.-f,. (11.11) ondary principal planes.
5. An eye has a 21-mm cornea-to-retina distance and
The nodal points are displaced away from the princi-
pal planes by a +59.00 D equivalent power. The primary princi-
pal plane is 1 mm behind the cornea, and the
H N = ^ ^ . (11.32) secondary principal plane is 2 mm behind the cor-
nea. The vitreous has a 1.336 index. What dis-
All object space distances are reduced with n 0 , and all tance vision spectacle correction is needed at a
image space distances are reduced with ^ 10 mm vertex distance?
6. An emmetropic eye has a vitreous of index 1.336,
a +62.00 D equivalent power, a primary principal
plane located 1.3 mm behind the cornea, and a
For Thick Lenses (or Two Thin Lenses)
secondary principal plane located 1.9 mm behind
the cornea. What is the cornea-to-retina length?
^,+,-,,, (11.41) Where are the nodal points? What is the retinal
image size (in microns) for a distance object that
,^, (11.40) subtends an angle of 5'?
7. A 12-cm thick glass (n = 1.63) rod has air in front
P = (11.42) and water (n = 1.33) behind. The front air-glass
1-dP, interface has a power of +5.00 D. The back glass-
The Gauss System 227

water interface has a power of +3.00 D. Find the power, the neutralizing power, and the back ver-
neutralizing, back vertex, and equivalent dioptric tex power? Where are the principal planes?
powers. Find the principal planes and the nodal 10. An optical system in air has a 14-cm central
points. thickness. The front focal length is - 8 cm and the
8. It is desired to use a plastic (n = 1.44) sphere for back focal length is +4 cm. A real object is lo-
clinic testing with many patients. The sphere has a cated 20 cm from the front vertex, and an inverted
radius of 10 mm. What is the equivalent power? real image 2.70 times larger than the object is
Where are the principal planes for the sphere? formed on a screen behind the system. What is the
9. An 8-mm thick ophthalmic crown (n = 1.523) equivalent power? Where must the screen be
spectacle lens has a +14.00 D front surface and a placed?
-3.00 D back surface. What is the equivalent
CHAPTER TWELVE

System Matrices

12.1 The Translation Matrix of the point from the optical axis. To define the
plus-minus signs, let the h values be positive for
Consider reduced centered refraction systems. Such distances above the axis and negative for distances
systems are equivalent to thin lenses in air. A centered below the axis, and let the angles be positive when
system is one that has a common optical axis for all they are above the horizontal (as in Figure 12.1a) and
lenses (or interfaces) in the system. For the following negative when they are below the horizontal (as in
algebraic derivations, all angles are expressed in ra- Figure 12.1b).
dians and all distances are expressed in meters. In Figure 12.1a, two planes perpendicular to the
Assume that the horizontal line in Figure 12.1a is axis are separated by a reduced distance t (expressed
the optical axis. A ray at a particular point relative to in meters). The ray in Figure 12.1a is propagating
the optical axis can be quantitatively described by two from the point A on the first plane to the point B on
parameters: the (reduced) angle q that the ray makes the second plane. The ray parameters at the first plane
with the optical axis, and the perpendicular distance h are q! and hj. At the second vertical plane, the ray

20 mm

b) hi- 50 cm (not to scale) FIGURE 12.1. Paraxial ray parameters for trans-
lation.

229
230 Geometric, Physical, and Visual Optics

parameters are q 2 and h 2 . Because of the law of Thus,


rectilinear propagation /qi\ / - 8 . 7 3 x K T 3 radians\
q 2 = q. (12)
vhi/ \+0.020 m /
From Figure 12.1a,
The translation matrix for 50 cm is
h 2 - h ! ^ttanq!. (12.2) 1 0
For paraxial rays and q{ in radians, we can make the
V0.50m 1/
small angle approximation:
The ray parameters 50 cm downstream are
hz-h^tqj, q2\ /l 0\/-8.73xl0~3 \
or 2
h2/ \0.50m l/\+2.00x10 m
h 2 = h 1 H-tq 1 . (12.3)
-8.73 x 10"
We can write Eqs. 12.1 and 12.3 as one matrix
equation. 2/ V + 1.56X10" m
At 50 cm downstream, the ray is 15.6 mm above
q2\ /i o\/qi the axis.
(12.4)
hj \t l/U
The top matrix multiplication gives Eq. 12.1, and the
bottom matrix multiplication gives Eq. (12.3). 12.2 The Refraction Matrix
The 2 x 2 matrix in Eq. 12.4 is called the transla-
tion matrix T, i.e. Reduced refraction systems consist of a series of thin
lenses. Figure 12.2 shows a ray being refracted at a
1 0 distance h from the optical axis of a thin lens of
T = (12.5) dioptric power P. The incident ray parameters are q^
t 1 and h l 5 while the outgoing ray parameters are q 2 and
where t is in meters. Then we can write Eq. 12.4 as h 2 . The ray is not displaced at the lens, and thus,
/q2\ /qo h2 = h 1 = h . (12.6)
= When refracted, the ray is bent down through the
h2/ \hx
angle d. The incident ray left the axial object point at
When t is zero in Eq. 12.5, then T equals the unit the angle q ^ The outgoing ray passes through the
matrix 1: conjugate image point at the angle b. The optical axis,
1 0 incident ray, and exiting ray make a triangle with
T= l = internal angles qx, b , and q. The sum of the internal
0 1 angles of a triangle equals 180 or radians. So
qx + b + q = 77.
EXAMPLE 12.1
A point source in air is located 20 mm vertically From the incident ray and its extension together with
above an optical axis. A ray traveling slightly the outgoing ray,
downward leaves the point source and makes an q + |d| = 7.
angle of 30' with the axis. Determine the ray
parameters at the point source, and at a distance It follows that
that is 50 cm along the axis from the point source. |d| = q i +b,
The ray is traveling downward so the angle is
negative. which is a well-known plane geometry theorem stating
that an external angle of a triangle is equal to the sum
30' = 0.5 = 8.73 x 10~3 radians, of the two opposite internal angles.
so Because the bend is down, let d be a negative an-
ql = - 8 . 7 3 x 10" 3 radians. gle. Then
The point source is 20 mm or 0.020 m above the d=-qi-b.
axis, so For paraxial rays the angle in radians equals the
h, =+0.020 m. tangent. From Eq. 12.6 and the (small angle) tangents
System Matrices 231

FIGURE 12.2. Paraxial ray parameters for refraction.

of qx and b in Figure 12.2 7 When P is zero in Eq. 12.10, then R equals the unit
matrix 1:
h h
d=- - > 1 0
V
R= l =
h h 0 1
d= + - - - .
u v
Then EXAMPLE 12.2
A ray leaves an axial point source 25 cm from a
d=+hU-hV, +6D lens and hits the lens 12 mm below the axis.
or from the thin lens vergence equation What are the ray parameters when it leaves the
lens?
d = -hl3 (12.7) The angle that the ray makes with the axis is
Dimensionally, h is in meters, P is in diopters, and d is -1.2cm
ql ~ tanq! = = -4.8 x 10 radians.
in radians. Equation 12.7 is a disguised version of a 25 cm
clinicially used equation called Prentice's rule. (Pren- From Eq. 12.9,
tice's rule is covered in Chapter 18). Note that from
/q 2 \ / l - 6 D W - 4 . 8 X 1 0 - 2
Figure 12.2, taking into account the signs of the
angles. 1 A-1.2xKT 2 m
d = q2-qi.
^2\ /+2.40xl0" 2 radians\
It follows from the above two equations that .-1.2xl0~2m /
q2 = q i - h P . (12.8) The lens bends the ray up.
We can write Eqs. 12.7 and 12.8 as one matrix
equation:
/q 2 \ /i -p\/<h\ 12.3 The System Matrix
(12 9)
MU l
vh2/ \o 1 /\hi/ Consider a system consisting of two lenses separated
The top matrix multiplication gives Eq. 12.8, while the by a reduced distance t. For a ray incident on the
bottom matrix multiplication gives Eq. 12.6. The 2 x 2 system with parameters q1 and h1? the ray parameters
matrix in Eq. 12.9 is called the refraction matrix R, when it leaves the first lens are
i.e.,
q2\ /qi
R=l j. (12.10)

Then we can write Eq. 12.9 as where R, is the refraction matrix for the first lens.
/qx\ The ray then travels across the distance between
R the lenses, and is incident on the second lens with
h, parameters q3 and h3 where
232 Geometrie, Physical, and Visual Optics

l<\2
Then from the matrix multiplications, s n and s 22 are
dimensionless, while s 12 is in diopters and s 21 is in
,=T' meters.
h3/ \h2 If the system consists of one lens of dioptric power
and Tj is the translation matrix for the distance t. P, then S is equal to the refraction matrix R of that
Next the ray is refracted by the second lens, and lens, and s 12 is equal to - P . If the system consists of
the outgoing ray has parameters q 4 and h 4 where an empty tube of length t, then S is equal to the
q4\ /q3 translation matrix T for the tube length and s21 is
= R equal to t.
h4
EXAMPLE 12.4
We can combine the above three matrix equations to
An axial object point is 40 cm from the front of the
obtain two-lens system in the last example. A ray leaves
the object point and hits the +4.00 D lens 14 mm
R2T,R, (12.11) above the axis. What are the ray parameters as it
leaves the back lens?
Here h, is +0.014 m, and
The product of the refraction and translation matrices
is a 2 x 2 matrix called the system matrix S, i.e., + 1.4 cm
q,
M1 tanq,
M1 = . = 3.5 x 10 radians.
40 cm
S = R2T1R1. (12.12)
From Eq. 12.13,
Then
/qb\ /1.3 - 2 . 2 D \ / + 3 . 5 x l O - 2 radians\
q
'UP (12.13) h
b/ VO.lm 0.6 + 1.4xl0"2m
;
h4/ W + 1.47 x 10 2
radians
The elements of the system matrix S are conventional-
ly labeled: + 1.19xl0_2m
Suppose the reduced system consists of three sepa-
S=l (12.14) rated lenses. Following exactly the same procedure
used to derive Eq. 12.12, we can show that the system
matrix is given by
The system matrix is a single 2 x 2 matrix that relates
S = R3T2R2T,R1: (12.15)
the parameters of the ray incident on the front of the
system to the parameters of the ray leaving the back of where R,, R 2 , and R 3 are the respective refraction
the system. Insofar as it relates the incident (or object matrices for the first, second, and third lens, T, is the
space) ray to the exiting (or image space) ray, the translation matrix for the reduced distance between
system matrix performs the same task as the Gauss the first and second lens, and T 2 is the translation
system. matrix for the reduced distance between the second
and third lens.
EXAMPLE 12.3 For k lenses, the system matrix is
A system consists of a +4.00 D thin lens 10 cm in
front of a -3.00 D thin lens. What is the system S - R k T k _,R k T^TJRJ. (12.16)
matrix S? Since matrix multiplication is not commutative, the
The distance must be expressed in meters. Then order of the product is very important. The refraction
from Eqs. 12.5, 12.10, and 12.13, matrix for the first lens always appears on the right
+3D
,.( )(' y -4D> because, as shown in Eq. 12.11, it is the first matrix to
operate on the incident ray parameters.
V0 1 /VO.lm l/\0 1 Since in general, the system matrix performs the
n +3DW, -4m same task as the Gauss system, and since s 12 has the
dimension of diopters, you might have made an edu-
V0 1 /\0.1m 0.6 cated guess that s 12 is the equivalent dioptric power
/1.3 -2.2D^ P e of the system. If so, you are correct. The proof
S = follows.
VO.lm 0.6 Figure 12.3 is a diagram of the reduced system
In the refraction matrix, P is expressed in diopters. showing a ray incident parallel to the axis. The ray can
In the translation matrix, t is expressed in meters. be extended in until it intersects the secondary princi-
System Matrices 233

^ -

H, FIGURE 12.3. Ray incident parallel to the axis.

pal plane H 2 . The distance from H 2 to F 2 is the Pv is reciprocally related to fb, so


reduced secondary equivalent focal length f2. The qb = - h b P v ,
exiting ray or its extension leaves the intersection
point on H 2 pointing toward F 2 of the system. For the or
incident ray, the angle ql with the axis is zero. For the P =_3b
v
exiting ray, the angle with the axis is q b , and the small hb"
angle tangent approximation results in Since qA is zero, the bottom matrix multiplication in
h l
Eq. 12.19 gives
n -
h b = s22h1.
where the minus sign is introduced because qb is The above two equations together with Eqs. 12.20 and
negative. Pe is reciprocally related to f2, so 12.21 give
q^-PA- (12.17) v (12.22)

b
22
Now the system matrix S relates the parameters qt and Similarly,
h{ of the ray incident on the front of the system to the
parameters qb and h b of the ray incident on the back (12.23)
Pn =
of the system:
EXAMPLE 12.5
=S (12.18) Find Pe, Pv, and Pn for the system in Example 12.3.
From Example 12.3,
or /1.3 -2.2D\
S =
(12.19)
\0.1m 0.6
Then from Eq. 12.21,
Since qA is zero, the top matrix multiplication gives Pe = +2.2D.
qb = s12hle (12.20) From Eq. 12.22,
A comparison of Eqs. 12.17 and 12.20 shows that P = +2.2 D = +3.67D.
0.6
Sl2 = " P e (12.21)
From Eq. 12.23,

In Figure 12.3, F 2 is a distance fb from the back of the
system. The exiting ray parameters at the back of the
system are qb and h b . Then from the tangent of q b ,
=4 2 = + 6 '
and the small angle tangent approximation, EXAMPLE 12.6
A system consists of a +3.00 D lens 5 cm in front of
hh a -4.00 D lens, which in turn is 7 cm in front of a
qb = +6.00D lens. Find Pe, Pv, and Pn.
fh
234 Geometrie, Physical, and Visual Optics

The system matrix is From Eq. 12.10, the determinant of the refraction
matrix R is:
detR = ( l ) ( l ) - ( - P ) ( 0 ) = + l .
\0 1 /\0.07m l / \ 0 1
1 0\/l -3D From Eq. 12.5, the determinant of the translation
matrix T is:
0.05 m l / \ 0 1 detT = ( l ) ( l ) - ( 0 ) ( t ) = + l .
1 -6DV1 +4D
There is a general matrix algebra theorem that
0 1 /\0.07m 1.28 states: given square matrices X, Y, and Z where,
1 -3D Z = XY,
then
0.05 m 0.85
detZ = detXdetY.
0.40 -4.87 D
S= Let us apply this theorem to the building block matrix
0.13 m 0.88 B. Then
Then
P =+4.87D. B = TR,
From Eq. 12.22, and
+4.87 D CCA^ detB = d e t T d e t R = ( + l)( + l) = + l .
P = = +5.54D.
0.88 We can check this result by calculating the determin-
From Eq. 12.23, ant of B directly from Eqs. 12.24 and 12.26:
detB = ( l ) ( l - t P ) - ( - P ) ( t ) ,
d e t B = l - t P + tP = +1.
The first multiplications performed in a system The system matrix is a product of refraction and
matrix product are typically the TR products. Let translation matrices. By the above theorem, the de-
B = TR. terminant of the system matrix must equal + 1 ,
Then det S = s n s 2 2 - s12s21 = +1. (12.27)
1 0\/l -P
B Thus, the four elements of the system matrix are not
vt l/\0 1, independent of each other.
We can use Eq. 12.27 as a check on our arithmetic
B=( . (12.24) in doing the matrix multiplications. Alternatively,
\t 1-tP/ given the dioptric power parameter P e , P v , and P n , we
If desired, B can be used as the building block matrix can use Eq. 12.27 together with Eqs. 12.21, 12.22, and
for a system. Note that when t is zero, B is equal to R. 12.23 to calculate the elements of S.
Conversely, when P is zero, B is equal to T. For a
system of k lenses, EXAMPLE 12.7
Does the system matrix of Example 12.3 satisfy the
S = B k B k _ 1 ---B 2 B 1 , (12.25) determinant condition?
The determinant of the matrix in Example 12.3
where t is set equal to zero for B k . is
detS = (1.3)(0.6)-(-2.2)(0.1)
det S = 0.78 + 0.22 =+1.00.
12.4 Determinant Condition So the answer is yes.
The determinant of a 2 x 2 matrix is the product of the EXAMPLE 12.8
diagonal elements minus the product of the off-diag- Does the system matrix of Example 12.6 satisfy the
onal elements. For the system matrix S, determinant condition?
To four digits the system matrix in Example 12.6
is
s= / 0.3960 -4.8680 \
Q e t iJ ^11 22 1221* (12.26)
\ 0.1340 0.8780/
System Matrices 235

and St = S 2 S,, (12.28)


det S = (0.3960)(0.8780) - (-4.8680)(0.1340), where Sj is the system matrix for the first subsystem
det S = 0.3477 + 0.6523 = +1.000, and S2 is the system matrix for the second subsystem
which does satisfy the determinant condition. (Figure 12.4).
However, when the system matrix is rounded to Again the order of the multiplication is very im-
two digits, as in Example 12.6, then the arithmetic portant, and Sj is the system matrix for the component
gives system that the light is incident on first.
det S = (0.40)(0.88) - (-4.87)(0.13), Consider a system with a system matrix S a . Sup-
det S = 0.35+ 0.63 = 0.98. pose we place a thin lens against the back of this
system. What is the system matrix St for the combina-
The difference between 0.98 and +1 is due to the tion? The system matrix for the thin lens is just the
error introduced in the numerical round-off.
refraction matrix R. The light is first incident on the
system and then on the thin lens. So Eq. 12.28 gives
EXAMPLE 12.9
A system has an equivalent dioptric power of St = RS a . (12.29)
-16.48 D, a back vertex power of -10.38 D, and a
neutralizing power of -10.83 D. What is S? Conversely, if we place the lens against the front of
From Eq. 12.21, the system, the total system matrix would be
s12 =+16.48 D. St = SaR. (12.30)
From Eq. 12.22,
= P , = -16.48D EXAMPLE 12.10a
S22
Pv -10.38D lD Consider a system with an equivalent dioptric
power of +8.80 D and a system matrix of
From Eq. 12.23,
/1.40 -8.80D^
= P e = -16.48 D S.=
S
" Pn -10.83 D \0.05m 0.40
From Eq. 12.27, When a +5.00 D lens is placed against the back of
the system, what is the equivalent dioptric power
s21 = [(1.52)(1.59) - 1.00]/( + 16.48 D) = 0.086 m. and the system matrix of the combination?
So From Eq. 12.29,
/1.52 16.48 D\ / l -5DW1.40 -8.80D\

\0.086m 1.59 \0 1 /\0.05m 0.40 /'


/1.15 -10.8D^
S.=
\0.05 m 0.40
12.5 Subsystem Matrices The equivalent power of the combination is
+ 10.8 D.
In analyzing a system, we can break it up into two or
more subsystems. Similarly, we can combine two or EXAMPLE 12.10b
more systems together to make a larger system. In When a -7.00 D lens is placed against the front of
either event the total system matrix is equal to the the original system used in Example 12.10a, what is
product of the component system matrices. For ex- the equivalent dioptric power and system matrix of
ample, the combination?

(+
1
incident light exiting light
s, s2 ^

FIGURE 12.4. Subsystems 1 and 2.


236 Geometrie, Physical, and Visual Optics

From Example 12.10a, refraction matrix for a +7.00 D lens, or we can


/1.40 -8.80 D write down the system matrix for each lens and do

M\0.05m 0.40
the matrix multiplications.
S = R 2 R
From Eq. 12.30, 1 -4D -3D
/1.40 -8.80DW1 +7D S=
St = SaR = 1 1
\0.05m 0.40 / \ 0 1 -7D
/1.40 +1.00 D
S,= 1
\0.05m 0.75 Consider the system matrix for either a thick lens
The equivalent power of the combination is or two separated thin lenses. The lens powers are Pj
-1.00 D. and P 2 , and the separation distance in the reduced
system is t (Figure 12.5). From Eq. 12.12,
S = RJTJRJ,
12.6 Two Lens Systems 0\/l -Pi
s=
Consider two thin lenses of power and P2 in contact vo 1
with each other. We know from the previous chapters
that the dioptric powers are additive for thin lenses in
contact with each other, so that the combination just s=
\0 1-tP,
acts like one lens with a power of plus P2. The
system matrix for the combination should also manif- l-tP2 -(,+,-,,
est the additivity relationship. s=
For the two thin lenses in contact with each other, \ t 1-tP, /
S = R2R1? From S and Eq. 12.21,
1 1 "Pi Pe = P1 + P 2 - t P 1 P 2 ,
S= which is identical in form to Eq. 11.41 for the equiva-
V0 1 /\0 1 lent dioptric power of a thick lens.
The matrix multiplication gives: From S and Eq. 12.22,
/i - ( P 1 + P2V P
P =
1-tR
\0 1
which is identical in form to Eq. 11.40 for the back
Indeed, S has the form of a refraction matrix with the vertex power of a thick lens.
power given by the sum of Pj and P2. From S and Eq. 12.23,
EXAMPLE 12.11
A +3.00 D thin lens is held in contact with a 1-tP,
+4.00 D thin lens. What is S for the two lenses?
We can either add the dioptric powers together which is identical in form to Eq. 11.42 for the neut-
to get +7.00 D, and then the system matrix is just a ralizing power of a thick lens.

P2
(+)
4

incident light exiting light

FIGURE 12.5. Two-lens system.


System Matrices 237

EXAMPLE 12.12 system and is the translation matrix for the distance
A +3.00 D lens is located 10 cm in front of a from the back vertex to the image plane.
+4.00 D lens. What is the system matrix, Pe, Pv, The object-image matrix is still a 2 x 2 matrix; it is
and Pn for the two lenses? similar to Eq. 12.19 in that it relates the ray parame-
S = R2T1R1, ters (q b , h b ) at the image plane to the ray parameters
-4D 0\/l ( q ^ h j at the object plane:
1 -3D
S=' /q b \ /q^
0 1 0.1m l/\0 1
1
= s,
1 --4D -3D hi
s= 0 1 0.1m 0.7
or
0.6 5.8D^
s= 0.1m 0.7
We can use the determinant condition, Eq. 12.27, The bottom matrix multiplication gives
to check the arithmetic:
h b ^ s ^ q ! +s 2 2 h!.
det S = (0.6)(0.7) - (-5.8 D)(0.1 m),
detS = 0.42 + 0.58=+l. For a ray leaving the axial object point, h1 is zero. The
Then from Eqs. 12.21 through 12.23, corresponding image space ray must go to the axial
image point (h b = 0) independent of the object space
Pe = +5.8D,
angle q^ This means that s21 in the above equation
must equal zero, and then
P = ^ = +8.29D,
0.7 hb = s22h1.
+ 5.8D
P = 0.6 = +9.68D. For an off-axis object point, hl is the size of the object
and hb is the size of the image. The element s22 is just
the lateral magnification m. Hence,

12.7 The Object-Image Matrix


Sio ~
0 m
The box in Figure 12.6 represents the reduced system
for a series of lenses with a front to back vertex system The determinant of any system matrix has to equal +1
matrix S b . The system has a real object in front of it (Eq. 12.26), so
and a real image behind it. Conceptually, we can detS I O = s n m - s 1 2 0 = + 1 ,
expand the system and consider it starting at the or
object plane and ending at the image plane. The + 1 =snm,
system matrix for the expanded system is called an
object-image matrix S I O . S IO is related to S b by j_
S n
" m*
S,o = T A T . , (12.31) The element s12 is the equivalent dioptric power of the
where Tw is the translation matrix for the distance expanded system. However, multiplication by transla-
from the object plane to the front vertex of the lens tion matrices does not change the equivalent power.

i 1 i
* j
I I

FIGURE 12.6. Expanded systemobject to image.


238 Geometric, Physical, and Visual Optics

Therefore, s12 is equal to the equivalent power Pe of ponents. (Remember the first system goes on the
the lens system. The object-image matrix can then be right.) Therefore,
written as =
(Sio)t (^10)2(^10)1-

e From the individual system matrices and the matrix


m (12.32) multiplication, we obtain
S=
\0 m / 1
EXAMPLE 12.13 (S IO ) t =
Consider the two lens system used in the last
example, namely, a +3.00 D 10 cm in front of a \ 0
m1m2/
+4.00 D lens. The equivalent dioptric power of the The second row/second column element is the total
two lenses is +5.80 D. A real object is one third of lateral magnification, and the above matrix reconfirms
a meter from the first lens, and a conjugate real that the total lateral magnification is the product of the
image is 25 cm behind the second lens. Find the
object-image matrix and the lateral magnification. individual lateral magnifications.
From Example 12.12, the front to back vertex
system matrix is
0.6 -5.8D
Sb = 12.8 The Front Vertex to Secondary Focal
VO.lm 0.7 Plane Matrix
In the translation matrix T a , the distance is from
the object to the lens, which is +1/3 m. This is When the object or the image is at optical infinity, the
opposite in sign to the object distance u for the first object-image matrix loses its usefulness. For an object
lens. Then from Eq. 12.31, at optical infinity, the image is in the secondary focal
I 0W0.6 -5.8D\ plane. It then becomes useful to consider a system
matrix for the components that start at the front
Sio
0.25 m l/\0.1m 0.7 vertex of the lens system and go to the secondary focal
plane (Figure 12.7).
0
Consider a lens system with a front to back vertex
0.3333 m 1 system matrix S b . Then the front vertex to F 2 system
matrix S1F is given by
1.33 -5.80 D
$ i o ~~ $IF = T F S b ,
0 -0.75
Then from Eq 12.32, the lateral magnification m where T F is the translation matrix for the back focal
is -0.75. Note that the equivalent dioptric power length. S1F is a 2 x 2 matrix that relates the ray
in the object-image matrix, +5.80 D, is identical parameters ( q ^ h , ) at the front vertex to the ray
to the equivalent dioptric power of the two lens parameters (q b ,h b ) in the secondary focal plane:
system.
When the image for one system becomes the object = Sn
for a second system, the total object-image matrix is hi
the product of the object-image matrices of the com- Since multiplication by a translation matrix does not


1
1
sb
1
F2

^ > >
^ *
FIGURE 12.7 Expanded systemfront vertex to sec-
1 ondary focal plane.
System Matrices 239

change the equivalent dioptrie power, s12 remains Then the determinant condition gives
equal to - P e of the lens system. detS I F = s n 0 - ( - P e ) ( s 2 I ) = + l ,
In terms of the elements of S 1F , the bottom matrix
multiplication gives or
a L
h = s 21 p 2'
b 2iqi +s 2 2 h 1 .
The object at optical infinity subtends the angle ql at EXAMPLE 12.14
the front vertex of the system, and all of the incident Figure 12.8b shows a model of Gullstrand's #2
rays from the top object point are parallel to each (simplified) schematic eye (unaccommodated). The
other. Each of these rays must go to the conjugate cornea and aqueous are both considered to have a
four thirds (i.e., 4/3) refractive index. The corneal
image point, which is at h b , independent of point h, at power is +42.735 D. The anterior chamber contain-
which the ray hits the front vertex of the system. From ing the aqueous is 3.6 mm thick. The aqueous-
the above equation, this means that s22 must be zero, crystalline lens interface has a power of +8.267 D.
and The crystalline lens is homogenous with a refractive
index of 1.416 and a central thickness of 3.6 mm.
hb = s21q1, The crystalline lens-vitreous interface has a power
-P. of +13.778 D, and the distance from this interface
to the retina is 16.7 mm. The vitreous refractive
s1F index is 4/3. Find the cornea to retina system
y s 21 0
matrix, and determine the eye's equivalent dioptric
where Pe is the equivalent power of the lens system. power.

FIGURE 12.8. Gullstrand's schematic eyes.


240 Geometrie, Physical, and Visual Optics

The respective reduced distances for the an- The system matrix is given by:
terior chamber, crystalline lens, and vitreous cham- S = T3R3T2R2TJRJ,
ber are:
3.6 mm The only change is in T 3 since now
t, = r^^r- = 2.700 mm, 14 mm
1.333 = 10.500 mm.
t, = -
_ 3.6 mm 1.333
2 2.542 mm, The matrix multiplications give
~ 1.416
16.700 mm / 0.9062 -60.48 D \
t, = - = 12.525 mm. S=
1.333
\ 0.01470 m 0.1223 y
The system matrix is given by:
The equivalent dioptric power is still +60.48 D
S = T 3 R 3 T 2 R 2 T,R 1 , because the eye's refracting system (cornea and
where crystalline lens) is unchanged. The change in the
1 -42.735 D vitreous chamber depth results in an ametropic eye,
which is apparent from the non-zero s22 element.
0 1
Suppose we want a thin contact lens of power P c
1 0
placed on the cornea such that a distant object is
T, =
0.002700 m 1 imaged clearly on the retina of the unaccommodated
eye. Assume the contact lens is fit on K. Then the
1 -8.267 D
total system matrix S t for the contact lens and eye
R,=
0 1 should have a zero s22 element. The total system
matrix is given by
1 0
T2 = S t = SR,
0.002542 m 1 where S is the cornea to retina system matrix and R is
1 -13.778 D the refraction matrix for the thin contact lens. Let the
1*3 = elements of S t be primed and those of S unprimed.
0 1 Then
1 0
T.= Kx s;2\ / s n s 1 2 \/i - p c
V0.012525 m \)
The matrix multiplications give the cornea to retina Wi s y \s 2 1 s 2 2 /\o l
system matrix. The matrix multiplication for s^ gives
/0.9062 -60.48 D \ i & = 0 = ( s 2 I ) ( - P c ) + (s 2 2 )(l),
S=
0.01653 m 0 or
The equivalent dioptric power of the eye model is P = -22 (12.33)
+60.48 D. As shown by the fact that s22 is zero, the
secondary focal plane is at the retina, and the In Eq. 12.33, s21 and s22 are elements of the cornea to
Gullstrand # 2 schematic eye is emmetropic.
retina system matrix, and P c is the dioptric power of a
distance vision thin contact lens correction.

EXAMPLE 12.16
12.9 Ametropic Eyes Given the eye model and system matrix of Example
12.15, what contact lens correction is needed?
When the eye is ametropic, then the element s 22 in the From the system matrix in the preceding ex-
cornea to retina system matrix is not zero. ample,
0.1223 0 ^
EXAMPLE 12.15 P =
0.01470 m = +8.32D.
Consider an eye model with the same parameters
used in the preceding example except that the We can check S t by computing the matrix product
vitreous chamber (the distance from the back of for the contact lens to retina system. The result is
the crystalline lens to the retina) is 14 mm. Find the 0.9062 -68.02 D
cornea to retina system matrix and the equivalent S.=
power of the eye. 0.01470 m
System Matrices 241

The fact that s22 is zero verifies the dioptric power 12.10 The Gullstrand #1 Schematic Eye
of the correction.
Example 12.14 referred to the Gullstrand #2 sche-
EXAMPLE 12.17 matic eye. Allvar Gullstrand (1862-1930), a Swedish
An eye model has a +47.50 D cornea, a 3.6 mm professor of ophthalmology, was one of the world's
anterior chamber depth, a +8.00 D aqueous-crys- leading figures in the study of the eye as an optical
talline lens interface, a 3.6 mm thick crystalline instrument. Gullstrand measured and compiled data
lens, a +14.00 D crystalline lens-vitreous interface, on the parameters of the living human eye. Based on
and a 17 mm vitreous chamber depth. The refrac- this data, Gullstrand developed some schematic repre-
tive index of the vitreous and the aqueous is 1.336, sentations of the typical or average living human eye.
and the refractive index of the crystalline lens is The models differ in complexity and are discussed
1.416. Find the cornea to retina system matrix and below.
specify what contact correction the eye needs.
The respective reduced distances for the an- In the course of Gullstrand's work on the eye, he
terior chamber, crystalline lens, and vitreous cham- discovered a number of widespread misconceptions
ber are: about image formation in general. Gullstrand worked
3.6 mm hard to rectify these misconceptions, and his work
included some very sophisticated mathematical anal-
*' 1.336 = 2.695 mm,
yses of image formation. As a result, he became highly
3.6 mm regarded in the physics community, as well as in the
= 2.542 mm,
* 2 ~ 1.416 ophthalmology community. (Incidentally, Gullstrand
17 mm was an advocate of the vergence approach to visual
= 12.72 mm.
* 3 ~ 1.336 optics.)
In both 1910 and 1911, Gullstrand was nominated
The systemi matrix is given by for the Nobel prize in physics. In 1911, the original
S = T3R3T2R2T,R suggestion of the Physics Nobel prize committee was
where that Gullstrand be given the prize for his work on
geometric optics. The Physiology and Medicine Nobel
/ I -47.50 D\ Prize committee had the same idea and independently
R.= ,
recommended Gullstrand for the 1911 Nobel prize in
\o 1 /
medicine for his work on the dioptrics of the eye.
/i \ Gullstrand declined the Physics Prize so that he could
,= \0.002695m 1/
accept the Physiology and Medicine Nobel Prize.
Gullstrand's eye models are simplified in the sense
/I -8.00 D \ that spherical surfaces are used and certain assump-
tions are made about the indices of refraction. Gull-
\0 : ) strand's # 1 schematic eye is shown in Figure 12.8a,
/I o\ and the corresponding data is in Table 12.1. In the #1
T2 = schematic eye, the cornea is given a uniform refractive
\ 0.002542 m 1/ index (1.376), which is different from that of the
aqueous humor (1.336). The crystalline lens does not
/ I -14.00 D\
have a uniform refractive index, and this is modeled in
\0 , ) the #1 eye by assigning a core with a 1.406 index,
while the outer part of the lens is assigned a 1.386
/I o\ index. The aqueous and vitreous humors are each
T3 = assigned a 1.336 index.
\ 0.01272 m 1/ The cornea to retina system matrix for Gullstrand's
The matrix multiplications give the cornea to retina #1 schematic eye is
system matrix. /0.921 -58.64 D^
/0.9059 -64.75 D S =
\0.0168m 0.0169 y
S =
\ 0.01671 m -0.09035/ From the system matrix and Eq. 12.33,
From Eq. 12.33, 0.0169
P,= 0.0168 m = +1.01D.
-0.0935 KJ .\J XVJ<J III

P,= = -5.41D. So Gullstrand's # 1 schematic eye is a 1 D hyperope.


0.01671 m
242 Geometrie, Physical, and Visual Optics

T A B L E 12.1
Gullstrand's # 1 Schematic Eye

Unaccommodated Maximum Accommodation

Refractive index
Cornea 1.376 1.376
Aqueous and vitreous 1.336 1.336
Lens 1.386 1.386
Equivalent core lens 1.406 1.406
Position (mm)
Anterior corneal surface 0 0
Posterior corneal surface +0.5 +0.5
Anterior lens surface +3.6 +3.2
Equivalent core lens anterior surface +4.146 +3.8725
Equivalent core lens posterior surface +6.565 +6.5275
Posterior lens surface +7.2 +7.2
Radius of curvature (mm)
Anterior corneal surface +7.7 +7.7
Posterior corneal surface +6.8 +6.8
Anterior lens surface + 10.0 +5.33
Equivalent core lens anterior surface + 7.911 +2.655
Equivalent core lens posterior surface -5.76 -2.655
Posterior lens surface -6.0 -5.33
Refracting power (D)
Anterior corneal surface +48.83 +48.83
Posterior corneal surface -5.88 -5.88
Anterior lens surface +5.0 +9.375
Anterior core surface +2.53 +7.53
Posterior core surface +3.47 +7.53
Posterior lens surface +8.33 +9.375
Corneal system
Equivalent power (D) +43.05 +43.05
Position primary principal point -0.0496 -0.049
Position secondary principal point -0.0506 -0.0506
Lens system
Equivalent power (D) + 19.11 +33.06
Position primary principal point +5.678 +5.145
Position secondary principal point +5.808 +5.255
Complete optical system of eye
Equivalent power (D) +58.64 +70.57
Position primary principal point + 1.348 + 1.772
Position secondary principal point + 1.602 +2.086
Position primary focal point -15.707 -12.397
Position secondary focal point +24.387 +21.016
Primary focal length -17.055 -14.169
Secondary focal length + 22.785 + 18.930
Position primary nodal point + 7.08 + 6.53
Position secondary nodal point +7.33 + 6.85
Position fovea centralis + 23.9 +23.9
Position near point (mm) -102.3

Figure 12.8c shows the so-called reduced eye. This S = TR,


model is misnamed in the sense that a reduced system
is an equivalent air system, and yet the reduced eye is 1.000 -60.00 D
really a single spherical refracting surface model of the
eye. The reduced eye has an index of four thirds, an
s= 0.0166 m 0
axial length of 22.22 mm, and a 60.00 D refracting
surface. The location of the refracting surface roughly From the zero two-two element, the reduced eye is
corresponds to the location of the principal planes in emmetropic. The corresponding equivalent air system
the #1 and #2 schematic eyes. The system matrix for is the +60.00 D thin lens in air model that we have
the reduced eye is often used in previous chapters.
System Matrices 243

12.11 Vertex Vergence Relations However, one can use Eq. 12.35 without knowing
where the primary and secondary principal planes are.
In Figure 12.6, is the reduced distance from the A comparison of Eqs. 12.32 and 12.34 provides
front of the system to the object and v is the reduced equations for the lateral magnification
distance from the back of the system to the image.
The front-to-back vertex system matrix was S b , and m + S22> (12.36)
from Eq. 12.31, the object-image matrix S IO is
vb
or
1 0\/sn sx 1 0 1 S
12
= S11 (12.37)
m
Sio ~~
w 1/ \s 2 1 s22/\-u 1 In terms of the equivalent dioptric power, Eqs. 12.36
and 12.37 are, respectively,
The matrix multiplications give
1 u\ / Su s12w Sj2\ m = - 77" + S9 (12.38)

V 1 / \^21 &22M S2 or
1 Pe (12.39)
Sn
m IV
(12.34)
EXAMPLE 12.18
where The system in Example 12.6 consisted of a +3.00 D
lens 5 cm in front of a -4.00 D lens, which in turn
21 11 12 21 22 * was 7 cm in front of a +6.00 D lens. Use Eq. 12.35
to find the system's image conjugate to a real
but from Eq. 12.32, s21 equals zero. So object 50 cm in front of the first lens. Use Eq. 12.39
to find the lateral magnification.
0 = s n i7 - sl2v + s21 s22w, From Example 12.6, the front-to-back vertex
system matrix is
or /0.40 -4.87 D\
s22w s21 = (s n - sl2u)v.
\0.13m 0.88 /
We can divide both sides of the above equation by u
and by v, and then substitute The vergence incident on the first lens is
U ^ - 2 . 0 0 0.
Ut1 = 4 and Vb b =4, From Eq. 12.35,
u v
where \] is the vergence of the light incident on the 4.87D + (0.40)(-2.00D)
VK =
front vertex and Vb is the vergence of the light leaving 0.88-(0.13 m)(-2.00D)'
the back vertex. The result is
= - S ^ + SnUt
V
b _ s TJ '

^22 a
21^1 The light leaving the back of the system is converg-
where the s's are the elements of the front-to-back ing, the image is real, and
vertex matrix S b . From Eq. 12.21, 100
+28.0 cm.
+ 3.57 D
From Eq. 12.39, the lateral magnification is
Pe + s11U1
V = (12.35) -4.87 D
b
22 ^21^1 m = + 3.57 D + 0.88 =-0.48.
Equation 12.35 is an equation that relates the ver-
gence Vb of the light leaving the back vertex of the EXAMPLE 12.19
system to the vergence of the light incident on the Does Eq. 12.23 for the neutralizing power follow
front vertex of the system and to the matrix elements from the front-to-back vertex vergence equation
(Eq. 12.35)?
of the front-to-back vertex system matrix. Equation When plane waves leave the back of the system,
12.35 is a little more complicated than the Gauss the neutralizing power is equal to \Jl. For plane
equation, waves leaving
VH2 = Pe + U H l . v b = o,
244 Geometrie, Physical, and Visual Optics

and Eq. 12.35 gives Problems


0=
S S
1. A +4.50 D lens is located 10 cm in front of a
22 ~ 21^1
-4.50 D lens. Find the system matrix, the equival-
or ent power, the back vertex power, and the neut-
0 = Pe + s11U1. ralizing power.
Then 2. A thin lens in air and screen eye model consists of
a +58.00 D lens located 18 mm in front of the
n 1 ' retina (screen). Find the system matrix for the eye
and the (thin) contact correction.
which agrees with Eq. 12.23. 3. A -8.00 D lens is located 6 cm in front of a +3.00
lens. Find the system matrix, the equivalent
power, the back vertex power, and the neutraliz-
ing power.
12.12 Derivation of the Gauss Equation 4. A +9.00 D lens is located 5 cm in front of a
-4.00 D lens, which in turn is located 7 cm in
The Gauss equation is easily derived from the front- front of a +6.00 D lens. Find the system matrix,
to-back vertex vergence equation. When we pretend the equivalent power, the back vertex power, and
that the front vertex of the system is the primary the neutralizing power. Be sure to check the
principal plane Hx and the back vertex of the system is determinant of the system matrix.
the secondary principal plane H 2 , then because of the 5. The system matrix for a multiple lens system is
conjugacy of the principal planes, the vertex system
matrix must have the form of an object-image matrix 0.72 -6.56 D
with a lateral magnification of +1. Thus, S=
0.26 m -0.98
+1 -P.
Find the back vertex power and the neutralizing
s = power.
0 +1
With the above vertex system matrix, the vertex ver- 6. A +4.00 D lens is placed against the back of the
gence equation (Eq. 12.35) becomes system in problem no. 5. Find the new system
matrix, the new equivalent power, the new back
P, + ( + l ) U , vertex power, and the new neutralizing power.
vb = + 1-(0)U1 ' 7. A -3.00 D lens is placed against the front of the
system in problem no. 5. Find the new system
or matrix, the new equivalent power, the new back
Vb = Pe + U,. vertex power, and the new neutralizing power.
In this case, since the back is H 2 and the front is 1? 8 The equivalent power of a multiple lens system is
we can change the subscripts to obtain +2.900 D. The back vertex power is +24.166 D
and the front vertex power is -1.261 D. Find the
VH = P e + U H i system matrix.
The Gauss equation for the lateral magnification fol- 9. Find the system matrix for three thin lenses in
lows from Eq. 12.38 with s22 equal to +1. contact. According to the system matrix, are the
dioptric powers of the three lenses additive?
P. 10. A real object is 40 cm in front of a system with a
m= + 1.
system matrix of
The subscript change gives 0.400 -6.480 D
S=
m= - + 1, 0.120 m 0.556
Find the conjugate image and the lateral magnifi-
-p. + yH 7 cation.
m: V,H 7 11. Derive Eq. 12.22 for the back vertex power from
the vertex vergence equation (Eq. 12.35).
or 12. The system in problem no. 10 is placed against the
back vertex of the system in problem no. 5. What
m= VH/ is the system matrix for the combination?
System Matrices 245

13. A reflection (catoptric power) matrix is identical find the back vertex and the neutralizing power of
in form to a refraction (dioptric power) matrix. the lens-mirror system. (Your matrix product
Given this, consider the lens-mirror system in should correspond to the lens refraction, a transla-
Example 10.19. Use the matrix-product method to tion over a positive distance to the mirror, the
find the system matrix. Your equivalent power reflection by the spherical mirror, a translation
from the system matrix should equal the power of over a positive distance back to the lens, and the
the equivalent mirror. From the system matrix, second lens refraction.)
CHAPTER THIRTEEN

Angular
Magnification

13.1 Relative Size Magnification ratio of the retinal image sizes is generically referred
to as the angular magnification M. As shown below,
The foveal area of the human retina has the highest the reason for the name is that the ratio of the retinal
density of cones, and is the area that provides the best image sizes is equal to the ratio of the nodal angles
resolution of detail. The peripheral retina has a smal- subtended by the objects.
ler density of cones, and provides poorer resolution of Figure 13.1 shows a thin lens in air eye model. The
detail. If a person's foveal area is impaired, the person retina (screen) is a distance ve from the lens. Figure
will have decreased visual acuity even when the retinal 13.1a shows an object that has a nodal angle of w1?
image is clear. and results in a retinal image size of II. Figure 13.1b
Such a person can be helped by increasing the size shows a larger object at the same distance from the
of the object. This results in an increase in retinal eye. The larger object has a nodal angle of w2 and
image size, and enables the retina to obtain more results in a retinal image size of I 2 . From Figure 13.1a,
information about the object's detail. For example,
consider a person who holds a regular print book at tan w, = .
40 cm and has a retinal image size that is clear but too v
small for the detail to be resolved. The same person From Figure 13.1b,
can perhaps hold a large print book at 40 cm and have
a clear retinal image that is large enough for the detail tan w? = .
to be resolved.
In the above situation, each book is held at 40 cm, By dfinition,
each image is in focus on the retina, and the lateral
magnification m of the observer's eye remains exactly M=^ (13.1)
the same when the books are switched. However, we Ii"
are not so much interested in the lateral magnification Since ve is unchanged, it follows from the above three
as we are in the increase in the retinal image size. The equations that

t ^ ^ ^ ^ ^
o2
^^^^^
fh
wT*
b) 1 ' FIGURE 13.1. Relative size magnifi-
cation.

247
248 Geometrie, Physical, and Visual Optics

tan w9 image of the balloon is too small. As the balloon drifts


M= (13.2) closer, the observer's retinal image of the balloon
tanw
grows in size until the pattern is resolvable.
In paraxial cases, the tangent can be replaced by the Alternatively, an observer can make out the pat-
angle, so tern on the distant balloon by looking through a
telescope. The telescope optically magnifies the image
M= (13.3) of the balloon, and the pattern then becomes resolv-
able. Similarly, when a near object is too small for
Equation 13.3 is the mathematical statement that M is
resolution of the fine detail, a magnifying lens or a
equal to the ratio of the nodal angles subtended.
microscope can be used to increase one's retinal image
size.
EXAMPLE 13.1
Represent an unaccommodated emmetropic eye by Thus, without changing the physical size of the
a +60 D thin lens located 16.67 mm in front of a object, people can increase their retinal image sizes by
screen (the retina). For a set of letters located bringing the object closer or by using a magnifying
40 cm from the lens, the ocular accommodative system, such as a converging lens, a telescope, or a
demand is 2.5 D and the lateral magnification is microscope. Just as in relative size magnification, the
16.67 mm/(-40 cm), which equals -0.042. What is increase is quantified by comparing the magnified
the increase in retinal image size when viewing a retinal image size to the original (unmagnified) retinal
letter that subtends an angle of 20' as compared to image size. Again, the magnification of interest is not
viewing a letter that subtends an angle of 5'? What lateral magnification, which is a comparision of an
is the change in the lateral magnification?
From Eqs. 13.1 and 13.3, image size to its conjugate object size, but rather is a
comparison of one image size to another image size.
20' As above, angular magnification is the generic name
M=-=4X. for the ratio of one retinal image size to another.
The retinal image size increases by a factor of 4. Figure 13.2 shows the comparison for an object at
Since the object and image distances are un- two different distances. The person's eye is repre-
changed, the lateral magnification is unchanged. sented by the thin lens and screen. The object has size
O and is initially at position 1, which is a distance of ux
As shown in the above example, angular magni- from the eye. The conjugate image size, determined
fication is conventionally labeled with an X, al- by the nodal ray, is I e . When the object is moved in to
though the X is a dimensionless unit. There are position 2 (which is a distance u2 from the eye) and the
other ways of increasing an observer's retinal image person accommodates appropriately, then the retinal
size without physically changing the object size, and image is clear and the size is I m . The angular magnifi-
these are considered in the next sections. The angu- cation M is specified by
lar magnification due to an object size change is
sometimes called relative size magnification. M=^. (13.4)

The angles we and wm subtended by the images at the


nodal point of the lens are given by
13.2 Angular Magnification without a
Physical Size Change tan w UvP
and
A distant hot air balloon appears as a tiny shape in the
tan we = I e / e .
sky. Even a visually normal person cannot make out
the colorful pattern on the balloon because the retinal When accommodation occurs, the power of the lens

FIGURE 13.2. Relative distance magnification.


Angular Magnification 249

changes but the lens-screen distance ve remains the Since the object size O is unchanged in the two
same. Then from the above three equations, different situations, one can use the above two equa-
tions together with Eq. 13.4 to obtain
M = Iig = tanw ni
(13.5)
I. tan w (13.7)
m,
In the small angle approximation, which is accurate in Equation 13.7 shows that the angular magnification is
the paraxial region, the tangents can be replaced by equal to the ratio of the lateral magnifications, pro-
the angles and then vided the object size remains unchanged.
In particular, the angular magnification M can be
(13.6)
large even when the lateral magnification m2 is small.
In order for this to happen, m1 has to be much smaller
Equation 13.6 shows that the ratio of retinal image than m2. Similarly, the angular magnification M can
sizes is equal to the ratio of the nodal angles. In effect, be small even when m2 is large. This occurs whenever
one can increase (magnify) the retinal image size by m1 is much larger than m2. Consequently, when deal-
increasing (magnifying) the angle of the nodal ray ing with angular magnification, it is important to keep
incident on the eye. the initial viewing conditions in mind.
Figure 13.3 shows the situation for a person view-
ing a near object directly, and then viewing the near
object through the magnifying lens. The retinal image 13.3 Relative Distance Magnification
size and the nodal angles are labeled the same as in
Figure 13.2. Equation 13.4 through 13.6 again give the As stated in Section 13.2, the retinal image size of an
angular magnification. object can be increased by bringing the object closer
It is important to note that angular magnification is to the eye. The angular magnification that results in
always a comparison of one situation versus another. this case is sometimes referred to as relative distance
To emphasize this comparison, consider the relation magnification.
between angular magnification and lateral magnifica- In Figure 13.2, the object is originally at a distance
tion for the case in which the physical size of the ux from the eye, and then is moved in to a distance u2
object is unchanged. For the situation shown in Figure from the eye. In each case, the angle subtended by the
13.3a, the lateral magnification \ for the direct view- object at the nodal point is equal to the angle sub-
ing is defined as tended by the image at the nodal point. From the
triangles involving the object,
m, i e /o,
tan w e = 0 / ( M j ) ,
while the lateral magnification m2 for the total mag-
nifying lens-eye system (Figure 13.3b) is defined as and
m2 = I m / 0 . tanw m = 0 / ( - w 2 ) .

FIGURE 13.3. a. Unaided eye. b. Eye aided by


simple magnifier.
250 Geometrie, Physical, and Visual Optics

Then from Eq. 13.5, the angular magnification is just is virtual, erect, and larger. Thus, one can use a plus
equal to the ratio of the object distances, or lens as a simple magnifying lens as shown in Figure
13.3.
M= (13.8) In order to determine the angular magnification,
(u2)' we must first establish the initial conditions. Assume
Again, the angular magnification M is dimensionless that initially the person directly views the object at a
but is conventionally labeled with an X. For example, distance of ue in front of the eye (Figure 13.3a). The
an angular magnification of 9X indicates a ninefold angle subtended by the object at the eye's nodal point
increase in the retinal image size (or in the angle of is equal to the angle subtended by the retinal image at
the eye's nodal ray). the nodal point, and both are labeled w e . Then
Note that as the object is moved closer to the eye, tan we = 0 / ( - e ) .
the person's accommodative demand increases. The
person's near point is the closest the object can get Figure 13.3b shows the person viewing the object
and still be seen clearly. through the magnifying lens. Note that the actual
object to eye distance is different from ue. The object
EXAMPLE 13.2 is a distance of u in front of the magnifying lens, and
G.O. Student has her notes propped on her desk the lens is a distance d in front of the eye. The virtual
100 cm from her eyes, and is having trouble reading image (size I) is a distance of i; from the lens, and um
the fine print. She then moves the notes in to 20 cm from the eye. From the nodal ray for the eye,
from her eyes. What angular magnification does
G.O. achieve? (i.e., What is the increase in G.O.'s tanw m = I/(-K m ).
retinal image size?)
From Eq. 13.8, From Eq. (13.5), the angular magnification M is then
(-100 cm)
M= = +5X. M=
(-20 cm) 0(0 '
G.O.'s retinal image increases by a factor of 5 but I/O is the lateral magnification of the magnifier.
when the notes are moved in. Further, the ratio of the eye's object distances is just
the angular magnification due to the distance changes.
EXAMPLE 13.3 Therefore, the total angular magnification M is equal
G.O.'s homework solutions are 25 cm in front of to the product of the lateral magnification m,ens of the
her eyes. What angular magnification does G.O. magnifying lens and the angular magnification Mdis
achieve when she brings the homework solutions in
to 20 cm in from of her eyes? due to the distance change, or
From Eq. 13.8, M = m Iens M dls . (13.9)
(-25 cm) EXAMPLE 13.4
M= + 1.25 X.
(-20 cm) Ranger Rick, an emmtrope, is examining a small
In each of the above two examples, the object ended insect 20 cm in front of his eye. Rick then uses a
up 20 cm from the eye, but the angular magnification +20.00 D thin lens located 7 cm in front of his eye
(or increase in retinal image size) is different because as a magnifier. Relative to the initial 20 cm viewing
the starting conditions are different. distance, what angular magnification does Rick
achieve when he holds the insect 4.5 cm in front of
As a final note, it should be obvious that if the the +20.00 D lens?
object is moved away from the eye instead of closer, For the +20.00 D lens:
the retinal image size will decrease and M will be less
than one. For example, when the object at 20 cm is ul = 4.5 cm,
moved to 25 cm from the eye, the angular magnifica- 10
tion due to the distance change is 0.80 X. An M less
u,= -4.5 cm =-22.22 D.
than one indicates a decrease in the retinal image size. From the thin lens vergence Eq. 5.1,
V, = +20.00 D + (-22.22 D),
V, = -2.22 D,
and
13.4 Plus Lens as a Simple Magnifier: vi = -45.0 cm.
Intuitive Method
The lateral magnification for the lens is
When an object is placed inside the primary focal U, -22.22 D
point of a plus (converging) lens, the image of the lens
m
.ens = y = +10.
-2.22 D
Angular Magnification 251

The virtual image formed by the magnifying lens is Hence, the lateral magnification of the lens gets larger
10 times larger than the insect. Since the magnify- while the angular magnification due to the distance
ing lens is 7 cm from Ranger Rick's eye, the virtual change gets smaller.
image is 52cm from his eye (i.e., 45 + 7). The In the limit of the object going to the primary focal
object distance for Rick's eye is then point, the lateral magnification mlens goes to infinity
um = -52 cm. and the angular magnification Mdis goes to zero. From
The initial (direct) viewing distance was 20 cm, so Eq. 13.9, the total angular magnification M is then the
the angular magnification due to the distance product of infinity times zero, which is an indetermi-
change is nate form. The value can be determined by numerical-
ly taking the limit. This involves calculating the result
for a series of cases as the object is moved closer and
The total angular magnification is closer to the primary focal point.
M = mlensMdis = ( + 10)(+0.38), Consider a person viewing an object 10 cm in front
of his eye. The person then takes a +20.00 D magnify-
M=+3.8X.
ing lens and holds the lens 3 cm in front of his eye.
When Ranger Rick uses the magnifying lens, the The object is originally placed 4 cm in front of the lens
virtual image that he sees is actually 10 times larger and then moved closer and closer to the primary focal
than the insect but is also farther away than the point (which is 5 cm in front of the lens). The calcula-
object initially was. The angular magnification due tions are done as in Examples 13.4 and 13.5; Table
to the distance change is 0.38 (a decrease). As a
result, Ranger Rick increases his retinal image size 13.1 shows the results.
by only a factor of 3.8 when using the +20.00 D Note that in Table 13.1, m,ens steadily increases
magnifying lens under the specified conditions. while Mdis steadily decreases, but their product M
approaches 2.00 X. Therefore, in the limit of the
EXAMPLE 13.5 object going to the primary focal point (-5.00 cm), M
Paula Bear, an emmtrope, holds the fine print of a goes to 2.00 X.
playbill program at a typical reading distance of An equation for the value of the angular magnifica-
40 cm in front of her eye. Since she still cannot read tion M when the object is at the primary focal point
the letters, Paula uses a +10.00 D thin lens as a can be determined from Figure 13.4. For the object at
magnifier. She holds the print 6 cm from the lens the primary focal point, plane waves leave the mag-
and the lens 4 cm from her eye. What angular nifying lens and all the exiting rays are parallel. The
magnification does Paula achieve? nodal ray for the eye again subtends the angle wm.
For the lens
However, the nodal ray for the eye is now parallel to
ux -6.00 cm, the nodal ray for the magnifying lens, and hence the
1^ =-16.67 D, object O subtends an angle equal to wm at the nodal
VY = +10.00 D + (-16.67 D) = -6.67 D, point of the magnifying lens. Therefore,
= -15.00 cm, tanw m = 0 / ( - f 1 ) .
m.ens = -16.67 D/(-6.67 D) = +2.5. The initial nodal angle we is given by
The object distance for the eye is
tanw e = 0 / ( - w e ) .
um = -(15 cm + 4 cm) = 19 cm,
and the angular magnification due to distance Then from Eq. 13.2,
changes is M = (-e)/(-f1).
Mdls
dis = ^ = 2.105.
-19 cm
The total angular magnification is
TABLE 13.1
M = mlensMdis = (2.5X2.105) = 5.26 X. Numerical Limit for Angular Magnification
ux (cm) vl (cm) m
iens Mdis M

13.5 Plus Lens as a Collimating Magnifier -4.00 -20 5 4.35 xlO"1 2.17
-4.50 -45 10 2.08 X101 2.08
-4.90 -245 50 4.03 x 10~2 2.02
As the object is moved closer to the primary focal -4.99 -2495 500 4.00 x 10" 3 2.00
point of the magnifying lens, the virtual image formed -5.00 00 00 0 ?
by the lens gets larger and moves farther away.
252 Geometrie, Physical, and Visual Optics

maanifying (F

\ 1
Tens s s,

^\ V
i
W^JK.
Fi \ In
% ,1
-^ FIGURE 13.4. Unaided eye and eye aided by a collimating mag-
nifier.

The dioptric power P of the magnifying lens is related EXAMPLE 13.7


to f ! by In the above example, the person initially viewed
the object 10 cm in front of his eye. Then the
P=-l/f1. person used the +20.00 D lens as a collimating
Therefore, magnifier held 3 cm in front of his eye. The person
increased his retinal image size by a factor of 2.
M = (-we)P. What is the angular magnification relative to the
initial 10 cm viewing distance when the person uses
For all real objects, we is negative and the +20.00 D lens as a collimating magnifier but
holds the lens 18 cm in front of his eye?
M = |tt e |R (13.10) Since for the collimating magnifier the person's
retinal image and angular magnification are in-
In Eq. 13.10, we must be in meters so that M is dependent of the distance between the magnifying
dimensionless; Eq. 13.10 gives the angular magnifica- lens and the eye, M should again be + 2 X . Indeed,
tion whenever the lens is used as a collimating mag- from Eq. 13.10,
nifier (i.e., whenever plane waves leave the lens). M = (0.10m)(+20.00D),
Note that Eq. 13.10 is independent of the distance M=+2.00X.
d between the eye and the magnifying lens. This can
be verified conceptually by noting that all the rays EXAMPLE 13.8
leaving the magnifying lens in Figure 13.4 are parallel Marylynn Minnow, an emmtrope, has grown old
and that no matter what the distance d is, the nodal and has a near point of 150 cm. Marylynn is trying
ray incident on the eye subtends an angle equal to w m . to read the menu at a Hollywood restaurant. Final-
Consequently, the retinal image size I m is also in- ly, she pulls out a +32.00 D thin lens and uses it as
dependent of the distance d. a collimating magnifier. What is Marylynn's in-
crease in retinal image size?
In this case, the best Marylynn can do without
EXAMPLE 13.6
the magnifier is to hold the menu at her near point,
Table 13.1 shows the angular magnification for a
and ue is then 150 cm in magnitude. From Eq.
+20.00 D magnifying lens held 3 cm in front of the
13.10,
eye. The initial condition was that the person di-
M = (1.5m)(+32.00D),
rectly viewed the object 10 cm in front of her eye.
According to Eq. 13.10, when the +20.00 D lens is M = 48 X.
used as a collimating magnifier, Note that the distance between the magnifying lens
M = (0.10m)(+20.00D), and the eye was not needed in the above example.
M=+2.00X.
EXAMPLE 13.9
The angular magnification of 2 X agrees with the Elmer Thugfuss, an 8-year-old emmtrope, is din-
value arrived at by the limit procedure shown in ing at the table next to Marylynn's. Elmer's near
Table 13.1. point is 6.5 cm in front of his eye. Elmer sees
Angular Magnification 253

Marylynn using the +32.00 D magnifying lens and Angular magnification was defined as the ratio of
wants to try it out. Will Elmer think it is as great as the magnified retinal image to an original unmagnified
Marylynn does? Assume that Elmer uses the lens retinal image. The actual angular magnification match-
as a collimating magnifier. es the value given by Eq. 13.11 only when the initial
Elmer achieves his largest retinal image without direct viewing distance is 25 cm. When the initial
the magnifier by holding the menu at his near
point. From Eq. 13.10, the increase in Elmer's direct viewing distance is not 25 cm, then the angular
retinal image size with the collimating magnifier is magnification achieved will not match the value given
by Eq. 13.11.
M = (0.065 m)(+32.00 D),
M = 2.08X.
Elmer does get a 2.08 increase with the magnifying EXAMPLE 13.10
lens, but that is a factor of 23.1 less than Sylvia Frost, an emmtrope with a 5 D amplitude
Marylynn's 48. of accommodation, views a stamp held at her near
Clearly, their different near points are respon- point. Since she cannot see the fine detail, Sylvia
sible for the different results. In fact, if Elmer held buys a 6X thin magnifier at a hobby shop. What
the menu at 150 cm (Marylynn's near point), and angular magnification does Sylvia achieve with the
then brought the menu in to 6.5 cm, he would 6 X magnifier relative to her near point?
achieve an angular magnification due to the dis- From Eq. 13.11, the 6X thin magnifier is a
tance change of +24.00 D lens. The default assumption is that Syl-
via uses the lens as a collimating magnifier. Sylvia's
M = 150 cm/6.5 cm = 23.1 X, near point is 20 cm in front of her eye. Therefore,
which is exactly the difference between their angu- M = (0.20 m)(+24.00 D) = +4.8 X.
lar magnifications.
Sylvia achieves an angular magnification of 4.8 X
relative to her near point. She does not achieve the
full 6 X because her near point is at 20 cm instead
of the conventional 25 cm reference distance. In
13.6 The Conventional Reference Distance fact, if Sylvia moved the object from 20 cm to
25 cm and then used the 6 X collimating magnifier,
As illustrated in Examples 13.8 and 13.9, different her total angular magnification would be the prod-
people get a different increase in retinal image size uct of the angular magnification due to the distance
change and the standard 6X of the magnifying
because of their different near points. This is extreme- lens. Specifically,
ly important to keep in mind when prescribing mag-
nifiers for a person with low vision. 2 Q m
C
/* VN

However, to avoid confusion in the general public, M=


25 cm (6X),
it is desirable to label lenses with one magnification or
value so that a person can go to a drug store or a M = (0.80)(6X) = 4.8X,
hobby shop and buy a specific magnifier (such as 4 X
which is the same value attained above.
magnifier). To do this a conventional, standard refer-
ence distance is needed.
The conventional standard reference distance is EXAMPLE 13.11
25 cm. This distance is sometimes referred to as the Foster Grant has a near point 40 cm in front of his
distance of most distinct vision, although the latter title eye. Foster uses a collimating magnifier and a-
is meaningless. The 25 cm is a point that might be a chieves an angular magnification of 9 X relative to
typical near point for a group of people that includes his near point. What is the standard magnification
both children, young adults, and some presbyopes. rating of the lens?
From Eq. 13.10, the power of the lens is given
However, other people have variously argued that by
either 40 cm or 100 cm would be a better choice.
Since 25 cm is 1/4 meter, the angular magnification P = M/|u e |,
for a collimating magnifier of dioptric power P refer- P = 9/0.40=+22.50 D.
enced to the standard distance is, from Eq. 13.10, From Eq. 13.11, the standard magnification is

4 (13.11)

Consequently, a +20.00 D thin lens would be marked +22.50 D


M= = 5.6X.
as a 5 X magnifier, while the +32.00 D thin lens would
be marked as an 8 X magnifier. The lens would be marked as a 5.6 X magnifier.
254 Geometrie, Physical, and Visual Optics

Equation 13.12 is the equation for a collimating mag-


13.7 Collimating Magnifier: Object to Eye
nifier under the restricted condition that the distance
Distance Unchanged
between the object and the eye is not changed when
the magnifier is added. Equation 13.10 is the general
Equation 13.10 is the general equation for a thin lens
equation. Historically, people have occasionally mis-
collimating magnifier. Under a restricted condition,
used Eq. 13.12 in situations where it was not ap-
we can derive another equation for a collimating
magnifier. The restricted condition is that the distance plicable.
between the object and the eye is unchanged when the
magnifier is put in place. EXAMPLE 13.12
Notice that in Figure 13.4, the distance between the Carmel Corn, an emmtrope with a neck cast, is
object and the eye changes when the magnifier is used. viewing a stamp on a board in front of her. Carmel
Figure 13.5 shows the restricted condition where the is unable to move her head or the board, but she is
distance between the object and the eye is unchanged able to use a 4X magnifier to examine the fine
when the magnifier is used. In Figure 13.5a, the detail on the stamp. Carmel places the 4X mag-
nifier in front of the stamp, notes that the image is
original viewing distance is ue. The magnifier is placed blurred, and moves the magnifier towards the
so that the object is at the primary focal point of the stamp until the image clears. At this point the
lens. The distance between the lens and the eye is magnifier is 30 cm from her eye. What angular
labeled d. From the restriction, magnification has Carmel achieved relative to her
initial direct viewing?
d = k|-|f,|, When Carmel initially looked through the mag-
or nifier, the light leaving the magnifier was converg-
ing, which is why she saw a blurred image. The
k l = |f,| + d. image clears when the light leaving the lens consists
Then the general equation (Eq. 13.10) gives of plane waves, so Carmel is then using the lens as
a collimating magnifier.
M = [|f1| + d]P, From Eq. 13.11,
or P = (4X)4=+16.00D.
Because the restricted conditions apply in this case,
M = |f1|P + dP. we can use Eq. 13.12,
For a thin lens in air, the focal length and the dioptric M = 1 + dP,
power are reciprocally related so that M = l + (0.30m)( + 16.00D),
M = 1 + dP. (13.12) M = 1 + 4.8 = 5.8 X.

a)

b) FIGURE 13.5. Special case where the distance between the


object and the eye is unchanged between the initial and
magnified viewing, a. Unaided eye. b. Eye viewing through a
luJ-d+IM collimating magnifier.
Angular Magnification 255

EXAMPLE 13.13 initially and when viewing through the magnifier. Fi-
Use the general equation (Eq. 13.10) to solve for gure 13.6 shows the situation. The object is initially
the angular magnification in the previous example. viewed at the distance ue. When the magnifying lens is
The general equation is used, the object distance ux is adjusted until the
M=| e |P, virtual image that the emmtrope sees is at a distance
where ue is the original direct viewing distance. um that is equal to we. From Eq. 13.9,
From the previous example the lens is a +16.00 D M = m lens M dis ,
lens used as a collimating magnifier and held 30 cm
from Carmel's eye. The focal length for a +16.00 D but since ue = um,
thin lens is 6.25 cm. Since Carmel is using the lens
as a collimating magnifier and holding it 30 cm from M dls = ^ = 1.
her eye, the physical distance between the stamp
and her eye is 36.25 cm. Under the conditions Then for this case
given, the 36.25 cm was Carmel's original viewing
distance. Thus, M = m lens ,
|we| = 36.25 cm = 0.3625 m,
and Eq. 13.10 gives
M = (0.3625 m)( +16.00 D), M = i; 1 (V 1 -P),
M = 5.8X. M= 1- L^P,
The general equation gives the same result as the
restricted equation used in the previous example. M = 1 + | 1 |P.

From Figure 13.6b,


kl = hl + d,
13.8 Magnifiers and Maximum or
Accommodation
kl = k l - d .
An emmtrope does not need to accommodate when Then we can substitute the above equation into the
looking through a collimating magnifier. However, the last equation for M to obtain
emmtrope might be able to achieve more magnifica-
M = l + (|Me|-d)P. (13.13)
tion through a magnifier by accommodating. In this
case, the light coming out of the magnifier is diverg- It is clear from Eq. 13.13 that the larger the distance d
ing, and Eq. 13.9 gives the angular magnification. Let from the magnifying lens to the eye, the smaller the
us investigate the particular case in which an emmet- angular magnification M. When d is zero, Eq. 13.13
rope uses the same amount of accommodation both gives a maximum.

FIGURE 13.6. Special case where the ocular accommoda-


tive demand is unchanged between the initial and the mag-
nified viewing, a. Unaided eye. b. Eye viewing through the
|u m | = |u e | = | v , | + d magnifier.
256 Geometrie, Physical, and Visual Optics

The d equal zero situation is not practical because 13.9 The Badal Principle
it implies that the magnifying lens is pressed against
(or in contact with) the eye. Nonetheless, let us as- The distance d between the magnifier and the eye does
sume d is zero and obtain the maximum M: make a difference even though we set it equal to zero
to obtain the mythical maximum. The d value tends to
M = l + |u e |P. (13.14) negate the value of ue, and thus bring M down from
the maximum value of Eq. 13.14. In fact, it appears
Equation 13.14 is sometimes compared to Eq. 13.10, that a large d can even make m smaller than the
which was the magnification for a collimating mag- collimating magnifier value of Eq. 13.10.
nifier. Note that for a finite we, the two equations This means that there is some d value for which the
describe different situations. Plane waves leave the angular magnification is the same irregardless of the
collimating magnifier and an emmtrope does not accommodative demand. We can find this d value by
have to accommodate. For the case of Eq. 13.14, the equating Eqs. 13.10 and 13.14. Then
light leaving the magnifier is diverging, and the person
needs to accommodate for the distance ue. By using l + (k|-d)P = k|P,
the accommodation, the person is able to add an
additional one unit of magnification over what he or or
she could obtain while unaccommodated (assuming 1 - dP = 0,
that d is zero). and
d = l / P = f2.
EXAMPLE 13.14
Sylvia Frost, the emmtrope in Example 13.10, had Thus, when d is equal to the secondary focal length
an amplitude of accommodation of 5 D. When of the magnifying lens, the angular magnification is
Sylvia used the +24.00 D lens as a collimating independent of the vergence leaving the magnifying
magnifier, the magnification rating was 6X but lens. This is called BadaVs principle.
Sylvia's actual angular magnification was 4.8 X. The optics of Badal's principle are shown in Figure
Relative to her near point, what is the maximum 13.7. Note that the lens representing the eye is placed
angular magnification that Sylvia can achieve?
The maximum angular magnification occurs at the secondary focal point of the magnifying lens.
when Sylvia accommodates maximally and when Three different object positions are shown. The ray
the distance d between the magnifier and the eye is shown is common to all three objects. The incident ray
zero. Equation 13.14 describes this case when we is is parallel to the axis, and the exiting ray passes
equal to the near point distance (which is 20 cm for through F 2 of the magnifier. Since that point is also
Sylvia), i.e., the nodal point for the lens representing the eye, the
M = l + (0.20)(+24.00D), ray proceeds straight through without bending and
specifies the retinal image size I m . Note that as the
M = 1 + 4.8 = 5.8 X.
object is moved, the vergence incident on the eye
The maximum angular magnification is 5.8 X com- changes, but if the eye accommodates correctly, the
pared to the 4.8 X that Sylvia gets when she uses retinal image will be clear and its size will not change.
the lens as a collimating magnifier. Note that it is
purely accidental that the 5.8 X obtained for Sylvia Badal's principle supplies a way to measure accom-
is close to the 6X rating of the magnifier. For a modative responses independent of retinal image size
different person (i.e., different set of numbers) changes. Badal's principle is also used in some other
these values may not be close. ophthalmic instruments such as a lensometer.

Equation 13.14 gives the mythical maximum value


of angular magnification available. Suppose this
amount is referenced to the conventional 25 cm or
1/4 m viewing distance. Then |we| in Eq. 13.14 equals
1/4 m and

M= l + (13.15)
4'
In conjunction with Eq. 13.15, the accommodative
demand for an emmtrope is 4 D both when directly
viewing the object and when viewing through the
magnifier. FIGURE 13.7. Badal principal.
Angular Magnification 257

13.10 General Equation for a shows that, in general, the angular magnification de-
Simple Magnifier pends on the initial observation distance we, the power
P of the magnifier, the distance d between the mag-
Section 13.4 described an intuitive method for obtain- nifier and the eye, and the directed distance x of the
ing the angular magnification of a simple magnifier. object relative to the primary focal point of the mag-
We can derive an equation that replaces the intuitive nifying lens.
method. The advantage of such an equation is that it There are, however, special circumstances, as pre-
conceptually ties much of the above discussion to- viously discussed, in which the dependence on some of
gether. The disadvantage of the equation is that the these variables drops out. For example, suppose the
optics involved are not as explicitly apparent as they magnifier is used as a collimating magnifier. Then the
were in the intuitive method. object is placed at the primary focal point of the
The derivation proceeds in the same manner as magnifier and x is zero. Equation 13.16 then simplifies
those in the previous sections. The result is that to Eq. 13.10 for any d, i.e.,
klP M = |we|P for any d.
M= (13.16)
l-[xP(l-dP)]' Alternatively, suppose that x is free to vary, but
where x is the directed distance from the primary focal that d is equal to the secondary focal length of the lens
point to the object (i.e., the extrafocal object dis- (Badal's principle). Then the term (1 - dP) is equal to
tance), d is the distance between the magnifying lens zero, and Eq. 13.16 again simplifies to Eq. 13.10 for
and the eye, P is the dioptric power of the magnifying any x, i.e.,
lens, and ue is the original viewing (or reference) M = |we|P for any x.
distance (Figure 13.3). All the distances in Eq. 13.16 Thus, the angular magnification is independent of the
must be expressed in meters. vergence leaving the magnifier.
EXAMPLE 13.15
Let us use Eq. 13.16 to rework Example 13.5 in-
volving Paula Bear.
From the data in Example 13.5, 13.11 The Compound Microscope
ue = 0.40 m,
d = 0.04 m. The simplest compound microscope consists of two
thin lenses. The lens closest to the object is called the
Since the object was situated 6 cm in front of a
+10.00 D lens, the x value is 4 cm, or objective lens, and its function is to form a magnified
real image inside the microscope (Figure 13.8). The
x=+0.04m. lens closest to the eye is called the ocular lens or the
Then eyepiece. The ocular lens serves as a simple magnifier
xP = (0.04 m)( +10.00 D) = 0.4, with the internal real image as its object. An un-
accommodated emmtrope or corrected ametrope
and would focus the microscope so that plane waves leave
( l - d P ) = l-(0.04)( + 10.00), the ocular lens.
= 1-0.4 = 0.6. The magnification of the microscope is expressed in
From Eq. 13.16, terms of the increase in retinal image size achieved
with the aid of the microscope. The comparison is with
(0.40)( +10.00 D) the retinal image size I e of an object viewed directly at
M=
1-[(0.4X0.6)] ' the standard reference distance of 25 cm. When one
4 looks through the microscope one has a larger retinal
M= image size I m . Thus, the magnification of the mi-
1 - 0.24 '
croscope is
=
76=526'
which is the same answer obtained in Example M=
13.5.
As before, the ratio of the retinal image sizes is equal
It is preferred to work this problem using the to the ratio of the angles subtended at the nodal point
intuitive method as shown in Example 13.5. However, of the eye. In Figure 13.4, we is the nodal angle in the
Eq. 13.16 does conceptually fit different aspects of initial viewing situation, while in Figure 13.8, wm is the
simple magnifiers together. Specifically, Eq. 13.16 angle subtended at the nodal point of the eye by the
258 Geometrie, Physical, and Visual Optics

objective ocular eye


4

Vi 1 FIGURE 13.8. Compound microscope.

rays leaving the microscope. Thus, So


tanwm U =-4.54D,
M =
tan w, and
From Figure 13.4, V2 =+45.45 D,
v2 = +2.20 cm.
tan w e = 0 / w e .
So in the forward direction, the object would be
For plane waves leaving the microscope, the internal placed 2.2 cm in front of the objective lens (just
image must be at the primary focal point of the ocular 0.2 cm outside the primary focal point of the objec-
tive lens), and the internal image is formed
lens. Then from Figure 13.8,
22.00 cm from the objective lens with a lateral
t a n w m = I1/f1. magnification m of 22/(-2.2) which equals - 1 0 .
From Eq. 13.17, the magnification of the mi-
Since croscope is
Ii=mobjO, + 80 D
(-10)
where m o b j is the lateral magnification of the objec-
tive, the above three equations give M = (-10)(+20X) = -200X.
Relative to the standard reference distance, the
MobjO/fl
M = = m
obj"eP magnification of the microscope is 200. The minus
sign indicates that the microscope's image is in-
The 25 cm standard reference distance gives verted.

P
M = m obj 4* (13.17) EXAMPLE 13.17
A microscope consists of a +30.00 D lens located
EXAMPLE 13.16 20 cm from the ocular lens. When focused for plane
A compound microscope consists of a +50 D objec- waves leaving, the object is 4.29 cm in front of the
tive lens located 23.25 cm in front of a + 8 0 D objective lens. What is the magnification of the
ocular lens. What is the magnification of the mi- microscope?
croscope when it is adjusted for plane waves In this case, let us think in terms of the light
leaving? traveling in the forward direction.
The most intuitive method to solve this problem M, = -4.29 cm,
is to reverse the light, in which case we consider
plane waves incident on the ocular lens. Then we U, = -23.31 D,
can follow the light back through the microscope V, = +30.00 D + (-23.31 D) = +6.69 D,
and find the object position.
In the reversed system, vx = +14.95 cm,
:
P_ + 0=+80D. -23.31 D
Then
The internal image is at the primary focal point of
vx = 1.25 cm, the ocular lens, so the focal length of the ocular
and lens must be
u2 = -(23.25 cm - 1.25 cm) = -22.00 cm. |f11 = 20 cm - 14.95 cm = 5.05 cm.
Angular Magnification 259

Then 100cm , Arfx 100cm


Poc = +19.80 D. d = +34.722
,_ D + 16.0cm + +72.00 D
From Eq. 13.17, the magnification of the mi- = 20.269 cm.
croscope is The image distance for the objective lens equals the
M = (-3.48)( + 19.80D/4) = -17.2X. secondary focal length of the objective plus the
extrafocal image distance x'. Therefore,
In the above examples, we obtained the lateral
magnification of the objective lens from the image v = +2.88 cm + 16.0 cm = +18.88 cm.
distance-object distance ratio or from the vergence Then
ratio. Eq. (9.7) was an alternative equation for the V= +5.297 D,
lateral magnification in terms of the extra-focal image
U = +5.297 D - 34.722 D = -29.426 D,
distance x':
and
m=-7-==-x'Pobj. u= -3.398 cm.
l
2

EXAMPLE 13.19
In the case of a microscope focused for plane waves
A 40 X microscope has a +25.00 D objective lens.
leaving, the distance x' is the distance between the What is the magnification when the objective is
objective lens' secondary focal point and the ocular switched to +93.75 D?
lens' primary focal point. The distance between these The default assumption is that the optical tube
two focal points is called the optical tube length of the length is unchanged. Then from Eq. 13.18,
microscope (Figure 13.8). Note that the actual tube M +93.75 D
length is the distance from the objective lens to the
ocular lens. When we substitute the above equation -40 X +25.00 D '
and
for m in terms of the optical tube length x' into Eq. M=-150X.
13.17, we obtain
When an emmtrope or corrected ametrope uses a
x'P P microscope, the focus condition for best visual comfort
M= r
objroc (13.18)
is to have plane waves leaving the microscope. How-
Equation 13.18 is really just a disguised version of Eq. ever, the microscope can be focused for either con-
13.17. verging or diverging light leaving. Equation 13.17 has
In many microscopes the objective lens or lenses the form that the angular magnification M of the
can be changed to increase the magnification. The microscope is equal to the lateral magnification m of
design is typically such that the optical tube length x' the objective times the angular magnification of the
remains the same when the objective lens is changed. ocular, or
Equation 13.18 explicitly shows that the magnification M = (m lat ) objective (M ang ) 0 (13.19)
increase is directly proportional to the increase in
Equation 13.19 is the general form for the magnifica-
objective lens power; 160 mm is a typical optical tube
tion of a two-lens microscope even when the mi-
length.
croscope is focused so that converging or diverging
light is leaving (as opposed to plane waves). In the
EXAMPLE 13.18
A 100 X compound microscope has an 18 X ocular latter case, the appropriate equation for the angular
lens and a 160 mm optical tube length. Find the magnification Mang of the ocular lens is used. For
power of the objective lens, the distance from the example, if the person focuses the microscope for an
objective lens to the ocular lens (the actual tube accommodative demand of 4 D , then Eq. 13.15,
length) when the microscope is focused for plane
waves leaving, and the distance from the objective Ma
lens to the object.
From Eq. 13.10, would be the appropriate equation for the angular
Poc = (18 X)(4 D) = +72.00 D. magnification of the ocular lens (instead of Mang =
P/4). Usually the dioptric power of the ocular lens is
Then from Example 13.18, high and the resulting change in magnification is very
-100 X = -(0.160 m)Pobj(18 X), small. In practical terms, this means that changing the
P obJ =+34.722 D. focus of a high magnification microscope does not
The distance between the two lenses equals the appreciably affect the magnification, so an emmtrope
sum of the two focal lengths plus the optical tube might as well focus for the comfortable situation of
length: plane waves leaving.
260 Geometrie, Physical, and Visual Optics

13.12 Equivalent Power Formulation ence distance. To use Eq. 13.19, we need to know
for Magnification where the object or the image for the front surface
is. Since the thick lens is serving as a collimating
The compound microscope consists of at least two magnifier, the object must be at the thick lens'
lenses. The microscope, like any system of spherical primary focal point. We can find the primary focal
point by reversing the light, in which case we
lenses, has three dioptric power parameters: the neut- consider plane waves incident on the back of the
ralizing power (also called the front vertex power), the lens. For the reversed case
back vertex power, and the equivalent power. In
terms of the Gauss system, the compound microscope U,=0,
acts like a single lens with the equivalent power. Vx = +16.00 D.
The derivation for the magnification of the com- The reduced thickness of the lens is 33 mm/1.50 or
pound microscope (relative to the 25 cm standard 22 mm. Then
reference distance) then proceeds exactly like that of U2 = +24.69 D,
Eq. 13.11 with P replaced by the equivalent power P e . V2 = +46.69 D,
The result is
v2= +2.14 cm.
(13.20) For light in the forward direction, the primary focal
M-*. point of the lens is 2.14 cm in front of the +22.00 D
surface. When the object is in the primary focal
EXAMPLE 13.20
plane (light in the forward direction),
The compound microscope in Example 13.16 con- ux = -2.14 cm,
sisted of a +50.00 D lens located 23.25 cm in front U t =-46.69 D,
of a +80.00 D lens. Find the magnification by using Vx = -24.69 D,
the equivalent power equation, and compare the
results to those in Example 13.16. U 2 = -16.00 D,
The equivalent power is V2 = 0.
Pe = +50.00 D + 80.00 D - (0.2325 m)(+50.00 D) The lateral magnification for the +22.00 D surface,
x (+80.00 D), which serves as the objective, is
Pe = -800D. Hi -46.69 D
= +1.89.
From Eq. 13.20, -24.69 D
-800 D The +16.00 D surface serves as the ocular, and
M= = -200X. from Eq. 13.19, the total magnification relative to
the standard reference distance is
The magnification of -200 X agrees with Example
13.16. M = ( + 1 . 8 9 ) ^ = + 7.56X.
9

Simple magnifiers are high plus lenses. In the real The alternative formulation is to calculate the equi-
world, these lenses are thick. The reduced system valent power and use Eq. 13.20. For the thick lens,
corresponding to a thick lens consists of two thin ^, + ,-dP^,
lenses in air, respectively representing the front and Pe = +22.00 D + 16.00 D - (0.022 m)
back surfaces of the thick lens. Thus, the magnifica- x ( +22.00 D)( +16.00 D),
tion theory of the thick lens exactly parallels that of
Pe = +30.26 D.
the two-lens compound microscope. By interpreting
the front surface as the objective and the back surface From Eq. 13.20,
as the ocular, we can use Eq. 13.19. Alternatively, we +30.26 D
can use the equivalent power of the thick lens and Eq. M= = +7.56X.
13.20. 4
The two different methods give the same result.
EXAMPLE 13.21 EXAMPLE 13.22
A thick (n = 1.50) biconvex lens has front and back This example considers a thick meniscus-convex
surface powers of +22.00 D and +16.00 D, respec- lens (n = 1.50) used as a collimating magnifier. The
tively, and a central thickness of 33 mm. When this lens has front and back surface powers of +22.00 D
lens is used as a collimating magnifier, what is the and -3.00 D, respectively, and a central thickness
magnification? of 21 mm. What is the magnification?
The default assumption is that the magnification Again, we can find the primary focal point by
requested is relative to the standard 25 cm refer- reversing the light and considering plane waves
Angular Magnification 261

incident on the back of the lens. For the reversed 13.13 Afocal Telescopes
case
Ux=0,
V ^ - 3 . 0 0 0. A slight modification of the compound microscope
yields the telescope. We use a telescope to magnify a
The reduced thickness is 21 mm/1.50, or 14 mm. distant object, whereas the microscope was used to
Then magnify a near object. The direct viewing of the
U2 = -2.88D,
distant object without the aid of the telescope consti-
V 2 =+19.12 D, tutes the initial situation. Then the distant object is
v2= +5.23 cm. viewed through the telescope. The (angular) magnifi-
When the object is in the primary focal plane (light cation of the telescope is the ratio of the retinal image
in the forward direction), size when the telescope is used to the retinal image
ul = -5.23 cm, size in the initial situation. When the telescope is
afocal, it gives plane waves out for plane waves in.
U! = -19.12 D,
Figure 13.9 shows a two-lens Keplerian or as-
V^+2.88 0, tronomical telescope. The Keplerian telescope has a
U 2 =+3.00D, converging objective lens and a converging ocular
V2 = 0. lens. The two lenses are separated by the reduced
The lateral magnification for the +22.00 D surface distance d. The converging objective lens forms an
is inverted, real image inside the telescope. The ocular
U, -19.12D lens is then used as a simple magnifier to observe the
m= = -6.64. internal image. In Figure 13.9, the incident rays be-
+2.88 D
long to one bundle that originated at a distant off-axis
The -3.00 D surface serves as the ocular, and from image point. When the telescope is not present, the
Eq. 13.19, the total magnification relative to the
standard reference distance is rays in the bundle subtend an angle we at the eye.
When the telescope is present, the incident rays sub-
M = (-6.64) ~ 3 ^ Q P = +4.98 X. tend the angle we at the objective lens of the telescope
(Figure 13.9). The bundle of rays leaving the tele-
Note that the -3.00 D surface actually decreases scope is a parallel bundle subtending the angle wm.
the magnification relative to the 25 cm reference Assume that the observer looking directly at the
distance. distant object has a retinal image of size I e , and that
The alternative formulation is to calculate the
equivalent power and use Eq. 13.20. For the thick when the observer looks at the distant object through
lens, the telescope, the retinal image size is I m . Then the
(angular) magnification M is given by
Pe = Pi + P 2 -P 1 P 2 ,
Pe = +22.00 D + (-3.00 D) - (0.014 m) M = I21 = tanw i!!
x (+22.00 D)(-3.00D), T tan w
Pe = +19.92 D.
From Eq. 13.20, The internal image has a size \x (which is negative)
+ 19 92 D and is formed at the secondary focal point of the
4 objective lens. According to the sign convention for
Again, the two different methods give the same ray angles, the angle is positive when the ray slopes
result. upward and negative when the ray slopes downward.

FIGURE 13.9. Keplerian telescope.


262 Geometrie, Physical, and Visual Optics

The nodal ray for the objective lens shows that EXAMPLE 13.24
A Galilean telescope consists of a +2.00 D objec-
I.
11 p tive lens and a -10.00 D ocular lens. Find the
tan w p = =T
magnification and length of the telescope.
(f 2 ) o b ) *obj From Eq. 13.21,
The nodal ray for the ocular lens shows that -10.00 D
M=- +5X.
tan w m = I, = -,, +2.00 D
(fl)o The magnification is positive, which is consistent
The above equations give with the fact that the image seen through the
telescope is erect.
M = - (13.21) Again, air is between the two lenses, so the
L
obj length is equal to the reduced length. Remember
that the secondary focal length of the diverging lens
Also from Figure 13.9, the reduced length of the
is negative.
telescope is equal to the sum of the reduced secondary
focal lengths of the ocular and objective lenses. In d = d = 50.00 cm + (-10.00 cm) = 40 cm.
equation form,
EXAMPLE 13.25
d = (f2)obj + (f2)oc. (13.22) A solid glass (n = 1.5) rod 21 cm long is made into
an afocal Galilean telescope. If the front surface of
EXAMPLE 13.23 the rod has a +5 D surface power, what is the
A Keplerian telescope has a +2.00 D objective lens power of the back surface of the rod and what is
and a +10.00 D ocular lens. What is the magnifica- the magnification of the telescope?
tion and length of the telescope? We can find the back surface power by following
From Eq. 13.21, incident plane waves through the telescope. The
back surface has to have the correct power to give
+ 10.00 D
M= = -5X. plane waves leaving the back. Alternatively, we can
+2.00 D use Eq. 13.22,
The minus sign indicates that the image seen
through the Keplerian telescope is inverted. A 21 cm
Air is between the two lenses, so the actual d= r = 14 cm.
1.5
length of the telescope equals the reduced length. Then
Equation 13.22 gives 100
14cm =
d = d = 50 cm + 10 cm = 60 cm. +5.00 D +(f 2 )oc
Figure 13.10 shows an afocal Galilean or terrestrial 14 cm = + 2 0 . 0 0 c m + (f 2 ) 0
telescope. The Galilean telescope has a converging
or
objective lens and a diverging ocular lens. The image
(f 2 ) oc = -6.00cm.
seen through the Galilean telescope is magnified and
erect. The equations for the Galilean telescope are the So
same as those for the Keplerian telescope. P oc = -16.67 D,

FIGURE 13.10. Galilean telescope.


Angular Magnification 263

and from Eq. 13.21, the deteminant condition simplifies to


-16.67 D
+3.33 X. _ J_
M = - +5.00 D S
2 2 ~~ c

The afocal telescope takes plane waves in and gives


The magnification of the telescope is equal to the ratio
plane waves out. When one turns the telescope
of the angles subtended by the rays, or
around, plane waves still leave when plane waves are
incident, but now the telescope minifies instead of
magnifies. The same equations (Eq. 13.21 and 13.22) q
still hold.
Since the matrix multiplications for q' give
EXAMPLE 13.26 q'=snq,
Samantha Smelthooker takes the telescope of Exam- it follows that
ple 13.24 and looks through it backwards at a distant
object. What magnification does Samantha achieve? M = sn.
The telescope in Example 13.24 was a 5 X Galilean Then the simplification of the determinant condition
telescope with a +2.00 D objective lens and a gives
-10.00 D ocular lens. When Samantha looks through
it backwards, the +2.00 D lens becomes the ocular 1_
S22
and the -10.00 D lens becomes the objective. ~M
Then from Eq. 13.21, Therefore, the system matrix for an afocal telescope is
+2.00D _ 1
/M (T
-10.00 D 5
Note that the 5 X telescope minifies by a factor of 5 S= (13.23)
when used backwards. * 21
Ml
Since an afocal telescope gives plane waves out for The system matrix is the product of the individual
plane waves incident, the equivalent power of the refraction and translation matrices. For the two lens
telescope is zero. Whenever the equivalent power is afocal telescope,
zero, the back vertex power is zero, and the neutraliz-
S = RCTRobj>
ing power is zero.

l l -Poc\ I1 0W1 obj 1

s= lo 1 / la 10
13.14 Alternate Afocal Telescope Equations

The length of an afocal telescope and the dioptric


powers of the two lenses are interrelated. Therefore,
we can derive alternate expressions for the afocal
-c ?)c -dPoc
~~ " o b j

1-dP

telescope magnification. One method is to use the


system matrix approach.
( '
a 1-
0
)
Let us use q as the ray angle and h as the perpen-
dicular distance from the axis to the ray. For paraxial Then
rays, the system matrix satisfies the equations, M=l -dP , (13.24)
*"*x oc>

or
1
\ (13.25)
and the determinant condition, 1-dP obj

S i 1 S-> S19S0 = +1. Equations 13.24 and 13.25 are alternate magnification
equations for an afocal telescope. Equations with the
Since form of (Eq. 13.25) are used considerably in spectacle
Pe = - S 1 2 = 0 , magnification.
264 Geometrie, Physical, and Visual Optics

EXAMPLE 13.27 Then


Use Eq. 13.25 to find the magnification of the
telescope in Example 13.25. In Example 13.25, a ^ = l-(0.25m)P o b j ,
solid glass (n = 1.5) rod 21 cm long was made into a
Galilean telescope with a front surface power of +0.0833 = 1-(0.25 m)Pobj,
+5.00 D. The reduced length of the rod is 14 cm. P o b j =+3.66D.
From Eq. 13.25,
From Eq. 13.21,
l-(0.14m)(+5.00D)' Poc = -MP o b j = - ( + 12X)(+3.66),
Poc = -44.00 D.
1 - 0.70 0.3 ' The Galilean telescope has a minus ocular lens as
expected. We can further verify the arithmetric by
M=+3.33X, checking whether or not the dioptric powers found
which agrees with the answer found in Example give the correct length. From Eq. 13.22,
13.25. - 100 100
+3.66 D + +44.00 D '
EXAMPLE 13.28
A 12 X Keplerian telescope has a length of 25 cm. d = +27.27 cm + (-2.27 cm) = 25.00 cm.
What are the dioptric powers of the objective and
ocular lenses?
The Keplerian telescope inverts so
M=-12X. 13.15 Converting a Telescope to a Microscope
From Eq. 13.25,
M
1 Some low-vision patients wear small spectacle
" 12X
-l-(0.25m)Pobj mounted telescopes to aid their distance vision.
When diverging light is incident on an afocal tele-
Then
scope, the light leaving can be considerably more
^ 2 = 1 -(0.25 m)Pobj, divergent. Hence, when a low-vision patient looks
-0.0833 = 1 - (0.25 m)Pobj, through the telescope at a near object, the accom-
modative demands can be large.
Pobj = +4.33D. The accommodation problem can be avoided by
From Eq. 13.21, placing a reading cap on the front of the telescope.
Poc = -MP obJ = -(-12X)(+4.33), In thin lens theory, the reading cap serves as a
collimating magnifier. Thus, the diverging light
Poc =+52.00 D. from the object is incident on the reading cap,
We can verify the arithmetic by checking whether which converts the light into plane waves. The
or not the dioptric powers give the correct length. plane waves then pass through the telescope and
From Eq. 13.22, are magnified in the normal manner. The total
- 100 100 magnification is then the product of the magnifica-
+4.33 D + +52.00 D ' tion of the magnifier and the magnification of the
d = +23.08 cm + 1.92 cm = 25.00 cm, telescope.
which is consistent. Note that it was crucial to use Mtot = McapMtel. (13.26)
-12X as the magnification. When +12X is used,
the lenses that make a Galilean telescope are Since we are again dealing with a near object, it
found.
is important to specify the initial conditions. If we
assume the standard 25 cm reference distance, then
EXAMPLE 13.29
A 12 X Galilean telescope has a length of 25 cm. the magnification is the ratio of the observer's
What are the dioptric powers of the objective and retinal image size that results from viewing through
ocular lenses? the telescope with the reading cap to the retinal
For the Galilean telescope, image size that the observer has when directly
M=+12X. viewing the near object at 25 cm. Equation 13.26
then becomes
From Eq. 13.25,

+ 12X= M tot = ^f- M tel . (13.27)


l-(0.25m)Pobj
Angular Magnification 265

The Galilean telescope with the reading cap is Examples 13.30 and 13.31 show numerically that
actually a microscope that gives an erect image. the telescope with a reading cap is equivalent to a
The microscope magnification equations in Section microscope. We can also start with Eq. 13.26 and
13.12 are general equations that should give the show this algebraically. Let Pobj and Poc be the respec-
same M t o t . This is explored in the following two tive objective and ocular lens powers of the telescope.
examples. From Eqs. 13.26 and 13.21,

EXAMPLE 13.30
M=-
A 3 X telescopic low vision aid has a 24 mm length. 4 Pobj
A +3.5 D reading cap is added to the telescope.
What is the resulting magnification? or
P P
From Eq. 13.27, M= ^ -2 (13.28)
P 4
+3.50 D r
obj ^
M= (3X), The reading cap and the objective lens of the tele-
scope act together like the objective lens of the mi-
M = (0.875)(3X) = 2.625X.
croscope. The lateral magnification of the micro-
Note that the magnification of the cap is less than scope's objective lens is given by
1, or relative to the standard 25 cm reference dis-
tance the cap actually minifies. U
m= V
where
EXAMPLE 13.31
(Pcap + Po b j ) + U .
Treat the above telescope as a microscope and find
the magnification from the microscope magnifica- When the cap is acting like a collimating magnifier,
tion equations.
There are several ways to solve this problem, U = -P.,
but, in any event, we need to find the objective and and then
ocular lens powers for the 3 X telescope. From Eq.
13.25, V=P,obj '
+3X= From the above equation, the lateral magnification
1 - (0.024 m)P obj ' becomes

m=
(0.024 m)Pobj, obj

and
l
Pobj =+27.78 D.
We can substitute the above equation into Eq. 13.28
and note that Poc + 4 is the angular magnification of
Then from Eq. 13.21, the ocular lens. Therefore,
Poc = - ( 3 XX+27.78 D) = -83.33 D. M = (m lat ) obj (M ang ) oc ,
The +3.50 D reading cap is placed against the which is Eq. 13.19 for the magnification of a mi-
+27.78 D objective. With the thin lens assumption, croscope.
these two lenses act like a single new objective lens
of power +31.28 D. The total system now looks
like a microscope with a +31.28 D objective, a
-83.33 D ocular lens, and a length of 24 mm. For
the correct focus, the object is placed 33.33 cm 13.16 Magnification Terminology
from the reading cap. We can now use any of the
microscope magnification equations to find the Unfortunately, there is no standardized magnification
magnification. terminology. Many authors use the name apparent
The equivalent power of a two-lens system is angular magnification and contrast it with the name
P = p +p actual angular magnification. Most of these authors
e i 2-dP2P1,
define apparent angular magnification as the ratio of
Pe = +31.28 D + (-83.33 D) - (0.024 m) the magnified retinal image size relative to the un-
x (-83.33 D)(+31.28D), magnified retinal image size when directly viewing the
Pe = +10.50 D. object at the standard 25 cm reference distance. They
Then from Eq. 13.20, further define actual angular magnification as the ratio
of the magnified retinal image size to the unmagnified
P=10^0D = retinal image size that exists when the magnification
4 4 system is removed without changing the distance be-
which is the same result as the preceding example. tween the eye and the object.
266 Geometrie, Physical, and Visual Optics

A number of other authors use the phrase magnify- front of her eye. Emmy then holds the stamp 3 cm
ing power in place of angular magnification. Most (but in front of a +20.00 D lens. Her eye is 10 cm from
not all) of these authors define magnifying power as if the lens. How much must Emmy accommodate in
it were apparent angular magnification. In some order to see the magnified stamp? What is the
books, magnifying power is defined as actual angular magnification relative to the original 15 cm dis-
magnification, but the equations used correspond to tance?
those for apparent angular magnification. As you 7. Barry Oilwater, an uncorrected 52-year-old em-
might expect, some people get very confused by these mtrope, uses a 4X magnifier to read the news-
terminology differences. paper. What is the angular magnification when
Here, use of the adjectives apparent and actual is Barry first holds the paper at 40 cm and reads
avoided. Instead, it is emphasized that angular magni- directly, and then holds the magnifying glass
fication is a comparison of two different retinal image 20 cm from his eye and views the collimated light
sizes. In particular, one needs to pay attention to coming through the magnifier?
exactly what unmagnified retinal image the magnified 8. A compound microscope has a +50.00 D objec-
image is being compared to. For generality, the dis- tive lens and a +40.00 D ocular lens. The mi-
tance between the object and the eye is allowed to croscope is adjusted so that the specimen is placed
change when the magnification system is put in place. 2.25 cm from the objective lens of the microscope.
When the latter distance does remain constant, What is the distance from the objective lens to the
specialized equations such as Eq. 13.12 can be used. ocular lens? What is the magnification? What is
However, Eq. 13.12 cannot be used when the distance the optical tube length of the microscope?
changes. 9. A Keplerian telescope has a +2.00 D objective
As noted in Section 1.1, models are used for lens and a +24.00 D ocular lens. What is the
economy and simplicity. In this chapter, a model magnification and the length of this telescope?
based on retinal image size is used and the retinal 10. What is the length and ocular lens dioptric power
image sizes are related to subtended angles. For sim- of a 10 X Galilean telescope with an objective lens
plicity, a thin lens and screen eye model is used. In of +3.00 D?
Chapter 14, a more complicated eye model is used. 11. A +8.00 D thin lens is used as an erecting lens in
A number of equations appeared in this chapter. A a 7 X Keplerian telescope with an objective lens of
subset of the most important equations consists of +2.00 D. The length of the telescope without the
Eqs. 13.9, 13.10, 13.11, 13.18, 13.19, 13.20, 13.21, and erecting lens is 57.14 cm. What is the length after
perhaps 13.25. the erecting lens is added?
12. A 4-cm thick biconvex glass (n = 1.50) lens has a
+10.00 D front surface and a +6.00 D back sur-
Problems face. As a collimating magnifier, what is the angu-
lar magnification relative to the conventional
1. Randy Vermin holds an object 50 cm from his 25 cm reference distance?
cornea and then moves the object closer until it is 13. A 3 X magnifying lens is held 5 cm in front of a
20 cm from his cornea. What gain in retinal image 4X magnifying lens. An object is placed so that
size does Randy obtain? What is the magnification the light that travels through both lenses emerges
name for this change? collimated. What is the angular magnification for
2. Joe Brown holds a stamp 16 cm from his eye and this system relative to the conventional 25 cm
views the stamp directly. He then picks up his reference distance?
+28.00 D magnifying lens, places the stamp at F{ 14. A 7X Galilean telescope has a +4.00D objective
of the magnifying lens, and holds the magnifying lens, a -28.00 D ocular lens, and a length of
lens 11cm from his eye. What is the angular 21.43 cm. The telescope is converted to a near
magnification in this situation? magnification device by placing a 3 X simple mag-
3. What magnification will be marked on a +28.00 D nifier (a reading cap) against the front surface.
magnifier in your local stamp collector's shop? What is the total magnification when plane waves
4. Suppose Joe Brown uses his left hand to hold a emerge from the ocular lens (relative to the 25 cm
stamp 16 cm from his eye and then without mov- conventional reference distance)?
ing either his left hand or his eye inserts his 15. Suppose the 3 X reading cap is placed against the
+28.00 D magnifying glass between the stamp and ocular lens of the 7X telescope in the previous
his eye so that the stamp is at of the magnifying problem, and the object adjusted until plane
glass. What is the magnification in this case? waves leave the last lens. What is the magnifica-
5. Could Eq. 13.12 be used for problem no. 2? tion relative to the conventional 25 cm reference
Could Eq. 13.12 be used for problem no. 4? distance? How far must the object be from the
6. Emmy Trope directly views a stamp held 15 cm in objective lens? (Warning: This is a poor design.)
CHAPTER FOURTEEN

Spectacle
Magnification
and Relative
Spectacle
Magnification
14.1 The Aperture Stop and the When the aperture stop is the first element in a
Entrance Pupil system, then it is easy to specify the angular size of
the bundle of incident rays that will get through the
As discussed qualitatively in Chapter 6, an ametrope's system (Figure 14.1a). When the aperture stop is not
retinal image size is dependent on whether the amet- the first element in the system, then it is not as
rope is corrected by a spectacle lens or a contact lens. straightforward to specify the bundle size of the object
This chapter considers these differences quantitatively. space rays that will pass through the system.
Spectacle magnification is the name given to the For the system shown in Figure 14.1b, one of the
change that occurs in the retinal image size when a easiest ways to find the angular bundle size of object
lens is placed in front of the eye. In spectacle magnifi- space rays that pass through the system is to consider
cation, one of the retinal images involved is usually the second lens C as an object and trace rays from it
blurred. The properties of a blurred image depend on backwards (left) through the first lens B, as shown in
both the lenses and the apertures in a system. Figure 14.2a. The rays emerging backwards appear to
Apertures are placed in optical systems to control be coming from the top of the virtual image D of the
the amount of light passing through the system. The second lens formed by the first lens. One of the rays
apertures also affect the field of view of the system. coming from the image point passes through the axial
Even when no apertures are present, the lenses or position A. Figure 14.2b shows the same system with
mirrors have finite sizes and act as apertures. the rays leaving an object point at the axial position A
For an axial object point, one aperture or lens rim and traveling forward (right) through the first lens. By
is most effective at limiting the amount of light that the principle of reversibility, the ray leaving the object
passes through the system. This particular aperture is point A pointing toward the top of the image D is bent
called the aperture stop of the system. Note that a by the first lens and passes through the top edge of the
system may have many apertures or stops, but only lens C. This ray is the top boundary ray for the bundle
one is called the aperture stop. that gets through the system. Any rays aiming higher
Figure 14.1a shows a two-lens system in which the miss the second lens. A similar boundary ray exists on
first lens is the aperture stop, while Figure 14.1b shows the bottom. Thus, when the lens C is the aperture
a two-lens system in which the second lens is the stop, the bundle size of the object space rays that
aperture stop. The aperture stop is object distance pass through the system is determined by the image D.
dependent, and a system's aperture stop may switch The image of the aperture stop formed by any
when the object is moved. For example, Figures 14.1a lenses in front of it is called the entrance pupil. When
and 14.1b might be the same system with a different there are no lenses in front of the aperture stop, the
object distance. aperture stop itself is called the entrance pupil.

267
268 Geometrie, Physical, and Visual Optics

a)

FIGURE 14.1. a. First lens is the aper-


ture stop. b. Second lens is the aperture
stop.

Theoretically, when one looks from the axial object In Figure 14.2b, the image D is the entrance pupil.
point toward the front of a system, all the possibilities The bundle size of the object space rays that get
for the entrance pupil are visible (either apertures, through the system is specified by D. An observer at
lens rims, or images of apertures and lens rims formed A looking toward the two lenses sees the first lens B
by the preceding lenses). The entrance pupil itself is and the image D of the second lens formed by the first
the candidate that subtends the smallest angle at the lens. The image D subtends a smaller angle at A than
axial object point. Because the entrance pupil sub- does the first lens B, so D (the entrance pupil)
tends the smallest angle, it specifies the bundle size appears smaller to the observer than B.
of the object space rays that ultimately pass through Figure 14.3a shows a two-lens system with an aper-
the system. Also, because the entrance pupil subtends ture in the middle. The aperture may or may not be
the smallest angle, it appears smaller than the other the aperture stop of the system. When the system is
candidates when viewed from the axial object point. viewed from the axial object point A, the candidates

FIGURE 14.2. a. D is the image of C formed by lens


B. b. D is the entrance pupil for this case.
Spectacle Magnification and Relative Spectacle Magnification 269

a)

T I
I D

red

red

::IU- 1 -

red
c)
red red
I -^A-

B
T- r
G
FIGURE 14.3. a. The system, b. The
entrance pupil candidates, c. Entrance
pupil candidates for changed sizes.
270 Geometrie, Physical, and Visual Optics

for the entrance pupil are visible. As shown in Figure EXAMPLE 14.1
14.3b, these candidates are the first lens B, the image A +12.00 D lens is located 8 cm in front of a
E of the aperture formed by the first lens, and the circular aperture with a 2 cm diameter. The lens
image G of the second lens formed by the first lens. Of itself has a circular shape and a diameter of 3 cm.
these candidates, the image E of the aperture formed An axial object point is 50 cm in front of the
by the first lens subtends the smallest angle, and so +12.00 D lens. Find the aperture stop of the
system.
would be the entrance pupil (and would visually ap- Remember to follow good problem-solving
pear smallest from the position of the axial object strategy and try to anticipate the answer before
point A). It follows that the aperture stop must be the working out the solution.
aperture C. The general approach is to first find the entrance
In Figure 14.3c, the entrance pupil candidates are pupil. The entrance pupil is the candidate that
again the first lens B, the image E of the aperture subtends the smallest angle at the axial object
formed by the first lens, and the image G of the point. The candidates are the +12.00 D lens (since
second lens formed by the first lens. Of these candi- it is directly visible from the axial object point) and
dates, the image G of the second lens formed by the the image D of the aperture formed by the
first lens subtends the smallest angle, and so would be +12.00 D (Figure 14.5a). Note that the aperture C
is not directly visible from the object plane, but
the entrance pupil (and would visually appear smallest that the aperture's image D formed by the
from the axial object point A). In this case, the second +12.00 D lens is directly visible from the object
lens is the aperture stop. plane.
An equivalent air (reduced) system representation The top half of the lens subtends the angle Wj at
of the human eye is an example of a two-lens system the axial object point A where
together with an aperture. The cornea serves as the 1.5 cm
first lens and the crystalline lens serves as the second tan w,1 = = 0.03.
50 cm
lens. The pupil serves as the aperture. Let us assume
that the person is reading a book. If you place your To find the image D, we take light coming from the
eye at the book's position and look back at the aperture and image it through the +12.00 D lens.
reader's eye (Figure 14.4), you can see the cornea B This light is traveling in the reversed direction (or
and the image E of the eye's pupil formed by the left in Figure 14.5a). Thus,
cornea. Of these two, the pupil's image E appears
ur = - 8 cm,
smaller. The remaining candidate for entrance pupil is
the image G of the crystalline lens formed by the Ur=-12.5D,
cornea, but this image subtends a larger angle than the Vr = +12.00 D + (-12.5 D) = -0.5 D,
pupil's image. Therefore the pupil's image E is the vr= -200 cm.
entrance pupil of the eye. You might think about this The lateral magnification is
while closely observing someone else's eyes. Typically,
the iris sits about 3.6 mm behind the cornea, and the -12.5 D
m =
image of the iris formed by the cornea is virtual and ^ 5 ^ = +250
located about 3.0 mm behind the cornea (Example The diameter of the image D is
7.19). (+25.0)(2cm) = 50cm.

A (book)

FIGURE 14.4. Entrance pupil representation for


an eye.
Spectacle Magnification and Relative Spectacle Magnification 271

a)
+ 12.00D
1
-*

IT
u By

8cm

b) (not to scale)

FIGURE 14.5. a. Entrance pupil candidates.


b. Axial bundle size. c. Entrance pupil candi-
dates and axial bundle size for a different object
distance.

At the object point A, the top half of the image D w1 < w 2 ,


subtends the angle w2 where
and the +12.00 D lens subtends the smaller angle
25 cm at the object point A. Thus, the +12.00 D lens is
tan w2 = (200 + 50) cm 0.10.
the entrance pupil of the system. Since the
Both angles are acute (i.e., less than 90), and, +12.00 D lens is not an image, it is also the aper-
therefore, the angle with the smaller tangent is the ture stop of the system.
smaller angle. Since You may have initially expected that the aper-
ture would be the aperture stop. However, the
tan w, < t a n w 2 , image conjugate to the axial object point A is only
it follows that 10 cm behind the lens, and the entire converging
272 Geometrie, Physical, and Visual Optics

axial bundle of image space rays gets through the To find the turning point set,
aperture (Figure 14.5b). So the lens is the element
tanwj = tan w 2 ,
that limits the amount of light from the axial object
point. or
1.5 25
EXAMPLE 14.2 -u 200- u '
For the same lens and aperture used in the preced- Then
ing example, how close must the axial object point
A be to the +12.00 D lens for the aperture to be 300-1.5w = - 2 5 M,
the aperture stop? 300=-23.5M,
Assume that the axial object point A is a dis-
tance u from the +12.00 D lens (Figure 14.5a). The and
top half of the lens then subtends the angle wx at A M = -12.77 cm.
where
1.5 cm The axial object point A would have to be closer
tan w, = . than 12.77 cm to the lens for the aperture to be the
1
u aperture stop (Figure 14.5c). At 12.77 cm, the
Note that u is negative, so minus u is positive. The + 12.00 D lens and the aperture tie for the aperture
image D of the aperture then subtends the angle w 2 stop (the dashed line from J to D in Figure 14.5c).
at A where When the axial object point A is farther than
25 cm 12.77 cm from the lens, the lens is the aperture
tan w2 = stop.
200n cm

entrance FIGURE 14.6. a. Marginal ray. b. Chief


pupil ray. c. Blurred image size.

b) entrance

entrance

blurred
image)
Spectacle Magnification and Relative Spectacle Magnification 273

The entrance pupil is conjugate to the aperture In Figure 14.7a, the virtual image D is the image of
stop. Therefore, to get a ray to just pass through the the first lens B formed by the second lens C. The first
edge of the aperture stop, one would aim the ray lens B is the aperture stop, so the virtual image D is
toward the edge of the entrance pupil as shown in the exit pupil. The angular bundle size ' of the image
Figure 14.6a. After refraction by any lenses in front of space rays is determined by the virtual image D. In
the aperture stop, the ray just passes the edge of the Figure 14.7b, the first lens B is again the aperture
aperture stop. Such a ray is called a marginal ray. stop. The real image D is the image of B formed by
Similarly, to get a ray to pass through the center of the second lens C. So the real image D is the exit
the aperture stop, one would aim the ray toward the pupil, and it determines the angular bundle size ' of
center of the entrance pupil. After the refractions, the the image space rays.
ray passes through the center of the aperture stop At the axial object point, the entrance pupil sub-
(Figure 14.6b). The ray from an object point that is tended the smallest angle relative to the other candi-
incident on the system pointing toward the center of dates for entrance pupil. The candidates for exit pupil
the entrance pupil is called the chief ray. After the are the last lens, any aperture after the last lens, and
appropriate refractions, the chief ray passes through the image of each lens and aperture formed by the
the center of the aperture stop. For a paraxial bundle, lenses behind it. At the axial image point, the exit
the chief ray is the central ray of the bundle (Figure pupil subtends the smallest angle of these candidates.
14.6c) and takes on special significance when dealing Theoretically, when viewed from the axial image posi-
with blurred image sizes. tion, the exit pupil appears to be the smallest of the
candidates.
The chief ray enters a system pointing toward the
center of the entrance pupil, and after the appropriate
14.2 The Exit Pupil and Blurred Image Sizes refractions actually passes through the center of the
aperture stop. The exit pupil is conjugate to the
The exit pupil is the image of the aperture stop formed aperture stop, and so when the chief ray leaves the
by any lenses behind it. When there are no lenses system, it appears to be coming from the center of the
behind the aperture stop, the aperture stop itself exit pupil.
serves as the exit pupil (e.g., Figure 14.6c). Just as the The system's entrance pupil is in object space, and
entrance pupil specifies the bundle size of the object the system's exit pupil is in image space. For the entire
space rays, the exit pupil specifies the angular bundle system, the exit pupil is conjugate to the entrance
size ' of the image space rays. pupil. For example, if the entrance pupil happens to

aperture
a)

FIGURE 14.7. a. System with a vir-


tual image as the exit pupil, b. System
with a real image as the exit pupil.
274 Geometrie, Physical, and Visual Optics

FIGURE 14.8. Schematic (not to scale) representa-


tion of the eye's cornea (C), crystalline lens (L), retina
(R), primary focal point (F,), iris (I), entrance pupil
(N), and exit pupil (X). The blur circle size (B) equals
the exit pupil size.

coincide with the primary principal plane H l 5 the exit the image of the cornea formed by the crystalline lens.
pupil is the same size as the entrance pupil (m = +1) Of these, the pupil's image would appear smallest, so
and coincides with the secondary principal plane H 2 . it is the exit pupil of the eye. Since the size of an axial
The chief ray is the center of any bundle of paraxial blur circle is defined by the exit pupil, a penlight held
rays coming from an off-axis object point. Thus, the at the primary focal point of your eye (about 1.67 cm
center of a blur circle resulting from such a bundle is in front of your cornea) results in a blur circle size on
specified by the chief ray. The size of the blur circle is your retina that is equal to the exit pupil diameter
specified by the size of the exit pupil (Figure 14.7a or (Figure 14.8).
14.7b). The system in Figure 14.9 consists of an aperture,
Consider the human eye again. A tiny organism which is the aperture stop, followed by a convex lens.
sitting on the retina and looking back toward the The system has a real object and a conjugate real
cornea would theoretically see the crystalline lens, the image. Since the screen is placed closer to the lens
image of the pupil formed by the crystalline lens, and than the conjugate image position, the image on the

FIGURE 14.9. Chief ray and blurred


image size.
Spectacle Magnification and Relative Spectacle Magnification 275

>+

object

FIGURE 14.10. The light from lens B to Q consists of a converging beam made up of diverging
bundles. Q is the exit pupil.

screen is blurred. When the diameter of the aperture passed. Despite this convergence, no image of the
stop is decreased (bottom figure), the blur circles system's object (the arrow) appears. The light be-
decrease in size and the image on the screen gets tween lens B and Q is an example of a converging
clearer (pinhole effect). If we neglect diffraction, we beam made up of diverging bundles, the fourth logical
could theoretically narrow the aperture until only the possibility mentioned in Section 2.8.
chief rays are left. The image on the screen would
then be clear, and the image size would be determined
by the chief rays. When the aperture stop diameter is
increased, the image of the square blurs (top view). 14.3 Thin Lens Spectacle Magnification
We want to distinguish between the size of the
blurred image and the amount of blur. An elementary Figure 14.11a shows a reduced Gauss system repres-
way to specify the amount of blur is in terms of the entation of an ametropic human eye viewing a distant
diameters of the individual blur circles. We specify the object. The chief ray determines the size Ie of the
size of the blurred image by the relative locations of blurred image. The spectacle magnification, Mspec is
the chief rays in the image plane. defined as the ratio of the retinal image size when the
Figure 14.10 shows a two-lens system with an aper- person is wearing a spectacle lens to the retinal image
ture in front of the lenses. The aperture is the aperture size when the person is not wearing a lens.
stop of the system. Two different pencils of rays are Figure 14.11b shows the eye corrected by a thin
shown, from the top and bottom object points, respec- spectacle lens of power P located a distance d in front
tively. Lens A forms a real image of the object. The of the eye's entrance pupil. The corrected retinal
real image serves as the object for lens B. The pencils image size is I m . So
leaving lens B are diverging, so lens B would have a
virtual image (which is not shown, but which you can *spec T (14.1)
locate by tracing the exiting rays in each pencil back-
wards until they intersect). By following rays from the In Figure 14.11a and 14.11b, the chief rays subtend
aperture to position Q, we see that the aperture is respective angles qe and qm at the center of the exit
imaged by lenses A and B at Q. So the aperture's pupils. The distance between the exit pupil and the
image at Q is the exit pupil. (Note that the aperture's retina is w; thus, w times the tangent of the subtended
image is not conjugate to the arrow object.) When a angle is equal to the image size. From Eq. 14.1, it
screen is placed behind lens B and then moved back, follows that
the illumination on the screen appears to converge _ w tan qff
M
until position Q is reached and then diverges once Q is w tan qe
276 Geometrie, Physical, and Visual Optics

FIGURE 14.11. Spectacle


magnification, a. Blurred image
size. b. Spectacle corrected reti-
nal image size.

For paraxial rays,

M
1V1
= \ 0 s22 /
spec ' '
He where Pe is the equivalent power of the system. The
We want to relate the above ratio of image space chief ray is defined in terms of the centers of the
angles at the exit pupil to a ratio of object space angles entrance and exit pupils. So in terms of the entrance
at the entrance pupil. In Section 12.7, we saw that a pupil-exit pupil system matrix, the chief ray object
front to back vertex system matrix could be expanded space angles qm and qe are related to the respective
into an object-image matrix, or into any other con- image space angles by
venient system matrix. Here it is convenient to use a
system matrix that goes from the entrance pupil (in
object space) to the exit pupil (in image space). Since
for the whole system, the entrance pupil and exit pupil and
are conjugate to each other, the entrance to exit pupil
system matrix has the general form of an object-image
matrix, or
Spectacle Magnification and Relative Spectacle Magnification 277

From the matrix multiplications, Equation 14.3 is the spectacle magnification equa-
q'm =
Snq m , tion for a distant object and a thin spectacle lens of
power P. The distance d is the distance from the
and spectacle lens to the entrance pupil of the eye.
For an emmtrope wearing thin zero power (i.e.,
ql = Snq e . piano) spectacle lenses, such as sunglasses, Mspec is
It follows that equal to one. This means that the emmetrope's retinal
image size is the same with or without the thin piano
qin _ qn (14.2)
M.. lenses.
qe qe Frequently, we talk about the percent gain Gspec
when putting the spectacle lens on. Since for no
Equation 14.2 shows that in the reduced system the
change, Mspec is equal to one, the percent gain is
spectacle magnification is equal to the ratio of the
defined as
object space angles subtended by the chief rays at the
center of the eye's entrance pupil. G spec = (M spec - 1)100%. (14.4)
In Figure 14.11b, the ray angle at the center of the
eye's entrance pupil is qm. This particular ray was
incident on the spectacle lens with an angle of qe at a EXAMPLE 14.3a
distance of h above the axis. From the refraction and Veronica Vein's distance Rx is a +6.00 D spectacle
translation matrices, the ray parameters of the ray lens at a 13 mm vertex distance. What is the differ-
incident on the lens are related to those at the en- ence between Veronica's uncorrected retinal image
trance pupil plane by size and her retinal image when wearing the specta-
cle lens?
qn\ /q e By definition, the difference is the spectacle
UTR magnification. The eye's entrance pupil is typically
0/ \h 3 mm behind the cornea, and then
l o\/i -p\/qe d = 13 mm + 3 mm = 16 mm.
From Eq. 14.3,
d l/\0
1
qm\ /i -P M pec
1 - (0.016 m)(+6.00D)'
0/ Id l-dP/\h
M
The matrix multiplications give -c=i4 = 1 1 0 6
Veronica's retinal image size increases by a factor
'q m \ / qe-hP of 1.106 when she puts her spectacle lenses on. The
percent gain is
0 / \dq e + ( l - d P ) h y
Gspec = (1.106 - 1)100% = +10.6%.
From the bottom equality,
-dqe EXAMPLE 14.3b
h=
1-dP' What is Veronica's spectacle magnification when
she switches to a thin contact correction?
Now substitute the above expression into the top Veronica's distance contact correction is +6.51
matrix equality to find D. The contact sits on the cornea, so d is the 3 mm
dPqe distance between the cornea and the entrance pupil.
q- = q. + y 3 d p .
1 = 1.020,
^ _qe-dPqe+dPqe M, pec
l-(0.003)(+6.51D)
1-dP
and
or
qe Gspec = ( 1 0 2 0 - 1)100% = 2.0%.
Mm ! _ dp Veronica has almost an 11% gain in retinal image
Then from Eq. 14.2, size with the spectacles vs a 2% gain with the
contacts. In others words, Veronica's retinal image
size with the spectacles is about 9% larger than
M .pec
_ = (14.3) with the contacts.
1 - dP '
278 Geometrie, Physical, and Visual Optics

EXAMPLE 14.4a The smaller x is, the less important are the higher
Herman Turnip's distance Rx is a -4.00 D specta- order terms (e.g., when x is 0.1, x 2 is 0.01; when x is
cle lens at a 12 mm vertex distance. What is the 0.01, x 2 is 0.0001). Frequently, we can neglect the
difference between Herman's uncorrected retinal higher order terms, in which case
image size and his retinal image when wearing the
spectacle lens? y ~ l + x.
Again, the difference is the spectacle magnifica-
tion. The eye's entrance pupil is typically 3 mm In the spectacle magnification equation (Eq. 14.3), d
behind the cornea, and is in meters and dP is frequently much smaller than
one. The approximation then yields
d = 12 mm + 3 mm = 15 mm.
From Eq. 14.3, M s p e c ~ l + dP. (14.5)

1 From Eq. 14.4,


M s _ec =
P 1 - (0.015 m ) ( - 4 . 0 0 D ) ' -(dP)100%. (14.6)
1 The approximation (Eq. 14.6) is most accurate for
Mspec = 0.943.
1.060 small values of G s p e c and gets worse as G s p e c increases.
Herman's retinal image size decreases by a factor The approximate equations clearly show the direct
of 0.943 when he puts his spectacle lenses on. The dependence of spectacle magnification on the power P
percent gain is of the lens and the distance d between the lens and the
G spec = (0.943 - 1)100% = - 5 . 7 % gain, entrance pupil. For a plus lens, the percent gain
increases with increasing power P of the lens (vertex
or
G
distance fixed) or with increasing vertex distance (P
s P ec = + 5 7 % ^SS- fixed). Of course, for a clear retinal image the image
Note that a negative gain is equal to a positive loss. of the lens must be at the far point of the eye. For a
minus lens, the percent loss gets larger with increasing
power (same vertex distance) or with increasing vertex
EXAMPLE 14.4b distance (P fixed).
What is Herman's spectacle magnification when he
The exact spectacle magnification, Eqs. 14.3 and
switches to a thin contact correction?
Herman's distance contact correction is - 3 . 8 2 14.4, is not very different from 1, and sometimes
D. The contact sits on the cornea, so d is the 3 mm people round it off to 1, in which case they have
distance between the cornea and the entrance pupil. discarded all the percent gain information. When the
approximation (Eq. 14.6) is used, the important infor-
M._ = -^ *, . n. ^ = 0.989, mation (the percent difference from 1) is directly
1 - (0.003)(-3.81 D)
available. In that sense, the use of the approximation
and (Eq. 14.6) sometimes actually improves the accuracy.
G spec = (0.989 - 1)100% = - 1 . 1 3 % gain,
G , p e c = + 1 . 1 3 % loss. EXAMPLE 14.5
Herman has about a 1% loss in retinal image size Rework Examples 14.4a and 14.4b using the ap-
with the contacts vs about a 6% loss with the proximate equation (Eq. 14.6).
spectacles. In others words, Herman's retinal image In Example 14.4a, Herman wore a thin -4.00 D
size with the contacts is about 5% larger than with spectacle lens at a distance of 15 mm from the
the spectacles. This is a noticeable difference, and entrance pupil of the eye. From Eq. 14.6,
is one of the advantages of a contact lens correction G
sPec = ( 1 5 m)(-4.00 D)100% = - 6 . 0 % ,
for a myope.
as opposed to the - 5 . 7 % exact answer.
In Example 14.4b, Herman wore a thin -3.82 D
contact lens correction. The distance d is the 3 mm
distance from the contact lens to the entrance pupil
14.4 Spectacle Magnification Approximation of the eye. From Eq. 14.6,
G spec = (0.003 m)(-3.82D)100% = - 1 . 1 5 % ,
For small x, an equation of the form
as opposed to the - 1 . 1 3 % exact answer.
1
y=
T^'
is approximated by the Taylor series expansion: EXAMPLE 14.6
Rework Examples 14.3a and 14.3b using the ap-
y ~ l + x + x2 + x 3 + . proximate equation (Eq. 14.6).
Spectacle Magnification and Relative Spectacle Magnification 279

In Example 14.3a, Veronica wore a thin the refractive error, the higher the correcting lens
+6.00 D spectacle lens at a distance of 16 mm from power, and the greater the minification. The larger the
the entrance pupil of the eye. From Eq. 14.6, length d, the more the minification. The smaller the
G
sPec = ( 0 0 1 6 m)(+6.00 D)100% = +9.6%, length d, the less the minification.
as opposed to the +10.6% exact answer.
In Example 14.3b, Veronica wore a thin
+6.51 D contact lens correction. The distance d is
the 3 mm distance from the contact lens to the 14.6 Magnification Properties of a Thick
entrance pupil of the eye. From Eq. 14.6, Piano Lens
G
sPec = (0003 m)(+6.51 D)100% = +1.95%, An ophthalmic lens typically has a converging front
as opposed to the +1.99% exact answer. surface and a diverging back surface. An ophthalmic
lens with an equivalent power of zero has a back
vertex power of zero and a front vertex or neutralizing
Examples 14.3 to 14.6 illustrate a general principle power of zero. In other words, the back surface just
already touched on in Chapter 6. For thin lens correc- neutralizes the front surface, and the lens is an afocal
tions, a hyperope has a larger retinal image size when system. Figure 14.12 shows several thick piano lenses,
corrected by spectacles (as opposed to a contact lens and the corresponding reduced systems. The reduced
correction), while a myope has a larger retinal image systems exactly match those of an afocal Galil-
when corrected by contacts as opposed to spectacles. ean telescope.
For the thick piano lens, the spectacle magnifica-
tion M is the ratio of the observer's retinal image size
when viewing a distant object through the lens to the
14.5 Telescope Concepts and Spectacle retinal image size of the person viewing the distant
Magnification object directly. Since the thick piano lens acts like an
afocal telescope, the spectacle magnification is just
The spectacle magnification equation (Eq. 14.3), equal to the afocal telescope magnification given by
Eq. 13.25;

-spec ! _ d P '
has the same form as the telescope magnification
equation (Eq. 13.25). In the telescope magnification where d is equal to the reduced thickness c of the lens,
equation, P is the objective lens power and d is the and Pobj is the front surface power Pj. Since the piano
length of the telescope. lens has zero power, we say that the spectacle magnifi-
For spectacle magnification, we can think of the cation is due to the shape of the lens. We can then
correcting lens as the objective lens of an afocal relabel the above equation as
telescope. The spectacle or contact lens is neutralizing
the refractive error as measured at the entrance pupil
of the eye. So we can think of the ocular lens of the
afocal telescope as the refractive error. The distance The corresponding percent gain is
between the correcting lens and the entrance pupil is G shape = (M s h a p e -1)100%. (14.8)
the length of the telescope.
A hyperope has a minus refractive error that is EXAMPLE 14.7a
corrected by a plus lens. The correcting lens and the An afocal glass (n = 1.5) lens has a 9 mm central
refractive error act like a Galilean telescope. The thickness and a front surface power of +10.00 D.
higher the refractive error (the ocular lens), the higher What is the magnification of the lens?
the correcting lens power (the objective lens), and the The reduced central thickness is
higher the magnification. For a fixed refractive error, 9 mm
c= =6 mm.
the larger the vertex distance, the longer the length d
of the telescope, and the higher the magnification. From Eq. 14.7,
A myope has a plus refractive error that is corrected
by a minus lens. The correcting lens and the refractive lvl
shape _ (0.006 m)( +10.00 D) '
error act like a reverse Galilean telescope (i.e., the x

person is looking backwards through the telescope).


-^3^060-1064.
Therefore, the retinal image is minified. The higher
280 Geometrie, Physical, and Visual Optics

Pv = zero

Pi P2 Pi P2
V
b) v
Pv = zero

FIGURE 14.12. a. Afocal lenses, b. Reduced


V A systems.

From Eq. 14.8, the percent gain is Equation 14.10 can also be written as
G shape = (1.064 - 1)100% = +6.4% gain.
0 ^ = ^,100%, (14.11)
EXAMPLE 14.7b where c is the actual central thickness and n is the
Find the back surface power of the afocal lens and index of refraction. For ophthalmic lenses, the central
use it to verify the above magnification. thickness c expressed in meters is small, and the shape
For plane waves incident on the front,
factor approximations are frequently accurate.
U,=0, and V1 = +10.00 D. Ophthalmic lenses typically have a converging front
The reduced central thickness is 6 mm, so the surface. For a converging front surface, Eq. 14.11
vergence arriving at the back surface is explicitly shows that the shape magnification increases
U 2 = +10.64 D. with increasing front surface power. Ophthalmic len-
ses with higher (plus) front surface power are said to
Since the lens is afocal, plane waves must leave the
be bent more. The higher the bend of the lens, the
back surface, or
greater the shape magnification. The piano lens on the
V2 = 0, right in Figure 14.12a has a higher magnification than
then the piano lens on the left. The shape magnification
P2 = V 2 - U 2 = - 1 0 . 6 4 D. also increases with increasing central thickness and
decreases with increasing refractive index.
The back surface power of the afocal lens is
-10.64 D.
The front surface of the afocal lens serves as the EXAMPLE 14.8
objective lens of the telescope, and the back sur- A piano photochromic (n = 1.523) lens has a
face serves as the ocular lens. From Eq. 13.21, the 2.3 mm central thickness and a +6.00 D front sur-
magnification is face. What is the spectacle magnification percent
gain?
M= P
~ - 1 0 6 4 D =1.064, The reduced central thickness is
obj +10.00 D
1
c= Y ^ = 1.51 mm = 1.51 x 1 0 ' 3 m.
which agrees with the magnification found in Ex-
ample 14.7a. Since c is small, we will use the approximation (Eq.
14.10):
The small magnification approximations to Eqs.
14.7 and 14.9 are G s h a p e = (1.51 x 10" 3 m)(+6.00 D)(100%)
= 0.91% gain.
Mshape = l + c P (14.9)
Since G s h a p e is small, the above answer should be
and very accurate. You may use the exact equations,
G s h a p e = (cP,)100%. (14.10) Eqs. 14.7 and 14.8, to check the results.
Spectacle Magnification and Relative Spectacle Magnification 281

14.7 Shape and Power Factors From the standard V= P + U equation.


PV = P2 + U 2 .
In the real world, all correcting lenses have thickness
and shape. Ametropic correcting lenses also have a Then
non-zero back vertex power Pv. Figure 14.13a shows a
spectacle correction in front of an ametropic eye; p 2 = pv u2,
Figure 14.13b shows the reduced representation of the or
spectacle lens. Pj and P2 are the spectacle's respective
front and back surface powers. p2=pv+pa, (14.12)
When plane waves are incident on the spectacle where
lens, the light is refracted by the front surface P l5
P. = - U , (14.13)
travels across the lens, is refracted by the back surface
P2, and exits the back surface with a vergence equal to It is easier to think in terms of the reduced system, in
the back vertex power Pv. The vergence U 2 incident which the back surface is represented by a thin lens of
on the back of the lens is given by the vergence power P2. Equation 14.12 shows that we can consider
effectivity relation: the thin lens P2 as the sum of thin lenses of respective
Pi powers, Pa and Pv (Figure 14.13c). As shown by Eq.
u,= 1 - cPj * 14.13, the lens Pa just neutralizes the vergence coming

FIGURE 14.13. Spectacle magnification shape


and power factors, a. Actual parameters, b. Re-
duced representation, c. P2 represented as the sum
of Pa and Pv, where Pa neutralizes Vx and Pv is the
back vertex power.
282 Geometrie, Physical, and Visual Optics

from the front surface power 1# Therefore, Pj and P a M tot = (1.143)(1.222) = 1.397,
are an afocal or telescopic system that has a magnifica-
tion given by the shape factor, Eq. 14.7.
G t o t = (1.397 - 1)100% = +39.7% gain.
The remaining thin lens of power Pv acts like a thin
lens correction, for which the spectacle magnification
EXAMPLE 14.10
is given by Eq. 14.3. Since the latter contribution
Victor Vegetable wears a -0.50 D ophthalmic
depends on the back vertex power, it is called the crown (n = 1.523) lens at an 11 mm vertex distance.
power factor. We can then relabel Eq. 14.3: The front surface power of the lens is +7.00 D and
1 the central thickness is 3 mm. What is the total
M (14.14) spectacle magnification?
1 d P '
u r
ben v Assume a 3 mm distance between the cornea
where d ben is the distance from the back vertex of the and the entrance pupil. Then
spectacle lens to the entrance pupil of the eye. d ben = 11 mm + 3 mm = 14 mm.
Figure 14.13c shows the ray angles for the reduced
The -0.50 D specified in the Rx is the back vertex
system. The incident rays subtend an angle q e . The power, so
shape factor Galilean telescope then magnifies the ray
angles, and the rays leaving P a make an angle q a where M = = 0 993
power
=
1 - (0.014 m)(-0.50 D)
qa Mshapeqe
The reduced central thickness is
The lens Pv magnifies the ray angles so that at the 3.3 mm
entrance pupil the chief ray subtends an angle q m , and c = 1.523 = 2.17 mm,
iV1 and the shape factor is
4m power4a*
1
The above two equations give M, = 1.015.
shape _ (0.00217 m)(+7.00 D)
=
4m MpowerMshape4e From Eq. 14.15,
Then from Eq. 14.2, the total spectacle magnification M tot = (1.015)(0.993) = 1.008,
is the product of the power and shape factors, or
or
^tot ^power^shape (14.15) G t o t = (1-008 - 1)100% = +0.8% gain.
The total percent gain is We expect a myope to have a percent loss from the
power factor, while the shape factor supplies a
G t o t = (M tl 1)100%. (14.16)
percent gain. Since Victor is such a low myope, the
power factor was overidden by the shape factor.
EXAMPLE 14.9
George Wholecomb, an aphake, wears a +14.00 D
CR-39 lens (n = 1.495) at a 10 mm vertex distance.
The lens has a +17.00 D front surface and an
11 mm central thickness. What is George's total 14.8 Shape and Power Factor Approximations
spectacle magnification?
Assume a 3 mm distance between the cornea When M s h a p e and M p o w e r are close to 1, then we can
and the entrance pupil. Then make the small magnification approximations:
d ben = 10 mm + 3 mm = 13 mm. Mshape-1+cPj,
The +14.00 D specified in the Rx is the back vertex and
power, so
MpOWer 1 "" dben"v
1
Mpower _ ^ 013 m ^ +14 00 D ) = 1.222. The individual percent gain equations become
The reduced central thickness is Gshape~cP,100%,
11 mm : Gpower~dbenPv100%.
7.36 mm,
1.495
The total spectacle magnification becomes
and the shape factor is
1
M.o.-a+cupja+cp,),
M., = 1.143.
-shape _ (0.00736 m)( +17.00 D) or
From Eq. 14.15, M,ot~l+dbenPv+cPI+dbeilcP1Pv.
Spectacle Magnification and Relative Spectacle Magnification 283

The approximations depended on both d b e n P v and cPj Consider an eye with an uncorrected retinal image
being small. Therefore, the last term, which is the size I e . Assume that one correcting lens results in a
product of the two small terms, is negligible. So clear retinal image of size \x and that a second correct-
ing lens results in a clear retinal image of size I 2 . The
comparison between the retinal image sizes of the
and same eye provided by the two lenses is quantified as
G t o t ~ ( d b e B P v + cP,)100%, the ratio I 2 to lx:

or M.
^tot ^power ^shape ' (14.17)
Then
Equation 14.17 shows that in the small magnification
approximation the individual percent gains are addi- M. Vi. Mo
"P Ij/Ie M/
tive. Note that when the small magnification approxi-
mation is not applicable, the individual percent gains or
are not additive.
M comp = ^ (14.18)
M, '
EXAMPLE 14.11
Dr. Michael Leaking wears a -3.00 D ophthalmic where M 2 and M1 are the respective total spectacle
crown spectacle lens at a 12 mm vertex distance. magifications for the two correcting lenses. The per-
The central thickness is 2.3 mm, and the front cent comparison between the spectacle magnifications
surface power is +4.00 D. What is the total specta- for the two lenses is
cle magnification percent gain (loss)?
Gcomp = ( M c o m p - 1 ) 1 0 0 % . (14.19)
The shape factor should be small, and the ap-
proximation (Eq. 14.11) gives Equations 14.18 and 14.19 are the exact equations for
2.3x10" the comparison. When the small magnification approx-
shape
(+4.00D)100% imations are accurate, we can make a derivation simi-
1.523
lar to that of Eq. 14.17 to obtain
~ +0.60% gain.
G
comp ~ G
2 G,. (14.20)
The back vertex of the lens is a distance of 15 mm
from the entrance pupil (i.e., 3 + 12). The power Within the accuracy of the small magnification approx-
factor approximation gives imation, the percent comparison between the two
G p o w e r ~ (15 x 10" 3 )(-3.00 D)(100%) lenses is just the difference between the two individual
4.5% gain. percent gains.
From Eq. 14.17,
EXAMPLE 14.12
G tot ~ +0.60% + (-4.5%)- -3.9% gain, Linda Swift-Spider wears a left -6.00 D highlite
(n = 1.701) spectacle lens at a vertex distance of
or
11 mm. The lens has a 2.3 mm central thickness and
+3.9% loss. a +2.00 D front surface. Linda has an aniseikonic
problem. With her spectacle lens on, she needs a
3% increase in her left retinal image. What specta-
cle lens at the same vertex distance will supply the
14.9 Comparison of Different Lenses for extra 3 % magnification?
the Same Eye Since the vertex distance is to remain fixed, the
back vertex power must remain fixed. Therefore,
the spectacle magnification power factor is fixed,
If the right retinal image differs in size from the left and we have only the shape factor to work with.
retinal image by a few percent (especially 5 % or Since the shape factors are small, let us use the
more), the human visual system has trouble fusing the approximation (Eq. 14.17).
two images. The problem of unequal retinal images is Linda's current lens has a percent gain given by
refered to as aniseikonia (an = not; isos = equal;
= 2.3xl0~3m
eikon = image [Greek]). (The neural part of the J
shape- (+2.00D)(100%)
1 7 0 1
human visual system can also be a source of
aniseikonia). The spectacle magnification shape and = +0.27%.
power factors are the tools for solving aniseikonic From Eq. 14.20, Linda's new lens must then have a
problems. The resulting lens is called an iseikonic lens. shape factor of 3.27%. Many lenses will work. Let
284 Geometrie, Physical, and Visual Optics

us select an ophthalmic crown lens (n = 1.523) with In the preceding discussions, spectacle magnifica-
a central thickness of 4.3 mm. Then tion has been formulated in terms of the chief ray that
4.3xl0"3m is used to specify the size of blurred images. The chief
G shape = 3.27% = ^(100%), ray passes through the center of the entrance pupil, so
1.523
the distance d ben to the entrance pupil appears in the
equations.
P1 = + 1 1 . 5 8 D. There are other formulations of spectacle magnifi-
cation in which other distances are used. With a
EXAMPLE 14.13 simple thin lens and screen eye model, the distance to
A +8.00 D ophthalmic crown (n = 1.523) lens has a the thin lens representing the eye is used. Sometimes
12 mm vertex distance, a +14.00 D front surface the vertex distance to the cornea is used. In comparing
power, and a 10 mm central thickness. A different clear images, an argument can be made for using the
correction is obtained for the same person. The distance to the eye's primary nodal point. In the case
new correction is ophthalmic crown, has an 8 mm of an eye that is turning to see a different object, the
vertex distance, a +9.00 D front surface power,
center of rotation of the eye is used (dynamic or
and a 6 mm central thickness. What is the change in
the spectacle magnification percent gain? moving eyes are treated in Chapter 18).
Let us use the exact equations. For the initial Each of the different distances results in a different
lens: absolute number for the spectacle magnification. How-
10 mm ever, when comparing the spectacle magnification of
one lens with the corresponding spectacle magnifica-
1.523 = 6.57 mm,
tion of another, the ratios of the spectacle magnifica-
1
M shape tions are always the same.
1 - (0.00657 m)( +14.00 D) For example, consider a thin +8.00 D spectacle
= 1.101. correction at a vertex distance of 14 mm. When refer-
enced to the entrance pupil, the spectacle magnifica-
The entrance pupil is 3 mm behind the cornea, so
tion is 1.157, but when referenced to the cornea, the
1 spectacle magnification is 1.126. Another correcting
M = -
power i - ( 0 . 0 1 5 m)(+8.00D) = 1.136, lens for the same eye is a thin +8.47 D lens at 7 mm
M 1 = (1.101)(1.136) = 1.251. vertex distance. When referenced to the entrance
pupil, the spectacle magnification for the +8.47 D lens
For the second lens, is 1.093; when referenced to the cornea, the spectacle
6 mm magnification is 1.063. The comparisons are:
= 3.94 mm,
1.523 referenced to the entrance pupil,
1
M = 1.037. 1.157
shape _ (0.00394 m)(+9.00 D) M com = 1.028;
P 1.126
At an 8 mm vertex distance, the vertex power is referenced to the cornea,
+8.26 D. The entrance pupil is 3 mm behind the
cornea, so 1.093
=
M,comp = 1.028.
1.063
Mpower =
1 - ( 0 . 0 1 1 m)(+8.26D) = 1 100,
" The point is that whichever set of absolute specta-
cle magnification is used, the lens to lens comparison
M 2 = (1.037)( 1.100) = 1.140.
(1.028 above) is the same. For a static eye, the use of
Then from Eq. 14.18, the entrance pupil is conceptually less restrictive be-
1.140 cause the retinal image does not have to be sharply
M, omp 1 > 2 51
= 0.912, focused.
and
G comp = (0.912 - 1)100% = - 8 . 8 % gain
= +8.8% loss. 14.10 Gauss Formulation for Spectacle
According to the exact equations, the second lens Magnification
provides 8.8% less spectacle magnification than the
first lens. The shape and power small magnification As opposed to using the shape and power factors, we
approximations together with Eq. 14.20 give 8.6% can make an alternate formulation of spectacle magni-
as the answer. fication in terms of the equivalent power and the
Spectacle Magnification and Relative Spectacle Magnification 285

principal planes. We can then treat the thick lens Since V2 must equal +14.00 D,
exactly like the thin lens. P2 = V2 - U2 = +14.00 D - 19.43 D = -5.43 D.
Just as in Section 14.3, the spectacle magnification
is defined as the ratio of the retinal image size with a The equivalent power is given by
p
lens in front of the eye to the retinal image size = Pi + P2 - cPjP, = +12.25 D.
e
without the lens. Figure 14.14 shows the principal From the back surface, the secondary principal
planes for a meniscus-convex correcting lens. The plane location is
distant object subtends an angle q e . The chief ray
subtends an angle qm at the eye's entrance pupil. L2H2 r
b h p p ,
The derivation exactly parallels the derivation of
the spectacle magnification equation for a thin lens
L2H2 = +71.42 mm - 81.63 mm = -10.21 mm.
(Eq. 14.3). The secondary principal plane H 2 is consi-
dered to be the back surface of the equivalent lens of Then
power P e . This distance dHen is the distance from H 2 to
the eye's entrance pupil. Then the total spectacle dHen = 10.21 mm + 10 mm + 3 mm = 23.21 mm.
magnification for the thick lens is From Eq. 14.21,
1
M (14.21) M = 1.397,
- = l - d (0.02321 m)( +12.25 D)
t o t = r

Equation 14.21 gives the same numerical results as the which agrees with the shape and power factor
shape and power factor equations (Eqs. 14.7, 14.14, answer obtained in Example 14.9.
and 14.15). In fact, they can be directly derived from
each other. When a plus spectacle lens is bent more, the spec-
tacle magnification increases. We explained this previ-
EXAMPLE 14.14 ously by noting that the shape factor increases as the
In Example 14.9, George Wholecomb wore a lens is bent. How can the increase be explained in the
+ 14.00 D CR-39 (n = 1.495) lens at a 10 mm vertex Gaussian formulation?
distance. The lens had a +17.00 D front surface Figure 14.15 shows a series of plus lenses with the
and an 11 mm central thickness. Use Eq. 14.21 to same back vertex power. The principal planes are
find the total spectacle magnification of the thick symmetrically located inside the equiconvex lens. For
lens. the convex-piano lens, the primary principal plane has
First, we need to find the equivalent power and moved to the front surface (which is the only surface
the location of the secondary principal plane. To
find the equivalent power we need to find the back with a non-zero dioptric power), and the secondary
surface power. The back vertex power of the lens is principal plane has also moved toward the front sur-
+ 14.00 D. So for plane waves incident on the front, face. For the meniscus-convex lens, the principal
the vergence V2 leaving the back surface is planes continue to move in the same direction, and
+14.00 D. Therefore, when the lens is bent enough, both of the principal
1^=0, planes are in front of the lens. The fact that H 2 is
moving away from the spectacle's secondary focal
Vj = +17.00 D, point (and away from the entrance pupil of the eye)
and explains the increase in magnification. (This same
U 2 =+19.43 D. general principle applies to a telephoto lens.)

FIGURE 14.14. Spectacle magnification in


terms of principal planes.
286 Geometrie, Physical, and Visual Optics

a)

Hi H2 Hi H 2
Hi H2

d)

Hi H2 FIGURE 14.15. Principal planes as plus lens is bent.

Figure 14.16 shows a series of minus lenses with the to the entrance pupil. As indicated in Eq. 14.15, the
same back vertex power. The principal planes are power factor must be multiplied by the shape factor to
symmetrically located inside the equiconcave lens. For obtain the total spectacle magnification. On the other
the planoconcave lens, the secondary principal plane hand, the Gauss equation, (Eq. 14.21) contains the
has moved to the back surface and the primary princi- equivalent power and the distance dHen from the sec-
pal plane has also moved toward the back surface. For ondary principal plane to the entrance pupil. The
the meniscus-concave lens, the principal planes con- shape and power information are encoded in the
tinue to move in the same direction, and when the lens equivalent power and in the principal plane location.
is bent enough, both of the principal planes are in When a corrected curve spectacle lens is properly
back of the lens. The fact that H 2 is again moving ordered, the refractive index, the back vertex power,
away from the spectacle's secondary focal point ex- front surface power, and the central thickness are all
plains the increase in magnification. specified. Hence, the shape and power factor ap-
It is important to note that the power factor equa- proach to spectacle magnification is easy to use. The
tion (Eq. 14.14) contains the back vertex power Pv equivalent power and the location of the secondary
and the distance d. from the back vertex of the lens principal plane are not specified on an order. There-

to)
V c)

Hi H2 Hi H 2 Hi H2

d)

FIGURE 14.16. Principal planes as minus lens is


Hi H2 bent.
Spectacle Magnification and Relative Spectacle Magnification 287

fore, they must be calculated before the Gaussian shape factor. Pv and the eye's refractive error consti-
equation for spectacle magnification can be used. tute an afocal telescope that is analogous to the power
factor. So the magnification Madj of the adjusted
telescope is

14.11 Telescopes for Ametropes Madj M.shapeM.power> (14.22)


or
Consider an afocal telescope. For incident plane
waves, plane waves leave the telescope. Suppose an M adj (14.23)
(l-cPobj) (l-dbenPv)'
uncorrected ametrope wants to use the telescope. The
ametrope can adjust the length of the telescope so that where c is the reduced length of the adjusted tele-
the vergence leaving the telescope corrects the refrac- scope, Pobj is the objective lens power, dben is the
tive error. distance from the ocular lens to the entrance pupil of
the eye, and Pv is the back vertex power of the
EXAMPLE 14.15 adjusted telescope.
A 5 X Kepler telescope consists of a +2.00 D lens
located 60 cm in front of a +10.00 D lens. J. Falk
Cerebrii needs a +7.00 D spectacle correction at a EXAMPLE 14.16
12 mm vertex distance. Assume that the telescope's For J. Falk Cerebrii's telescope in Example 14.15,
ocular lens sits 12 mm from the eye. What length the magnification is
should the telescope have for J. Falk to clearly see
a distant object through it without accommodating? Madj =
1-(0.8333 m)(+2.00D)
The light leaving the +10.00 D ocular lens must
have a vergence of +7.00 D, so the vergence inci- 1
dent on the +10.00 D must have a vergence of 1-(0.015 m)(+7.00D)'
-3.00 D, or the ocular lens must be 33.33 cm away
from the internal image. For plane waves incident Ma d j (-1.50)(1.117) = -1.676.
on the +2.00 D objective lens, the internal image is The adjusted telescope provides a retinal image
50 cm from the objective. The new length of the that is inverted relative to the uncorrected retinal
telescope is then 83.33 cm. image and 1.676 times larger than the uncorrected
retinal image.
Once the length has been adjusted, what is the new Actually, the shape factor, -1.50, is the com-
magnification? First, we need to be certain about what parison between the retinal image size Im obtained
we mean by magnification. Angular magnification is a with the adjusted telescope and the retinal image
comparison of retinal image sizes between two differ- size I a obtained with a thin +7.00 D correcting lens
ent situations. One possible situation would be to at a 12 mm vertex distance.
compare the retinal image size Im that the observer
has with the adjusted telescope to the retinal image EXAMPLE 14.17
Racquel Mulch needs a -4.00 D spectacle correc-
size Ia that the observer has when wearing their tion at an 11 mm vertex distance. Racquel takes an
correction. A second possible comparison is between afocal 5 X Galilean telescope with a +2.00 D objec-
Im and the observer's uncorrected retinal image size tive lens, a -10.00 D ocular, and a 40 cm length,
I e . Let us use the second possibility so that the and adjusts it so that the vergence leaving the
magnification M of the adjusted telescope is given by telescope is -4.00 D. What magnification does the
adjusted telescope provide when the ocular lens is
M= 11 mm from the cornea?
For plane waves incident on the adjusted tele-
We can analyze the adjusted telescope in the same scope, a -4.00 D vergence must leave the ocular
lens. The ocular lens is a -10.00 D lens, so the
way that we analyzed the shape and power factors for vergence incident on it must be +6.00 D. The
a thick lens. In particular, we can consider the ocular length of the adjusted telescope is then 50.00 cm
lens Poc on the adjusted telescope to be the sum of two minus 16.67 cm, or 33.33 cm. The magnification is
thin lenses:
p1 = pL + Ap Mad =
oc a ' v? J 1-(0.3333 m)(+2.00D)
where Pa is a lens that neutralizes the objective and Pv 1
is a lens that is equal to the back vertex power of the
adjusted telescope. Then the objective lens and Pa 1-(0.014 m)(-4.00D)'
constitute an afocal telescope that is analogous to the Madj = (+3.00X0.947) +2.84 X.
288 Geometrie, Physical, and Visual Optics

Racquel's retinal image size with the magnifier is are blurred and their size is specified by the chief ray.
erect relative to the uncorrected retinal image and The blurred ametropic retinal image sizes are all equal
is 2.84 times larger. The +3.00 X shape fact,A to I s . For the ametropes, the correcting lenses clear
indicates that Racquel's retinal image with the ad- the retinal images and change their size. From the
justed telescope is 3 times larger than Racquel's spectacle magnification equation (Eq. 14.3), we know
retinal image with a thin -4.00 D correction worn that a thin myopic correction decreases the retinal
at an 11 mm vertex distance.
image size and a thin hyperopic correction increases
the retinal image size.
The above two examples indicate that a Kepler
telescope loses magnification when lengthened and Figure 14.17b shows the chief ray for an un-
a Galilean telescope loses magnification when shor- corrected axial hyperope, an uncorrected axial myope,
tened. We gain magnification when a Kepler tele- and the standard emmtrope. The chief ray specifies
scope is shortened and when a Galilean telescope is the blurred image sizes for the ametropes. The clear
lengthened. Note that since an adjusted telescope is retinal image size for the +60.00 D standard emmet-
no longer an afocal system, we could also use the rope is I s . The uncorrected retinal image size for the
Gaussian formulation, Eq. 14.21, of spectacle mag- axial myope is larger than I s , which in turn is larger
nification to find the results. (For the two previous than the retinal image size for the axial hyperope.
examples H 2 is behind the eye, and d Hen is From the spectacle magnification relations, we know
negative.) that a thin correcting lens decreases a myope's retinal
image size. However, the fact that the retinal image
size for the uncorrected axial myope is larger than Is
raises the possibility that the retinal image size for the
14.12 Relative Spectacle Magnification corrected axial myope might be larger than I s . Similar-
ly, the fact that the retinal image size for the un-
Spectacle magnification dealt with a comparison of the corrected hyperope is smaller than Is raises the pos-
corrected retinal image size to the uncorrected retinal sibility that the retinal image size for the corrected
image size for the same eye. For aniseikonia of optical axial hyperope might still be smaller than I s .
etiology, we want to compare the image size of the Relative spectacle magnification (RSM) is the ratio
right eye to the image size of the left eye. Two of the retinal image size Ia in the corrected ametropic
different eyes are then involved. The right eye may or eye to the retinal image size I s in the standard emmet-
may not have the same axial length as the left eye, and ropic eye, or
the right eye may or may not have the same refractive
error as the left eye. RSM=^. (14.24)
For instructional purposes, it is easier to establish a
standard eye and to compare each eye (right and left) In Eq. 14.24 both I a and Is are clear retinal images and
to the standard eye, just as it is easier to establish a the nodal points of the eye can be used to specify the
standard distance, such as the meter, and to compare subtended angles. In a reduced system, the nodal
individual distances to the standard (e.g., distance A points are located at the principal planes. Remember
equals 13 m, distance B equals 26 m). Let us use a also that the image sizes in a reduced system are the
+60.00 D emmetropic eye as the standard eye. same as those in an actual system.
Individual eyes can differ from the standard eye in In the reduced system, the standard +60.00 D em-
axial length, power of the refracting elements, or metropic eye has a 16.67 mm distance between the
both. It is relatively easy to treat differences in the secondary principal plane and the retina (i.e., the
refractive elements when the axial length is the same secondary equivalent focal length is 16.67 mm). A
as the standard eye (i.e., a refractive ametrope). It is refractive ametrope has the same 16.67 mm distance
also relatively easy to treat differences in axial length between the eye's secondary principal plane and the
when the refractive elements are the same as the retina, but the equivalent power is not +60.00 D. An
standard eye (i.e., an axial ametrope). It is harder to axial ametropic eye has a +60.00 D equivalent power,
treat the general case in which both the refractive but the distance between the eye's secondary principal
elements and the axial length differ (i.e., a mixed plane and the retina is not 16.67 mm.
ametrope). The discussions here are confined to axial For a refractive ametrope, the fact that the
and refractive ametropes. 16.67 mm length is the same as the standard eye
Figure 14.17a shows the chief ray for a refractive means that the RSM equation can be derived in
hyperope, a refractive myope, and the standard em- exactly the same manner as the spectacle magnifica-
mtrope. The standard emmetrope's clear retinal tion equation (Eq. 14.3). (Especially note that the
image size is I s . The retinal images for the ametropes distance w was the same in the derivation of Eq. 14.2.)
Spectacle Magnification and Relative Spectacle Magnification 289

blurred -
ametropic images

b)

FIGURE 14.17. Ametropic vs emmetropic re-


tinal image sizes, a. Refractive ametropes. b.
Axial ametropes.

It follows that the RSM equation for a refractive In Figure 14.19a, the eye's principal planes are
ametrope with a thin correcting lens of power Pv has dashed, while the principal planes for the correcting
the same form as Eq. 14.3 except that now the dis- lens and eye system are solid. From the nodal ray for
tance d is from the correcting lens to the eye's primary the total system, the retinal image size I a is given by
principal plane Ht. In Figure 14.18, d is labeled as
dbH, and then I. = f.tanq,
RSMr = (14.25) where q is the nodal angle and fa is the secondary
1-<WV equivalent focal length for the corrected ametropic
On the other hand, one cannot proceed the same eye.
way with the RSM relations for the axial ametropes For the standard emmetropic eye viewing the same
because the lengths are different, in which case Eq. object (Figure 14.19b),
14.25 does not apply. Let Ps be the +60.00 D equival-
ent power of the standard emmtrope. Consider the Is = 16.67 mm tan q.
total Gauss system for the corrected axial ametrope. Then from the above two equations and Eq. 14.24,
The thin lens power of the axial ametrope's eye is
equal to Ps. Assume a correcting lens of power Pv RSM = 16.67 m m '
located at a distance dbH from the eye's primary
principal plane. The equivalent power Pa for the cor- or
rection together with the ametropic eye is
RSMa=^. (14.26)
P =P +P d b H P v P.
290 Geometrie, Physical, and Visual Optics

Hi H 2 FIGURE 14.18. Thin lens correction for refrac-



16.67 mm tive myope.

FIGURE 14.19. a. Principal planes


and nodal ray for correction-eye system
of previous figure, b. Nodal ray for stan-
dard emmtrope.

Equation 14.26 shows that RSMa is equal to the ratio 0.01667 m to obtain
of the equivalent dioptric powers for the two systems. 1
(It can also be shown that the relative spectacle RSM =
(0.01667m)Pv + l - d b H P /
magnification for refractive ametropes, RSM r , is equal
to the ratio of equivalent powers.) It follows that or
1
RSlVi (14.27)
RS1V = 1 - (dbH - 0.01667 m)Pv '
P +P AHPVPS Equation 14.27 for an axial ametrope is different
V S

We can divide both the numerator and the denom- than Eq. 14.25 for a refractive ametrope. For thin
inator by Ps and use the fact that the reciprocal of Ps is correcting lenses, Eq. 14.25 shows that for refractive
Spectacle Magnification and Relative Spectacle Magnification 291

myopes I a < I S , while for refractive hyperopes I a > I S . than the standard emmetropic eye, the corrected
For axial ametropes, the situation depends on whether myope has a 3.8% larger retinal image than the
d bH is longer or shorter than the primary equivalent standard emmtrope. Figure 14.21 shows a super-
focal length 16.67 mm. position of the reduced emmetropic eye and the
When the thin correcting lens for an axial ametrope corrected reduced ametropic eye. The solid ray is
the nodal ray for the emmtrope, and the dashed
is placed at the primary focal point of the eye, d bH
nodal ray is for the myopic eye. The two rays come
equals 16.67 mm, and then Eq. 14.27 becomes in parallel to each other. The dashed ray is bent up
RSM a = l. by the minus spectacle lens, so at the primary nodal
point of the eye it subtends a smaller angle than the
This particular phenomenon is known as Knapp's law. solid ray. However, since the axial myope has a
The corresponding ray diagram is shown in Figure longer eye, I a turns out to be larger than I s .
14.20. The nodal ray for the thin correcting lens is also
the ray that passes through the primary focal point of EXAMPLE 14.19
the eye, so the image space ray in the eye is parallel to A refractive myope needs a -10.00 D correction at
the axis, and I a is equal to I s for both the axial myope a distance of 13 mm from the primary principal
and hyperope. plane of the eye. What is the relative spectacle
For either the axial or the refractive case, the per- magnification?
From Eq. 14.25,
cent gain G of the corrected retinal image I a relative
to the standard emmetropic image I s is 1
RSM =
1 - (0.013 m ) ( - 1 0 . 0 0 D ) '
G = (RSM-1)100%. (14.28)
1
RSM r = = 0.885.
EXAMPLE 14.18 1 + 0.130
A myopic axial ametrope needs a -10.00 D thin The percent gain is
spectacle correction at a distance of 13 mm from Hl
of the eye. What is the relative spectacle magnifi- G = (0.885-1)100%,
cation? G = - 1 1 . 5 % gain,
From Eq. 14.27, or
1 G = + 1 1 . 5 % loss.
RSM =
(13 mm - 16.67 mm) While the axial myope in the previous example had
1 (-10.00 D)
. 1000 mm/m a percent gain relative to the standard emmtrope,
1 the refractive myope here has a large percent loss.
RSM = = 1.038.
1 - 0.0367
The percent gain is EXAMPLE 14.20
What is the relative spectacle magnification for the
G = (1.038 - 1)100% = 3.8%. axial myope in Example 14.18 when wearing a
In this case, since the axial myopic eye is longer (thin) contact correction?

16.67 mm 16.67 mm FIGURE 14.20. Knapp's law.


292 Geometrie, Physical, and Visual Optics

2 jS
8 (0
"<D
sJ o]

myope's
CO

retina
a>
E
E
a>

a
* bH ^
/
Hi H2

v FIGURE 14.21. Superposition of standard


16.67 mm emmetropic eye and corrected axial myope.

From Gullstrand's #1 schematic eye, the pri- G = (0.988-1)100%,


mary principal plane sits 1.348 mm behind the cor- G = - 1 . 2 % gain,
nea. From the vergence effectivity relations, the
correction at the cornea is -8.96 D. From Eq. or
14.27, G = + 1 . 2 % loss.
RSlvt The contact corrected refractive myope has only a
small percent loss relative to the standard emmet-
rope, while the spectacle corrected refractive
["(1.348 m m - 1 6 . 6 7 mm) myope has a larger percent loss.
(-8.96 D)J
L 1000 mm/m
1 Table 14.1 summarizes the comparisons of retinal
RS1VI = = 1.159. image sizes relative to the standard emmetropic retinal
1-0.137
image size I s . Small percent differences are labeled as
The percent gain is about the same. The myopic information corresponds
G = (1.159-1)100%, to the above four examples.
Aniseikonia is usually associated with anisomet-
G = +15.9% gain.
ropia, which is the condition of significantly different
The contact corrected axial myope has a large refractive errors in the right eye vs the left eye. When
percent gain. one eye is myopic and the other eye is hyperopic, the
anisometropia is called antimetropia. From spectacle
magnification relations, one might initially think that
EXAMPLE 14.21 the retinal image size in a spectacle corrected myopic
What is the relative spectacle magnification for the
refractive myope of Example 14.19 when wearing a eye would be smaller than the retinal image size in a
(thin) contact lens correction? spectacle corrected hyeropic eye, but the RSM rela-
Again, the primary principal plane is 1.348 mm tions show this may not be true especially if the
behind the cornea, and the power of the correction refractive errors are due to axial causes.
at the cornea is -8.96 D. From Eq. 14.25, In working in the real world, one should keep in
1 mind that the refractive error may not be just axial or
RSMr = just refractive, but might be mixed. Also in the real
1-(0.001348 m)(-8.96D)
world, the correcting lenses are thick. The RSM equa-
1 tions are modified for a thick correcting lens by simply
RSMr = = 0.988.
1 + 0.0121 multiplying the appropriate RSM equation by the
The percent gain is shape factor, Eq. 14.7, of the thick lens.
Spectacle Magnification and Relative Spectacle Magnification 293

TABLE 14.1
Retinal Image Sizes Relative to the Standard Emmtrope

Spectacle Lens
Correction at Contant Lens
Uncorrected Ft of the Eye Correction

Axial myope Larger Same Larger


Axial hyperope Smaller Same Smaller
Refractive hyperope Same Larger About the same
Refractive myope Same Smaller About the same

14.13 Summary of Important Equations Problems

Thin Lens Spectacle Magnification 1. A +5.00 D lens of diameter 4 cm is located 8 cm


in front of an aperture of diameter 2 cm. What is
M = ^ - (14.3) the entrance pupil for a real object located 40 cm
iV1
spec _ d p from the lens?
2. A +9.00 D lens of diameter 4 cm is located 10 cm
Thick Lens Spectacle Magnification in front of an aperture of diameter 2 cm. What is
the entrance pupil for a real object located 50 cm
M =M M (14.15) from the lens?
1YX
tot 1YA
shape 1 Y X power> 3. A +6.50 D lens is located 10 cm in front of a
1 (14.7) 3.50D lens. A small aperture is located midway
M shape 1 _ cPj '
between the lenses. For a distant object, the
1 aperture is the aperture stop. Find the entrance
M power (14.14)
I _ H P ' pupil and the exit pupil.
1 U
benrv
Alternatively, 4. A 9X Keplerian telescope has a +4.00D objec-
tive lens located 27.78 cm in front of a +36.00 D
M
tot 1 _ H P * (14.21) ocular lens. If the objective lens is the aperture
1 U
Henre stop, find the entrance pupil and the exit pupil.
5. An eye has a +44.00 D cornea located 3.7 mm in
Relative Spectacle Magnification: Thin Correction front of a 3 mm pupil. Assume the cornea and
aqueous humor both have a 1.336 refractive
1 index. Locate the entrance pupil and give its size.
RSMr = (14.25)
i-cWV 6. A hyperope has a +5.75 D spectacle correction
worn at a 10 mm vertex distance. What is the
1
RSM = (14.27) spectacle magnification in terms of the percent
1 - (dbH - 0.01667 m)Pv gain or loss? (Assume the entrance pupil is 3 mm
behind the cornea.)
Percent Gain 7. A myope has a 7.25 D spectacle correction worn
at a 12 mm vertex distance. What is the spectacle
G = (M-1)100%. magnification in terms of the percent gain or loss?
(Assume the entrance pupil is 3 mm behind the
cornea.)
Small Magnification Approximation 8. An afocal lens (n = 1.50) has a +12.00 D front
surface, a 5 mm central thickness, and a -12.50 D
1 back surface. Verify that the equivalent and back
M= = l+x,
1-x vertex powers are zero. Find the shape factor of
G = (x)(100%). the spectacle magnification for this lens. Report it
as a percent gain or loss.
For small magnifications, the percent gains are ad- 9. Mary Proudfoot needs a -5.00 D spectacle lens at
ditive. a 10 mm vertex distance. The lens (n = 1.58) used
294 Geometric, Physical, and Visual Optics

has a +4.00 D front surface power and a 4 mm magnification? Compare this result to that of a
central thickness. Give the shape factor, power refractive hyperope with the same correction.
factor, and total spectacle magnification. (Assume 13. Millie Yuppie is an axial myope with a -9.00 D
the entrance pupil is 3 mm behind the cornea.) spectacle correction at a 10 mm vertex distance.
10. Antonio Termite wears a +6.00 D spectacle cor- Millie's entrance pupil is 3 mm behind her cornea,
rection (n = 1.53) with a +12.50 D front surface and her primary principal plane is 1.44 mm behind
power and an 8.0 mm central thickness. For a her cornea. What is the relative spectacle magnifi-
10 mm vertex distance, give the shape factor, cation? Compare this result to that of a refractive
power factor, and total spectacle magnification. myope with the same correction.
(Assume the entrance pupil is 3 mm behind the 14. Consider a thin lens and screen model for an
cornea.) emmtrope eye. The entrance pupil and the prin-
11. Rowena Meadowbrook has a -5.00 D spectacle cipal planes are both located at the thin lens
correction (n = 1.523) for her left eye. The vertex representing the eye. For a distant object the
distance is 10 mm, the front surface power is retinal image is clear without any lenses in front of
+2.00 D, and the central thickness is 2.3 mm. Due the eye. Suppose a -2.00 D lens is placed 14 mm
to an aniseikonia problem, Rowena needs an in- in front of the lens representing the eye. By
crease of 3% in her spectacle magnification for accommodation, the retinal image would again be
this lens. Specify a lens fit at the same vertex clear. How much larger or smaller (percent gain
distance that could provide the additional 3% or percent loss) is the retinal image for the accom-
gain. You may use the small magnification approx- modated state vs the retinal image for the un-
imations. accommodated state. (Hint: You could consider
12. Sam Fox is an axial hyperope with a +8.00 D the accommodation as a refractive error and the
spectacle correction worn at a 10 mm vertex dis- -2.00 D lens as the correction.) This example
tance. Sam's entrance pupil is 3 mm behind his gives an indication of retinal image size changes
cornea, and his primary principal plane is 1.34 mm when a patient is over-minused.
behind the cornea. What is the relative spectacle
CHAPTER FIFTEEN

Astigmatism:
On Axis

15.1 Introduction Figure 15.2 shows a series of lenses made with


cylindrical and/or flat surfaces. The cylindrical lenses
The paraxial optics of spherical systems is built on the each have an axis meridian and a perpendicular power
concept of point objects and conjugate point images. meridian. We identify the axis meridian by the flat
Systems that form point images are sometimes called meridian on the cylindrical surfaces. The power meri-
stigmatic systems. An astigmatic system forms line dian is the meridian in which the cylindrical surfaces
images instead of point images. have circular cross sections.
Astigmatism is quite common in the human eye. In When plane waves are incident on the cylindrical
some groups over 80% of the people have some lens, the paraxial wavefront leaving the surface is
astigmatism. Astigmatism is optically corrected by cylindrical. Converging cylindrical wavefronts form a
cylindrical or spherocylindrical lenses. real line image that coincides with the axis of the
exiting cylindrical wavefronts. In Figure 15.3, the re-
fracting cylinder has a vertical axis meridian and forms
a real vertical line image. Note that all the converg-
ence occurs in the horizontal power meridian, and no
15.2 Cylindrical Lenses convergence occurs in the axis meridian. (A cylindrical
beer or soft drink helps in visualizing the optics of a
Figure 15.1 shows a cylinder. Geometrically, we can cylindrical lens.)
generate a cylinder by taking two parallel lines, and Figure 15.4a shows a planoconvex cylindrical lens
rotating one around the other while maintaining the with a vertical axis meridian and a horizontal power
parallelism. The line that serves as the center of the meridian. For analyzing the ray behavior, the lens is
rotation is called the axis of the cylinder. Each point divided into a series of thin horizontal segments (Fig-
on the cylinder's surface is equidistant from the axis. ure 15.4b). The front surface of each horizontal seg-
A point on the line that is rotating traces out a ment is a cross section of the power meridian and has
circle around the axis. The radius of this circle is the a circular shape. Each segment's back surface is flat.
distance from the point to the axis. A cylinder is flat Figure 15.5a shows incident rays coming from a
along any meridian parallel to the axis, and has a near object point. The thin top segment of the lens
circular cross section along any meridian perpendicu- converges the rays horizontally and forms a point
lar to the axis. For a cylindrical refracting or reflecting image at A. The thin bottom segment converges the
surface, the flat meridian is referred to as the axis rays horizontally and forms a point image at A. Each
meridian, and the perpendicular meridian with the thin segment converges the rays horizontally and
circular cross section is referred to as the power forms its own point image. Collectively, all the in-
meridian. The axis meridian and the power meridian dividual point images make a vertical line image. Even
are the principal meridians of the cylindrical surface. though the real line image is parallel to the axis

295
296 Geometrie, Physical, and Visual Optics

meridian, it is the power meridian that is responsible


for the convergence that forms the line.
Figure 15.5b shows a planoconvex cylindrical lens
with a horizontal axis meridian. The lens is segmented
vertically along the power meridian. The left thin
segment converges the rays vertically and forms a
point image at B. The right thin segment converges
the rays vertically and forms a point image at B'.
FIGURE 15.1. A cylinder with a vertical axis. Again, each thin segment forms its own point image.

Equi- Bi- Piano- Concavo- Equi- Bi- ^ Piano-'


convex convex convex convex concave concave concave FIGURE 15.2. Cylindrical lenses.

vertical line image

cylindrical wave front

cylindrical lens

indicent
plane
wave

FIGURE 15.3. Wavefronts passing through a cylindrical


lens.

Collectively, all the point images form a real horizon-


tal line image.
a) ( ^ When the plus cylindrical lens with a vertical axis is
b)
rotated clockwise toward a horizontal axis, the real
vertical line image rotates clockwise with the lens. In
particular, the real line image always remains parallel
to the axis meridian and perpendicular to the power
meridian.
In Figure 15.4a, the axis meridian has flat surfaces
of zero dioptric power. Since the power meridian of
FIGURE 15.4. a. Power meridian curvature, b. Sections of a the cylindrical front surface has a circular cross sec-
cylindrical lens. tion, its power is given by
Astigmatism: On Axis 297

a) object
point
_ vertical
line image

b) object
point
horizontal
line image

FIGURE 15.5. a. Line image formation by sectionvertical


axis meridian, b. Line image formation by sectionhorizontal
axis meridian.

The incident vergence is -2.00 D and from Eqs.


p.= (15.1) 15.2 and 15.3,
V= +5.00 D + (-2.00 D) = +3.00 D.
where is the radius of curvature of the power
meridian, nx is the refractive index of the medium in The (real) line image position is
front of the surface, and n2 is the refractive index of 100cm/m
the lens. Since the back surface is piano, its power, P 2 , y=
T^D" = +33 3cm
-
is zero. For a thin cylindrical lens, the total power in
the power meridian is Power in the vertical meridian forms horizontal
lines. The horizontal line image can be observed on
P = P,+P2 (15.2) a screen 33.3 cm behind the lens. Note that the real
line image is parallel to the axis meridian.
From the power meridian, the location of the line
image can be found from For a thin cylindrical lens, Figure 15.6a shows the
ray behavior in the axis meridian and Figure 15.6b
V = P + U. (15.3) shows the ray behavior in the power meridian. There
is no convergence in the axis meridian, and the rays go
EXAMPLE 15.1 straight through without bending. The segment shown
A real point source is located 50 cm in front of a by the power meridian converges the rays and forms
planoconvex cylindrical lens with a horizontal axis one point (A) on the line image. For Figure 15.6b, the
meridian. The refractive index of the lens is 1.5, line image itself would be perpendicular to the plane
and the power meridian radius of the cylindrical of the paper.
surface is 10 cm. Where is the real line image? Once the real line image position v is determined
What is its orientation? from the power meridian, the line image length can be
determined from the similar triangles in the axis meri-
In the vertical power meridian, Eq. 15.1 gives dian. Thus, while the axis meridian does not form the
(1.50 - 1.00)100 c m / m = + 5 0 Q D line image, it does determine the length of the line
1
+10 cm image.
298 Geometrie, Physical, and Visual Optics

u
FIGURE 15.6. Cross-sectional ray diagrams.
a. Axis meridian, b. Power meridian.

EXAMPLE 15.2 We can watch the formation of the line image by


A point image is located 25 cm in front of a thin placing a screen directly behind the lens and then
+10.00 D cylindrical lens with a vertical axis meri- moving the screen toward the line image position.
dian. The lens has a round aperture 40 mm in Figure 15.7 shows the cross-sectional illumination pat-
diameter placed against it. What is the location, terns that appear on the screen for the lens in the
length, and orientation of the real line image?
previous example. Initially, the illumination pattern is
The line image is parallel to the axis meridian,
or vertical in this case. The incident vergence U is circular and the same size as the aperture. As the
-4.00 D. In the power meridian, screen is moved back, the illumination pattern gets
longer in the vertical meridian (since the vertical
V = P + U, meridian is the axis meridian and supplies no converg-
V = +10.00 D + (-4.00 D) = +6.00 D, ence), and shorter in the horizontal meridian (since
and the horizontal meridian is the power meridian that
v = +16.67 cm, supplies the convergence). The result on the screen is
a vertically oriented elliptical blur pattern. As the
which is the location of the line image.
screen is moved farther back, the ellipse continues to
The axis meridian (vertical) has a ray diagram
just like Figure 15.6a. In Figure 15.6a, the triangle grow vertically and to shrink horizontally. The vertical
xLL' is similar to the triangle xAA'. Then the line line image occurs at the position where the ellipse has
image length AA' is given by shrunk to a zero horizontal width. As the screen is
moved past the line image position, the light starts to
AA' _ LL'
diverge in the horizontal meridian, and a vertically
\u\ + v ir oriented blur ellipse again appears.
The effective lens diameter is 40 mm or 4 cm. So At any position, the lengths of the major and minor
AA 4 cm axes of the blur ellipse can be determined from the
25 cm + 16.67 cm 25 cm ' similar triangles in Figures 15.6a and 15.6b. The ellip-
and tical blur patterns occur because of the circular aper-
ture. If a square aperture with horizontal and vertical
A A = 6.67 cm. sides is used, the blur patterns would be rectangular
The line image is 6.67 cm in length. (see Figure 15.14).

(circular)
directly
behind ?
line FIGURE 15.7. Image space illumination patterns for a point
lens
focus source.
Astigmatism: On Axis 299

It is sometimes helpful to analyze the optics of a looking at the patient (Figure 15.8). This view is the
cylindrical lens in terms of a power cross. The power so-called contra-ocular view (i.e., against the view of
cross shows the dioptric powers of the lens in the the patient's eye). Each meridian could be identified
principal meridians. The power cross for the +10.00 D by two different angles (e.g., the vertical meridian
cylinder with a vertical axis meridian has zero power could be labeled 90 or 270). We use the smaller of the
vertically and +10.00 D horizontally: two angles except in the case of the horizontal meri-
dian. The horizontal meridian is labeled as the 180
0 meridian instead of the 0 meridian. (Zero is not used
in order to avoid confusion between a cylindrical lens
with a horizontal axis meridian and a spherical lens
that does not have an axis meridian.)
+10.00 D
As seen by the examiner, the first quadrant angles
label the meridians clockwise from the top, and the
second quadrant angles label the meridians counter-
clockwise from the top. Figure 15.9a shows the 30
The power cross for a +7.00 D cylindrical lens with a
degree meridians, while Figure 15.9b shows the 150
horizontal axis meridian has zero power horizontally
degree meridians.
and +7.00 D vertically:
The conventional way of specifying the orientation
+7.00 D of a cylindrical lens is in terms of the axis meridian,
although it might have been more logical to do in
terms of the power meridian. A +3.00 D cylindrical
lens with a vertical axis is labeled:
0 +3.00 x 90.
Diopters are assumed for the power and degrees are
assumed for the meridian specification. Usually the D
and the degree sign are not written. The x stands for
the axis meridian, so the above mathematical expres-
15.3 Standard Axis Notation sion literally stands for a +3.00 D cylindrical lens with
an axis meridian of 90 degrees. The expression is
A spherical lens has rotational symmetry around the typically read: +3 axis 90. The corresponding power
optical axis of the lens. In other words, the spherical cross is
lens can be rotated clockwise or counterclockwise
around the optical axis with no optical effects. A 0
cylindrical lens does not have rotational symmetry.
When the cylindrical lens is rotated clockwise or coun-
terclockwise, the resulting line image rotates with it. +3.00 D
Consequently, when a cylindrical lens is used to cor-
rect astigmatism, the orientation of the lens must be
specified together with the power.
The coordinate system used to specify the orienta- Consider a patient with a +2.00 x 30 lens in front
tion is set up in terms of the examiner's view when of each eye. In Figure 15.10, the drawing for the

FIGURE 15.8. Coordinate system for cylinder orientation.


300 Geometrie, Physical, and Visual Optics

FIGURE 15.9. a. 30 degree meridians, b. 150 degree meri-


dians.

Consider two thin cylindrical lenses with axis meri-


dians that are aligned. When these lenses are placed in
contact with each other, the cylinder power adds. For
example, a +3.00 x 60 combined with a +2.00 x 60 is
equivalent to a +5.00x60. Similarly, a +7.00x135
combined with a -3.00 x 135 is equivalent to a
+4.00 x 135.

FIGURE 15.10. +2.00 x 30 lens in front of patient's left eye.


Power cross in front of the right eye.
15.4 Perpendicularly Crossed
Cylindrical Lenses
Example 15.2 considered a +10.00 D cylindrical lens
with a vertical axis meridian (i.e., +10.00x90) to-
gether with a point source 25 cm in front of the lens.
Suppose a +7.00 D cylindrical lens with a horizontal
axis meridian (i.e., +7.00 x 180) is placed in contact
with the +10.00 D cylindrical lens. The +7.00x180
has zero dioptric power in the horizontal meridian.
Thus, the total power in the horizontal meridian is
provided by the +10.00 x 90. Similarly, the +10.00 x
90 provides zero power in the vertical meridian, and
FIGURE 15.11. Same as previous figure for a +4.50 x 150. the total power there is provided by the +7.00 D x
180.
In terms of the power crosses for the individual
patient's left eye shows how the lens would look if lenses, we can add respective horizontal and vertical
made with cylindrical surfaces, and the drawing for the powers:
patient's right eye shows the power cross. Figure 15.11
shows a similar case for a +4.50 x 150 lens in front of 0 +7D +7D
each eye.
Besides the symbol x, which represents the axis
meridian, the symbol @ is used to specify the power in + 10D + 0 = + 10D
a meridian. For the +3.00x90, there is +3.00 D @
180 and 0 D @ 90.
Astigmatism: On Axis 301

For a point source 25 cm in front of a +10.00 x 90


lens, a real vertical line image is formed 16.67 cm
behind the lens. When the +7.00 x 180 is combined
with the +10.00x90, the horizontal power remains
+10.00 D and the vertical line image is still formed
16.67 cm behind the lenses.
For a point source 25 cm in front of a +7.00 x 180 length of
lens, V= P + U = +7.00 D - 4.00 D = +3.00 D, and a horizontal
real horizontal line image is formed 33.3 cm behind line image
the lens. When combined with the +10.00x90, the
power in the vertical meridian is still +7.00 D, and a
horizontal line image is still formed 33.3 cm behind the length of
lens. We can keep track of the vergences and image vertical
line image
distances with the power cross technique.

+7.00 D -4.00 D
FIGURE 15.12. Cross-sectional ray diagrams for perpendicu-
larly crossed lenses.

+10.00 D + -4.00 D
Again, consider a circular aperture of diameter
40 mm placed against the two perpendicularly crossed
cylindrical lenses. When a screen is placed directly
+3.00 D behind the lenses, the illumination pattern on the
screen is circular. As the screen is moved back away
from the lenses, convergence occurs in both meri-
+6.00 D dians. Since convergence in the horizontal meridian
occurs faster than the convergence in the vertical
meridian, the resulting illumination pattern is a verti-
cally oriented blur ellipse in which both the major and
minor axes are less than 40 mm in length (Figure
The +6.00 D vergence in the horizontal meridian is 15.13). As the screen is moved back to the 16.67 cm
responsible for the formation of the vertical line image position, the convergence continues in both meridians,
at 16.67 cm, while the +3.00 D vergence in the vertical but since the horizontal meridian is converging faster,
meridian is responsible for the formation of the its length shrinks to zero first, and the vertically
horizontal line image at 33.33 cm. oriented blur ellipse turns into the vertical line image.
Figures 15.12a and 15.12b show the cross-sectional Once the screen is moved past the vertical line image
ray diagrams. Figure 15.12a shows the horizontal position, the light starts diverging in the horizontal
meridian in which the rays converge to a point at meridian, while at the same time it is still converging
16.67 cm, and then diverge. The point is actually on in the vertical meridian. The result is again a vertically
the vertical line image that would be perpendicular to oriented blur ellipse. However, as the screen is moved
the plane of the paper. Figure 15.12b shows the farther back, the growing horizontal length becomes
vertical meridian in which the rays converge to a point equal to the shrinking vertical length and the result is
at 33.33 cm. This point is actually a point on the a blur circle called the circle of least confusion. Once
horizontal line image. We can determine the length of the screen is moved past the circle of least confusion,
the vertical line image from the similar triangles in the growing horizontal length becomes greater than
Figure 15.12b, the cross-sectional view of the vertical the shrinking vertical length and the result is a
meridian. Similarly, we can determine the length of horizontally oriented blur ellipse. As the screen moves
the horizontal line image from Figure 15.12a, the toward the 33.33 cm position, the vertical meridian
cross-sectional view of the horizontal meridian. Note shrinks to zero length resulting in the horizontal line
that the line image lengths are determined by the image. Once the screen is past the 33.33 cm position,
meridian perpendicular to the meridian that forms the light in the vertical meridian starts to diverge, and
them. again a horizontally oriented blur ellipse appears.
302 Geometrie, Physical, and Visual Optics

O0lo?,=
X^__^/ circle of
FIGURE 15.13. Cross-sectional illumination pat-
terns for a circular aperture.
least confusion

We have been considering only one point source in position. Once past the last line image position, both
front of the lenses, so only one bundle is incident on principal meridians are diverging and a horizontal
the lens. In the astigmatic bundle leaving the lens, the rectangle again results.
two line images are perpendicular to each other and Figure 15.15 shows the solid conoid that results for
are aligned with the principal meridians of the perpen- the square aperture with a single point source in front.
dicularly crossed lenses. Note that all the light in the Again note that all the light passes through the first
bundle passes through the first line image and then line image, and then proceeds to the second line
through the second line image. At the second line image.
image position in the bundle, there is no blur from the The horizontal and vertical lengths of either the
first line image. blur ellipses, the blur rectangles, or the line images
The entire three-dimensional astigmatic bundle is can be calculated from the similar triangles shown in
called the conoid of Sturm after J.F.C. Sturm who Figure 15.12. In particular, the following example
appears to have made the first systematic investigation compares the line image lengths in the perpendicularly
of astigmatic bundles (1838). The distance between crossed cylinder case to the line image length in the
the two line images is called the interval of Sturm. case of the single plus cylinder in Example 15.2.
The position of least confusion lies between the two
line images, and is characterized by equal lengths of
the illumination patch in the principal meridians. EXAMPLE 15.3
When the aperture is circular, a blur circle occupies In Example 15.2, the point source was 25 cm in
the position of least confusion. However, when the front of the +10.00 x 90. The circular aperture had
a diameter of 40 mm. The real line image was
aperture has a different shape, the blur pattern at the vertical, occured at 16.67 cm, and had a length of
position of least confusion is no longer a circle. 6.67 cm. What happens to this line image when the
Consider a square aperture with horizontal and +7.00 x 180 cylindrical lens is placed against the
vertical sides placed against the two perpendicularly + 10.00x90?
crossed cylindrical lenses. The line image orientations From the previous discussion, the vertical line
and positions remain the same. A screen placed direct- image is still formed at 16.67 cm. However, its
ly behind the lenses has a square illumination pattern length is now different, which can be observed by
on it (Figure 15.14). As the screen is moved back, the comparing Figures 15.6a and 15.12b. In Figure
convergence is faster in the horizontal meridian than 15.12b, the triangle LL'x' is similar to the triangle
in the vertical meridian, so the illumination pattern AA'x'. The length LL' is equal to the diameter of
the aperture, which is 40 mm or 4 cm. The length
becomes a vertically oriented blur rectangle. The AA' is equal to the length of the vertical line
horizontal width shrinks to zero at the vertical line image. From the similar triangles,
image position and then starts growing. The light in
the vertical meridian is still converging, and a vertical- 33.3 cm _ 16.67 cm
ly oriented blur rectangle results. At the position of 4 cm AA
least confusion, the shrinking vertical length is equal and
to the growing horizontal length, and a square illumi-
nation pattern is formed. Past the least confusion AA = 2 cm.
position, the horizontal length is longer, and the il- So while the vertical line image remained in the
lumination pattern becomes a horizontal rectangle that same position, it did get shorter when the +7.00 D
shrinks to zero vertically at the horizontal line image cylindrical lens was added.

0 FIGURE 15.14. Cross-sectional illumination patterns


for a square aperture.
Astigmatism: On Axis 303

and
+5.0C

The individual cylindrical lenses are +2.50 x 90 and


+5.00 x 180, respectively.

15.5 Collapsing the Conoid by Equalizing


the Cross Cylinders
Figure 15.16a shows the ray diagrams for the image
space rays of the previous example. Suppose the
+2.50 x 90 is replaced with a +4.00 x 90. What hap-
FIGURE 15.15. Solidfigureof the conoid of Sturm for a square pens to the line images?
aperture. The new power in horizontal meridian is +4.00 D,
and since the power in the horizontal meridian forms
the vertical line image, it now occurs at +25 cm and is
shorter. The power in the vertical meridian is still
EXAMPLE 15.4 +5.00 D, so the horizontal line image remains at
Plane waves from a distant point source are inci- +20 cm. However, since the line images are now
dent on two perpendicularly crossed cylindrical lens- closer together, the horizontal line image is also shor-
es. A real horizontal line image is found 20 cm ter. (Compare Figures 15.16a and 15.16b).
behind the lenses, and a real vertical line image is If the +4.00 x 90 is now changed to a +4.50 x 90,
found 40 cm behind the lenses. Draw the resultant
power cross for the perpendicularly crossed cylin- the vertical line now moves to +22.2 cm and gets
drical lenses and specify their power and orien- shorter. The horizontal line image remains at +20 cm,
tation. and it also gets shorter. The closer the line images are
The horizontal line image is formed by the to each other, the shorter they get.
convergence in the vertical meridian, and the verti- If the +4.50 x 90 is changed to a +5.00 x 90, both
cal line image is formed by the convergence in the the vertical and the horizontal meridians form their
horizontal meridian. Since plane waves are inci- line images at 20 cm, but now both lines have zero
dent, U = 0, and P = V = l / i ; . Thus, the power length. In effect, the conoid has collapsed and a point
cross for the lenses is image has formed at 20 cm. Furthermore, since the
convergence is equal in the horizontal and vertical
meridians, blur circles are formed instead of blur
ellipses. The two +5.00 D perpendicularly crossed
cylindrical lenses converge the light exactly as a
+5.00 D spherical lens converges it.
We can repeat the above argument for any two
perpendicularly crossed cylindrical lenses of equal
power. Hence, we have a general principle: in the par-
axial approximation, two thin perpendicularly crossed
cylindrical lenses each of power P are equivalent to a
The two individual cylindrical lenses have respec- single spherical lens of power P, and vice versa.
tive power crosses of The power cross

0 +5.00 D

+2.50 D +5.00 D
304 Geometrie, Physical, and Visual Optics

a)

-20cm/
v
40cm
Horizontal
Vertical

b) FIGURE 15.16. a. Cross sections of the conoid for a


+ 2 . 5 0 x 9 0 0 + 5 . 0 0 x 1 8 0 . b. Cross sections when the
+2.50 x 90 is changed to +4.00 x 90. The vertical line
image moves closer to the horizonal line image and gets
shorter. The horizontal line image stays in the same
Horizontal Vertical place and gets shorter.

represents a +5.00 D spherical lens. Since the equival- +5.00 D


ence is good for any two perpendicularly crossed
spherical lenses, the power cross
+5.00 D
+5.00 D

We then add the powers in the respective meridians to


+5.00 D obtain the resultant power cross:
also represents the +5.00 D spherical lens. Similarly, +7.00 D
any other orientation of the perpendicularly crossed
lenses is equivalent to a sphere. The different possible
orientations manifest the fact that a spherical lens is
+.00 D
rotationally symmetric around its optical axis.

For the incident plane waves, the +7.00 D of power in


15.6 Transposition and Equivalent the vertical meridian forms a real horizontal line
Combinations image at +14.3 cm, and the +5.00 D of power in the
horizontal meridian forms a real vertical line image at
Consider plane waves from a distant point source +20 cm.
incident on a +5.00 D spherical lens placed in contact The sphere-cylinder combination gives a typical
with a +2.00 x 180 cylindrical lens. We can analyze conoid of Sturm. The same conoid can be generated
the effects of the lenses by use of the power crosses. with perpendicularly crossed cylindrical lenses. We can
The power cross for the cylindrical lens is determine the lens by noting that the above power
cross is equal to
+2.00 D
0 +7.00

+5.00 D +

As discussed in the previous section, a possible power The left power cross is for a +5.00 x 90 and the right
cross for the spherical lens is power cross is for a +7.00 x 180.
Astigmatism: On Axis 305

The sphere-cylinder combination above consists of powers (+5.00 D and +7.00 D) each serve as the
a spherical lens combined with a plus cylindrical lens. spherical power in a sphere-cylinder combination, and
There is a second sphere-cylinder combination that that the difference between the +7.00 D and the
also gives the same conoid of Sturm. The second +5.00 D gives the magnitude (2.00 D) of the cylindri-
sphere-cylinder combination consists of a spherical cal lens power.
lens together with a minus cylindrical lens. The powers Consider a spherical lens of power S combined with
in the resultant power cross are +7.00 D and +5.00 D, a cylindrical lens with parameters C x , i.e.,
and the spherical lens in the first combination is
+5.00 D. The spherical lens in the second combination SCCX0.
has the other power, namely, +7.00 D. The above When 0 is a first quadrant angle, the power cross for
power cross is then equal to the sum of the power the combination is
cross for the +7.00 D sphere and the power cross for
the unknown cylindrical lens. S+C

+7.00 D

+5.00 D
The other equivalent sphere-cylinder combination has
to have a spherical power equal to S + C. Then

+7.00 D

+7.00 D +

Subtraction in each principal meridian gives gives a power cross for the cylindrical lens of
P,=0
P =0

P =-2.00D,

and the cylindrical lens parameters are - C x (0 + 90).


0 was shown as a first quadrant angle in the above
which is the power cross for a -2.00 x 90 cylindrical power crosses. When 0 is a second quadrant angle, the
lens. Thus, the following three lens combinations are cylinder axis meridian of the equivalent sphere-cylin-
equivalent in the sense that they all give the same der combination is (0 - 90).
conoid of Sturm: The process of going from the sphere-plus cylinder
combination to the equivalent sphere-minus cylinder
+5.00 0 + 2 . 0 0 x 1 8 0 , combination or vice versa is called transposition. A
+7.00 0 - 2 . 0 0 x 9 0 , synopsis of transposition is:
+ 5 . 0 0 x 9 0 O + 7 . 0 0 x 180. Given a sphere-cylinder combination
The resultant power cross is the same for each of the S OCX 0,
three combinations.
When comparing the equivalent sphere-cylinder the transposed combination is
combinations, note that the magnitude of the cylinder (S + C ) O - C x ( 0 9 O ) .
power is the same (2.00 D), but the sign is different.
Note also that the cylinder axis meridians are 90 EXAMPLE 15.5
degrees different. When comparing the perpendicular- For a +4.000-3.00x42, find the equivalent
ly crossed cylinder combination to the equivalent sphere-cylinder and perpendicularly crossed cylin-
sphere-cylinder combinations, note that the respective der combinations. Also, write down the resultant
306 Geometrie, Physical, and Visual Optics

power cross, and for incident plane waves specify which corresponds to +4.00 x 35. The equivalent
the orientation and location of the line images. combination is then
The combination given is a sphere-minus cylin-
der. The equivalent combination is a sphere-plus +5.50 0 + 4 . 0 0 x 3 5 .
cylinder. From the transposition rule, the equival-
ent sphere-plus cylinder is
Suppose in the above example you picked the
[+4.00 + (-3.00)] C + 3.00 x (42 + 90), +9.50 D for the sphere power. If you work the rest
or correctly, you would obtain +9.50 0 - 4 . 0 0 x 1 2 5 .
+ 1.00 C+3.00 x 132. Then you can simply transpose to obtain +5.50 O
+4.00 x 35.
The resultant power cross for the original combina-
tion can be obtained by adding the power cross-
es. We can proceed more quickly by noting from
the original + 4 . 0 0 O - 3 . 0 0 x 42 lens that in the 42
meridian only the sphere contributes power, and in
the 132 meridian the sphere and cylinder both 15.7 Toric Surfaces
contribute. Then the resultant power cross is
+ 1.00 A conoid of Sturm can be generated by a single toric
surface. Figure 15.17a shows a torus. An inflated inner
+4.00 tube (used for tires, or for floating down rivers) is a
common example of a torus. A perfect doughnut
would be another example. The line in Figure 15.17b
is the symmetry axis of the torus. The small circle has
The resultant power cross is equal to a radius r 1? and its center is a distance g from the
symmetry axis. A rotation of the circle about the
+ 1.00 0
symmetry axis forms the torus.
+4.00 Consider cross sections that contain the point B on
the outer surface of the torus. Two of the cross
sections are circular in shape, and these two are
perpendicular to each other. One is the small circle of
so the equivalent perpendicularly crossed cylinder radius r,. The other is the circle of radius r 2 (where
combination is
r 2 = g + r j formed by the rotation of point B around
+4.00x132 0 + 1.00x42. the symmetry axis of the torus (Figure 15.17c). The
For incident plane waves, the +4.00 D of power in two circular cross sections are the principal meridians
the 42 meridian gives a 132 degree line at 25 cm, through point B.
and the +1.00 D of power in the 132 meridian gives Just as a spherical surface is part of a sphere, a
a 42 degree line image at 100 cm. toric surface is part of a torus. A refracting or reflect-
ing surface can be toric in shape. The characteristics of
EXAMPLE 15.6 the toric surface are the symmetry, as defined above,
Find the sphere-plus cylinder combination that is and the two perpendicular principal meridians with
equivalent to +9.50 x 35 O + 5.50 x 125. circular cross sections. Toric surfaces can have two
In the sphere-plus cylinder combination, the
sphere must have the least plus power, which is convex principal meridians, two concave principal
+5.50 D. Then the power crosses are: meridians, or one concave and one convex principal
meridian.
Consider a long glass (n = 1.5) rod with a toric
front surface. The principal meridians are both con-
+5.50 = +5.50 vex, and aligned with the vertical and horizontal meri-
dians (Figure 15.18). The vertical meridian has a
20 cm radius of curvature, and the horizontal meridian
and the power cross for the cylindrical lens must be has a 25 cm radius of curvature. We can represent the
toric surface by a power cross aligned with the princi-
Pj = +4.00 pal meridians, and calculate the refracting power in
each principal meridian from
P7 = 0

P = n,-n,
Astigmatism: On Axis 307

a)


Symmetry axis

FIGURE 15.17. a. A torus, b. Symmetry


parameters, c. Radii of curvature. (View is per-
pendicular to the symmetry axis.)

Then nation. As a result, the equations describing the toric


dioptric parameters have the same form as those for
sphere-cylinder combinations. For example, the toric
P V =H = +2.50D surface with the above power cross can be described as
a +2.00 0 + 0 . 5 0 x 1 8 0 surface. Alternatively, the
same toric surface might be described as +2.50 O
P = | = +2.00D -0.50 x 90, or as +2.00 x 90 O+2.50 x 180. All three
h
of these equations give the toric surface's power cross.
The wavefronts leaving the toric surface are them-
selves toric in shape. In general, toric wavefronts form
The power in the principal meridians forms a conoid a conoid of Sturm.
of Sturm in the glass rod. For a distant point source, A cylindrical surface is a special case of a toric
U = 0 and the vergence V in the vertical meridian surface in which one of the radii goes to infinity, and
equals +2.50 D. The power in the vertical meridian the corresponding meridian is flat. A spherical surface
forms the horizontal line image, which is at a distance is a special case of a toric surface in which the two
of 150/+2.50 D or +60 cm from the surface. The radii in the principal meridians are equal.
power in the horizontal meridian forms the vertical A lens made with a toric surface, or surfaces,
line image, which is a distance of 150/+2.00 D or generates a conoid of Sturm just like a sphere-cylinder
+75 cm from the surface. combination or the equivalent perpendicularly crossed
The power cross for a toric surface has the same cylinder combination. A lens with one or more toric
form as the power cross for a sphere-cylinder combi- surfaces is called a spherocylindrical lens. A typical
spherocylindrical lens has a spherical front surface and
a toric back surface.

EXAMPLE 15.7
A point source is 100 cm in front of a thin
spherocylindrical lens that has a +8.00 spherical
front surface, and a -3.00 O-2.00 x 45 toric back
surface. What are the line image positions and
orientations? What is the sphere-minus cylinder
equation that describes the lens?
Let us find the power cross for the lens first. The
spherical front surface can be represented by a
power cross with +8.00 D of power in both the 45
and 135 meridians. The powers in the respective
FIGURE 15.18. Toric interface on a long rod. principal meridians add, so
308 Geometrie, Physical, and Visual Optics

Pi + To write a sphere-minus cylinder equation for


the lens, consider the power cross for the lens. The
highest plus power was +5.00 D in the 45 meridian.
,+3 v+8 .-5 The sphere-minus cylinder equation is then +5.00
0 - 2 . 0 0 x 4 5 . (You can verify this with power
'+5 = \^^+8 + \^""-3 crosses if you want.) Note that since the lens is thin,
we could just add the spherical front surface power,
+8.00 D, to the spherical power in the toric back
surface equation, -3.00 O-2.00 x 45, to obtain the
equation for the total lens, + 5 . 0 0 0 - 2 . 0 0 x 4 5 .
The power cross for the lens has +3.00 D of power
in the 135 meridian and +5.00 D of power in the 45
meridian. We can now use a power cross representa-
tion to obtain the vergences in the principal meri-
dians of the exiting toric wavefront: 15.8 Extended Objects and Multiple Conoids
The above discussion of astigmatism dealt only with a
v u single point source and a single astigmatic bundle
(i.e., a single conoid of Sturm). Extended objects
consist of many point sources. Consider a target con-
sisting of the five paraxial point sources shown in
Figure 15.19a. One point source is centrally located.
va = +5 + The others are above, below, left, and right of the
center, respectively. The object plane is 50 cm in front
of a thin +7.00O-3.00X 180 spherocylindrical lens.
What does the image of the points look like at various
So the vergence cross for the wavefront leaving the positions behind the lens?
lens is
We start by asking how the central point is imaged
Vb = +2 in the conoid. The power cross for the lens is
V =+4
+4.00 D

The vergence of +2.00 D in the 135 meridian forms


a 45 degree line image at +50 cm. The vergence of +7.00 D.
+4.00 D in the 45 degree meridian forms a 135 line
image at +25 cm.

center
O0io c=> o
point
source 20cm
y
50cm

c)

/
FIGURE 15.19. a. Array of point sources, b.
20cm A conoid for one point source, c. Line images in
the respective conoids from two different point
50cm sources.
Astigmatism: On Axis 309

U is - 2 D , so the vergence cross for the toric wave- for the five points is simulated by drawing a vertical
front leaving the lens is line centered on the chief ray of each bundle (Figure
15.20b). The overlapping of the individual vertical line
+2.00 images gives the image the appearance of having a
solid vertical line in the middle.
At the position of the circle of least confusion, the
+5.00 D. five-point image is simulated by drawing the blur circle
centered at each chief ray (Figure 15.20c). At the
position of the horizontally oriented blur ellipse, the
five-point image is simulated by drawing a horizontally
The +2.00 D vergence in the vertical meridian forms a oriented blur ellipse at the position of each chief ray
horizontal line image +50 cm behind the lens, and the (note some overlapping in Figure 15.20d). At +50 cm,
+5.00 D vergence in the horizontal meridian forms a the position of the horizontal line image, the five-point
vertical line image +20.0 cm behind the lens. Figure image is simulated by drawing a horizontal line image
15.19b shows the line images and the corresponding centered on each chief ray position. As shown in
blur ellipses. Figure 15.20e, the resulting extended image appears to
For the central point source, the chief ray in the have a solid horizontal line.
astigmatic bundle lies on the optical axis. For the point Assume the five point sources are replaced by the
source below the center, a different astigmatic bundle letter T. Figure 15.21 shows the results of the simula-
is formed, and its chief ray travels slightly upward tion process for astigmatically imaging the letter T. In
relative to the axis (Figure 15.19c). In the paraxial Figure 15.21, the top row shows the cross-sectional
approximation, the vertical line image in the latter illumination patterns in the conoid of Sturm for a
conoid is still in the +20 cm plane and the horizontal single astigmatic bundle. The second row shows the
line image is still in the +50 cm plane. For the point simulated illumination patterns for the letter T. To
source above the center, still another astigmatic bun- construct the simulations, first remember that the chief
dle is formed and its chief ray travels slightly down- rays cross over the axis, so the image of the letter T is
ward relative to the axis. In fact, an astigmatic bundle inverted.
is formed for each of the paraxial point sources, and At the position in the single conoid of the vertically
the vertical line image in each astigmatic bundle is in oriented blur ellipse, a vertically oriented blur ellipse
the +20 cm plane and the horizontal line image is in is drawn for each chief ray (i.e., at each point in the
the +50 cm plane. inverted T). The result shows that the horizontal bar
Once we know how a single point is imaged, we in the T image is blurred more than the vertical bar.
can simulate how the collection of points constituting At the vertical line image position (+20 cm), a vertical
the extended target is imaged. In Figure 15.19b, the line image is drawn for each chief ray or each point in
blur ellipse nearest the lens is vertically oriented. To the inverted T, i.e., the image of the horizontal bar in
simulate the illumination pattern for the five-point the inverted T consists of a series of vertical line
extended target, we center a vertically oriented blur images that results in a blurred horizontal bar.
ellipse on each of the five chief ray positions. As (Whether the blurred horizontal bar is visible or not
shown in Figure 15.20a, some or all of these blur depends on the amount of blur.) The image of the
ellipses may overlap. vertical bar in the inverted T consists of a series of
For each astigmatic bundle, the vertical line image overlapping vertical lines. The result is a clear vertical
occurs at +20 cm. The illumination pattern at 20 cm bar. At the circle of least confusion position, a blur

o
a)
0Q0 b) c) O o o
o

O
d) OOO e)

o FIGURE 15.20. Illumination patterns at various


locations for the object points of Figure 15.19a.
310 Geometrie, Physical, and Visual Optics

0 0 io
object

FIGURE 15.21. Illumination patterns at


various locations for a single point object and
simulation of illumination patterns for an ex-
tended object.

circle is drawn for each point in the inverted T, and in The resulting illumination patterns at the position of
this case the horizontal and vertical bars are equally one line image do give information about different
blurred. The procedure is the same at the position of a bars or lines in the target. The amount of information
horizontally oriented blur ellipse, in which case the depends on the amount of blur, which depends on the
vertical bar is blurred more than the horizontal bar. orientation of the bar. The only bars that have a clear
At the horizontal line image position, the horizontal image somewhere in the conoid are those that are
line images at each chief ray position give a blurred aligned with the principal meridians of the conoid.
vertical bar in the inverted T, while the overlapping (Incidentally, the process of simulating the illumina-
horizontal line images give a clear horizontal bar. tion patterns on a point by point basis can be formal-
Figure 15.22 shows a similar simulation in which ized mathematically and is called a convolution.)
the letter A is the object. The astigmatic images
correspond to an inverted A. For the inverted A, the
horizontal bar is clear at +50 cm, which is the horizon-
tal line image position in a single astigmatic bundle.
The diagonal lines in the inverted A are not aligned 15.9 The Circle of Least Confusion and
with a principal meridian of the spherocylindrical lens, the Spherical Equivalent
and as a result are not clear anywhere.
For most image positions, there are differing Since the circle of least confusion has some special
amounts of blur for the different bars depending on significance, let us figure out its exact position in the
the bar orientation. The only position where the conoid. Consider a lens with a circular aperture of
amount of blur is equal for all orientations is at the diameter h. Assume the exiting wavefront is converg-
circle of least confusion. For a complex target, the ing and toric with a vergence Vj in one principal
circle of least confusion then supplies equal informa- meridian and a vergence V2 in the other principal
tion, regardless of orientation. In this case, the circle meridian. Figure 15.23 gives the corresponding ray
of least confusion is regarded as the best image posi- diagrams. The first line image occurs at (which
tion for a complex target. However, the amount of equals 1/VJ, and the second line image occurs at v2
information supplied by the circle of least confusion is (which equals 1/V2).
less than a clear image supplies. The circle of least confusion occurs between the
Remember that for a single astigmatic bundle, all two line images at the position vc in which the cross
of the light first passes through one line image and sectional length g is the same for each principal meri-
then proceeds to the second line image. At one line dian (dashed line in Figure 15.23). From the similar
image position, there is no information in the illumina- triangles in the left diagram.
tion pattern about the other line image (the top row in
Figure 15.21). However, for an extended object, such h _ g
as the letter T, there are multiple astigmatic bundles. v, - '

object

A FIGURE 15.22. Simulated il-


lumination patterns for the letter
A as an object.
Astigmatism: On Axis 311

FIGURE 15.23. Geometry to find circle of least con-


v2 fusion location.

From the similar triangles in the right diagram, The incident vergence U is -2.50 D. Then
h _ g
V = P + U

We can solve each of the above two equations for g/h


and set the results equal to each other to obtain
v -

or

gives
Then
V V
^ = +5.00 0
c c
2= +
vl v2 ^V7 = +1.00D
and

1 =1 1 From Eq. 15.5,


Vr V, V,
Let Vc be the dioptric location of the circle of least _ +5.00 D +1.00 D
+3.00 D.
confusion:
Note that 3 is halfway between 1 and 5. The linear
location of the circle of least confusion is
where vc is in meters. 100
u +3.00 D = +33.3 cm.
Then
(15.4) The first line image is oriented at 35 and is located
2VC=V,+V2, at
and finally
^ = T?ro = + 2 0 0 c m
(15.5) The second line image is oriented at 125 and is
located at
Equation 15.5 says that the circle of least confusion is
dioptrically located halfway between the line images.
This does not mean that the circle of least confusion is
located linearly halfway between the line images. Figure 15.24 summarizes the results. The linear
halfway point between the two line images is the
EXAMPLE 15.8 +60 cm position. Note that while the circle of least
A point source is located 40 cm from a thin confusion is dioptrically halfway between the two
+7.50 0 - 4 . 0 0 x 1 2 5 spherocylindrical lens. Find line images, it is definitely not linearly halfway
the circle of least confusion position and the line between. Instead, it is much closer to the first line
image positions and orientations. image.
312 Geometrie, Physical, and Visual Optics

In this sense, the +10.00 0 - 6 . 0 0 x 1 2 5 lens acts


in a rough fashion like a +7.00 D spherical lens.
\ Note that for a spherocylindrical lens with parame-
ters S C C x 0, the spherical equivalent and the diop-
r tric location of the circle of least confusion depend
20cm 100cm
_y only on S and C, and are independent of the cylinder
33.33cm axis meridian 0.
FIGURE 15.24. Circle of least confusion location. The so-called Jackson cross cylinder (JCC) is a
spherocylindrical lens with a zero spherical equivalent.
Jackson cross cylinders are widely used in determining
When plane waves are incident on a spherocylindri- a person's refractive error. A JCC consists of a
cal lens of parameters S O C x 0, the vergence in each spherocylindrical lens with plus power in one principal
principal meridian equals the lens' dioptric power in meridian and minus power of equal magnitude in the
that meridian, i.e., other principal meridian, e.g., a + 0 . 2 5 0 - 0 . 5 0 x 9 0
or the equivalent +0.25 x 180O-0.25 x 90. You can
verify from Eqs. 15.5 or 15.6 that the spherical equi-
V, = S + C valent of a JCC is zero.

15.10 Virtual Line Images


Equation 15.5 gives the dioptric location for the circle
of least confusion A toric wavefront that is converging in both principal
meridians yields two real line images. Once the toric
v=i sc)is=s+c wavefront is past the second line image position, it is
diverging in both meridians. An observer viewing the
For a complex target, the circle of least confusion is real line images as aerial images has the diverging toric
the best image position for the spherocylindrical lens. wavefront incident on his or her eye.
The spherical lens, which gives a point image at the Exiting toric wavefronts can also be diverging in
same position as the circle of least confusion, is called both principal meridians. These diverging toric wave-
the spherical equivalent (Sp Eq) of the spherocylindri- fronts yield two virtual line images. To a downstream
cal lens. The dioptric power of the spherical equiva- observer (Figure 15.25) it makes no difference optical-
lent is halfway between the principal meridian dioptric ly whether the line images are virtual images or aerial
powers of the spherocylindrical lens. Alternatively, images. A downstream observer cannot focus on both
from the above equation, line images simultaneously.
An exiting toric wavefront can also be diverging in
Sp Eq = S + y (15.6) one principal meridian while converging in the other.
Such a wavefront yields one virtual line image and one
For a spherical lens, C is zero and the Sp Eq is just the real line image.
sphere power S. For a cylindrical lens, S is zero and Consider collimated light incident on a thin -5.00
the spherical equivalent is equal to C/2. O-3.00 x 180 spherocylindrical lens. The power cross
representation for the exiting vergence is
EXAMPLE 15.9
A point source is located 15 cm in front of a thin -8.00 D
+ 10.00 C-6.00X 155 spherocylindrical lens. What
is the Sp Eq?
The Sp Eq is defined in terms of the dioptric
parameters of the lens, so the point source location -5.00 D
is irrelevant. From Eq. 15.6,

Sp Eq = +10.00 D + ~ 6 ^ = +7.00 D.
A +7.00 D spherical lens gives a point image at the The resulting virtual line images are vertical at -20.0
same position as the circle of least confusion given cm and horizontal at -12.5 cm.
by the +10.00 O-6.00 x 125 spherocylindrical lens. Consider the power cross analysis for a +6.00 x 180
Astigmatism: On Axis 313

converging
tone
tc wave front

real aerial
line images diverging toric
wave front

FIGURE 15.25. Converging toric wavefronts that form real


aerial line images and then diverge to eye compared to a
spherocylindrical lens that forms two virtual line images and gives
virtual
line images the same diverging toric wavefront incident on the eye.

lens with a point source 50 cm in front. From V = to optical infinity. Thus, a plus cylindrical lens with
P + U, plane waves incident has a real line image parallel to
the axis and at a finite distance from the lens, together
V +6.00 -2.00 D with a virtual line image parallel to the power meri-
dian at optical infinity.
Hence, every thin spherocylindrical lens with non-
--2.00D zero cylindrical power forms two perpendicular line
V. = 0 +
images. The line images might both be real, both
virtual, or one real and one virtual.

or
V =+4.00D 15.11 Through a Glass Astigmatically

Suppose an unaccommodated emmtrope holds a


V =-2.00D +5.00 x 180 cylindrical lens directly in front of her eye
and looks through it at a distant point source. What
does the emmtrope see?
Before rushing to answer, let us first ask what an
The vergence cross shows that the wavefront leaving unaccommodated emmtrope would see through a
the lens is toric with one converging meridian and one +5.00 D spherical lens. By itself, the +5.00 D spheri-
diverging meridian. As expected, the +4.00 D ver- cal lens from a real point image at +20 cm. When held
gence in the vertical meridian gives a real horizontal directly in front of the eye, the +5.00 D spherical lens
line image at +25 cm. The -2.00 D vergence in the pulls the eye's point image 5.00 D forward into the
horizontal meridian gives a virtual vertical line image vitreous humor leaving a blur circle on the retina.
at - 5 0 cm, the distance at which the original point Thus, the +5.00 D lens simulates 5 D of myopia. Note
source is located. Figure 15.26 shows the rays for the that the observer sees the blur circle, and not the point
virtual line image. image (Figure 15.27a).
When the object point for a plus cylindrical lens By itself, the power meridian of the +5.00 x 180
goes to optical infinity, the virtual line image also goes cylindrical lens forms a real horizontal line image.

real
horizontal line
image

position of FIGURE 15.26. Real and virtual line images for


virtual line image a converging cylindrical lens.
314 Geometrie, Physical, and Visual Optics


+ 5.00x180

b)

FIGURE 15.27. a. Observer viewing through a +5.00 D


point spherical lens. b. Observer viewing through a +5.00 x 180
image lens.

When placed directly in front of the eye, the +5.00 x lens is held directly in front of the emmetropic eye, it
180 lens simulates a myopic condition in the vertical pulls the eye's horizontal line 5 D in front of the retina
meridian while leaving the horizontal meridian em- and leaves the vertical line image on the retina. The
metropic. The horizontal line image, which is formed patient thus reports seeing a vertical line image when
by the myopic meridian, is in the vitreous humor. The looking through the +5.00 x 180 lens.
vertical line image, which is formed by the emmet- A third strategy is to note that while, for a distant
ropic meridian, is on the retina and therefore is the point source, the +5.00 x 180 cylindrical lens forms a
line observed (Figure 15.27b). real horizontal line image at +20 cm, it also forms a
We can show the same thing numerically by assum- virtual vertical line image at optical infinity. An un-
ing that +60 D is emmetropic, and represents the accommodated emmtrope looking through the +5.00
emmetropic eye by a power cross that has + 6 0 D in x 180 cylindrical lens at the distance point source
both vertical and horizonal meridians. Neglecting ef- focuses on the virtual vertical line image.
fectivity, the +5.00 x 180 together with the eye gives a
total power cross of EXAMPLE 15.10a
Teddy Roosterbird is an emmtrope. Teddy looks
V, +5D +60 D through a +0.25 0 - 0 . 5 0 x 9 0 Jackson cross cylin-
der (JCC) at a distant point source. When the JCC
is flipped about a meridian 45 degrees from the
principal meridians, the power meridians exchange
Vv = 0D + +60D and the new parameters are + 0 . 2 5 O - 0 . 5 0 X 180.
Does flipping the JCC make any optical difference
to Teddy?
The power cross for the + 0 . 2 5 0 - 0 . 5 0 x 9 0
or JCC is

+65 D +0.25 D

+60 D -0.25 D

The myopic vertical meridian forms a horizontal Together with the eye, the JCC sets up a conoid of
line image that is 5 D in front of the retina. The Sturm in which the line image formed by the
emmetropic horizontal meridian forms a vertical line vertical meridian is pulled 0.25 D in front of the
image that is on the retina and observed by the retina and the other line image is "pushed" 0.25 D
emmtrope. Note that a conoid of Sturm is set up in behind the retina. The spherical equivalent of the
the emmetrope's eye and that all the light first goes JCC is zero, and the circle of least confusion is on
through the horizontal line image and then moves on the retina. When the JCC is flipped, the line images
to form the vertical line image. Thus, when the em- change position, but the circle of least confusion
remains on the retina (Figures 15.28a and 15.28b).
mtrope looks through the +5.00 x 180 lens, the em- Since the circle is unchanged during the flip, there
mtrope sees no evidence of the horizontal line image is no optical difference for Teddy.
in the vitreous humor. (See Figure 16.2a if necessary.)
Another analysis strategy is to consider the emme- EXAMPLE 15.10b
trope's point image as zero length vertical and Franklin Roosterbird has 0.25 D of uncorrected
horizontal line images. By itself, the +5.00 x 180 lens myopia and is viewing a distant point source. The
forms a horizontal line image. When the +5.00 x 180 + 0 . 2 5 O - 0 . 5 0 x 90 JCC is placed directly in front
Astigmatism: On Axis 315

+ 0.25 c -0.50x90 + 0.25 C-0.50x180

a)

G
+ 0.25 S-0.50x090
b)

+ 0.25C-0.50x180

d)

<3 FIGURE 15.28. a. ft. Flipped JCC positions for an


emmtrope, c. d. Flipped JCC positions for an uncorrected
myope.

of Franklin's eye and then flipped. Does Franklin behind the retina (Figure 15.29b). (We usually say the
notice any difference? second line image is "behind" the retina.)
By itself, Franklin's eye forms a point image A compound myopic astigmat (CMA) has both line
0.25 D in front of the retina. The power cross for images in front of the retina (Figure 15.29c). A com-
the +0.25 0 - 0 . 5 0 x 9 0 JCC is the same as in
pound hyperopic astigmat (CHA) has both line images
Example 15.10a. For Franklin's eye and the JCC,
the power in the horizontal meridian pushes the behind the retina (Figure 15.29d).
vertical line image 0.25 D back placing it on the A mixed astigmat (MA) has one line image in front
retina. The power in the vertical meridian pulls the of the retina and one line image behind the retina
horizontal line 0.25 D forward leaving it 0.50 D in (Figure 15.29e). As a special case, an equal mixed
front of the retina. Franklin then sees a vertical astigmat (EMA) has each line image dioptrically the
line. same distance from the retina, in which case the circle
When the JCC is flipped, the powers in the of least confusion is on the retina (Figure 15.29f).
principal meridians exchange, and the line images Assuming that a typical emmetropic meridian has a
also exchange (Figures 15.28c and 15.28d). Thus, power of +60.00 D , Figure 15.30 shows power crosses
the line image on the retina changes from vertical
for the cases considered in Figure 15.29. Obviously,
to horizontal, and Franklin clearly notices the dif-
ference. the SMA has one myopic and one emmetropic meri-
By comparing the situation for Teddy and Fran- dian (Figure 15.30a), the SHA has one hyperopic and
klin, we can see how the JCC is used to refine the one emmetropic meridian (Figure 15.30b), the CMA
power needed in the correction. has two myopic meridians (Figure 15.30c), the C H A
has two hyperopic meridians (Figure 15.30d), and an
M A has a myopic and a hyperopic meridian (Figure
15.30e). In the E M A , the dioptric difference between
15.12 Classification and Correction the myopic meridian and the emmetropic meridian is
of Ast ig mats equal in magnitude to the difference between the
hyperopic meridian and the emmetropic meridian (Fig-
The cornea is a major source of astigmatism in the ure 15.30f).
human eye. This happens when the middle of the Astigmatism is also classified by orientation. The
corneal surface is toric instead of spherical. The crys- major meridian of the eye is the principal meridian
talline lens is also a possible source of astigmatism. with the most plus power. When the eye's major
When astigmatism is present, a conoid of Sturm is set meridian is vertical or 30 degrees on either side, the
up inside the eye. astigmatism is called with-the-rule (WTR in Figure
Astigmatism classification by power depends on 15.31a). When the eye's major meridian is horizontal
where the conoid's line images are relative to the or 30 degrees to either side, the astigmatism is called
retina in the unaccommodated eye looking at a distant against-the-rule (ATR in Figure 15.31b). When the
point source. The classifications are simple myopic, eye's major meridian is in either one of the remaining
simple hyperopic, compound myopic, compound hy- zones, the astigmatism is called oblique (Figure
peropic, and mixed. 15.31c). One oblique zone is centered on 45 degrees
A simple myopic astigmat (SMA) has one line and the other on 135 degrees.
image on the retina and one in front (Figure 15.29a). WTR is the most common type of astigmatism, and
A simple hyperopic astigmat (SHA) has one line A T R is the next most common. As people age, WTR
image on the retina and one that wants to be formed astigmatism tends to change to A T R astigmatism.
316 Geometrie, Physical, and Visual Optics

a)

O SMA
O SHA

d)

CMA CHA

e)

MA
O EMA FIGURE 15.29. Classification of astigmatism.

+ 62.00 + 60.00

a) b)
-+60.00 + 58.00

SMA SHA

+ 59.00
+ 61.00
d)
-+63.00 -+57.00

CHA
CMA

+ 62.00
+ 59.00

+58.00
-+63.00

FIGURE 15.30. Ocular power crosses and


EMA
MA astigmatism classification.
Astigmatism: On Axis 317

a) WTR

v b)
ATR

A
FIGURE 15.31. Classification by orientation of the major ocu-
OBLIQUE lar meridian.

Oblique astigmatism is the least common. In Figure -3.50 D lens is removed, the vertical line image
15.30, a, b , d, and f are W T R , while c, and e are A T R . must then be 3.50 D in front of Belinda's retina and
The minus cylinder correction for W T R astigmat- the horizontal line returns to the retina (Figure
ism has its axis meridian within 30 degrees of 180 15.32b).
(e.g., - 1 . 0 0 x 1 8 0 , - 1 . 0 0 x 2 2 , - 1 . 0 0 x 1 6 4 ) . The The lens needed to correct Belinda's vision must
move the vertical line image 3.50 D back without
minus cylinder correction for A T R astigmatism has its
moving the horizontal line image. This lens is a
axis meridian within 30 degrees of 90 (e.g., - 1 . 0 0 x -3.50 x 90. Belinda is an ATR SMA.
90, - 1 . 0 0 x 1 2 0 , - 1 . 0 0 x 7 5 ) . We can also determine Belinda's correction
A correction of - 5 . 0 0 0 - 0 . 7 5 x 3 4 is for an ob- from a power cross analysis. Assume that +60.00 D
lique compound myopic astigmat. A correction of is emmetropic. Power in the vertical meridian
+ 0 . 5 0 C - 1 . 0 0 x 90 is for an A T R E M A . The spheri- forms horizontal line images. Since Belinda sees a
cal equivalent of a correction for an E M A is zero. horizontal line when uncorrected and unaccommo-
dated, she must be emmetropic in the vertical
meridian. Since power in the horizontal meridian
EXAMPLE 15.11 forms vertical lines and Belinda sees a vertical line
When uncorrected and unaccommodated, Belinda with the -3.50 D spherical lens in place, her
Smiley looks at a distant point source and sees a horizontal meridian must be myopic. A power cross
horizontal line. When a -3.50 D spherical lens is for Belinda's eye is
held directly in front of her unaccommodated eye,
Belinda sees a vertical line. What type of astigmat- +60.00 D
ism does Belinda have? What is her spherocylindri-
cal prescription in minus cylinder form?
When uncorrected and unaccommodated, Be-
linda has a horizontal line image on her retina. A +63.50 D
spherical lens held in front of the eye moves both
of the line images the same dioptric amount. With
the -3.50 D spherical lens in place, the horizontal
line image moves 3.50 D behind the retina. Neglecting effectivity, the correcting lens together
According to the given information, with the with the eye must add to +60.00 D of power in
-3.50 D lens in place, Belinda has a vertical line both meridians, so the correcting lens power cross
image on her retina (Figure 15.32a). When the is

-3.50DS
Y

a) b)
3.50 D
FIGURE 15.32. Line image locations for Belinda Smiley.
318 Geometrie, Physical, and Visual Optics

EXAMPLE 15.13
Stacy Zelp needs a - 5 . 0 0 0 - 1 . 0 0 x 1 4 4 distance
vision correction. Classify Stacy's astigmatism and
show where her line images are when Stacy is
-3.50 D uncorrected and unaccommodated.
The power cross for Stacy's correction is

-5.00
which is the power cross for a -3.50 x 90 cylindri-
cal lens. -6.00

EXAMPLE 15.12
Dr. Ulysses Bonemeal looks at a distant point
source through a +4.75 D spherical lens and sees a The major meridian is oblique, and Stacy's correct-
horizontal line image. When the +4.75 D lens is ing lens has minus power in both principal meri-
replaced by a +6.75 D spherical lens, Dr. Bone- dians. Stacy is an oblique CMA.
meal sees a vertical line. Assuming no accommoda- We can work from the power cross to determine
tion, what is Dr. Bonemeal's astigmatism classifica- the line image location in the uncorrected eye. The
tion and sphere-minus cylinder correction? -5.00 D in the 144 meridian means that in the
Since Dr. Bonemeal sees a horizontal line when uncorrected eye, a line image at 54 is 5.00 D in
the +4.75 D sphere is in place, the horizontal line front of the retina. The -6.00 D in the 54 meridian
image in Dr. Bonemeal's uncorrected eye must be means that in the uncorrected eye, a line image at
4.75 D behind the retina. Similarly, the vertical line 144 is 6.00 D in front of the retina (Figure 15.34).
image in Dr. Bonemeal's uncorrected eye must be
6.75 D behind the retina (Figure 15.33a).
When the +6.75 D spherical lens is in place, the EXAMPLE 15.14
vertical line image is on the retina and the horizon- Nancy Block-Head holds a +7.00 D spherical lens
tal line image is 2.00 D in front of the retina directly in front of her eye. With the spherical lens
(Figure 15.33b). Then a -2.00 x 180 lens moves in place, Nancy can clearly see a grating target
the horizontal line back to the retina without mov- consisting of vertical lines when the target is 50 cm
ing the vertical line. This collapses the conoid and in front of the lens. When the target is rotated so
leaves a point image (Figure 15.33c). The sphere- that it consists of horizontal lines, Nancy can clear-
minus cylinder correction is +6.75 0 - 2 . 0 0 x 1 8 0 . ly see the chart when it is 33.3 cm in front of the
Dr. Bonemeal is a WTR CHA. lens. Assuming no accommodation, classify Nan-
Note that with the +4.75 D spherical lens in cy's astigmatism and give her correction in sphere-
place, the horizontal line is on the retina and the minus cylinder form.
vertical line is 2.00 D behind the retina. Then a For a single point source, power in the horizon-
+2.00 x 90 cylindrical lens would pull the vertical tal meridian forms vertical line images. The fact
line forward and collapse the conoid. The resulting that the vertical lines are clear when the +7.00 D
Rx, +4.75 0 + 2 . 0 0 x 9 0 , is the transpose of the spherical lens is in place and the grating is at 50 cm
sphere-minus cylinder Rx. ( U = - 2 . 0 0 D ) indicates that the eye's horizontal

+ 6.75DS

b)
a)

+ 6.75 - 2 . 0 0 x 1 8 0

point
c) . image

FIGURE 15.33. Line image locations for Dr. Bonemeal.


Astigmatism: On Axis 319

15.13 The Clock Dial Chart


Extended targets are much easier to use clinically than
point sources. We can use an extended target consist-
ing of radial lines to check for the orientation of the
principal meridians as well as the amount of astigmat-
FIGURE 15.34. Line image locations for Stacey Zelp. ism. Such a chart is called a clock dial chart (Figure
15.35). In situations where accommodation remains
active, the clock dial chart is typically used as follows.
meridian needs a +5.00 D vergence leaving the The patient views the reasonably distant (i.e., 4 to
correcting lens. 6 m) chart through a lens holder (usually a phoropter).
Power in the vertical meridian forms horizontal Plus spherical lenses of increasing power are placed in
line images. The fact that the horizontal line images the holder until the patient reports the chart is blurred,
are clear when the +7.00 D spherical lens is in and the examiner is sure the patient's image is in front
place and the grating is 33.3 cm away means that of the retina. In this case the patient is said to be
the eye's vertical meridian needs a +4.00 D ver- fogged. Now accommodation blurs the image more,
gence leaving the correcting lens.
and that presumably eliminates any stimulus to accom-
For distance vision, the power cross of the
correction is modate. (For simplicity, a possible non-zero resting
state for accommodation is neglected.)
+4.00 The plus spherical power is then decreased (or
minus spherical power is added). When the person has
no astigmatism, all the radial lines clear up simulta-
neously. When the person has astigmatism, one of the
+5.00
radial lines clears first. At that point, one line image is
on the retina and the other line image is in front of the
retina, just as in simple myopic astigmatism. The
From the power cross, the correction is + 5 . 0 0 0 examiner then starts adding minus cylindrical lenses of
-1.00 x 180. Nancy is a WTR CHA. increasing power with the axis of the minus cylinder
lined up perpendicularly to the clear line on the retina.
EXAMPLE 15.15 This minus cylinder moves the other line back onto
The principal meridians of Duke Fred Overboot's the retina collapsing the conoid.
eye are 55 and 145. In the 55 meridian, the power Consider Ella Mentary. Ella sits in front of a
in Duke's eye is +62.00 D, while in the 145 meri- distant clock dial chart and is fogged. As the fog is
dian, the power is +58.00 D. The power cross for decreased, Ella reports that a line at 63 degrees comes
Duke's eye is clear first. The spherical power in place is +3.50 D.
+62.00 Minus cylinder axis 153 is then added to the sphere.
Ella reports that all lines clear up when - 0 . 7 5 D of
cylindrical power is in place. Ella's Rx is +3.50 O
- 0 . 7 5 x 153.
Note that with this procedure, the Rx automatically
' +58.00 comes out as a sphere-minus cylinder combination as
opposed to a sphere-plus cylinder combination. The
Assuming +60.00 D is emmetropic, what is the sphere-minus cylinder combination controls accommo-
classification of Duke's astigmatism, and what is his dation while the sphere-plus cylinder combination
correction? does not.
The power cross for Duke's correction is

, -2.00

+2.00

From the power cross, the correcting spherocylin-


der is + 2 . 0 0 O - 4 . 0 0 x 145.
Duke is an oblique EM A. FIGURE 15.35. Clock dial chart.
320 Geometrie, Physical, and Visual Optics

The standard coordinate system is set up in terms those for a spherical wavefront. However, the work
of the view of the examiner looking toward the patient has now doubled since there are two principal meri-
(i.e., the contra-ocular view). However, the patient dians.
viewing the clock dial is looking in the other direction For a S C C x spherocylindrical lens, the powers
and tends to set up a different coordinate system. To in the principal meridians are S and (S + C). These
avoid this, we simply label the examiner's standard are the powers that must be used in the effectivity
coordinate system as it would appear from the ocular calculations. In general, C is the difference between
view (Figure 15.36a). In the ocular view, zero occurs the powers in the two meridians and cannot be used
on the left and the first quadrant angles are clockwise by itself in the effectivity calculations. An apparent
up from the zero. You can investigate these two views exception occurs when S is zero, in which case (S + C)
by drawing a standard examiner's coordinate system is equal to C.
on a sheet of paper (i.e., contra-ocular view), and
then viewing it from the back side (the ocular view). EXAMPLE 15.17
Harry Ape wears a -9.00O-3.00x 180 distance
EXAMPLE 15.16 vision correction at a vertex distance of 13 mm.
Anna Brown is fogged, and as the fog is reduced, What is Harry's correction at the cornea?
the 2 to 8 o'clock line comes clear first. The The power cross for Harry's correction is
spherical lens is place is -4.25 D. Then correctly
oriented minus cylinders are added, and all lines -12.00 D
come clear at a -1.25 D cylinder power. What is
Anna's Rx in minus cylidner form?
From Figure 15.36b, the 2 to 8 o'clock line is the
150 degree line. The line image still in front of the -9.00 D
retina is the 60 degree line. A -1.25x60 lens
moves the 60 degree line back without moving the
150 line. This collapses the conoid and all lines
clear. The Rx is -4.25 O-1.25 x 60.
We then do the effectivity calculations in each
For the minus cylinder refraction, we can derive a meridian exactly as in Chapter 6. For the vertical
clinically used rule called the rule of thirty. As the fog meridian a -12.00 D wavefront traveling right has
is reduced with the spherical lenses, one of the clock its center of curvature at 100/(-12), or -8.33 cm.
lines clears first. According to the rule, take the From the cornea, the center of curvature is
-(8.33 + 1.3) cm or -9.63 cm. The vergence at the
smaller clock number of the first clear line (e.g., 2 cornea is 100/(-9.63) cm or -10.38 D. The ver-
o'clock), multiply it by 30 degrees, and set the minus gence changes show that the vertical meridian's
cylinder axis at the resulting angle (e.g., 2 x 30 = 60). diverging wavefront loses curvature as it moves
Then add minus cylinder power until all lines are from the spectacle plane to the cornea. The wave-
clear. front center of curvature in the vertical meridian is
at the eye's far point for that meridian. The eye's
vertical meridian is myopic and the far point is real.
The same procedure for the horizontal meridian
gives -8.06 D at the cornea (again, a loss in curva-
15.14 Effectivity and Spherocylindrical Lenses ture). The horizontal meridian is also myopic and
its far point is real and at the center of curvature of
In the principal meridians of a toric wavefront, the the wavefront in the horizontal meridian. The ver-
vergence effectivity considerations are identical to gence cross at the cornea is

150

180

ocular view ocular view FIGURE 15.36. Ocular view of a clock dial.
Astigmatism: On Axis 321

s
-10.38 D
a) Ui . V,

-8.06 D

From the vergence cross, Harry's Rx at the cornea distant object


is - 8 . 0 6 O - 2 . 3 2 x 180.
Note that the minus cylinder axis did not
change. Also, note that the effectivity calculations
were done for the -12.00 D and - 9 . 0 0 D principal
meridian powers, and not for the -3.00 D cylinder
power. If you use a -3.00 D in the spectacle plane,
the result at the cornea is -2.89 D, which does not
equal the -2.32 D of cylinder power in the answer.
near object
EXAMPLE 15.18 FIGURE 15.37. Ocular accommodative demand.
Mary Ape has a + 1 5 . 0 0 0 - 5 . 0 0 x 1 4 6 correction
at the cornea. What is Mary's spectacle correction than a spectacle corrected myope viewing the same
at an 11 mm vertex distance?
object at the same distance. Because of vergence
The power cross at the cornea is
effectivity, a spectacle corrected astigmat may have
+ 15.00 + 10.00 different ocular accommodative demands in the princi-
pal meridians. Since accommodation is spherical, this
means that the astigmatism at near would no longer be
completely corrected. This effect is noticeable only for
high astigmats. Fortunately, in most astigmats the
For the 146 meridian, the center of curvature of the amount of astigmatism is low (3.00 D or less).
wavefront is 100/( + 15D), or +6.67 cm. Clearly,
the 146 meridian of the eye is hyperopic and the far EXAMPLE 15.19
point is virtual. From the center of curvature, the Cornelius Hooper wears a distance vision +12.00
spectacle plane is +(6.67 + 1.1) cm, or +7.77 cm. C - 9 . 0 0 x 180 spectacle correction at a 12 mm ver-
The power in the 146 meridian of the spectacle tex distance. How much uncorrected astigmatism
plane is 100/(+7.77 D ) , or +12.88 D. A converging does Cornelius have when threading a needle held
wavefront moving from the spectacle plane to the 10 cm in front of his lens?
cornea gains curvature, and +12.88 D at the spec- We need the ocular accommodative demand in
tacle plane to +15.00 D at the cornea is consistent each principal meridian. From Chapter 6, the
with this. spherical equation is
The same procedure for the 56 meridian gives a
power of +9.01 D in the spectacle plane. The An = U f , Uv
resulting power cross in the spectacle plane is where U fp is the vergence incident on the eye when
the person looks through the lens at a distant
+ 12.88 +9.01 object (Figure 15.37a) and U x is the vergence
incident on the eye when the person looks through
the lens at the near object (Figure 15.37b).
For the distant object, plane waves are incident
in each principal meridian, and the vergence leav-
ing the lens equals the dioptric power. There is a
and the Rx is + 1 2 . 8 8 0 - 3 . 8 7 x 1 4 6 . (Again, the vergence effectivity change as the wavefront moves
minus cylinder axis is the same, and the -5.00 D of to the cornea. The effectivity calculations give
cylinder power cannot be used directly.)

+3.00 D + 3.11 D
15.15 Accommodation and Spherocylindrical
Corrections
+ 12.00 D +14.02 D
As discussed in Chapter 6, vergence effectivity also
influences the ocular accommodative demand. A spec-
tacle corrected hyperope has to accommodate more spectacle plane cornea (U f )
322 Geometrie, Physical, and Visual Optics

For the near object, the incident vergence U on the mand in the horizontal meridian is 11.97 D. The
spectacle lens is -10.00 D. The vergence cross for difference in the accommodative demands between
the light leaving the spectacle lens is found from the principal meridians is 11.97 D minus 9.57 D,
V= P + U for each principal meridian. which equals 2.40 D. Corny, as he is called, will
have 2.40 D of uncorrected astigmatism when look-
ing through his distance correction at the needle. If
V P + U Corny accommodates +11.97 D, the vertical line
image in each conoid will be on the retina. If Corny
+3.00 D -10.00 D accommodates +9.57 D, the horizontal line image
in each conoid will be on the retina. If Corny
accommodates halfway between (i.e., +10.77D),
the circle of least confusion in each conoid would
+ 2.00 D + -10.00 D be on the retina.

-7.00 D 15.16 Meridional Telescopes

A 5X afocal Galilean telescope with a 40cm length


has a +2.00 D objective lens and a -10.00 D ocular
-2.00 D
lens. The telescope magnifies five times in all meri-
dians (refer to Section 13.13).
Now consider an afocal telescope consisting of a
The vergence effectivity gives +2.00 x 90 objective lens 40 cm in front of a -10.00 x
90 ocular lens. This is a meridional telescope that
-7.00 D -6.46 D magnifies horizontally but not vertically. When plane
waves are incident on the objective, plane waves leave
the ocular, and all meridians will be seen clearly
through the meridional telescope.
+2.00 D +2.05 D
Remember that magnification is not a function of a
single bundle. Magnification is a function of the rela-
tionship of one bundle to another. In other words,
spectacle plane cornea ( UJ magnification is a property of the beam, and not a
property of the individual bundles.
Then The power crosses for the meridional telescope's
objective and ocular lens are
A0 u fP - u
0 0
+3.11 D -6.46 D

+2.00 D -10.00 D
+14.02 D - +2.05 D

objective ocular

+9.57 D The power in the vertical meridian of both lenses is


zero. As a result, there is no magnification in the
vertical meridian. Figure 15.38a shows the vertical
+ 11.97D meridian ray diagram and Figure 15.38b shows the ray
diagram for the horizontal meridian. The magnifica-
tion occurs in the horizontal meridian because that is
where the power is. Consequently, a square with
horizontal and vertical sides viewed through the
The ocular accommodative demand in the vertical meridional telescope would appear to be a horizontal
meridian is 9.57 D, while the accommodative de- rectangle (Figure 15.39).
Astigmatism: On Axis 323

wave. However, the five plane waves are each travel-


ing at a slightly different angle relative to the axis. The
+2.00 x 90 objective lens, by itself, forms five con-
verging cylindrical wavefronts, and each wavefront
forms a real vertical line image (Figure 15.40b). The
-10.00x90 ocular lens changes each converging
vertical meridian
cylindrical wavefront back into a plane wave, and in
the process increases the horizontal angular separation
of the wavefronts. The result is a horizontal magnifi-
cation.
Consider an afocal meridional telescope with a
+2.00 x 180 objective lens and a -6.00 x 180 ocular
lens. The power meridians are vertical, and the 3X
magnification occurs in the vertical meridian. Through
the telescope, a square with horizontal and vertical
sides is imaged as a vertical rectangle (Figure 15.41a),
and a circle is imaged as a vertical ellipse (Figure
15.41b).
horizontal meridian Suppose the square with horizontal and vertical
FIGURE 15.38. Cross-sectional rays for a meridional telescope. sides has diagonal bars in it. Each diagonal bar makes
an angle of 45 degrees with the horizontal. Through
the above telescope, the square appears as a vertical
rectangle. The horizontal and vertical bars in the
square remain horizontal and vertical in the rectangle.
However, the diagonal bars are deviated, and no
longer make an angle of 45 degrees with the horizon-
object image tal. For a unit square object, the apparent rectangle
FIGURE 15.39. Object and conjugate image. has a three-unit vertical side together with the unit
horizontal side (Figure 15.42a). The apparent diagon-
als then make an angle of 71.6 degrees with the
For the above 5 X meridional telescope, the power horizontal (i.e., tanw = 3/1w = 71.6degrees). In
in the horizontal meridian forms a vertical line image, the above case, the apparent diagonals are rotated
and yet that same horizontal power magnifies horizon- (like scissors) in opposite directions relative to the
tally. Consider the five-point target in Figure 15.40a. diagonals in the object. The bars aligned with the
Diverging spherical waves leave each point. After principal meridians made no rotational movement.
traveling a long distance, each wave becomes a plane Consider a 3 X afocal telescope with a +2.00 x 45

a)

+ 2.00x090

b)
FIGURE 15.40. a. Object points, b. Real line images formed by
objective lens. The magnification relates to the horizontal separa-
tion of the vertical line images.
324 Geometrie, Physical, and Visual Optics

objective lens and a 6.00 x 45 ocular lens. The mag-


nification is in the 135 meridian. Figure 15.42b shows a
distant clock as viewed through the telescope. Re-
a) member that this is the ocular view, and the 2:30 to
8:30 line is the 135 meridian. The 2:30 to 8:30 line is
object
magnified three times while the 10:30 to 4:30 line is
unmagnified.
image
Figure 15.42c shows how a square with horizontal
and vertical sides would appear through the 3 X x 45
telescope. The 135 meridian is magnified and the 45
meridian is unmagnified. Since the diagonal bars are
aligned with the principal meridians, they are not
b)
deviated. However, the horizontal and vertical bars
are not aligned with the principal meridians, and they
now appear to have made a scissors movement.
object
In general, meridional telescopes cause deviations
image in meridians that are not the principal meridians. The
deviation increases with rotational distance from the
FIGURE 15.41. Objects and conjugate images.
principal meridians, so the worst meridians are those
45 degrees away from the principal meridians. No
deviation occurs in the principal meridians.

EXAMPLE 15.20
a)
W = 71.6 An afocal telescope has a 20 cm length, and a
+ 4 . 0 0 0 - 1 . 0 0 x 9 0 objective lens. What is the
magnification of the telescope? What are the
parameters of the ocular lens?
Either answer may be found first. The appropri-
ate telescope magnification equations are:

M=
obj

b) 1
M=
1 - d P obj
The power cross for the objective lens is

+4.00 D
ocular view

+3.00 D

The second equation applied to each principal


meridian gives 5 X vertically and 2.5 X horizon-
tally:

5X

ocular view 2.5 X


FIGURE 15.42. a. Scissors effect, b. Oblique distortion of a
clock dial. c. Oblique distortion of a square.
Astigmatism: On Axis 325

From vergence effectivity calculations, the ver- EXAMPLE 15.21


gence incident on the ocular lens is Marjorie Rigamarole has a thin +2.00O-5.00X
180 correction at a 10 mm vertex distance. What is
+20.00 D the percent gain or loss?
The power cross is

-3.00 D
+7.50 D
+2.00 D

The ocular lens power cross is then


The distance to the entrance pupil is 13 mm, and
-20.00 D the small magnification approximation gives

-3.9%
-7.50 D
+2.6%

and the parameters are -7.50 O-12.50 x 180. The


magnification can now be double-checked with the There is a 3.9% loss in the vertical meridian and a
first magnification equation. 2.6% gain in the horizontal meridian. A square
with horizontal and vertical sides is imaged as a
Near meridional magnifiers can also be developed. horizontal rectangle.
However, a simple near magnifier introduces astig-
matic blur as well as meridional magnification. Again, For small percent differences, the human visual
deviation occurs in all meridians other than the princi- system perceptually adapts to horizontal or vertical
pal meridians. In streak retinoscopy, the latter fact is magnifications, and once adapted Marjorie would still
used to identify the principal meridians of a patient's perceive a square even though her retinal image is
eye. actually a horizontal rectangle. The human visual sys-
tem does not perceptually adapt to oblique magnifica-
tions such as might occur with a strong spectacle
correction for an oblique astigmat. Clinically, one
15.17 Spectacle Magnification: Power Factor might wish to use contact lenses for such a patient, or
undercorrect the astigmatism, or rotate the cylinder
A thin spherocylindrical spectacle lens corrects the axis of the spectacle correction toward horizontal or
astigmatic refractive error. Thus, the spectacle lens vertical. The latter two methods sacrifice some clarity
together with the refractive error constitutes an afocal of the retinal image in order to provide less distortion.
system. The equation
1
Mn
l-dbenP 15.18 Spectacle Magnification: Shape Factor
can be applied to each principal meridian to obtain the
spectacle magnification in that meridian. Consider spectacle lenses with one toric and one
The distance dben is the distance from the lens to spherical surface. When the spherical surface is on the
the entrance pupil. Technically, the entrance pupil front, the shape factor magnifies equally in all meri-
position may be slightly different for each principal dians. When the toric surface is on the front, the
meridian. However, it is usually a good approximation shape factor gives a different magnification in each
to assume that the entrance pupil is 3 mm behind the principal meridian. The total magnification is the
cornea regardless of the principal meridian powers. product of the shape and power factors in each princi-
Furthermore, for small percent magnifications, we pal meridian.
can use the equation For a principal meridian, the shape factor magnifi-
cation is
S p ower = d b e n P v ( 1 0 0 % ) ,

Mclshape
in each principal meridian. 1 _ c^ '
326 Geometrie, Physical, and Visual Optics

where Px is the front surface power and c is the 2. A point source is located 50 cm in front of a
reduced central thickness. +4.50 0 + 3 . 0 0 x 9 0 spherocylindrical lens. Find
The small magnification approximation for the per- the line image positions and specify the orienta-
cent gain is tions. Also, find the position of the circle of least
confusion.
SshaPe = cP, (100%).
3. Specify the power cross for a +5.75 O - 3.35 x 64
When the small magnification approximations are lens. Also, specify the spherical equivalent and
valid, the shape and power factor percent gains are the equivalent sphere-plus cylinder combination.
additive. 4. Each of the following patients has horizontal and
vertical principal meridians. The eye is repre-
EXAMPLE 15.22 sented by a thin spherocylindrical lens in air lo-
Sue Ellen Shape wears a +4.00C-3.00 x 165 oph- cated 16.67 mm in front of the screen (retina). For
thalmic crown (n = 1.523) spectacle lens at an a distant point source, specify the part of the
11mm vertex distance. The lens has a + 10.00 0 conoid on the retina for the uncorrected eye,
+3.00x75 front surface and a 7.2 mm central classify the astigmatism, and specify the correction
thickness. Use the small magnification approxim- as a sphere-minus cylinder combination. You may
ations to figure the total spectacle magnification
percent gain. neglect any vertex distance effects.
For the 75 meridian: a. Susan Vosvertical power +62.00 D, horizon-
tal power +60.00 D.
Sshape = 7 , 2
^ 2 3 3 m ( + 1 0 0 0 )(100%) b. Al Hootvertical power +62.00 D, horizontal
power +58.00 D.
-4.7%, c. Sammy Bavisvertical power +58.00 D, hor-
Spower = 14 x 10"3 m ( + 1.00 D)(100%) izontal power +56.00 D.
d. Bonnie Ardenvertical power +58.00 D, hor-
-1.4%, izontal power +60.00 D.
Stot = Sshape + Spower = 4.7% + 1.9% e. Henry Kissingvertical power +64.00 D, hor-
-6.1%. izontal power +62.00 D.
f. Wayne Porkvertical power +54.00 D, hor-
For the 165 meridian: izontal power +59.00 D.
5. For each of the patients in the previous problem
Sshape = ? ' 2 1.5^3 3m ( + 13.00D)(100%) specify what lines on a distant clock dial would
-6.1%, appear darkest through the following lenses. Lens
1: +2.00x180 0 - 2 . 0 0 x 9 0 . Lens 2: -2.00 x
Spower = 14 x 10"3 m (+4.00 D)(100%) 1800+2.00x90.
-5.6%, 6. Specify the sphere-minus cylinder combination
that is equivalent to -0.25 x 180 O+0.25 x 90.
Stot = S,hape + Spower = 6.1% + 5.6%
7. When looking at a distant point source through a
-11.7%. -5.00 D lens, Gerald Weakman (unaccommo-
In a power cross format, the results are dated) clearly sees a line parallel to the 60 meri-
dian. When looking through a 3.00D lens,
/6.1% Gerald clearly sees a line at 150. What is Gerald's
correction in minus cylinder form?
8. Mickey Rookie (unaccommodated) looks at a dis-
tant point source through a +2.00 D lens and sees
/ 11.7% a clear line parallel to the 160 meridian. When
Mickey looks through a -1.00D lens, he sees a
The exact equations give 6.5% for the 75 meridian clear line parallel to the 70 meridian. What is
and 12.9% for the 165 meridian. Mickey's correction in minus cylinder form?
9. Joyce Jinked (unaccommodated) looks at a distant
point source through a -4.50 D lens and reports
Problems that the 11 to 5 o'clock line is clear. When Joyce
looks through a 2.50 D lens, she reports that the
1. A point source is located 25 cm in front of a 2 to 8 o'clock line is clear. What is Joyce's correc-
+6.00 x 180 cylindrical lens. Where is the real line tion in minus cylinder form?
image? What is its orientation? 10. Kevin Wallflower (unaccommodated) looks at a
Astigmatism: On Axis 327

distant point source through a +2.25 D lens and how a distant square would appear through the
reports that the 10 to 4 o'clock line is clear. When telescope.
Kevin looks through a +3.25 D lens, he reports 17. Ryan O'Coughin has a +10.00 O-4.00 x 90 spec-
that the 1 to 7 o'clock line is clear. What is tacle correction at a 12 mm vertex distance. What
Kevin's correction in minus cylinder form? is the spectacle magnification (percent loss or
11. When uncorrected, Kevin Wallflower of the previ- percent gain) in each principal meridian? When
ous problem looks at the distant clock dial chart. spectacle corrected, Ryan looks at a distance
What accommodation could make all lines equally square. Is Ryan's retinal image a square, a rec-
visible? tangle longer vertically than horizontally, or a
12. Tim Battleaxe looks at a distant target consisting rectangle that is longer horizontally then vertical-
of vertical and horizontal lines. Tim (unaccommo- ly. (Assume the entrance pupil is 3 mm behind the
dated) sees the horizontal lines clearly when a cornea.)
-6.50 D lens is in front of his eye. The vertical 18. Mike Keats wears a - 7 . 5 0 o - 6 . 0 0 x 180 specta-
lines are blurred in this case. A slit aperture (a cle correction at a 13 mm vertex distance. What is
stenopeic slit) is now placed against the -6.50 D the spectacle magnification percent gain or loss in
lens. What orientation of the stenopeic slit each principal meridian? Is Mike's retinal image
(horizontal or vertical) clears the vertical lines of a distant square a rectangle? If so, which side is
without changing the horizontal lines? larger? (Assume Mike's entrance pupil is 3 mm
13. Betty Peg (unaccommodated) looks at a distant behind his cornea.)
target consisting of vertical and horizontal lines. 19. A distant square has horizontal and vertical sides.
Betty clearly sees all lines when looking through a Cheri Tree views the square through her +2.00 O
horizontal stenopeic slit combined with a +4.50 D -5.00 x 135 spectacle lens. Cheri's vertex distance
lens. When the stenopeic slit is vertical, Betty is 12 mm, and the entrance pupil is 3 mm behind
clearly sees all lines when a +6.00 D lens is in the cornea. Give the spectacle magnification per-
place. What is Betty's correction in minus cylinder cent loss or percent gain. Is Cheri's retinal image a
form? square, a rectangle, or a parallelogram? (Hint:
14. June Trucks has a -7.25 0 4 . 5 0 x 1 5 8 spectacle Consider the magnification for the diagonals of
correction at a 13 mm vertex distance. What is the square.)
June's correction at the cornea? 20. Clem Obtuse has a +11.00O-7.00 x 180 specta-
15. Fib Corn has a +8.75 0 - 5 . 0 0 x 5 6 correction at cle correction worn at a 13 mm vertex distance.
the cornea. What is Fib's spectacle correction at a For a thin lens eye model, find the ocular accom-
12 mm vertex distance? modative demand in each principal meridian when
16. A meridional telescope consists of a +2.00 x 90 Clem views an object 17.4 cm from his spectacle
objective lens located 33.33 cm in front of a lens. Since accommodation is spherical, how much
-6.00 x 90 ocular lens. List the magnification in uncorrected astigmatism does Clem have when
the horizontal and vertical meridians. Describe viewing the near object through his spectacles?
CHAPTER SIXTEEN

Dioptric Power and


Off-Axis Meridians

16.1 Curvature and Torsion single plane, and so the principal meridians have zero
torsion. The optical effects of the principal meridians
Curvature has been an important component in the are then due only to the curvature of these meridians.
optics of the previous chapters. In particular, the However, when we deal with off-axis meridians, the
dioptric power of a single spherical refracting surface torsion is non-zero. For the off-axis meridians, both
equals the product of the refractive index difference the torsion and the curvature contribute to the optical
times the curvature of the surface. effects.
In general, a curve in space is completely specified
at each point by two parameters, the curvature k and
the torsion w. The torsion w is a measure of the rate at
which the curve twists out of a plane. An example of a 16.2 Sturm's Conoid and Twisting
curve with a constant non-zero curvature and a con- Sheets of Light
stant non-zero torsion is a circular helix, such as the
curve formed by the thread on a bolt. The non-zero Consider a paraxial astigmatic bundle of light that has
torsion means that the helix does not stay in a single passed through a square aperture. The astigmatic bun-
plane. Conversely, the torsion is zero for a curve that dle forms the conoid of Sturm (Figure 16.2a) with its
stays in a single plane. For example, the torsion at any two perpendicular line images aligned with the princi-
point on an ellipse is zero. pal meridians. All the light (hence all the rays) in the
Consider a curve on a cylinder such that at each conoid first goes through line image 1 (horizontal),
point the curve makes an angle with the axis meri- and then proceeds and passes through line image 2
dian. For 0< <90, this curve spirals up the cylin- (vertical).
der and in fact is a circular helix which has a non-zero The ray on the upper right corner (marked J)
torsion (Figure 16.1). The off-axis meridian defined by moves down and to the left as it moves through the
making a constant angle with the axis meridian lies conoid. Similarly, the ray on the lower left corner
along this circular helix and has a non-zero torsion. moves up and to the right as it moves through the
Similarly, on a torus, an off-axis meridian specified conoid. When a slit aperture (clinically called a
by a curve that makes a constant angle with a principal stenopeic slit) is used to isolate this diagonal meridian,
meridian spirals around the torus and does not stay in these rays form the end rays of a sheet of light that
a single plane. (An inflated inner tube can be used to twists around as it moves through the conoid. The
illustrate this principle.) So the off-axis meridian of a diagonal line in Figure 16.2b shows the orientation of
torus is also helical in nature and has a non-zero the twisting sheet at various positions in the conoid. In
torsion. particular, the sheet twists around so that it is aligned
The curves that specify the principal meridians of a with the first line image when it passes through that
torus or of a cylinder are curves (circles) that stay in a position, and then continues to twist so that it is

329
330 Geometrie, Physical, and Visual Optics

behavior of the isolated sheets of light is the fact that


iQ the off-axis (or non-principal) meridians have a non-
zero torsion component, while the principal meridians
have no torsion. The torsion causes the off-axis sheets
to twist as they move through the conoid, while the
principal meridian sheets do not twist. The torsion
also prevents the formation of a point image by the
off-axis sheets, while the on-axis sheets do form a
FIGURE 16.1. The helix that makes a constant angle with the
axis meridian. point image.
The meridians of a spherical lens, or the on-axis
meridians of a spherocylindrical lens, have zero tor-
sion and a dioptric power that is due only to the
curvature of those meridians. An off-axis meridian of
a spherocylindrical lens has a dioptric power compo-
nent that is due to the curvature of the off-axis
meridian and another dioptric power component that
is due to the torsion of the off-axis meridian. The
torsional component is responsible for the off-axis
twisting sheet of light.

16.3 Dioptric Power Components


For a single toric refracting interface with parameters
a) S ^ C x 0, the dioptric power curvature component Pk
for a meridian that makes an angle with the axis
meridian is given by
Pk = S + C sin2 . (16.1)

J J 10
J J LI
In fact, when we factor the refractive index difference
out of Eq. 16.1, we have left the sine squared form of
the differential geometry equation for the (normal)
b) curvature of that meridian.
The magnitude of the dioptric power torsional com-
FIGURE 16.2. a. Conoid of Sturm for a square aperture. J ponent Pt for this meridian is given by
marks the ray in the upper right corner, b. Cross sections
showing the twisting sheet of light. |P t | = |Csin cos |. (16.2)
When we factor the refractive index difference out of
Eq. 16.2, we have left the differential geometry equa-
aligned with the second line image when it passes tion for the magnitude of the (relative) torsion for that
through that position. Clearly, the twisting sheet does meridian. The sign of the torsional component de-
not stay in the same plane. Furthermore, the twisting pends on the coordinate system in use. For a specific
sheet does not form a point image. Actually, the coordinate system, the dioptric power equations are
twisting sheet was already present before the stenopeic derived in Section 16.5.
slit was introduced. The stenopeic slit simply blocks The 45-45-90 right triangle and the 30-60-90 right
the other rays so that the twisting sheet is readily triangle are helpful in developing intuition about the
apparent. effects of the sine and cosine terms (Figure 16.3). In
When the stenopeic slit is used to isolate a principal the 45-45-90 right triangle, the sides adjacent to the
meridian of the conoid, the light shows a different right angle are equal, and the Pythagorean theorem
behavior. There is no twisting, the rays stay in the gives side ratios of 1, 1, /2. Then
principal meridian, and they converge to form one of 2
the points on the line image that is perpendicular to sin 45 = cos 45 = 1 /V2, so sin 45 = 1 /2.
that meridian. From the 30-60-90 triangle, the side opposite the 30
The underlying geometric basis for the varying degree angle is one half the hypotenuse. The Pyth-
Dioptrie Power and Off-Axis Meridians 331

principal meridians and fastest for the meridian half-


way between the principal meridians ( =45).
Figure 16.4b shows a plot of the dioptric power
/AS 90| torsional component's magnitude as a function of the
1 angle from the axis meridian. The torsional compo-
a) b)
nent is zero on the principal meridians ( = 0 and
FIGURE 16.3. . 45-45-90 right triangle, b. 30-60-90 right = 90), and has its maximum magnitude of one half
triangle. the cylinder power at the meridian halfway between
the principal meridians (=45). Note that if the
cylinder power C is zero, the torsional component is
agorean theorem gives side ratios of 1, V3, 2. Then zero for any meridian.
sin 30 = 1 / 2 and sin2 30 = 1 / 4 , A lens clock can be used to measure the curvature
of any meridian on a toric surface. With the proper
sin60 = cos30 = V3/2 and sin2 60 = 3/4. refractive index conversion, the lens clock reading is
the dioptric power curvature component for that meri-
Figure 16.4a shows a plot of the curvature compo-
dian. A lens clock can be used to indirectly get the
nent as a function of the angle with the axis
magnitude of the dioptric power torsional component
meridian. For the axis meridian, equals zero. It
for a meridian (assuming zero experimental error).
follows that the sine term is zero, and the curvature
One procedure is to take lens clock readings on
component is equal to S. For the power meridian,
meridians that are 45 on either side of the meridian in
equals 90, the sine term is 1, and the curvature question, subtract one reading from the other, and
component is S + C. These are the extreme values of divide by 2, i.e.,
the dioptric power curvature component. The vari-
ation of the sine squared term is slowest around the |P.I = l(P b + 4 5 - P b - 4 5 ) | / 2 , (16.3)

s+c

a)

FIGURE 16.4. a. Dioptric power curvature


component as a function of the angle with the
axis meridian, b. Magnitude of dioptric power
torsional component.
332 Geometrie, Physical, and Visual Optics

where Pt is the torsional component for meridian b


and the terms in the parentheses are the lens clock
readings (with the appropriate conversion for index) at
the meridians indicated. Equation 16.3 can be derived
from Eqs. 16.1 and 16.2.

EXAMPLE 16.1a
What does a lens clock read on the 60, 90, 105,
120, and 150 meridians of a +6.00 O - 4 . 0 0 x 150
toric surface (n = 1.53)?
The principal meridians are 60 and 150, and the
respective powers are +2.00 D and +6.00 D. Since
no conversion for index is needed, these are the
respective lens clock readings in the principal meri-
dians. All other lens clock readings fall between FIGURE 16.5. Lens clock readings on a + 6 . 0 0 C - 4 . 0 0 x 150
+2.00 D and +6.00 D. toric surface.
The vertical (90) meridian is closest to the 60
principal meridian, so we expect the lens clock
reading to be closer to +2.00 D than to +6.00 D.
From Eq. 16.1,
Pk = +6.00 D + (-0.27 D) = +5.73 D.
Pk = +6.00 D + (-4.00 D) sin 2 (90 - 150),
For the 45 meridian,
Pk = +6.00 D + (-4.00 D)(3/4),
Pk = +6.00 D + (-4.00 D) sin 2 (45 - 150)
and
Pk = +6.00 D + (-3.00 D) = +3.00 D. = +2.27D.

The 105 meridian is halfway between the principal Then


meridians, so we expect the lens clock reading to |P t | = |(5.73 - 2 . 2 7 ) | / 2 = 1.73 D.
be halfway between +2.00 D and +6.00 D. From For the check, Eq. 16.2 gives
Eq. 16.1,
|P t | = K-4.00 D) sin(90 - 150) cos(90 - 150)|,
Pk = +6.00 D + (-4.00 D) sin (105 - 150),
|P t | = |(-4.00D)(-0.87)(0.50)|,
or
|Ptj = 1.73D.
Pk = +6.00 D + (-4.00 D ) ( l / 2 ) = +4.00 D.
The 120 meridian is closer to 150, so we expect the In conjunction with the sine squared curvature
reading to be closer to +6.00 D. From Eq. 16.1, variation of a toric surface, there is a relationship
known as Euler's law. In terms of lens clock readings,
Pk = +6.00 D + (-4.00 D) sin2(120 - 150),
Euler's law states that for a toric surface the sum of
any two perpendicular lens clock readings is a con-
Pk = +6.00 D + (-4.00 D ) ( l / 4 ) = +5.00 D. stant. Consider a toric surface described by S O C x .
The readings are recorded on the power cross in Then Euler's constant E is given by
Figure 16.5. E = [S + C sin 2 ] + [S + C sin2(< 90)].

EXAMPLE 16.1b Since


From lens clock readings, what is the magnitude of sin 2 (0 90) = cos 2 ,
the dioptric power torsional component in the ver-
tical meridian of the same surface? Does this check we have
with Eq. 16.2?
We expect the magnitude of the torsional com- E = S + C sin 2 + S + C cos 2 </>,
ponent to lie between 0 and half the cylinder power
or
(2.00 D). The vertical meridian is near the halfway
meridian between the principal meridians, so we E = 2S + C ( s i n 2 0 + c o s 2 0 ) .
expect the torsional component to have a mag-
Since
nitude close to 2.00 D.
From Eq. 16.3, we need to take lens clock sin 2 + cos 2 = 1,
readings at 45 on either side of the vertical meri-
dian. The lens clock reading for the 135 meridian is it follows that
2
Pk = +6.00 D + (-4.00 D) sin (135 - 150), E = 2S + C. (16.4)
Dioptrie Power and Off-Axis Meridians 333

According to Eq. 16.4, the sum of any two perpen- not stay in one meridian. The optical axis and a
dicular lens clock readings is equal to the constant E. meridian form a plane. A ray that lies in that plane is
For a spherical surface, the curvature is the same in all called a meridional ray, while a ray that is not in that
meridians, and E equals 2S. For a cylinder, E is equal plane is called a skew ray. In general, the emergent
to C. rays in an off-axis pencil that has passed through a
spherocylindrical lens are skew rays.
EXAMPLE 16.2 Chapter 12 specified a paraxial meridional ray in
Use the 90 and 180 meridians to find Euler's con- terms of its distance h from the optical axis and its
stant for a -5.00O-2.00 x 30 surface. (reduced) angle q with the optical axis. Figure 16.6
The lens clock reading in the 90 meridian is shows a ray together with a horizontal (x) and vertical
Pk = -5.00 + (-2.00) sin2(90 - 30), (y) coordinate system. The optical axis lies along the z
Pk = -5.00 + (-2.00X3/4) = -6.50 D. axis. A paraxial skew ray can be described in terms of
The lens clock reading in the 180 meridian is its orthogonal components in the x-y system. At the
first x-y plane, the ray position components relative to
Pk = -5.00 + (-2.00) sin2(30 - 0), the optical axis are hx and h y , and the angular direc-
Pk = -5.00 + (-2.00)(l/4) = -5.50 D. tion components relative to the optical axis are qx and
The sum is qy. At the second x-y plane, which is labeled x'-y' and
is a reduced distance t downstream, the ray compo-
E = -6.50 D + -5.50 D = -12.00 D. nents are hx, hy, qx, and qfy
We can use Eq. 16.4 as a check: For translation purposes, both the x and the y
E = 2(-5.00) + (-2.00), meridians behave just like the meridians in Chapter
E = -10.00 + (-2.00) = -12.00 D. 12. The only difference is that now the work is dou-
We can also apply Euler's law to the principal bled since we have to keep track of two meridians.
meridians. Here, the principal meridian powers are During the translation, the ray angles do not change,
-5.00 D and -7.00 D, and their sum is indeed so
-12.00 D.

It is possible to express the dioptric power equa- and


tions in terms of an angle g with the power meridian as
opposed to the angle with the axis meridian. The
form of the torsional component equation does not From Eq. 12.3, the position components are related
change, while the form of the curvature component by:
becomes h
x = hx + tq x ,
Pk = S + Ccos 2 g. (16.5) hy = h y + t q y .
Both sine squared and cosine squared terms appear in
the following formalism.

16.4 The Translation Matrix

We can derive the equations for the off-axis dioptric


power components either from ray tracing or from
vergence considerations. Sections 16.4 through 16.7
/
(h'x.h'y)
/ I
7\ t
y
y'

/ 1
give the ray tracing derivation expressed in terms of
the matrix formalism. (It is assumed that the reader is /
ry/ / 1

familiar with the matrix formalism of Chapter 12.) The

l_
rest of the chapter gives some applications of the / / ""
formalism. q
The rays in an astigmatic bundle may be part of a
\l///^ *
pencil that is aligned with a principal meridian, or they (hx,hy)
may be part of a pencil that is aligned with an off-axis
meridian. The off-axis pencils twist as they pass FIGURE 16.6. Translation of a paraxial skew ray in a horizon-
through the astigmatic bundle, and thus the rays do tal and vertical coordinate system.
334 Geometrie, Physical, and Visual Optics

The four parameters for the ray in the x-y plane can 1 0
be put into a 4 x 1 ray matrix A: 1= (16.11)
0 1
0 is the 2 x 2 null (or zero) matrix

(16.6) (0 0\
A= (16.12)
0=
0 0
\h y / and
A similar matrix, A', exists for the ray components at 't 0\
the x'-y' plane. Then the four equations describing the (16.13)
translation can be written as a single matrix equation: \0 X,

jl 0 0 0\ /q,\ Note the similarity between the partitioned form of


the 4 x 4 translation matrix Eq. 16.10, and the 2 x 2
0 10 0 qyl translation matrix Eq. 12.5 for a spherical system.
(16.7)
K t 0 1 0 U
\o t o 1/ w 16.5 The Refraction Matrix

The matrix multiplications in Eq. 16.7 simply give The translation matrix gets a ray from one plane to
the four translation equations. Equation 16.7 can be another. We also need a refraction matrix to get a ray
written in a shorthand version by using the ray mat- through a thin spherocylindrical lens. (Remember that
rices defined in Eq. 16.6: in a reduced system, each single refracting surface is
represented by a thin lens.) Since the principal meri-
A' = TA, (16.8) dians of the spherocylindrical lenses may not be
aligned with the coordinate system, we need to look at
where T is the translation matrix,
the relations between different coordinate systems.
/l 0 0 0\ Figure 16.7 shows a u-v coordinate system that is
rotated around the z axis by an angle relative to the
0 1 0 0 x-y coordinate system. A ray can be described in the
T = (16.9) u-v system by its position components hu and h v , and
t 0 1 0 by its angular components qu and qv. A 4 x 1 ray

\o t o 1/
The 4 x 4 translation matrix, Eq. 16.9, can be par-
titioned into four 2 x 2 blocks.
/I 0 0 0\
I

10 1 0 01
T =
1 * 1 0|

\o t 0 1/
and then written as

T = (16.10)

FIGURE 16.7. The u-v coordinate system is rotated through


where 1 is the 2 x 2 unit matrix the angle relative to the x-y (horizontal and vertical) system.
Dioptrie Power and Off-Axis Meridians 335

matrix in the u-v system is change so

h' = h .
qv
A. 1 1 V (16.14) The above four equations can be written as the single
matrix equation

Kl
lq'"\ /l 0 -s o \
/q
There are well-established mathematical relations be-
qv 0 1 0 - ( S + C) q
tween coordinate systems that differ only by a rota-
tion. From these relations, h
K 0 0 1 0
hu = hx cos 0 + hy sin 0,
hv = - h x sin 0 + hy cos 0. \KI \o o 0 1 / \h
(16.18)
In the small angle approximation, the angles q are
directly related to the directions h. Therefore, the
angles qu and qv are related to the angles qx and qy by or
the equations ' = R A (16.19)
^ uv uv '
qu = qx cos 0 + qy sin 0, where Ruv is a paraxial refraction matrix for the u-v
qv = - q x sin 0 + qy cos 0. coordinate system:
Then the above four equations can be written as one /l 0
matrix equation relating the parameters of the same
-s
0
ray expressed in the u-v system to those expressed in \
0 1 0 - ( S + C)
the x-y system: R..v
(16.20)
AUV = GA, (16.15) 0 0 1 0
where \o o 0 1 /

cos 0 sin 0 0 0 Equations 16.15 and 16.17 can be substituted into Eq.
16.19 to give
-sin0 cos0 0 0 GA' = RUVGA.
G=
0 0 cos0 sin0 Then we can multiply both sides from the left by G"1
to obtain
0 0 -sin0 COS0/
(16.16) G GA' = G RUVGA.
_1
Since G G equals the 4 x 4 unit matrix, we can write
Similarly, for the image space ray, this as
AV = GA. (16.17) (16.21)
RA,
Assume that a thin spherocylindrical lens with para-
where
meters S O C x 0 has its principal meridians aligned
with the u-v coordinate system shown in Figure 16.7. R = G R G. (16.22)
Then the dioptric power in the u meridian is S and the Note that Eq. 16.21 relates the object and image space
dioptric power in the v meridian is S + C. The refrac- ray parameters expressed in the horizontal and vertical
tion relations for each meridian are identical to those in x-y meridians even when the principal meridians of the
Chapter 12 for a spherical lens. In particular, from spherocylindrical lens are not horizontal and vertical.
Eq. 12.8, In that sense, the 4 x 4 matrix R is the refraction
qu = q u - h u s , matrix expressed in a horizontal and vertical coordi-
nate system for a spherocylindrical lens in which the
q; = q v - h v ( S + C). principal meridians are not necessarily horizontal and
At the thin lens, the ray's position coordinates do not vertical.
336 Geometrie, Physical, and Visual Optics

The matrix G is refraction matrix for a spherical lens. In fact, the


similarity shows that, for a spherocylindrical lens of
parameters S c C x 0, the 2 x 2 dioptric power matrix
/cos0 -sin 0 0
\ P is the generalization of the dioptric power for a
spherical lens.
sin cos 0 0
G = Consider a ray coming in parallel to the optical
0 0 cos0 -sin0 axis. From Eq. 16.6, the ray matrix for the incident
ray is
\ 0 0 sin 0 cos 0 I
(16.23)
l\
which can be verified by showing that the product 0
G G 1 equals the 4 x 4 unit matrix. We can substitute A=
Eqs. 16.16, 16.20, and 16.23 into Eq. 16.22, and do
the 4 x 4 matrix multiplications to obtain
V
/ o -P X \ From Eq. 16.21, the matrix multiplications for the
0 1 -Pt angles of the exiting ray give
R (16.24)
q; = - P x h x - p t h y ,
0 0 1 0
q; = - P t h x - P y h y .
\o o o 1 /
When the horizontal meridian is isolated by a
where stenopeic slit aperture, hy is zero and then
Px = S + C sin2 0, (16.25) qx = -P x h x ,
2
Pv = S + C cos 0, (16.26) q; = -p t h x -
and The above two equations show that Px and Pt are the
P = - C s i n 0 cos 0. (16.27) dioptric power components of the horizontal meridian.
For a single toric surface, a comparison with the
The refraction matrix R can also be partitioned in general differential geometry equations for curvature
four 2 x 2 matrices: shows that Px is the dioptric power curvature compon-
ent and Pt is the dioptric power torsional component.
-px -pt\
/' The curvature component causes a ray to bend (re-

R=
'o i -pt
-v fract) in that meridian, while the torsional component
causes the ray to bend (refract) perpendicular to the
meridian. In that sense, the torsional component cou-
0 0 1 0 j
ples the two perpendicular meridians x and y, and the
perpendicular refraction is the basis for the twisting
\0 0 0 1 /
sheet of light.
or A similar argument shows that Py is the dioptric
power curvature component and the other off diagonal
element Pt is the dioptric power torsional component
(16.28) for the vertical meridian. Note that for a thin sphero-
0 1 cylindrical lens, the torsional component in the verti-
where 1 is the 2 x 2 unit matrix, 0 is the 2 x 2 null cal meridian equals the torsional component in the
matrix, and P is a 2 x 2 dioptric power matrix horizontal meridian.

'P. R
P= (16.29)
,P, P, 16.6 The System Matrix

Note the similarity between the partitioned form of Chapter 12 discussed the 2 x 2 spherical system mat-
the 4 x 4 refraction matrix R and Eq. 12.9, the 2 x 2 rix, which was the appropriate product of the individu-
Dioptrie Power and Off-Axis Meridians 337

al 2 x 2 refraction and translation matrices. For combi- lens. For a thin spherocylindrical lens with parameters
nations of spherocylindrical lenses, there is a 4 x 4 S C C x , the matrix P expressed in a horizontal and
system matrix S, which is just the product of the vertical coordinate system is given by Eqs. 16.25
individual 4 x 4 refraction and translation matrices, through 16.29. For convenient reference, these equa-
i.e., for j separated spherocylindrical lenses, tions are repeated here:
s^RjT,- .j-
TaR^R,. (16.30) 'P. Pt
The principal meridians of the spherocylindrical P= (16.29)
lenses in the system do not have to be aligned. When iP. P.
the principal meridians are not aligned, the lenses are
referred to as obliquely crossed. where
Consider two obliquely crossed spherocylindrical Px = S + C sin2 0, (16.25)
lenses in contact. For the zero separation, the transla- 2
tion matrix equals the unit matrix, and the system Pv = S + C cos 0, (16.26)
matrix is given by and
S = R2R1, (16.31) Pf= - C s i n 0 c o s 0 . (16.27)
or Px and Py are the diagonal elements in the dioptric
-P2\/I Pi
power matrix. (The diagonal elements are those for
S= (16.32) which the row number equals the column number:
0 1 0 1 such as first row-first column or second row-second
column.) The diagonal elements are given by sine
In partitioned form, the 4 x 4 matrix multiplications squared and cosine squared terms characteristic of the
give curvature components. The off-diagonal elements
have the sine-cosine product characteristic of the tor-
(P 2 + P,) sional components. Since P is expressed in a horizon-
S= tal and vertical coordinate system, the angle used is
1 just the cylinder axis 0.
For a thin spherocylindrical lens, the two off-diag-
or
onal elements both equal P t . Thus, while the dioptric
power matrix P has four elements, only three of them
S= (16.33) are independent for a thin spherocylindrical lens. Note
that S, C, and 0 are a set of three independent
parameters. Similarly, P x , Py, and Pt are a set of three
where independent parameters.
Pe=P 2 + Pl (16.34) Consider a +5.00C-3.00 x 180 lens. Figure 16.8a
shows the power cross. The dioptric power matrix is
The above equations show that, just as the dioptric
powers are additive for two spherical lenses in contact, +5.00 D 0
the dioptric power matrices are additive for two thin
spherocylindrical lenses in contact, even when the 0 +2.00 D
lenses are obliquely crossed. The resultant matrix Pe is
the dioptric power matrix of a single spherocylindrical The off-diagonal elements Pt equal zero, while the first
lens that is equivalent to the obliquely crossed combi- row-first column element Px is the power in the
nation. The additivity of the dioptric power matrices is horizontal meridian and the second row-second col-
important since the spherocylindrical parameters umn element Py is the power in the vertical meridian.
S OC x 0 are not additive for obliquely crossed lenses. In this case, the dioptric power matrix is just a dis-
guised power cross.
The dioptric power matrix for a - 1 . 0 0 x 9 0 cylin-
drical lens is
16.7 The Dioptric Power Matrix -1.00 D 0
P=
The previous sections show that a thin spherocylindri- 0 0/
cal lens has a 2 x 2 dioptric power matrix P, which is
the generalization of the dioptric power for a spherical As shown in Figure 16.8b, P is again just a disguised
338 Geometrie, Physical, and Visual Optics

As one might expect, the above two matrices add to


give the matrix for a +5.00 C-3.00 x 180:
+5.00 D 0
+ 5.00D
0 +2.00 D
The dioptric power matrix for a spherical lens of
+ 2.00D power S is easy to recognize. Pt is zero, and the
diagonal elements both equal the sphere power. The
dioptric power matrix for a spherocylindrical lens with
horizontal and vertical principal meridians is also easy
to recognize. Again, Pt is zero, but the diagonal
elements are not equal to each other.
In general, Eq. 16.27 shows that any spherical lens
or any spherocylindrical lens with horizontal and verti-
cal principal meridians has zero off-diagonal compo-
nents. As long as the dioptric power matrix P is
-1.00D
expressed in a horizontal and vertical coordinate sys-
tem then spherocylindrical lenses with principal meri-
dians that are not horizontal and vertical will have a
non-zero off-diagonal element.
b) 0.00D
Consider a +5.00 C-3.00 x 170 lens. The axis 170
lens differs from the +5.00C-3.00 x 180 used above
by a 10 rotation. Figure 16.8c gives the power cross
for the axis 170 lens. The dioptric power matrix
elements are computed as follows. From Eq. 16.26,
Py = +5.00 + (-3.00) cos2(170),
+ 5.00D Py = +5.00 + (-3.00X-0.985) 2 = +2.09 D.
From Eq. 16.25,
+ 4.91D
Px =+5.00 +(-3.00) sin2(170),
Px = +5.00 + (-3.00X+0.174) 2 = +4.91 D.
From Eq. 16.27,
C)
Pt = - ( - 3 . 0 0 ) sin(170) cos(170),
+ 2.00D Pt = -(-3.00)(0.174)(-0.985) = 0.51.
FIGURE 16.8. a. Power cross for a +5.00 C - 3 . 0 0 x 180 lens. So
b. Power cross for a - 1 . 0 0 x 9 0 lens. c. Power cross with
horizontal and vertical dioptric power curvature components for +4.91 D -0.51 D
a + 5 . 0 0 O - 3 . 0 0 X 180. P=
-0.51 D +2.09 D
The off-diagonal elements, the torsional components,
power cross. Similarly, for a -3.00 x 180, are no longer zero. The vertical component Py is close
to the principal meridian value of +2.00 D, and the
/0 0 horizontal component Px is close to +5.00 D.
P= When the principal meridians are horizontal and
0 -3.00 D vertical, they are independent of each other, or uncou-
For a +5.00 D spherical lens, the dioptric power pled. Then a pencil of incident rays in the vertical
matrix is (horizontal) meridian passes through the lens and
stays in the vertical (horizontal) meridian. When the
+5.00 D 0 principal meridians of the lens are not vertical and
P= horizontal, the vertical and horizontal meridians are
0 +5.00 D coupled together by the torsional component P t . Then
Dioptrie Power and Off-Axis Meridians 339

a pencil of incident rays in the vertical or horizontal off-diagonal element P t . For the +5.00 0 - 3 . 0 0 x 4 5 ,
meridian twists out of that meridian after passing Pt is positive, while for the +5.00 O - 3 . 0 0 x 135 lens,
through the lens. Pt is negative.
The further the principal meridians are from In general, for a first quadrant minus cylinder axis
horizontal and vertical, the more the off-diagonal (i.e., less than 90), Pt is positive, while for a second
element Pt differs from zero. The largest difference quadrant minus cylinder axis (i.e., greater than 90), Pt
occurs when the principal meridians are at 45 and 135, is negative. The difference occurs because the cosine is
in which case the off-diagonal element has a mag- negative in the second quadrant. This information
nitude of C/2. provides a quick way to check the sign of P t .
Consider a + 5 . 0 0 0 - 3 . 0 0 x 1 3 5 lens. From Eq.
16.26, EXAMPLE 16.3
2 Make some estimates for the elements of P for a
Py = +5.00 + (-3.00) cos (135), -3.00 0 - 2 . 0 0 x 2 0 lens, and then compute the
Py = +5.00 + (-3.00X0.5) = +3.50 D. matrix.
Figure 16.9 shows the power cross. Based on the
From Eq. 16.25, power cross, Px is between -3.00 D and -5.00 D,
Px = +5.00 + (-3.00) sin2(135) = +3.50 D. and is much closer to -3.00 D. An estimate of
-3.50 D seems reasonable. Similarly, Py is between
From Eq. 16.27, -3.00 D and -5.00 D, and is much closer to
Pt = - ( - 3 . 0 0 ) sin(170) cos(170), -5.00 D, so an estimate of -4.50 D seems
reasonable.
Pt = -(-3.00X0.174X-0.985) Pt is positive, and is between C/2 and zero, so
= -0.51. an estimate of +0.75 seems reasonable. Equations
16.25 through 16.27 give
So /-3.23D +0.64 D\

/+3.50D -1.50D\ 1+0.64D -4.77 D/

1-1.50D +3.50D/ The estimates were a little off, but making the
estimates provided a strong intuitive feeling for
For the +5.00 0 - 3 . 0 0 x 4 5 lens: From Eq. 16.26, whether or not the numerical values for the matrix
elements are calculated correctly.
Py = +5.00 + (-3.00) cos2(45) = +3.50 D.
The dioptric power matrix does not depend on
From Eq. 16.25,
whether the lens parameters are expessed in plus or
Px = +5.00 + (-3.00) sin2(45) = +3.50 D. minus cylinder form. For example, the matrix for a
+ 1.00 O - 4 . 0 0 x 140 lens is computed as follows.
From Eq. 16.27, From Eq. 16.26,
Pt = - ( - 3 . 0 0 ) sin(45) cos(45), Py = +1.00 + (-4.00) cos2(140),
Pt = -(-3.00)(0.707)(0.707) = -(-3.00)(0.5) Py = +1.00 + (-4.00X-0.766) 2 = -1.35 D.
= +1.50D.
So

/+3.50D +1.50D\

1+1.50D +3.50D/
For the axis 45 or 135 lenses, Pt is equal in mag-
nitude to C/2, while the diagonal elements are equal
to each other, and are dioptrically halfway between
the two principal meridian powers. (In fact, equal
diagonal elements with a non-zero off-diagonal ele-
ment means that the principal meridians must be at 45
and 135.)
The only difference between the above matrices for FIGURE 16.9. Power cross together with horizontal and verti-
the axis 45 and axis 135 lenses is in the sign of the cal meridians for a -3.00 -2.00 x 20 lens.
340 Geometrie, Physical, and Visual Optics

From Eq. 16.25, properties are another manifestation of Euler's law.


Px = +1.00 + (-4.00) sin2(140), When the matrix P is describing a toric surface, the
sum of Px and Py corresponds to the sum of a horizon-
Px = +1.00 + (4.00X+0.643)2 = -0.65 D. tal and a vertical lens clock reading. Therefore, t must
From Eq. 16.27, equal Euler's constant E (Eq 16.4).
Pt = - ( - 4 . 0 0 ) sin(140) cos(140), The determinant d of a 2 x 2 matrix is the product
of the diagonal elements minus the product of the
Pt = -(-4.00)(0.643)(-0.766) = -1.97. off-diagonal elements. For P
So
d = PxPy-P?. (16.37)
-0.65 D -1.97 D
P= From Eqs. 16.25 through 16.27
-1.97D -1.35D
d = (S + C sin2 0)(S + C cos2 0)
The transposed equation for the +1.00O-4.00 x 140
is - 3 . 0 0 0 + 4 . 0 0 x 5 0 . From the plus cylinder equa- -(C2sin20cos20),
tion, the calculations for the matrix are:
which, after using a little algebra, gives
Py = -3.00 + (+4.00) cos2(50) = -1.35 D,
.2/
Px = -3.00 + (+4.00) sinz(50) = -0.65 D, d = S(S + C). (16.38)
Pt = -(+4.00) sin(50) cos(50) = -1.97D.
Equation 16.38 shows that the determinant d is also
So independent of the cylinder axis 0, and is equal to the
-0.65 D -1.97 D product of the powers in the two principal meridians.
P= The minus cylinder equation, the dioptric power
-1.97 D -1.35D matrix, and the trace and determinant are shown
below for three orientations of a -5.00 O-1.00 x 0
which is identical to the previous matrix. Thus, neither lens. From Eq. 16.36, the trace should equal -11.00.
the power cross nor the dioptric power matrix depends From Eq. 16.38, the determinant should equal
on whether the parameters are expressed in plus or +30.00 D . For a -5.00 0 - 1 . 0 0 x 1 8 0 ,
minus cylinder form. Given a thin spherocylindrical
lens, it has a unique dioptric power matrix P. The /-5.00D 0
\
spherocylindrical lens has two sets of spherocylindrical
parameters, the plus cylinder set and the minus cylin-
P=
\ 0 -6.00 D /
L
der set (related by transposition).
t- = (-5.00) + (-6.00)i =-11.00 D,
d~-= ( - 5 . 0 0 ) ( - 6 . 0 0 ) - 02 =+30.00 D2.

For a -5.00 0 - 1 . 0 0 x 7 0 ,
16.8 The Trace and Determinant of a Dioptric /-5.88D 0.32 D \
Power Matrix P=
\ 0.32 D --5.12D/
L
The trace, t, of a matrix is defined as the sum of the
diagonal elements. For the dioptric power matrix P, t == (-5.88) + (-5.12)=-11.00 D,
d == ( - 5 . 8 8 ) ( - 5 . 1 2 ) - (0.32)2,
i = Px + Py. (16.35)
d == (+30.10)-(0.10) =+30.00 D2.
From Eqs. 16.25 and 16.26,
For a - 5 . 0 0 O - l . 0 0 x 120,
t = S + C sin2 0 + S + C cos2 0,
i = 2S + C(sin 2 0 + cos 2 0), /-5.75D --0.43 D \
or
P=
\-0.43D --5.25 D /
L
t = 2S + C. (16.36)
t = (-5.75) + (-5.25) = -11.00 D,
Equation 16.36 shows that the trace is independent
of the cylinder axis 0 and depends on the sphere d= (-5.75X-5.25) - (-0.43) 2 ,
power S and the cylinder power C. Actually, the trace d = (+30.19) - (0.19) = +30.00 D2.
Dioptrie Power and Off-Axis Meridians 341

16.9 Finding the Spherocylindrical Parameters EXAMPLE 16.4


from the Matrix Given the dioptric power matrix,

Suppose one wishes to determine the spherocylindrical /-0.73 -1.68\


parameters S c C x from the dioptric power matrix \-1.68 +0.23/
P. When Pt is zero, the principal meridians are
horizontal and vertical. Then we can determine the find the minus cylinder S, C, and 0.
spherocylindrical parameters just as we do from a The torsional component Pt is negative, so the
power cross. minus cylinder axis is a second quadrant angle.
When Pt is not zero, the principal meridians are not From Eq. 16.35, the trace is
horizontal and vertical. However, the trace and the t= -0.73 + 0.23 =-0.50.
determinant equations supply two equations in two From Eq. 16.37, the determinant is
unknowns for S and C. From Eq. 16.36, d = (-0.73X+0.23) - (-1.68) 2 = -2.99 D2.
S = ^ . (16.39) Then from Eq. 16.40,
C = - V(-0.50) 2 -4(-2.99) = -3.49 D.
We can substitute Eq. 16.39 into 16.38 and solve for C From Eq. 16.39,
to obtain:
-0.50-(-3.49)
C=Vi2-4d. (16.40) *~ 2
+2.99
Given the matrix P, we can compute its trace and S = ^ = +1.50D.
determinant and then use Eq. 16.40 to find C. Once C
(Be careful with the arithmetic in the numerator.)
is known, Eq. 16.39 gives S. Then from Eq. 16.41,
Given S and C, we can find the cylinder axis from:
1.50-(-0.73)
*= -1.68 --1-33
tan0 = ^ - ^ , (16.41)
or
or 0 = tan _ 1 (-1.33)=-53.
0 = tan _ 1 [(S-P x )/P t ]. (16.42) In standard axis notation,
0 = -53 + 180 = 127.
You can verify Eq. 16.41 using Eqs. 16.25 and 16.27. The answer is +1.50O-3.49 x 127.
For a first quadrant cylinder axis, Eq. 16.42 returns The axis is a second quadrant axis, which we
the axis 0 directly. For a second quadrant cylinder expected. Also, we can use S and C and compute
axis, Eq. 16.42 returns the negative angle 0m shown in the trace and determinant as checks. From Eq.
Figure 16.10. Then 0 in standard axis notation is found 16.36,
by adding 180 degrees to 0m. t = 2( + 1.50) + (-3.49) = -0.50 D.
The plus or minus sign in Eq. 16.40 is an arbitrary From Eq. 16.38,
choice, and simply determines whether the sphero-
d = ( + 1.50)(-2.00) = -3.00 D2.
cylindrical parameters will be in plus or minus cylinder
form. The plus or minus cylinder answers are related Both the trace and determinant check.
by transposition.
EXAMPLE 16.5
Given the dioptric power matrix,
I-9.66 + 1.12\
I+1.12 -6.34/
find the minus cylinder S, C, and 0.
Since the torsional component is positive, the
minus cylinder axis is a first quadrant angle.
From Eq. 16.35, the trace is
t= -9.66 + (-6.34) = -16.00.
From Eq. 16.37, the determinant is
FIGURE 16.10. Relation of negative angle 0m to standard axis
notation. d = (-9.66X-6.34) - ( + 1.12)2 = +59.99 D2.
342 Geometrie, Physical, and Visual Optics

Then from Eq. 16.39, Figure 16.11 shows the power cross addition. The
answer is -3.00 O - 1 . 0 0 x 150.
C = - V ( - 1 6 . 0 0 ) 2 - 4 ( + 5 9 . 9 9 ) = -4.00 D.
From Eq. 16.40, When the principal meridians of the two sphero-
-16.00-(-4.00) cylindrical lenses in contact are not aligned, then the
S
~ 2 powers in the principal meridians are not additive.
However, the derivation of Eq. 16.34 showed that the
dioptric power matrices are additive, and that there is
a single spherocylindrical lens that is equivalent to the
Then from Eq. 16.41,
obliquely crossed combination. Therefore, we can
-6.00-(-9.66) compute the dioptric power matrix for each lens, add
t a n = = + 3 2 7
H12 ' the matrices, and then compute the spherocylindrical
or parameters of the equivalent lens from the resultant
0 = t a n - 1 ( - 3 . 2 7 ) = 73. matrix. In the following discussion, the single lens
equivalent to the combination is referred to as the
The answer is -6.00 -4.00 x 73. resultant lens.
The axis is a first quadrant axis, which we
expected. Also, we can use S and C and compute
the trace and determinant as checks. EXAMPLE 16.7
From Eq. 16.36, The -3.00 x 60 lens in the above example is ro-
tated 30 degrees to -3.00 x 90. The two lenses in
t = 2(-6.00) + (-4.00) = -16.00 D. contact are then -3.00 x 90 and -4.00 x 150. What
From Eq. 16.38, are the resultant spherocylindrical parameters?
If the cylinder axis of the two lenses were
d = (-6.00)(-10.00) = +60.00 D 2 .
aligned, then the answer would be 0 O - 7 . 0 0 x 0.
The trace checks and the determinant shows a 0.01 In the previous example, the two lenses were per-
difference. The 0.01 difference is due to round-off pendicularly crossed, and the result was -3.00
error. When the calculations are done to three -1.00 x 150. So in the obliquely crossed case, the
digits after the decimal, the determinant a- resultant sphere power is somewhere between the 0
grees exactly. and -3.00 D, while the resultant cylinder power is
somewhere between the -7.00 D and -1.00 D. To
find the solution we add the dioptric power mat-
rices.

16.10 Adding Obliquely Crossed


Spherocylindrical Lenses Lens Matrix

Consider two spherocylindrical lenses in contact. -3.00 x 90 (-30 )


When the principal meridians of two lenses are
aligned, we can add the powers in the respective vi/vwicn /-1.00 -1.73\
principal meridians (as considered in Chapter 15). -4.00X150 (_173 _300)
o H /-4.00 -1.73 \
EXAMPLE 16.6 Resultant [_1J3 _3m)
A -3.00 x 60 lens is combined with a -4.00 x 150
lens. What are the spherocylindrical parameters of From the resultant matrix, Eqs. 16.35 and 16.37
the single lens that is equivalent to the combi- give the trace and determinant.
nation?
The principal meridians of these two lenses are / = -4.00 + (-3.00) = -7.00,
aligned, and so we can add the respective powers. d= (-4.00X-3.00) - (-1.73) 2 = +9.01 D 2 .

-4.00

FIGURE 16.11. Power cross addition for aligned


lenses.
Dioptrie Power and Off-Axis Meridians 343

Then Eqs. 16.38 through 16.41 give the sphero- For the above example, the spherical equivalent of
cylindrical parameters. the -3.00 x 90 is -1.50 D, while the spherical equival-
ent of the -4.00 x 150 is -2.00 D. The spherical
C = - V(-7.00) 2 -4(+9.01) = -3.60,
equivalent of the resultant lens must be
s = -7.00 - (-3.60) = _L70> - 1 . 5 0 + (-2.00) = - 3 . 5 0 D.

-1.70-(-4.00) From the computed parameters,


tan=
or
^V^=-L33' Sp Eq = -1.70 + = -3.50 D,
= -53. which checks.
In standard axis notation, The second rule concerns the resultant cylinder
axis. The rule can be stated either in terms of plus or
= -53 + 180 = 127.
minus cylinder axis. Consider the two minus cylinder
The single lens equivalent to the combination is axes of the obliquely crossed lenses. One of the angles
- l . 7 0 c - 3 . 6 0 x 127. between the two minus cylinder axes is an acute angle
(angle w in Figure 16.12a). The rule is:
Several rules can be developed to help us judge
whether the final answer makes sense or not. The first The minus cylinder axis of the resultant lens lies in the
rule concerns the additivity of the spherical equival- interval occupied by this acute angle. When the minus
ents. Whenever two matrices add, it is easy to show cylinder powers of the obliquely crossed lenses are equal,
that the traces also add. The trace t of a spherocylin- the resultant minus cylinder axis lies exactly in the middle
drical lens equals 2S -I- C. The spherical equivalent of the interval occupied by the acute angle (H in Figure
(Sp Eq) equals S + C/2. Therefore, 16.12b). When one of the minus cylinder powers is grea-
ter in magnitude than the other minus cylinder power, the
SpEq = i/2. resultant minus cylinder axis is closer to the axis of the
stronger lens. The amount closer depends on the differ-
Since the traces add, the spherical equivalents also
ence in the powers.
add. Thus, the rule is:

For any two spherocylindrical lenses in contact, even In Example 16.7, the obliquely crossed lenses were
obliquely crossed, the spherical equivalent of the resultant a - 3 . 0 0 x 9 0 and a -4.00x150. According to the
lens is equal to the sum of the spherical equivalents of the rule, the interval occupied by the 60 acute angle is
individual lenses. from 90 to 150. If the two cylinder powers were equal,

a) b)

00x90

FIGURE 16.12. a. The acute angle w between two obliquely


crossed minus cylinder axis meridians, b. The midway meri-
dian H. c. The resultant minus cylinder axis is tilted from H
toward the minus cylinder axis meridian of the lens with the
most cylinder power.
344 Geometrie, Physical, and Visual Optics

then the resultant axis would be exactly halfway in d = (-13.803X-11.197) - (2.164) 2


between or 120 (H in Figure 16.12c). Since the = +149.85 D 2 .
powers are not equal, the resultant minus cylinder axis
is closer to the axis of the - 4 . 0 0 D cylinder. The From Eq. 16.39,
calculated minus cylinder axis was 127. C = - V(-25.00 D) 2 - (4X + 149.85 D 2 ),
C=-5.06D.
EXAMPLE 16.8
From Eq. 16.38,
A -8.00 0 - 2 . 0 0 x 2 2 lens is combined with a
- 1 . 0 0 0 - 5 . 0 0 x 7 2 . First, find the spherical equi- s = -25.00 D - ( - 5 . 0 6 D ) _ 9 9 7 D
valent of the resultant lens and estimate the minus
cylinder axis of the resultant lens. Then calculate
From Eq. 16.41,
the resultant lens parameters.
The spherical equivalents add: -9.97 D - ( - 1 3 . 8 0 D)
t a n
*= 2I6D U + 1
7 6

or
0 = t a n - 1 ( + 1.76) = 6O.5.
Lens Sp Eq
The resultant lens parameters are - 9 . 9 7 O - 5 . 0 6 x
-8.00 0 - 2 . 0 0 x 2 2 -9.00 D 60.5.
-1.000-5.00x72 -3.50 D The axis agrees well with the initial estimate.
Resultant -12.50 D The spherical equivalent of the resultant lens is

Sp Eq = -9.97 + ^ y ^ = -12.50 D,

The spherical equivalent of the resultant lens is which also agrees.


-12.50 D.
The acute angle interval between the minus
cylinder axis is 50 (i.e., from 22 to 72). If the
cylinder powers were equal, the resultant minus 16.11 Over-refraction
cylinder axis would be exactly halfway between, or
47. Here the resultant minus cylinder axis lies
closer to that of the -5.00 D lens. Let us guess that Over-refraction consists of performing a clinical re-
the resultant minus cylinder axis is around 60. We fraction while the person is wearing a spectacle or
do not know if that is correct, but we have some contact lens. The over-refraction lens determines what
intuition that tells us whether or not the calculated additional power that person needs. Over-refraction is
answer makes sense. particularly useful clinically in dealing with aphakes or
To calculate the resultant lens parameters, cal- other high ametropes. Sometimes the over-refraction
culate the dioptric power matrix for each lens and lens is obliquely crossed relative to the spectacle or
add the matrices. contact lens that the person is wearing.

EXAMPLE 16.9
An aphake is wearing a spectacle Rx of + 1 5 . 0 0 0
Lens Matrix 3.00 x 170. An over-refraction shows that he
needs an additional -0.50 0 - 0 . 5 0 x 1 5 0 . Assum-
/-8.28 +0.70\ ing thin lenses, what is the patient's new Rx?
-8.00 0 - 2 . 0 0 x 2 2 In this case, the over-refraction results in an
\+0.70 -9.72/ obliquely crossed lens problem. First make your
/-5.52 + 1.47\ expectations about the spherical equivalent and the
-1.000-5.00x72 resultant cylinder axis.
V + 1.47 -1.48/
/-13.80 +2.17
Resultant lens
\ +2.17 -11.20
Lens Sp Eq

From the sum matrix, the trace and determinant +15.00 O-3.00X 170 +13.50 D
are: -0.50 O-0.50 x 150 -0.75 D
Resultant +12.75 D
t = (-13.803) + (-11.197) = -25.00 D,
Dioptrie Power and Off-Axis Meridians 345

The spherical equivalent of the resultant lens is EXAMPLE 16.10


+ 12.75 D. A 4 D ocular myope is wearing an unknown contact
The acute angle between the minus cylinder axis lens fit on K. An over-refraction yields a -0.50 D
is 20 (from 150 to 170). If the cylinder powers lens. Neglecting vergence effectivity, what lens was
were equal, the resultant cylinder axis would be at placed on the eye?
160. Here the resultant cylinder axis will be closer The unknown contact lens and the over-refrac-
to that of the -3.00 D cylinder. Let us make an tion lens must add to give the -4.00 D correction:
estimate of 166.
-0.50D + P = - 4 . 0 0 D ,
The calculations are:

P=-3.50D.

EXAMPLE 16.11
Lens Matrix A patient needs a - 4 . 0 0 C - 2 . 0 0 x 90 contact cor-
rection fit on K. However, the contact is not orient-
+ 14.91 -0.51 ing correctly on the cornea. An over-refraction is
+15.00 O-3.00X 170 done and the result is a +0.45 0 - 0 . 9 0 x 1 2 8 . 5 .
-0.51 +12.09 Neglecting vergence effectivity, how is the contact
-0.63 -0.22 lens orienting on the eye?
-0.50O-0.50x 150 Here, just as in Example 16.10, the over-refrac-
-0.22 -0.88 tion lens together with the contact lens must be
equivalent to -4.00 0 - 2 . 0 0 x 9 0 . Since the mat-
+ 14.29 -0.73
Resultant rices are additive:
-0.73 +11.21 p __| p = p
over-refraction contact correction '

p =p - p
From the matrix sum, the trace and determinant contact correction over-refraction '

are -6.00 0
t= +14.29+ 11.21 = +25.50 D, 0 -4.00
d = ( + 14.29X + 11.21) - (-0.73) 2 = 159.80 D 2 . -0.10 -0.44
From Eq. 16.40,
-0.44 +0.10
C = - V(25.50 D) 2 - (4)(159.90 D 2 ),
-5.90 +0.44
C=-3.40D.
From Eq. 16.39, +0.44 -4.10
o +25.50-(-3.40) Then from Eqs. 16.39 through 16.41, the lens
S= ^ = +14.45 D. parameters are -4.00 0 - 2 . 0 0 x 7 7 . The minus
From Eq. 16.41, cylinder axis of the contact lens is supposed to be
90, but when placed on the eye the contact rotates
+ 14.45-14.29 13 to axis 77.
tan0 : -0.23,
-0.73
0 = tan - 1 (-O.23) = -12.7.
In standard axis notations, the minus cylinder axis 16.12 The Vector Addition Method
is -12.7 + 180=167.3.
Based on the estimate of 166 for the axis, the Besides the dioptric power matrix, there are other
calculated value of 167.3 makes intuitive sense. The methods for solving the obliquely crossed lens prob-
Sp Eq of the resultant lens is: lem. One of the other methods is a vector addition
approach.
Sp Eq = S + - , The vector addition approach can be derived from
the additivity of the dioptric power matrices as fol-
Sp Eq = +14.45 + i z M l = + Uj5 D, lows: For a spherocylindrical lens with parameters
SoCx0,
which also agrees.
To the nearest quarter diopter, the new Rx is / S + Csin20 -Csin0cos0\
+ 14.50 0 - 3 . 5 0 x 1 6 7 compared to the +15.00 0 P =
-3.00 x 170 that the aphake was wearing. ,-Csin0cos0 S + Ccos 0
346 Geometrie, Physical, and Visual Optics

From trigonometry, The addition of the dioptric power matrices meant


.2 1 - cos 20 that the three elements, Px, P y , and P t , are additive. In
sin2 = 2 ' the above equation for P, the matrix on the far right
has only two independent numbers: C cos 20 and
and C sin 20. The third independent number is the SpEq.
sin 20 Thus, the addition of the dioptric power matrices also
sin 0 cos 0 = -. means that the three numbers SpEq, Ccos20, and
Note that the argument of the sine and cosine on the C sin 20 are additive.
right side of the above two equations is two times the We discussed the additivity of the Sp Eq in Section
cylinder axis 0. The above two trigonometry equations 16.10. The new information is that C cos 20 and
can be substituted into the matrix equation for P to C sin 20 are also additive. The latter two numbers act
obtain like the components of a vector of length C and
direction 20. The additivity of the components means
C cos 20 C sin 20 \ that the cylinders satisfy vector addition rules provided
/s + 2 2 the angles used are 2 times the cylinder axis (i.e., 20).
P= This vector addition can be represented graphically
C C sin 20
- C sin 20 using a coordinate system in which the angles are
S+ w + doubled.
2 2 Figure 16.13a shows such a doubled angle coordi-
The above matrix can then be rewritten as the matrix nate system. The vector representing a +2.00 x 45
sum lens is plotted at 90 (i.e., the vector points straight
up). The vector representing a 4-2.00x90 lens is
plotted at 180 (i.e., pointing left in Figure 16.13b).
1 0 1 - C cos 20 - C sin 20
P = Sp Eq The vector representing a +2.00 x 135 lens is plotted
0 1 - C sin 20 +Ccos20 at 270 (or straight down in Figure 16.13c). The vector

a)
135

135

45 30

90 l
- )l80

FIGURE 16.13. Vector representations of cylinder power


and axis (double angle plots), a. +2.00 x 45. b. +2.00 x 90. c.
135 +2.00 x 135 (or -2.00 x 45). d. +2.00 x 180. e. +2.00 x 30.
Dioptrie Power and Off-Axis Meridians 347

representing a +2.00 x 180 lens is plotted at 360 (or Thus, the single lens equivalent to the combination
right in Figure 16.13d). The vector representing a has parameters +0.80 0+5.40x67. (You can
+2.00 x 30 lens is plotted at 60 (Figure 16.13e). check these results by adding the dioptric power
The vector for a minus cylindrical lens points in the matrices.)
opposite direction relative to the vector for a plus
cylindrical lens with the same axis. For example, the A nice feature of the double angle vector addition
vector for a +2.00 x 45 lens points up, and so the method for obliquely crossed lenses is that one can
vector for a -2.00 x 45 lens points down. Note that make a quick graphical sketch to estimate the re-
the vector pointing down also represents a +2.00 x sultant cylinder power and axis (Figure 16.14).
135 cylindrical lens. This is a manifestation of the
standard transposition properties (since a 2.00 x 45
lens transposes to a - 2 . 0 0 0 + 2 . 0 0 x 1 3 5 lens). Re-
member that the vector is representing only the cylin- 16.13 The Residual Refractive Error
der power and axis. The spherical power is hidden in
the SpEq. The correct ophthalmic lens (either spectacle or con-
On a double angle plot, the obliquely crossed tact) neutralizes the patient's refractive error. When
cylinder powers add vectorally to give the resultant the ophthalmic lens is not correct, there is a residual
cylinder power and axis. Then the Sp Eq sum can be refractive error. The dioptric power matrix for the
used to obtain the resultant sphere. residual refractive error equals the sum of the matrices
for the ophthalmic lens and the refractive error.
EXAMPLE 16.12 The sum of the matrices for the correct ophthalmic
Consider a +4.00x50 combined with a +3.00 x lens and the refractive error is the zero or null matrix.
90. What single spherocylindrical lens is equivalent As a by-product, the sum of the spherical equivalents
to the combination? of the correct ophthalmic lens and the refractive error
In Figure 16.14, the +4.00x50 is plotted at is also zero. Suppose the correct ophthalmic lens is
100, the +3.00x90 points to the left, and the
vector sum gives the resultant cylinder parameters rotated clockwise or counterclockwise relative to the
C x 0. The values are +5.40 x 67 (so the resultant patient's eye. Then the rotated lens and the refractive
vector C is plotted at twice 67, or 134). The error are obliquely crossed, and the matrix sum is no
resultant sphere value can then be computed from longer zero. Consequently, there is a non-zero residu-
the fact that the spherical equivalents add to al refractive error, and a conoid of Sturm exists in the
+3.50 D: patient's eye. However, the rotated lens still has the
same spherical equivalent, and therefore, the sum of
Sp Eq = S + the spherical equivalents of the rotated lens and the
2'
refractive error is still zero. Thus, in the case of the
gives
rotated ophthalmic lens, the residual refractive error
+3.50 = S+ (5.40/2), has a zero spherical equivalent.
or The correcting lens for an equal mixed astigmat
S=+0.80D. also has a zero spherical equivalent (i.e., the circle of
least confusion is on the retina of an uncorrected equal
mixed astigmat). So the residual refractive error in the
case of the rotated ophthalmic lens mimics the refrac-
tive error of an equal mixed astigmat. This means that
rotating the correct ophthalmic lens leaves the circle of
least confusion on the patient's retina. (This is man-
ifested in Example 16.11 where the +0.45 C - 0 . 9 0
x 128.5 over-refraction lens had a zero spherical
equivalent.)
For a small rotation of the correct ophthalmic lens,
180 the axis of the residual refractive error is approximate-
ly 45 away from the axis of the lens. We can show this
with the vector addition method.
In Figure 16.15, the vector A pointing up and to
the right represents the refractive error. The vector for
the correction should have the same length and point
FIGURE 16.14. Vector sum for cylinder power. in the opposite direction (along the dashed line).
348 Geometrie, Physical, and Visual Optics

/I
/
/
n
/
/

FIGURE 16.15. A is the refractive error, B is the


tentative correction, and E is the residual refractive
error.

Suppose the ophthalmic lens is rotated slightly so that We can write the parameters of a JCC in
its vector B points straight down. The vector sum E of spherocylindrical form (e.g., 4-0.25 0 - 0 . 5 0 x 1 8 0 ) .
the refractive error A and the rotated lens B repre- The JCC can be rotated clockwise or counterclockwise
sents the residual refractive error. From Figure 16.15, (e.g., a 17 rotation gives +0.25 0 - 0 . 5 0 x 1 7 ) . The
E is almost perpendicular to B. Since this is a double JCC can also be flipped about a meridian that is 45
angle plot, this means that the axis of the residual away from the principal meridian. The flip simply
refractive error is almost 45 from the axis of the lens exchanges the powers in the principal meridians.
B. For even smaller rotations of the lens, the axis of When flipped, a +0.25 O-0.50 x 180 JCC becomes a
the residual refractive error gets closer to being 45 +0.25 0 - 0 . 5 0 x 9 0 JCC.
from the lens axis. The JCC is particularly useful for refining a tenta-
tive cylinder power and axis in a correction. The axis
refinement is done as follows. A tentative correction is
determined in some manner (clockdial, retinoscopy,
16.14 Axis Refinement with a Jackson old Rx, etc.). If the tentative cylinder power and axis
Cross Cylinder are not quite right, a conoid of Sturm still exists in the
patient's eye. The spherical lens power is then adjus-
The Jackson Cross Cylinder (JCC) was introduced in ted to give the best acuity for a letter chart. This
Sections 15.9 and 15.11. The JCC is frequently used in adjustment places the circle of least confusion on the
the determination of the lens that corrects a person's retina. Then a JCC is placed in front of the tentative
refractive error. Conceptually, a JCC consists of a plus correction with the minus cylinder axis of the JCC 45
cylinder perpendicularly combined with a minus cylin- away from the minus cylinder axis of the correcting
der that has a power equal in magnitude (e.g., lens (e.g., for a tentative correction of -1.50 x 80, the
-0.25 x 180 combined with +0.25 x 90). Because the JCC is placed at a minus cylinder axis of either 35 or
powers are equal in magnitude but opposite in sign, 125). Since the JCC has a zero spherical equivalent, it
the spherical equivalent of a JCC is zero. leaves the circle of least confusion on the patient's
Dioptrie Power and Off-Axis Meridians 349

retina. The JCC is then flipped (i.e. if position #1 is +2.00x80 (i.e., he or she needs a correction of
+0.25 0 - 0 . 5 0 x 3 5 , the flipped position, #2, is -2.00 x 80). Suppose the tentative correction given is
+0.25 0 - 0 . 5 0 x 1 2 5 ) . The circle of least confusion a -1.50 x 90. In Figure 16.16a, A is the vector for the
remains on the retina during the flip, but may increase refractive error and B is the vector for the -1.50 x 90
or decrease in diameter. The patient looks through the lens. The residual refractive error is the vector sum E
lenses at an extended target (probably letters) and of the refractive error and the tentative correction.
reports which flipped position gives the better image. (The cylinder and axis part of the residual refractive
The better image corresponds to the flipped position error is +0.78 x 59.5.)
that results in the smaller circle of least confusion on A +0.25 O-0.50 x 45 JCC is now placed in front
the patient's retina. The minus cylinder axis of the of the correcting lens. The JCC together with the
correcting lens is then rotated toward the minus cylin- residual refractive error give the effective refractive
der axis of the JCC position that gives the clearest error for the patient's vision through the JCC and the
image. The JCC is then readjusted so that its minus tentative correction. On a double angle plot, the
cylinder axis is again 45 from the minus cylinder axis vector for the minus cylinder axis 45 JCC points down,
of the new correction, and the process is repeated. while the vector for the flipped position (minus cylin-
The process continues until the flipped positions give der axis 135) points up. Figure 16.16b shows the
equally blurred results (i.e., one is not better than the vector sum #1 of the residual refractive error E
other). This endpoint occurs when the size of the combined with the JCC minus cylinder axis 45 and the
circle of least confusion is unchanged during a flip. vector sum #2 for E combined with the flipped JCC
The correcting cylinder axis at the endpoint is the (i.e., minus cylinder axis 135). The vector sum #1 for
refined or improved cylinder axis. the JCC pointing down is smaller in magnitude than
In the axis refinement process, the JCC and the the vector sum #2 for the JCC pointing up. The
correcting lens are obliquely crossed lenses. The smaller magnitude #1 means that the effective refrac-
theory of the refinement process can be demonstrated tive error is less, and the patient has a smaller circle of
graphically with the vector addition approach. least confusion on the retina. The patient will report
Consider a person with a refractive error of #1 is better than #2. The position #1 occurs for the

FIGURE 16.16. a. A is the refractive error, B is the


tentative correction, and E is the residual refractive
error, b. # 1 and # 2 are the vector sums of the residual
refractive error with the two flipped positions of the
JCC. c. Same as the previous part for the correct
a) cylinder axis.
350 Geometrie, Physical, and Visual Optics

b)

FIGURE 16.16. (Cont'd)


Dioptric Power and Off-Axis Meridians 351

JCC minus cylinder axis 45. Once the patient reports


#1 as better, the examiner moves the tentative axis of
the correction from 90 towards 45.
v.(v-v)
Suppose the tentative correction axis is placed at where
80. Now the refractive error is + 2.00 x 80 and the Vx = Sv + C v sin 2 0 v ,
tentative correction is -1.50 x 80. The residual refrac-
tive error is +0.50 x 80, which is represented by vec- Vy = Sv + C v cos 2 0 v ,
tor E in Figure 16.16c. The JCC positions now have
minus cylinder axis 35 and minus cylinder axis 125. On Vt = -C v sin0 v cos0 v .
the double angle plot, the JCC vectors are now per- Here Vx is directly related to the horizontal curvature
pendicular to E. The vector sums #1 and #2 repre- of the toric wavefront, Vy is directly related to the
sent the effective refractive error through the JCC and vertical curvature, and Vt is directly related to the
the tentative correction. Here # 1 and #2 are equal in wavefront torsion in the horizontal and in the vertical
magnitude, which means that the circle of least confu- meridians. Given the matrix V, the spherocylindrical
sion stays the same when the JCC is flipped. Thus, the parameters can be extracted from the same trace and
patient reports that the two positions are equal (actu- determinant equations that work for the dioptric
ally equally blurred). Since we have equality, the power matrix P.
minus cylinder axis 80 is the refined or true axis of the Consider a toric wavefront incident on a toric
correction. surface (or thin spherocylindrical lens). The matrices
In summary, the endpoint of the test occurs when are additive even when the wavefront and the surface
the axis is true, and, thus, the JCC vector is perpen- are obliquely crossed, i.e.,
dicular to the residual refractive error vector. Then
V = P + U, (16.43)
the two flipped positions of the JCC give the same
length for the resultant vectors # 1 and #2 that repre- where U is the 2 x 2 vergence matrix for the incident
sent the effective refractive error, and the circle of wavefront V is the 2 x 2 matrix for the emergent
least confusion on the patient's retina is unchanged in wavefront, and P is the 2 x 2 matrix for the surface.
size when the flip occurs. The patient reports that #1
and #2 are equal. When the tentative axis is not true, EXAMPLE 16.13
the vector for the JCC is not perpendicular to the What is the back vertex power for a 6 cm thick
residual refractive error vector, and then one of the bitoric glass (n = 1.50) lens with a + 10.00C-2.00
JCC positions gives a shorter sum (either #1 or #2). x90 front surface and a -6.00 0 - 5 . 0 0 x 1 3 5
back surface?
The circle of least confusion on the patient's retina is The back vertex power is the vergence leaving
smaller for the shorter sum, and the patient reports the back of the system when plane waves are
that choice as better. The tentative minus cylinder axis incident on the front. The front surface creates a
of the correction is then turned toward the better toric wavefront with horizontal and vertical princi-
minus cylinder JCC and the process is repeated. A pal meridians. This wavefront then propagates ac-
tremendous advantage of the JCC process is that the ross the interior of the lens and is incident on the
cylinder axis can be refined even when the cylinder back surface.
power is not quite correct. Once the axis refinement is For incident plane waves, the vergence leaving
made, then the power can be refined. (In the power the front surface is +10.00 D in the vertical meri-
refinement case, the principal meridians are aligned, dian, and +8.00 D in the horizontal meridian.
and so the theory follows that of Chapter 15.) Downstream vergence calculations in each meri-
dian show that the vergence incident on the back
surface is +16.67 D in the vertical meridian and
+ 11.76 D in the horizontal meridian. The sphero-
cylindrical parameters for the wavefront incident
16.15 The Vergence Matrix on the back surface are then +16.670-4.90
x 90. The vergence matrix is
The generalization of the dioptric power for a toric / + 11.76D 0 \
surface is the 2 x 2 dioptric power matrix P. Similarly,
the generalization of the vergence for a toric wave- \ 0 +16.67 D/
front is a 2 x 2 vergence matrix V.
Consider a toric wavefront with spherocylindrical The dioptric power matrix for the back surface is
vergence parameters S v c C v x v . Then the vergence /-8.50D -2.50 D\
matrix is obtained with equations analogous to 16.25
through 16.29: \-2.50D -8.50 D/
352 Geometrie, Physical, and Visual Optics

Here refraction procedures, the dioptric power torsional


P V =V=P + U, component typically makes the curvature component
harder to accurately determine.
and the matrix sum gives One simple example of meridional refraction would
+3.27 D -2.50 D be to use spherical lenses together with a stenopeic
P = slit. The best spherical lens is then found for each
-2.50 D +8.17 D orientation of the stenopeic slit. Another simple exam-
ple would be to use a grating target (a series of lines or
Then from the trace and determinant equations,
the spherocylindrical parameters for the back ver- bars) together with spherical lenses. Then the best
tex power are +9.22C-7.01 x 113. spherical lens for each orientation of the grating target
We can find the front vertex or neutralizing is found. Other methods such as laser refraction tech-
power by reversing the light and considering plane niques (Chapter 21) can also be used.
waves incident on the back surface. In terms of
spherocylindrical parameters, the front vertex
power turns out to be +4.31 C-4.10 x 158. Note,
in particular, that the principal meridians are not
the same for the front and back vertex powers. Problems

For the horizontal meridian, Eq. 16.43 shows that 1. What is the dioptric power curvature component
curvature components are additive, for the 160 meridian on a +5.00 0 - 4 . 0 0 x 1 2 5
toric interface? What is the magnitude of the
V X = PX + U X , dioptric power torsional component in the 160
meridian?
and that the torsional components are additive,
2. What is the dioptric power curvature component
Vt = Pt + U , for the 70 meridian on a -2.00 C - 5 . 0 0 x 115 toric
interface? What is the magnitude of the torsional
So the torsional component behaves exactly as the component in the 70 meridian?
curvature component. The vertical meridian compo- 3. For the following lenses find the dioptric power
nents behave in the same way. matrix:
The dioptric power matrix appears again in Chap-
ter 18 in connection with prism power in the off-axis a. + 1.00C -4.00 x 90.
meridians of spherocylindrical lenses. The matrix for- b. + 1.00 C -4.00 x 120.
malism can also be used to further consider the equi- c. + 1.00C -4.00 x 60.
valent dioptric power matrix for systems of thick d. + 1.00O -4.00 x 135.
obliquely crossed lenses. However, this consideration e. + 1.00 O -4.00 x 45.
is outside the scope of this book. Also, compare b with c and d with e.
As mentioned in Section 16.2, when an oblique 4. For each of the lenses in the previous problem
meridian is isolated by a stenopeic slit aperture, a find the trace and determinant of the dioptric
twisting sheet of light is left in the conoid, and the power matrix.
twisting sheet does not form a point image. Because of 5. Given the following dioptric power matrices, find
this, some authors have stated that the off-axis meri- the spherocylindrical parameter of the lens in
dian does not have dioptric power. However, the minus cylinder form.
contention of this chapter is that the off-axis meridian
has two dioptric power components, the curvature +0.53 D -1.29 D
component and the torsional component, which can be a.
defined either from the refraction of a ray or from the -1.29 D -2.53 D
vergence changes of the wavefront. In the off-axis -0.84 D +1.35 D
meridian, the non-zero torsional component prohibits b.
the formation of the point image. + 1.35D -2.16D
The process of meridional refraction consists of
measuring a number of predetermined meridians to -3.50 D 0
determine the refractive error. In general, these pre- c.
determined meridians may not be the principal meri- 0 -5.50 D
dians. Most, if not all, of these meridional refraction + 11.00D -1.00D
methods determine the dioptric power curvature com-
ponent of the correction. In the subjective meridional -1.00D +11.00D
Dioptrie Power and Off-Axis Meridians 353

6. Given the matrix, 12. A person needs a -4.00 O-2.00 x 180 correction
/ - 2 . 7 5 D -0.06 D\ at the cornea. A hard contact lens of unknown
parameters is placed on the eye. An over-refrac-
\ - 0 . 0 6 D -4.50 D/ tion shows that the person needs an additional
+0.45 O-0.90 x 141.5. What are the parameters
First, estimate the spherocylindrical parameters by of the contact lens? What is wrong with the
assuming that the torsional component is zero. contact lens? With the contact lens in the eye,
Then solve for the exact spherocylindrical para- what part of the conoid of Sturm is on the retina?
meters.
13. A person needs a correction of -3.00 x 80. The
7. What single lens is equivalent to a -3.5 x 60 person is given a correction of -3.00 x 90. What
combined with a -6.25 x 20? are the principal meridians of the residual refrac-
8. A patient is wearing a +12.00 O-3.00 x 110 lens. tive error? (Note, from the axis estimation rules,
An over-refraction shows that the patient needs that you can answer this without having to work
an additional + 0 . 5 0 0 - 1 . 0 0 x 1 3 5 . What is the out the matrices.)
patient's Rx?
14. What does a lens clock read on the 50 meridian of
9. A cornea is represented by a +44.00 O+3.00 x a +5.00O-4.00 x 10 glass (n = 1.53) surface?
180 lens. The crystalline lens is represented by a 15. What would be the maximum magnitude of the
+ 19.00 0 + 2 . 0 0 x 6 5 lens. The reduced distance torsional component for a +12.00 0 - 4 . 0 0 x 1 5 4
between the cornea and crystalline lens is 4.5 mm, surface?
and the reduced distance between the crystalline
16. A meridional refraction is done with a stenopeic
lens and the retina is 15.0 mm. What correction
slit and spherical lens. For a person who needs a
(minus cylinder form) does this eye need at the
correction of +7.00 O-3.00 x 120, what spherical
cornea?
lens gives the best results for a vertical orientation
10. What is the spherical equivalent of the resultant of of a stenopeic slit?
a -5.00 O-2.00 x 170 combined with a -2.00 O 17. A meridional refraction is done with a grating
-4.00 x 124? consisting of horizontal lines. For a person who
11. A +5.00 0 - 3 . 5 0 x 7 0 lens is combined with a needs a correction of - 5 . 5 0 0 - 3 . 5 0 x 3 5 , what
spherocylindrical lens of unknown parameters. A spherical lens gives the best results for the
+8.00 O-2.50 x 90 is equivalent to the combina- horizontal grating?
tion. What are the spherocylindrical parameters of
the unknown lens (in minus cylinder form)?
CHAPTER SEVENTEEN

Prisms

17.1 Introduction wavelengths (Figure 17.Id). This is the phenomenon


of dispersion and results in a rainbow-like spectrum.
Figure 17.1a is a cross-sectional view of a glass plate For thin prisms, the amount of dispersion is small.
with parallel sides. Plane waves pass through the plate The fovea is the area on each retina with the
without a change in vergence. When a collimated highest density of cones. In order for a person to
pencil is incident at an angle i, the beam emerges at an resolve detail in the object, the conjugate retinal
angle i but is displaced from the original path. image must be on the fovea. This is normally achieved
Figure 17.1b shows a cross-sectional view of a piece when the person looks directly at the object.
of glass in which the sides are inclined at an angle A. For humans to enjoy the full benefits of binocular
In optics, this general shape is called a prism. The vision, the two eyes must both look at the object so
apex of the prism is the point at which the sides meet, that the conjugate retinal image of each eye is on the
and the angle A is called the apex angle. The base of fovea. Consider an emmtrope looking at a centrally
the prism is the side opposite the apex. Plane waves located near object (point E in Figure 17.2a). Not only
pass through the prism without a change in vergence. does the ciliary muscle have to provide the needed
However, the direction of travel is changed by the accommodation, but the extra-ocular muscles have to
prism, so that a collimated pencil is deviated through correctly aim each eye. Relative to the straight ahead
an angle d by the prism. position, the demand on each eye is to turn through
Figure 17.1c shows an observer looking through a the angle w e . Problems in the extra-ocular muscle
thin prism at an apple. Because of the prismatic system can cause stress (itching, burning, tearing), or
deviation, the apple appears to be deviated toward the even double vision (when the eyes do not aim at the
apex of the prism. In effect, the prism forms a virtual same point).
image that is displaced in the direction of the apex. Prisms are a tool that help us diagnose and solve
The fact that a glass prism (in air) deviates light problems with the extra-ocular muscle system. The
toward its base while the virtual image appears de- so-called base in prisms shown in Figure 17.2b can
viated toward the apex is crucial to the understanding help a convergence insufficiency problem by bending
of prism theory. A quick sketch, such as Figure 17.1c, the rays so that the eyes only need to converge (turn
helps keep the relationship clear. in) to point G (instead of to E). Relative to the
If the rays in Figure 17.1c are reversed, then straight ahead position, the demand for point G is the
converging light traveling left is incident on the prism. angle wg, which is less than the demand for point E
In this case, the prism has a real image, and the real without the prisms. So the base in prisms reduce the
image is deviated toward the base. convergence demand. The base out prisms in Figure
When white light is incident on a thick prism, the 17.2c increase the convergence demand and can be
short wavelengths are deviated more than the long useful in overconvergence problems.

355
356 Geometrie, Physical, and Visual Optics

apex

normal
normal

base
a) b)

FIGURE 17.1. a. Plane waves travel-


ing through a flat glass slab. b. Plane
waves bent toward the base of a prism, c.
Virtual image appears deviated toward
d) the apex. d. Dispersion.

17.2 Total Internal Reflection (Figure 17.3b). In both cases, a percentage of the
incident light is reflected at the surface. For paraxial
Before we launch into prism theory, we need to again rays, Fresnel's law, (Eq. 10.2) gives the percentage
consider SneWs law and an associated effect called reflected. For nonparaxial rays, the percentage re-
total internal reflection. flected increases.
For light incident on an interface between mediums For light initially in the lower index medium, the
n, and n 2 , Snell's law is percentage reflected approaches 100% at grazing inci-
dence. For light in the higher index medium, the
nl sin 0; = n2 sin 0r. (17.1)
percentage reflected approaches 100% much more
When light is initially in the lower index medium, quickly. The boundary angle for 100% reflection oc-
the ray bends (is refracted) towards the normal (Fi- curs when the angle of refraction 0r is 90 (Figure
gure 17.3a). When the light is initially in the higher 17.4). The critical angle 0C is the incident angle { for
index medium, the ray bends away from the normal which 0r is 90. For light in the higher index medium,
Prisms 357

top view

a)

FIGURE 17.2. a. Convergence demand


with object point at E. b. With base in
prisms, eyes do not need to converge as
much. c. With base out prisms, eyes have to
converge more.

normal

rii>n2
< n 2

a) b) FIGURE 17.3. Snell's law.


358 Geometrie, Physical, and Visual Optics

normal or
ft. = 41.8.
Any light in the glass incident on the surface at an
angle greater than 41.8 undergoes total internal re-
flection. In effect, this light is trapped in the glass
(unless it escapes at the next surface).
The higher the index of a medium in air,-the
smaller the critical angle. For light in diamond (n =
2.4) incident on a diamond-air interface, the critical
angle is only 24.6. For light in water incident on a
water-air surface, the critical angle is 48.7.
FIGURE 17.4. The critical angle. When the air is replaced by a medium, the critical
angle changes. For light in glass (n = 1.50) incident on
a glass-water (n = 1.33) interface, Eq. 17.2 gives:
any incident angle greater than the critical angle ftc sin ftc = 1.33/1.50 = 0.89,
results in 100% reflection. This effect is called total
internal reflection. ftc = 62.4.
The critical angle for an interface can be found Thus, total internal reflection at certain angles can be
from Snell's law. When ftr equals 90, equals the thwarted by changing the second medium.
critical angle ftc, and The sine function has a maximum value of 1. In
nl sin ftc = n2 sin(90). refraction angle computations with Snell's law, a com-
puted sine greater than 1 indicates that the ray under-
Since the sine of 90 is + 1 , it follows that goes total internal reflection.
sin ftc = n 2 /n 1 . (17.2)
For light in glass (n = 1.50) incident on a glass-air EXAMPLE 17.1
boundary, the critical angle is: What is the angle of refraction for light in diamond
(n = 2.4) incident on a diamond-air surface at a 35
sin ftc = 1.00/1.50 = 0.67, incident angle?

4
b)

o
FIGURE 17.5. a. Total internal reflection from
the anterior chamber angle, b. A Koeppe lens. c. A
Goldman lens.
Prisms 359

From Snell's law, escapes into the contact lens. A Koeppe lens allows
2.4sin(35) = l.Osin0r, direct viewing, while a Goldman lens incorporates a
mirror and allows indirect viewing (Figure 17.5b and
or 17.5c).
Small fibers of glass or plastic are used to pipe light
sin 0 = 1.38. around corners. Fibers with diameters in the micron
Since the sine function cannot be greater than 1, region work by total internal reflection (Figure 17.6a).
this ray undergoes total internal reflection and When the fiber is bent, the light hits the sides, but the
there is no angle of refraction. incident angles are greater than the critical angle.
Hence, total internal reflection occurs and the light
The aqueous humor flow in the eye is such that the remains trapped in the fiber. For fibers with micron
aqueous is completely replaced every four hours. The diameters, the corners can be surprisingly sharp before
aqueous drains through the trabecular meshwork that any light leaks out.
is located at the angle made by the iris and the sciera. Dust or scratches on the surface can frustrate the
If this angle closes up, the aqueous flow is impeded, total internal reflection and cause leakage from a fiber.
and the pressure inside the eye rises (a condition To combat this, the fibers are either a step index fiber
called glaucoma). This can eventually lead to retinal or a gradient index fiber. In the step index fiber, an
damage, tunnel vision, and even blindness. However, internal fiber is surrounded by an outer fiber with a
one cannot see the anterior chamber angle of a pa- lower index (Figure 17.6b). Similarly, in the gradient
tient's eye because the light coming from the angle index fiber, the index of refraction is highest in the
undergoes total internal reflection at the cornea (Fi- middle and lowest at the edge. The gradient index
gure 17.5a). This problem is solved by the use of actually causes a curved ray path. When the ray
special contact lenses placed on the eye. The index of wanders toward the edge, the gradient index causes it
the contact lens is higher than that of the eye, so total to curve back toward the middle (Figure 17.6c).
internal reflection can no longer occur, and the light An endoscope is a fiber optics instrument that

t.i.r.

end view

b)

FIGURE 17.6. a. Total internal re-


flection in an optical fiber, b. A step
index fiber, c. Curved ray in a gradi-
c) ent index fiber.
360 Geometrie, Physical, and Visual Optics

f/
FIGURE 17.7. Schematic representation of an
endoscope.

enables people to see around multiple bends and down even when the alpha particle is trapped inside the
into dark cavities (Figure 17.7). Some of the fibers in nucleus. If the gap is narrow enough, the alpha parti-
the endoscope carry light down into the cavity to cle tunnels across the gap in a manner analogous to
illuminate the object of regard. Other fibers carry the light above.
diffusely reflected light back to the observer. Endo-
scopes have made it possible for the medical profes-
sion to examine the insides of stomachs, bladders,
etc., without having to cut them open. 17.3 Deviation by a Thick Prism
The light transported by fiber optics bundles can
carry much more information than electrical signals Figure 17.8a shows a thick prism. A principal section
transported by copper wire. The copper wires used for of the prism is the section that contains the normals of
telephone communications in large cities need to be both the front and the back surfaces. Figure 17.8b
enormous in size. These enormous copper cables are shows a cross-sectional view of a principal section of a
now being replaced by relatively small fiber optics thick prism with index n2 and apex angle A. The
bundles. media in front of and behind the prism have respective
Total internal reflection is involved in the physical indices of nl and n3.
phenomenon called tunneling. Consider light in glass. Light incident on the prism is refracted at the first
When the light is totally internally reflected, a light surface and again at the second surface. The deviation
wave with an exponentially decaying amplitude actual- angle d for a ray incident on the prism at an incident
ly penetrates for a short distance (several wavelengths) angle of i, can be found by successively applying
into the lower index media. Even though this wave Snell's law at the surfaces.
penetrates into the lower index medium, the energy The derivation of an equation for the deviation
carried by the exponentially decaying wave propagates angle uses a geometry theorem, which says that the
back into the glass, and becomes part of the reflected
beam. When a second piece of glass is slowly brought
closer to the first piece, it encounters the exponentially
decaying wave before it touches the first piece of glass.
Then the flow lines change and some light will enter
the second piece even though contact is not yet estab-
lished. It appears that the light that should have been
trapped in the first piece of glass has tunneled across
the air gap. Of course, when contact is established
between the two pieces of glass, the total internal
reflection condition is removed, and the light is trans-
mitted.
The phenomenon of tunnelling also shows up in
quantum mechanics. Ordinarily, alpha particles are
trapped or bound in an atomic nucleus by the nuclear
forces. The strange thing about alpha decay is that the
alpha particles appear to escape the nucleus even
though they do not have enough energy to cross the
nuclear energy gap. The quantum mechanical explana- b)

tion is that the alpha particles have an exponentially FIGURE 17.8. a. A thick prism, b. Principal section of a thick
decaying wavefunction that extends out into the gap prism.
Prisms 361

exterior angle of a triangle equals the sum of the two Given the incident angle i1? we can then apply Eqs.
opposite interior angles, or from Figure 17.9 17.3, 17.5, and 17.4 in sequence to obtain r2.
a + b = d. Figure 17.10 shows the deviation angle between the
incident ray and the emerging ray. The ray extensions
We can prove the theorem easily. The three angles and the internal ray form a triangle for which d is the
inside a triangle add to 180, or external angle, while sl and s2 are the two opposite
a + b + c = 180. internal angles. Thus,
d = Sj + s 2 .
Angles c and d are supplementary angles, so they also
add to 180, or Now
c + d = 180.
We can then subtract the above two equations to and
obtain
a + b - d = 0,
or We can solve the above two equations for Sj and s2
and then substitute the results into the equation for d
a + b = d.
to obtain
The derivation also uses the familiar theorem that d = i1+r2-A. (17.6)
angles with perpendicular sides are equal. In Figure
17.8b, angle B (made by the normals) has sides that Given the incident angle i,, we can find the angle
are perpendicular to the sides of angle A; therefore, A of refraction r2 from Eqs. 17.3 through 17.5, and then
and B are equal angles. compute the deviation angle d from Eq. 17.6.
For the ray in Figure 17.8b, the incident and If the ray is reversed, it retraces the same path back
refraction angles at the first surface are ij and r l5 through the prism and has the same deviation angle.
respectively. At the second surface, the incident and For the reversed ray, r2 is the initial incident angle and
refraction angles are i2 and r 2 . Snell's law at the first i1 is the final angle of refraction. Simply exchanging
surface is and r2 in Eq. 17.6 algebraically confirms that the angle
of deviation is the same for the reversed ray.
nl sin^ = n2 sinr^ (17.3)
In computing the deviation d, we do need to
Snell's law at the second surface is distinguish between a ray incident on the base side of
the normal vs a ray incident on the apex side of the
n2 sini 2 = n3 sinr 2 . (17.4)
normal. When the ray is incident on the base side of
The normals to each surface intersect, making the the normal ij is positive (Figure 7.10). The sign con-
angle B. Since the normals are perpendicular to the vention for r2 is the same as that for i, (since the
sides, B is equal to A. B is the external angle to the reversed ray has an incident angle r 2 ).
triangle made by the two normals and the ray crossing It follows from Snell's law (Eq. 17.3) that and ij
interior of the prism. The two opposite interior angles always have the same sign. Therefore, when \i is
are rx and i2. Therefore,
i2 + r1 = B ,
and since A equals B,
2 = - . (17.5)

ii>0 r2>0
FIGURE 17.9. External angle d is the sum of the two opposite
internal angles a and b. FIGURE 17.10. Deviation angle.
362 Geometrie, Physical, and Visual Optics

and
r , = 35.53.
ii<0
n<o From Eq. 17.5,
r 2 >0 i 2 = 50-35.53 =+14.47.
i 2 >0
From Eq. 17.4,
1.617 sin( + 14.47) = 1.00 sin r 2 ,
0.404 = sin r 2 ,
a)
and
r2 = 23.83.
From Eq. 17.6,
d = +70 + 23.83 - 50 = +43.83
The prism bends this ray 43.83 toward the base.
ii>0 By reversibility, a ray with an incident angle of
23.83 will have a 70 final angle of refraction and a
i2<0 43.83 deviation angle.

EXAMPLE 17.2b
What is the deviation angle for a ray that has
grazing incidence on the base side of the normal?
b) In this case i2 is +90 and the four-step method
gives:
FIGURE 17.11. Sign convention.
1.00sin(+90) = 1.617 sin r,,
sinr 1 = 1/1.617 = 0.618,
positive, so is rx, and when il is negative so is r ^ This and
means that I*! is positive when the internal ray leaving r1 = +38.20.
the first surface is on the apex side of the normal
(Figure 17.10). When the internal ray leaving the first (As an aside, note that 38.20 is the critical angle
surface is on the base side of the normal r{ is negative for light in a medium of 1.617 incident on a glass-
air interface. Why?)
(Figure 17.11a).
From Eq. 17.5,
It follows from Snell's law (Eq. 17.4) that i2 and r 2
always have the same sign. When r 2 is positive, so is i2 = 50 -38.20 =+11.8.
i 2 , and when r 2 is negative so is i 2 . The angle i2 is Then
positive when the ray incident on the second surface is 1.617 sin( + 11.8) = 1.00 sin r 2 ,
on the apex side of the normal (Figure 17.11a), and
0.331 = sin r 2 ,
negative when the incident ray is on the base side of
the normal (Figure 17.11b). Note that the angles r, r2=+19.31,
and i2 are related by reversibility. and
The deviation angle d is positive when the bend is d = +90 + 19.31 - 50 = 59.31.
toward the base and negative when the bend is toward
Since grazing incidence gives the largest incident
the apex. The above sign conventions and the succes- angle possible, the 59.31 deviation is the maximum
sive use of Eqs. 17.3, 17.5, 17.4, and 17.6 constitute a deviation that this prism gives. By reversibility, the
four-step method for finding the deviation angle d of a 59.31 maximum deviation will also occur for an
thick prism. incident angle of 19.31.

EXAMPLE 17.2a Figure 17.12 shows a plot of the deviation angle vs


A flint glass (n = 1.617) prism in air has a 50 apex incident angle for this prism. The graph has a distorted
angle. What is the deviation for a ray with an U shape. At +90 the deviation angle is the 59.31
incident angle of 70 on the base side of the maximum. Note that at an incident angle of +19.31
normal? the deviation angle is again 59.31. Similarly, from
From Eq. 17.3, Example 17.2a, incident angles of 70 or 23.83 both
1.00 sin(+70) = 1.617 sin , give a 43.83 deviation.
sinr, =0.940/1.617 = 0.581, Starting at +90, as the incident angle is decreased
Prisms 363

90

LU 2,
_l LU
O
total I
z internai
< reflection Z
z <
g
> <
LU
O >

S S 8 S S S o
O)
o
CO
o
CO
o o
CO
o
CD
o
O)
+ + + I I I + + + I I I
INCIDENT ANGLE i1 (in degrees) INCIDENT ANGLE i, (degrees)
FIGURE 17.12. Deviation for a prism of index 1.617 and apex FIGURE 17.14. Deviation for a prism of index 1.523 and
angle 50. apex angle 26.

Then
i2 = 50 - 6 . 1 7 =+43.84,
and
1.617 sin (+43.84) = 1.00 sin r 2 ,
1.12 = sinr 2 .
The sine function cannot be greater than + 1 , so the
above equation says the incident angle of 43.84 is
greater than the critical angle and total internal
reflection occurs.

Figure 17.14 shows the deviation angle as a func-


FIGURE 17.13. Total internal reflection at the prism's back tion of the incident angle for a prism of index 1.523
surface. and apex angle 26. Again, the shape is that of a
distorted U. The maximum deviation is 40.72, and
occurs for incident angles of +90 or -23.28. The
the deviation falls to a minimum and then rises back to minimum deviation is 14.07, and the deviation at
the maximum at an incident angle of 19.31. For normal incidence (i1 = 0 ) is 15.89. Total internal re-
incident angles of less than 19.31 (including incident flection occurs for \x< -23.28, which means the rays
angles above the normal), total internal reflection are incident on the side of the normal at angles greater
occurs at the second interface (Figure 17.13). than 23.28 in magnitude. The figures for an incident
angle of - 1 5 are worked out in the following ex-
EXAMPLE 17.2c ample.
For the same prism, what is the deviation angle for
an incident angle of 10 on the base side of the
EXAMPLE 17.3
normal?
Find the deviation angle for a ray incident at -15
From the four-step method,
on a prism (n = 1.523) with a 26 apex angle.
1.00sin( + 10) = 1.617 s i n r ^ From the four-step method:
0.107 = sin r,, 1.00 sin(-15) = 1.523 sin r1?
or -0.170 = sin r19
^ = +6.17. r, = -9.78.
364 Geometrie, Physical, and Visual Optics

Then 1.534
Telescopic Crown
i2 = 2 6 - ( - 9 . 7 8 ) = +35.78, 1.532
and j 1.530
note expanded vertical scale

1.528.
1.523sin(+35.78) = 1.00 sin r 2 .
1.526-
So 1.524
r2 = 62.94,
and -+- 543 589 656
488
H
-+-h

d = (-15) + 62.94 - 26 = 21.94.


Whenever the prism is the higher index media,
such as a glass or plastic prism in air, the deviation is FIGURE 17.16. Index of refraction for telescopic crown as a
function of incident light wavelength.
towards the base of the prism (d > 0). When the prism
is the lower index media, such as a prism shaped air
cavity in a piece of glass, the prism deviation is toward
the apex ( d < 0 ) . Finally, Eq. 17.4 gives
n 2 s i n ( A / 2 ) = 1.00sin[(A + d m i n ) / 2 ) ] ,
which is typically written as
17.4 Minimum Deviation sin[(A + d m i n )/2]
n9 = (17.7)
sin(A/2)
The graph of the deviation as a function of the inci-
dent angle has the distorted U shape. Each deviation Equation 17.7 represents a classical method to
angle is given by two incident angles, and the two determine the refractive index of a material. A prism
incident angles are connected by reversibility (when with a known apex angle is made from the material.
one is i l 5 the other is r 2 ) . Near the minimum the two Then the minimum deviation angle is measured, and
connected angles (il and r 2 ) are almost numerically the index is calculated from Eq. 17.7.
equal. Because of dispersion, the minimum deviation
At the minimum, the two connected angles are angle is slightly different for each wavelength of inci-
numerically equal (ii = r 2 ) . For a prism in air (r^ = dent light. According to Eq. 17.7, the refractive index
n 3 = 1), this occurs when the internal ray crossing the n 2 of the prism is a function of the wavelength of the
prism is perpendicular to the bisector of the apex incident light. Figure 17.16 is a graph of the refractive
angle (Figure 17.15). In this case, the four-step index of telescopic crown glass as a function of the
method for the deviation simplifies to one equation. incident wavelength. So far we have been neglecting
For il equal to r 2 , Eq. 17.6 gives this slight dependence. The index dependence on
wavelength is responsible for chromatic aberration of
d min = 2r 2 A,
lenses and prisms. Chromatic aberration is treated in
or Chapter 20.
A + d

From Snell's law, it is easy to show that the internal EXAMPLE 17.4
angles and i2 are equal. Then from Eq. 17.5, What is the minimum deviation angle for a prism
with a 60 apex angle and index of 1.31?
A
12 = Note that sin[(A + d m i n )/2] is not equal to
T sin(A/2) + sin(d min /2). Therefore, to minimize
computational mistakes, let us rename the argu-
ment of the sine function in the numerator of Eq.
17.7. For example, let
n 3 = 1.00
q = (A + d m i n )/2.
Then Eq. 17.7 becomes
sin q
n, =
sin(A/2)'

sin q = 1.31 sin(30) = 0.655,


FIGURE 17.15. Minimum deviation ray. q = 40.9.
Prisms 365

Then we can solve for dmin: calm air just like leaves. In this case the majority of
the crystals are aligned horizontally and the deviated
light is concentrated at 22 horizontally from the sun.
These are the sun dogs of folklore. The sun dogs are
or
sometimes present together with the entire 22 halo,
A + dmin = 81.8, and at other times are visible even when the rest of the
dmin = 21.8~22. halo is not.
Ice crystals that form in the atmosphere have a Since the ice crystals have some depth, it is also
six-sided symmetry that is characteristic of their crystal possible for a ray to enter a horizontal surface and
lattice (Figure 17.17a). When light passes through the leave a vertical surface (Figure 17.17c). In this case
section of the crystal shown, the crystal acts like a the crystal acts like a prism with a 90 apex angle. The
prism with a 60 apex angle. The index of the ice minimum deviation angle for a 90 ice crystal (n =
crystals is 1.31. The previous example used the num- 1.31) prism is 46. Though it is less common than the
bers for an ice crystal, and the minimum deviation 22 halo, the 46 halo is occasionally observed and has
angle was about 22. been photographed.3
For a collection of randomly oriented ice crystals,
all incident angles would be present. In the vicinity of
maximum deviation, the deviation angles change very
quickly with the incident angle (for example, see 17.5 Deviation at Normal Incidence
Figure 17.12). In the vicinity of minimum deviation,
the deviation angles change slowly with the incident The four-step method also simplifies when a ray is
angle. Because of this, the light is concentrated at or incident normally on a prism (\l = 0 ) . In this case, the
near the 22 minimum deviation angle. The visual only refraction occurs at the second surface (Figure
result is the surprisingly common 22 ice crystal halo 17.18). From Eq. 17.6,
(Figure 17.17b). The 22 halo is frequently visible
when looking at the sun or a full moon through thin dnor = r2 - A,
cirrus clouds and sometimes visible when ice crystals or
are present without any noticeable clouds. r2 = dnor + A.
Consider looking through a bunch of ice crystals at 1
For even more fascinating ice crystal halos see "Rainbows, Halos,
either the rising or setting sun. When the ice crystals and Glories" by Robert Greenler, Cambridge University Press,
are shaped flat like a plate, they fall slowly through 1980.

FIGURE 17.17. a. Prism effect of an ice crystal. Dashed lines


show effective 60 degree apex angle, b. 22 degree halo. c. Ray
b) path for 90 degree apex angle.
366 Geometrie, Physical, and Visual Optics

17.6 Thick Prism Comparisons


Figure 17.19 is a plot of the deviation angle vs the
incident angle for a series of prisms with the same
index (1.50) but different apex angles. Each U-shaped
curve is marked with the corresponding apex angle.
As the apex angle is increased, the (distorted) U-
FIGURE 17.18. Ray path for normal incidence. shaped plot moves up and to the left. This indicates a
smaller range of incident angles for which light can get
through the prism.
From Eq. 17.5, Eventually, an apex angle is reached for which only
one ray will get through the prism. For larger apex
i 2 = A. angles, all light is totally internally reflected. In this
Then Eq. 17.4 gives case, a solution for d min from Eq. 17.7 results in
n 2 s i n A = sin(A + d n o r ) , n2sin(A/2)>l.

which is typically written The apex angle that lets only one ray through the
prism can be found from
sin(A + d n o r )
n7 = (17.8) n 2 sin(A/2) = 1,
sin A
Equation 17.8 gives the deviation for normal inci- or
dence. It is similar in form to Eq. 17.7 for minimum
sin(A/2) = l / n 2
deviation. By reversibility, Eq. 17.8 also gives the
deviation for normal emergence (r 2 = 90). A comparison with Eq. 17.2 shows that A / 2 equals
the critical angle 0C. Therefore, when the apex angle
equals twice the critical angle, only one ray gets
EXAMPLE 17.5
What is the deviation at normal incidence for a through the prism.
plastic (n = 1.44) prism with a 20 apex angle? As the apex angle is decreased, the (distorted)
Just as for minimum deviation, we can minimize U-shaped plot moves down and to the right and
the possibility of a computational error by letting flattens out. The minimum deviation angle eventually
q = A + dno,
Then from Eq. 17.8,
sin q = 1.44 sin(20) = 0.493,
q = 29.5,
and
d nor = q - A = 9.5.

EXAMPLE 17.6 o
z<
What is the deviation for normal incidence for a
flint glass (n = 1.617) prism with a 50 apex angle?
Just as in the preceding example, let <
>
q = A + d nor . LU
Q
Then 4 (thin)
sin q = 1.617 sin 50.

I I | I 1 1 I I | I I | I I | I I |
sin q = 1.239.
The sine function cannot be greater than 1, so total + I
internal reflection occurs at the back surface. The paraxial
deviation for this prism was plotted in Figure 17.12,
and, indeed, that graph shows that total internal
reflection occurs for all incident angles less than INCIDENTANGLE i1 (Degrees)
+ 19.3. FIGURE 17.19. Deviation angles for a series of apex angles.
Prisms 367

approaches the deviation at normal incidence. For an EXAMPLE 17.7a


apex angle of zero, the deviation is a straight line at What is the paraxial thin prism deviation for a glass
zero for all incident angles. prism (n = 1.5) with a 6 apex angle?
For a fixed apex angle, the prismatic deviation can From Eq. 17.9,
be increased by increasing the refractive index. This d = (0.5)6 = 3.
would also make the U-shaped plot move up and to
the left. EXAMPLE 17.7b
What is the paraxial thin prism deviation for a
highlite (n = 1.7) prism with a 6 apex angle?
From Eq. 17.9,
17.7 Thin Prisms d = (0.7)6 = 4.2

Thin prisms are prisms with small apex angles (Figure EXAMPLE 17.7c
17.20a). For thin prisms, the U-shaped deviation plot For a plastic (n = 1.44) prism with a 20 apex angle,
is very flat on the bottom (as shown by the bottom the exact deviation at normal incidence was 9.5
curve in Figure 17.19). In particular, for all paraxial (see Example 17.5). What does the thin prism
rays (i.e., rays for which i, is much smaller than 20), approximation give?
From Eq. 17.9,
the deviation angle is approximately a constant.
For paraxial incidence on thin prisms, we could d = (0.44)20 = 8.8.
determine the constant deviation from Eq. 17.8 for This is a 7% error, which is not bad since the 20
normal incidence, but since the angles involved are apex angle is pushing the limits of the paraxial
small, we can simplify further with a small angle region.
approximation.
The small angle approximation consists of replacing Figure 17.20b shows a prism (apex angle A) in bent
the sine function by the angles expressed in radians. form. The front surface is a +6.00 D spherical surface,
Let c be the conversion factor for degrees to radians, while the back surface is a - 6 . 0 0 D surface. For a thin
then Eq. 17.8 gives prism, the dioptric power is the sum of the surface
powers, and is still zero. The prismatic deviation is still
(A + d)c given by Eq. 17.9.

or
An 2 = A + d,
17.8 Ophthalmic Base Directions
and finally,
d = (n2-l)A. (17.9) The meridian occupied by a principal section of a
prism is called the base-apex meridian. The orienta-
For a thin prism in air, Eq. 17.9 shows that the tion of an ophthalmic prism is specified by the direc-
deviation depends only on the prism's refractive index tion of the base in the base-apex meridian (Figure
n 2 and apex angle A; increasing n 2 or A increases the 17.21a). A base down (BD) prism deviates the light
deviation angle. Since the paraxial deviation is also the vertically down, while a base up (BU) prism deviates
minimum deviation, Eq. 17.9 can also be derived from the light vertically up. The nose is used as a horizontal
the minimum deviation equation (Eq. 17.7). reference. A base in (BI) prism deviates the light
horizontally in toward the nose, while a base out (BO)
prism deviates the light out away from the nose.
To specify oblique base-apex meridians, we use the
cylinder axis coordinate system. Figure 17.21b is a
contra-ocular view of a prism in front of each eye. For
the right eye, a prism that deviates the light up and in
+ 6.00/ / - 6.00 along the 30 meridian is called a base up and in at 30
(BU & I @ 30). For the left eye, a prism that deviates
the light down and out along the 120 meridian is
called a base down and out at 120 (BD & O @ 120).
a) b) The above system is a redundant system in that the
FIGURE 17.20. a. Thin prism with plane sides, b. Thin prism parameters are overspecified; however, it is a good
with spherical sides. system to use while learning. Note that strictly speak-
368 Geometrie, Physical, and Visual Optics

apex
apex

r
base apex
base

a)
FIGURE 17.21. Base directions (contraocular
view).

90 90 a wall. The position at which the beam hits the wall is


60 120 60
120 marked. Then the thin prism is inserted in the beam at
150/ \30 150
/ \30 a distance of 100 cm from the wall. The deviated beam
now hits the wall at a distance y away from the initial
180 4- 4-0 180 4- 4-0 spot (Figure 17.23).
Given the distance y in centimeters, the deviation
21 oV ^330
21\
7*330 angle is found from
240 300 j 240
270
300 tand = y/100cm.
270
It turns out to be convenient to use the distance y
directly in rating the prisms. In this sense, we consider
the distance y at the fixed distance of 100 cm as a
measure of the deviation power of the prism. By
FIGURE 17.22. Coordinate system (contraocular). analogy to lenses, we refer to y as the (deviation)
power of the prism in prism diopters; Z designates the
prism diopter value. The prism diopter unit is repre-
ing, a BD prism is a BD @ 90, while a BI prism is a sented by the Greek capital letter delta used as a
BI @ 180. Similar statements hold for BU and BO. superscript. Then
In the overspecified system, there are some prism Z =100 tan d, (17.10)
base directions that are not allowed. For the right eye,
we can have a BU & I @ 45 or a BD & O @ 45, but where the 100 actually stands for 100 cm.
we cannot have a BU & O @ 45 or a BD & I @ 45. Consequently, a 3 prism shifts a beam 3 cm on a
Because of this we could, for the right eye, just specify wall 100 cm from the prism. A 7 prism shifts a beam
a prism @ 45 with a BU component, i.e., a BU @ 45. 7 cm on a wall 100 cm from the prism, etc.
The latter is a less redundant system. A third system is Suppose the wall is not 100 cm from the prism.
to specify the prism base directions by a 360 system Then the prism diopter value is found by a simple
(Figure 17.22). For the right eye, a BD & O @ 45 similar triangle ratio (Figure 17.24a). If the deviation
would then simply be a base @ 225, etc. is y on a wall a distance x from the prism, then
Remember that the virtual image formed by a
Z
prism appears deviated toward the apex. For example, (17.11)
a candle appears to be deviated down when looking 100
through a BU prism.

17.9 Prism Diopters

For paraxial purposes, we can classify thin prisms by


their deviation angles. Consider how the deviation
angle would be measured. One method is to take a
well-defined light beam (a laser beam works great),
and have it incident normally on a flat surface such as FIGURE 17.23. Linear vs angular deviation.
Prisms 369

17.10 Thin Prism Combinations

When the deviation angle is small,


t a n d ~ d (expressed in radians).
Then from Eq. 17.10,
Z = 100 d (in radians).
Consider two thin prisms, both BD, held together.
The first prism deviates the light down by an angle dx,
virtual and the second prism deviates the light down further
^ image
by an angle d2 (Figure 17.25). The resultant deviation
is the sum:
dr = d! + d 2 .
Then
100dr = ( l O O d ^ (100 d 2 ),
or
b)
Z r = Zj + Z 2 (17.12)
FIGURE 17.24. a. Similar triangles for prism diopter values.
b. Location of virtual imge. Equation 17.12 says that prism diopters are addi-
tive for small deviation angles. For example, a 3 BD
prism combined with a 4 BD prism is equivalent to a
single 7 BD prism. Essentially, this means that, on
EXAMPLE 17.8a the wall at 100 cm, the 3 BD prism deviates the light
On a wall 40 cm from the prism, the prism shifts 3 cm down, and the 4 BD deviates the light 4 more
the light down 2 cm. What is the power of the cm down. Therefore, the resultant deviation is 7 cm
prism? down.
From Eq. 17.11, the power is By the same reasoning, prism diopters subtract
Z = 2 when the base directions are opposite. For example, a
100 " 40' 3 BD prism combined with a 4 BU prism is equiva-
lent to a single 1 BU prism. Essentially, this means
(100)2 that, on the wall at 100 cm, the 3 BD prism deviates
z= 40
= 5* the light 3 cm down, and the 4 BU prism deviates the
light 4 cm back up. Therefore, the resultant deviation
For a prism of power Z, the shift is Z cm at 100 cm.
is 1 cm up.
For a closer distance, the shift is less than Z, and for a
Suppose a 3 BI prism is combined with a 4 BU
wall farther away the shift is greater than Z.
prism. For the wall at 100 cm, the 3 BI deviates the
light 3 cm horizontally in, while the 4 BU deviates
EXAMPLE 17.8b the light 4 cm up (Figure 17.26a). The Pythagorian
A glass (n = 1.5) prism has a 6 apex angle. What is theorem gives the magnitude of the resultant de-
the power of the prism? viation:
From Eq. 17.9,
V(3) 2 + (4)2 = 5.
d = (0.5)6 = 3.
Then from Eq. 17.9,
Z=100tan3 = 5.2A
We can also use Eq. 17.11 to determine the loca-
tion of the prism's virtual image. Since the prism has
zero dioptric power, the image distance is equal to the
object distance. The object point in Figure 17.24b is a
distance x from the prism, and the virtual image is a
distance y above the object point. Equation 17.11 then
relates x and y to the prism power Z. FIGURE 17.25. Thin prism combination.
370 Geometrie, Physical, and Visual Optics

T h e quick graph is

Zv = 4

Zx = 3

An estimate of the result from the graph is 5.5 BU


Zv = Zsin0 & I @ 120. Equation 17.13 gives
Z r = V(2) 2 + (5) 2 = 5.4A
In this case, Eq. 17.14 gives the supplementary
angle:
Zx = Zcos6
b) tan = -,
FIGURE 17.26. a. Horizontal and vertical prism combination. = 68.
b. Resolution into components.
So in standard notation, the angle is 180-68 or
112.
The single equivalent prism is 5.4 BU & I @
Thus, a single 5 prism oriented correctly gives the 112. Note that the simple guess from the graph was
same resultant deviation. If the prisms are held in not quite right, but it supplied a tremendous
front of the right eye, amount of intuition about the correct answer.
4
tan0 = The oblique vector in Figure 17.26b represents a
3'
prism of power Z. The horizontal and vertical compo-
and
nents of this prism are found by drawing perpendicu-
0 = 53.1 lars to the horizontal and vertical axis. The magnitude
of the components can be estimated from the graph,
Thus, for the right eye, the single prism equivalent to
or found from the equations
a 3 combined with a 4 BU is a 5 BU & I @
53.1. Zx = Z c o s 0 , Zy = Zsin0. (17.15)
The above process is just vector addition. So for
paraxial incidence on thin prisms, prism-diopters add EXAMPLE 17.9b
vectorally. The above equations are formalized as A 6 BU & O @ 154 is placed in front of a
patient's right eye. What are the horizontal and
Z r = V ( Z x ) 2 + (Z y ) 2 , (17.13) vertical components of the prism?
and The quick graph is

tan0 = (17.14)
Zv
While Eq. 17.14 sometimes returns the standard meri-
dian angle at other times it returns the supplementary
angle. A quick graph enables you to tell the differ-
ence. In fact, you can estimate the magnitude of all
the parameters from a quick graph. The estimates from the graph are: 5 BU and 2
BO.
Equation 17.15 gives
EXAMPLE 17.9a
Find the single prism equivalent to a 2 com- Z x = 6cos26 = 5.4A BO,
bined with a 5 BU in front of the left eye. Z v = 6 sin 26 = 2.6 BU.
Prisms 371

The graph and the estimates should be made quick- 17.11 Centrads
ly. These provide the intuition about the exact
answer. Strictly speaking, prism diopters are additive only for
small angles. On the other hand, the deviation angles
When two oblique prisms are combined, the prism are always additive. An angle is defined as an arc
powers add vectorally. Again, the result may be esti- length on a circle divided by the circle's radius. It
mated by a quick graph. One method of computing follows that the arc lengths are additive.
the result is to resolve each prism into its horizontal The centrad is a unit of arc length deviation. A
and vertical components. Then the horizontal compo- prism is placed at the center of curvature of a sphere
nents are added together to give the resultant horizon- that is 100 cm in radius. Then the arc length deviation
tal component, and the vertical components are added yc produced by the prism on the sphere is measured
together to give the resultant vertical component. If (Figure 17.27a). This deviation in centimeters is called
need be, the components can be recombined to give the centrad rating of the prism. For example, a 1
the parameters of the resultant prism. (This method centrad prism gives a 1 cm arc length deviation on the
corresponds to clinical procedure, where we usually screen. A 2 centrad prism gives a 2 cm arc length, etc.
work and think in terms of the horizontal and vertical For small deviation values, the arc length deviation
components.) yc is approximately equal to the deviation y on the flat
wall (Figure 17.27b). Thus, for small values, centrads
EXAMPLE 17.9c and prism diopters are essentially equal. However, for
For the left
a
eye, a 7 a BD & I @ 35 is combined large values, the arc length deviation differs from the
with a 3 BO. What single prism is equivalent to
the combination? deviation on a flat wall, so the centrad value differs
The quick graph is / from the prism diopter value.
The fact that centrads are additive is an advantage
over prism diopters. However, for the larger deviation
values (where the biggest differences would occur),
the deviation angle is a function of the incident angle,
\J so that a single unit (such as the centrad) may not
adequately describe the prism. Nevertheless, centrads
are used in the "graphical analysis" of binocular
vision.

From the graph, an estimate for the answer is 4.4


BD & I @ 47.
For the calculated value, first find the horizontal
and vertical components for the 7 BU & I @ 35.
From Equation 17.15,
Z x =7 A cos35 = 5.7ABI,
Zy = 7 A sin35 = 4.0ABD.
The 3 BO has a zero vertical component, so the
resultant components are:
horizontal
5.7 + 3 0 = 2.7 ,
vertical
4.0 + 0 = 4.0 BD.
From Eqs. 17.13 and 17.14, the parameters of the
single prism are:
Z = V(2.7)2 + (4.0)2 = 4.9,
4.0 b)
tan = = 56,
or FIGURE 17.27. a. Arc length for centrad values, b. Equival-
4.9 BD & I @ 56. ence of small arc length and linear deviations.
372 Geometrie, Physical, and Visual Optics

17.12 The Risley Prism For a counter-rotation of 30,


Z = 2(10A)cos30 = 17.3ABD.
A pair of counter-rotating prisms can be used to
The minimum (zero) occurs when one prism is base
generate continuously varying prism power. Consider
left and the other is base right (Figure 17.28d). For
two combined BD prisms each of power Z. The
each component, the angle of rotation amin specified
resultant prism power is 2Z BD, which is the max-
from the minimum position is the complement of amax
imum power obtainable from the combination (Figure
(Figure 17.28b), i.e.,
17.28a). When the prisms are counter-rotated through
an angle amax from the maximum position (Figure a
min + a
max = 90.
17.28b), the horizontal components are equal in mag- Therefore,
nitude and opposite in direction. So the resultant
prism is still BD (i.e., BD @ 90) as shown in Figure
17.28c. Each prism has a vertical component of and we can write the Risley equation as
Z y = Zcosa m a x , Z r = 2Zsina m i n . (17.17)
and the resultant is EXAMPLE 17.10b
Z r = 2 Z cos a (17.16) The Risley prism of the previous example is init-
ially set at zero. Then a 15 counter-rotation is
EXAMPLE 17.10a made. What is the prism power?
Each component of a Risley prism is a 10 prism. Z r = 2(10)sinl5 = 5.2 A BD.
What is the maximum prism power? What is the Since a counter-rotation of 15 from the minimum
power when they are counter-rotated by 30 from is equal to a counter-rotation of 75 from the
the maximum. maximum, we could solve the above problem as
The maximum is 2Z or 20 BD.
Z r = 2(10)cos75 = 5.2A BD.
To obtain other base directions with the Risley,
both prism components are rotated together. For ex-
ample, one might start with both components BI, and
then counter-rotate.

17.13 Prism Effectivity

Consider an object point located straight ahead of an


eye. When this object point is viewed through a prism,
it appears deviated toward the apex of the prism.
Hence, the eye must turn to keep the image of this
a) ?f b)
object point on the fovea. The amount the eye needs
to turn is referred to as the effect of the prism on the
eye.
Figure 17.29a is a thin prism with a deviation angle
d. To see the virtual image formed by the prism, the
eye rotates through the angle e about its center of
rotation, which is typically about 13.5 mm behind the
cornea.
The ray coming down the axis is bent down by the
angle d and misses the eye. The backwards extension
of the outgoing ray passes through the virtual image
point, which is a distance y above the object point.
The ray traveling upward at an angle g to the axis is
c) bent down again by the angle d, and when extended
FIGURE 17.28. Risley prism component vectors, a. Maximum passes through the eye's center of rotation. The angle
setting, b. Intermediate setting, c. Resultant prism for b. d. Zero e is the angle that the eye must rotate in order to
setting. image the object point on the fovea. This ray path,
Prisms 373

FIGURE 17.29. a. Prism effectivity ( e < d ) for a


near object in front of a thin prism with deviation
angle d. b. Prism effectivity for a distant object
(e = d).

together with the straight ahead axis, forms a triangle or


for which d is an external angle with g and e as the two tand
opposite internal angles. Thus, tane =
i-(c r o t /)
g + e = d, The prism has zero dioptric power, so
and we see that d is greater than e. V = U = l/w,
In Figure 17.29a, the object point is a distance u
from the prism, and the virtual image is a distance y and
above the object point. From the triangle involving tand
the object point, the straight ahead ray to the prism, tane = 1 - C V
r t
and the image point,
Let
tand = Z = 100 tane, (17.18)
where Z e is the prism diopter value corresponding to
or the angle the eye turns through. Then
y = u tand. 100 tan d
100 tan e
The distance from the prism to the center of rotation i-c r o t v
of the eye is labeled C rot . Since |w| = -w, or
tane : Z = (17.19)
-K + C rot " l - c v
Equation 17.19 is the prism effectivity equation for
The above two equations give
a prism of power Z placed a distance C rot from the
- u tan d center of rotation of the eye. The vergence V is the
tane = -w + C , vergence leaving the prism. (The equation is written in
r
374 Geometrie, Physical, and Visual Optics

terms of the vergence V instead of the incidence EXAMPLE 17.11c


vergence U because later on it generalizes in that An object is 33.3 cm from a 7 BU prism that is in
form.) the spectacle plane at a 12 mm vertex distance.
How much and in what direction does the eye turn?
The center of rotation is typically 13.5 mm be-
EXAMPLE 17.11a hind the cornea, so
An 8 BD prism is held 20 cm in front of an eye's
center of rotation. The person looks through the Cr = 12 + 13.5 = 25.5 mm.
prism at an object that is 2 m from the prism. How Here
much does the eye have to turn to see the object V=U = -3D.
(or what is the effective prism power)?
From Eq. 17.9, From Eq. 17.19,
8 A BD 7 A BU
Z = ze = 1 - (0.0255)(-3 D) '
-(0.20m)(-0.5D)'
BD 7 A BU
Z = = 7.3*BD. z= 1.077
= 6.5 .
1.10
The BD means the apparent object is deviated up, The eye turns down through an angle correspond-
and so the eye must turn up. However, it turns up ing to 6.5 prism diopters.
by 7.3 as opposed to 8 From Eq. 17.18, the
angle e is For a prism in the spectacle plane, the distance to
tane = 7.3A/100, the center of rotation is very small. In this case, we
could use the standard approximation:
e = 4.2.
1
1 + x, when x is small.
EXAMPLE 17.11b 1-x
As the person holds the prism 20 cm in front of the
eye's center of rotation, the person walks forward For Eq. 17.19, the approximation gives
until the prism is only 25 cm from the object. What
is the effective prism power? Z e = Z(l + C rot V). (17.20)

8 A BD Equation 17.20 shows that the difference between


Z = Z e and Z increases with an increase in distance (C rot )
l-(0.20m)(-4.0D)'
between the prism and the eye as well as an increase
8 A BD in the magnitude of the vergence V. In the clinic, the
= 4.4 BD.
1.80 difference is frequently small enough to be neglected.
The BD prism means that the object appears deviated
up. As the person walks toward the object, the effec-
tive prism power changed from 7.3 to 4.4. This is a
very noticeable effect as the object actually appears to 17.14 Fresnel Prisms
move down as the person approaches it.
As the person backs away from the object, the Figure 17.30a shows a collimated bundle of rays inci-
effective prism power increases. For an object at dent on a prism. Each ray in the bundle is deviated
optical infinity, U = V = 0 , and Eq. 17.19 gives through an angle d by the prism. Even though a ray
near the apex passes through a thin part of the prism,
ze = z. it has the same deviation as a ray that passes through a
thicker part.
This means that the angle e the eye turns through A Fresnel prism is a collection of prism sections
equals the deviation angle d of the prism. Figure from the apex area (Figure 17.30b). A ray traveling
17.29b shows the ray diagram. The incident rays are through any section is deviated by the angle d. How-
all parallel, and are all deviated by the angle d. For ever, the composite prism remains thin everywhere.
the ray that passes through the eye's center of rota- Note that a "thin" prism has a small apex angle. A
tion, angle e and angle d are opposite internal angles Fresnel prism in which the sections have large apex
between two parallel lines, and are equal. Compared angles would act like a thick prism even though the
to the above two examples, when a person looks actual thickness of the Fresnel remains small.
through an 8 prism at a distant object, the effective The boundary lines between the different sections
power of the prism is 8. of a Fresnel prism are perpendicular to the base-apex
Prisms 375

In Figure 17.31, the light exits the prism at the bot-


tom. The total deviation is 90. The internal reflection
gives a left-right inversion just like that of a flat
mirror. So when the object is the palm of a right hand,
the prism's virtual image is the palm of a left hand.
A Porro prism is a right angle prism that is turned
around so that the light enters the side opposite the
right angle (Figure 17.32a). The light then undergoes
two total internal reflections inside the prism and
reemerges, traveling back toward the incident direc-
tion (i.e., the total deviation angle is 180). Thus, the
Porro prism can be used as a retroreflector. For
example, the reflectors on bicycles are made up of
many little Porro prisms.
Each reflection in the Porro prism introduces a
left-right inversion, and two inversions cancel each
b) other. The geometry also causes the image to be
flipped in one meridian. So when the object for a

->-
* -

Front

c) Side FIGURE 17.31. Right angle deviation by total internal re-


flection.
FIGURE 17.30. a. Indepedence of deviation angle for different
ray paths, b. Fresnel prism, c. Front and side views of base down
Fresnel prism.
rh

/f^-^
meridian. Figure 17.30c shows front and side views of
a BD Fresnel prism.
Fresnel prisms can be made out of glass, hard
plastic, or plastic membranes. The plastic membrane
.-f-
Fresnel prisms adhere to spectacle lenses, and have a)
ophthalmic uses for temporary corrections or for high
prism powers. Light is scattered at the section boun-
daries of a Fresnel prism. This scattered light causes a
reduction in contrast, and hence, some loss of acuity
compared to a regular prism.

17.15 Reflecting Prisms


The internal reflections inside a prism are sometimes
beneficial in optical system design. The simplest exam-
ple is the right angle prism. Light enters the prism b)
normal to the first surface. The light undergoes total FIGURE 17.32. a. Porro prism as a retroreflector. b. Double
internal reflection at the side opposite the right angle. Porro prism erecting system.
376 Geometrie, Physical, and Visual Optics

truncated silvered

-Mrh

FIGURE 17.34. Penta prism.

FIGURE 17.33. a. Dove prism, b. Upside-down Dove prism.

Porro prism is an erect right hand, the virtual image is


an inverted right hand. Common binoculars consist of
a Keplerian telescope with a double Porro prism
erecting system (Figure 17.32b).
A Dove prism is just a right angle prism that has FIGURE 17.35. Amici prism.
been truncated (to reduce size and weight). The Dove
is used with collimated light. At the proper incident
angle, the emerging light travels in the same direction
as the incident light. When the object is an erect right
hand, the image is an inverted left hand (Figure Problems
17.33a). The Dove prism has an interesting symmetry
in that as it is rotated around its axis, the image 1. A glass (n = 1.617) prism has a 40 apex angle.
rotates twice as fast as the prism. Thus, when the What is the deviation angle for light incident
Dove prism is rotated 180 (an upsidedown Dove), the below the normal at an incident angle of 90
image rotates 360 or is back to the initial orientation (grazing incidence)? At 60? At 45? At 23.145?
(Figure 17.33b). What is the angle of minimum deviation? What
A Penta prism deviates the light 90 without intro- happens to light incident normally on the prism?
ducing a left-right inversion (Figure 17.34). In the 2. What is the critical angle for light in glass with a
Penta prism, the two internal reflections are not total, 1.57 refractive index?
so those sides must be silvered. The Penta prism is 3. What is the deviation at normal incidence for light
used in small rangefinders and in single lens reflex incident on prism of index 1.63 and apex angle
cameras. 29. What is the minimum deviation angle?
The Amici prism is a truncated right angle prism 4. A prism with a 60 apex angle has a minimum
with a roof section added onto the side opposite the deviation of 58. What is the refractive index of
right angle (Figure 17.35). The roof section inter- the prism?
changes the right and left portions of the image so that 5. A thin prism has an index of 1.53 and an apex
no left-right inversion occurs. The angular tolerances angle of 5.4. What is the deviation angle and
on the roof are very precise, so the Amici prism is not prism diopter rating?
easy to manufacture. 6. What linear deviation does a 6.7 prism give on a
There is a whole host of other reflecting prisms. wall 40 cm from the prism?
Some erect, some displace, some introduce left-right 7. A prism gives a 13.2 cm linear deviation on a wall
inversions, some serve as beam splitters, some serve as 330 cm from the prism. What is the prism diopter
polarizers.3 value?
8. Two thin prisms have an apex angle of 6.8. One
a
See, for example, W. Smith, Modern Optical Engineering, has a 1.49 index, and the other a 1.58 index. What
McGraw-Hill, 1966. is the difference in prism diopter values?
Prisms 377

9. A 3.2 BD prism is combined with a 4.7 BU 15. What single prism is equivalent to a 4 BU & O
prism. What single prism is equivalent to the @ 64 combined with a 5 BD?
result? 16. Jeremy Walk holds an 8 prism 10 cm in front of
10. A 3.2 BD prism is combined with a 4.7 BD his cornea (11.3 cm in front of his center of rota-
prism. What single prism is equivalent to the tion) and looks at an object 32 cm from the prism.
result? How much does he turn his eye (in prism diop-
11. For the right eye, a 3.2 BD prism is combined ters)? For a distant object?
with a 4.7 BO prism. What single prism is equi- 17. Nancy Dragon wears a 10 BO prism at a 12 mm
valent to the result? vertex distance. Nancy's entrance pupil is 3 mm
12. For the left eye, a 2.7 BD prism is combined with from her cornea, and Nancy's center of rotation is
a 5.3 BO prism. What single prism is equivalent 13 mm from her cornea. When Nancy looks at an
to the result? object 40 cm straight ahead of her lenses, how
13. What are the horizontal and vertical components much does she turn her eye (in prism diopters)?
of a 6 BU & I @ 122 prism. Which eye must this 18. A Risely prism consists of two 10 components.
prism be for? What is the resultant prism power when each
14. What are the horizontal and vertical components prism is counter-rotated 25 from the zero posi-
of a 7 BD & O @ 25. Which eye must this prism tion? When each prism is counter-rotated 30 from
be for? the maximum position?
CHAPTER EIGHTEEN

Prism Properties
of Lenses

18.1 Prentice's Rule We can determine the signs for a horizontal or a


vertical base direction from the same coordinate sys-
Consider a thin lens of dioptric power P. When a tem. Thus, BU is positive and BD is negative. For the
paraxial ray passes through at a distance h from the right eye, BI is positive while BO is negative (e.g., 3
optical center, it is deviated by an angle d (Figure B O = - 3 , while 3 BI = +3). The BI and BO signs
18.1). The deviation angle d is related to h and P by are reversed for the left eye ( e . g . , 5 A B O = + 5 , while
Eq. 12.7: 5 = - 5 ) .
A spherical lens is symmetric about its optical axis.
d=-Ph, (18.1)
For a plus spherical lens, the thickest part of the lens
where d is in radians, and h in meters. Then occurs at the optical center. The associated prism base
directions point toward the optical center, and a ray
(100 cm/m) d = -(100 cm/m) Ph.
passing through bends toward the optical center (Fi-
For small deviation angles, the left side mimics the gure 18.3a). For example, at a point above the optical
deviation power Z (in prism diopters) of a thin prism. center ( h > 0 ) on the plus lens ( P > 0 ) , Eq. 18.2 gives
So a BD result ( Z < 0 ) .
For a minus spherical lens, the thinnest part of the
Z = -Ph, where h is in cm. (18.2)
lens occurs at the optical center. The associated prism
Equation 18.2 is called Prentice's rule. It states that base directions point away from the optical center,
we can associate a different prism power with each and a ray passing through bends away from the optical
point on the lens. Dimensionally, when a distance center (Figure 18.3b). For example, at a point above
multiplies a dioptric power, the distance usually needs the optical center ( h > 0 ) on the minus lens ( P < 0 ) ,
to be in meters. Prentice's rule is an exception because Eq. 18.2 gives a BU result ( Z > 0 ) .
prism diopters are dimensionally equal to centimeters.
In Eq. 18.2, h is a directed distance from the EXAMPLE 18.1a
optical center to the point under consideration. The What is the prism power at a point 15 mm above
signs for a horizontal or vertical h can be determined the optical center on a +4.00 D lens?
From Eq. 18.2,
from a standard contra-ocular coordinate system in
front of each eye (Figure 18.2). For example, when a Z=-(+4.00D)(1.5cm)=-6 A ,
ray hits 2 cm above the optical center, h = + 2 cm. or
When the ray hits 2 cm below the optical center = 6 BD.
h = - 2 cm. For the right eye, a ray hitting 3 cm out The thickest part of the lens (at the optical center)
from the optical center gives h = - 3 c m . For a left is below the point under consideration and the
eye, a ray hitting 3 cm out from the optical center thinnest part (the edge) is above, so it makes sense
gives h = -1-3 cm. that the prism is BD.

379
380 Geometrie, Physical, and Visual Optics

FIGURE 18.1. Deviation angle d for a ray incident at a dis-


tance h from the optical center on a thin lens of power P.

Right Left

FIGURE 18.2.
\J
Plus-minus sign conventions for h (contraocular
view).

EXAMPLE 18.1b
For the same lens, what is the prism power at a Plus Lens
Minus Lens
point 5 mm below the optical axis? c) d)
From Eq. 18.2,
FIGURE 18.3. a. Deviation by plus lens. b. Deviation by minus
Z = - ( + 4 . 0 0 D ) ( - 0 . 5 cm) = +2 , lens. c. Base directions for plus spherical lens. d. Base directions
or for minus spherical lens.
= 2 BU.
In this case, the thickest part of the lens is above
the point.

EXAMPLE 18.2
Given a - 1 5 D lens in front of the left eye, what is
the prism power at a point 6 mm in from the optical
center.
From Eq. 18.2,
Z=-(-15.0D)(-0.6cm) = -9A
For the left eye, the coordinate system in Figure
18.2 shows that
= 9 BI.
The BI is consistent with the fact that the thinnest
part of the lens (the optical center) is out from the FIGURE 18.4. Prism at a point up and out from the optical
point under consideration. center of a minus lens.
Prism Properties of Lenses 381

+ 4.00D. -3.00D.
V
9

+ 4*BD 3

+ 4*BU + 3*BD
c)

+ 8 + 6*BD

+ 12 t 9*BD

a) b)

FIGURE 18.5. a. Prism powers and base directions for 1 cm steps on a +4.00 D lens. b. Same
for a 3.00 D lens. c. Contraocular viewisoprism curves.

For a spherical lens, the prism base directions lie 18.2 Prism-Lens Combinations
on radial lines that pass through the optical center
(Figures 18.3c and 18.3d). For any point on the lens, Consider a 2 BU prism combined with the +4.00 D
the magnitude of the prism power can be obtained lens. We can represent each point on the prism as
without worrying about the plus-minus signs, i.e., having 2 BU, and we can represent the lens as having
|Z| = |Ph|, where h is in cm. (18.3) the variable prism field described above. To find the
composite field, we simply add the prism diopter
Then the base direction can be determined by check- values at each point (Figure 18.6).
ing where the optical center is relative to the point At a point 5 mm below the optical center of the
under consideration. +4.00 D lens, the resultant prism power is 4 BU
(i.e., 2 BU + 2A BU). At the optical center of the
EXAMPLE 18.3 original +4.00 D lens, the resultant prism is 2 BU
What is the prism power at a point 12 mm up and (i.e., 2 BU + 0). At a point 5 mm above the optical
out from the optical center along the 128 meridian center of the original +4.00 D lens, the resultant prism
on a - 3 D lens in front of the right eye? is 0 (i.e., 2 BU + 2 BD). At a point 10mm above
From Eq. 18.3, the optical center, the resultant prism is 2 BD (i.e.,
2 + 4 BD).
|Z| = |(-3.00D)(1.2cm)|=3.6 A
The prism diopter values for the combination are
For the -3.00 D lens, the base directions point identical to the prism field of a +4.00 D lens that has
away from the optical center. From Figure 18.4, been shifted (decentered) up by 5 mm. Thus, a de-
Z = 3.6 A BU & O @ 128. centered spherical lens is equivalent to a prism-lens
combination.
Figure 18.7a, shows three centered lenses. The
We can consider a thin spherical lens as a field of optical axis of the system passes through the optical
radial prism powers. Those points close to the optical center of each lens. Suppose the middle (centered)
center have small prism diopter values, while those lens is the +4.00 D lens. When a 2 BU prism is
points far from the optical center have large prism placed against the +4.00 D lens, it shifts the prism
diopter values. Figure 18.5a shows some values for the field at each point by 2 BU. In particular, the straight
+4.00 D lens, while Figure 18.5b shows some values line that passes through the optical centers of the first
for the -3.00 D lens. and third lenses now passes through a point on the
Figure 18.5c shows that the iso-prism curves on a prism-lens combination that has prism power of 2 BU
spherical lens are circular and centered on the optical (Figure 18.7b). However, we can achieve the same
center of the lens. This is a manifestation of the shift in prism values without using the prism. We
rotational symmetry of a spherical lens about its opti- simply take the original centered +4.00 D lens, and
cal axis. decenter it up by 5 mm. (Figure 18.7c).
382 Geometrie, Physical, and Visual Optics

2 A BU + 4.00D.
A
-P-2 BU 4-6 4 A BD
4-f2^BU 1-4
+ 4 2 AB U 4- 2
2 0 =>
\jr 2 B U + 2iBU 6
44- 2 BU 4-4
+ 2 4- 6 8
! | FIGURE 18.6. Prism-lens combination.

For an uncorrected ametropic eye looking straight Figure 18.8a is a thin BD ophthalmic crown glass
ahead, the point at which the line of sight intersects prism with a +7.00 D front surface and a -7.00 D
the spectacle plane is called the Major Reference back surface. Figure 18.8b shows a +4.00 D ophthal-
Point (MRP). (To balance far and near vision, the mic crown glass thin lens with a +7.00 D front surface,
MRP is sometimes placed lower than the straight and a -3.00 D back surface. Figure 18.8c shows the
ahead position.) When the spectacle Rx for the eye prism-lens combination. The ophthalmic crown prism
does not call for any prism power, the lens is placed so and lens can be fused together in which case the
that its optical center is at the MRP. When the specta- boundary line disappears (Figure 18.8d). Alternative-
cle Rx is a prism-lens combination, the prism field of ly, the lens in Figure 18.8d could be ground from a
the lens is shifted, and the resultant prism power at piece of ophthalmic crown glass.
the MRP is the prism power called for in the Rx. By Prentice's rule, the final configuration for the
above lens-prism combination is equivalent to a de-
centered +4.00 D lens. When the prism is to be
obtained by decentration, the lens is decentered be-
fore it is edged to fit in the spectacle frame.

" * J r We can use the coordinate system in Figure 18.2 for


horizontal and vertical decentrations. As a directed
distance, the amount of decentration (DEC) is oppo-
site in sign to the h value in Prentice's rule. For
example, decentering a lens down ( D E C < 0 ) causes
the line of sight to pass through a point above the
optical center ( h > 0 ) . Thus,
h=-DEC,

y and from Prentice's rule,

4
1 i

Z = P(DEC), DEC in cm. (18.4)

1 HI
o.e. ^

b) ' f t
-3D

r
c)
FIGURE 18.7. a. Three centered lenses, b. Base up prism FIGURE 18.8. a. Prism with -7.00 D back surface, b. Lens
combined with middle lens. The numbers are the resultant prism with +7.00 D front surface, c. Prism-lens combination, d. Prism
powers at the indicated positions, c. Same effect achieved by fused with lens. (The result can also be achieved by using a prism
decentering middle lens up. wedge during surfacing).
Prism Properties of Lenses 383

For a plus spherical lens ( P < 0 ) , D E C has the The lens is plus, so we expect the prism to be
same sign as Z , so the decentration is always in the BO. From the coordinate system for the right eye,
same direction as the prism base. For a minus spheri- h = - 0 . 8 cm.
cal lens ( P < 0 ) , D E C is opposite in sign to Z , so the From Eq. 18.4,
decentration is always in the opposite direction as the
prism base. Z = (+6.00 D ) ( - 0 . 8 cm) = - 4 . 8 ,

EXAMPLE 18.4 = 4.8 BO.


Cilia Peachseed's left Rx is -7.00 D combined with
3 . What decentration relative to the MRP gives In Figure 18.9b, the optical center of the +6.00 D
the correct prism power? lens is out from the MRP. The prism base direc-
The lens is minus, and so (according to the tions point toward the optical center on a plus lens,
above comment) we expect the decentration to be so the prism at the MRP is BO, which confirms the
out. calculated base direction.
From the coordinate system for the left eye
(Figure 18.2),
=-3
18.3 Prism in Cylindrical Lenses
From Eq. 18.4,
- 3 A = (-7.00D)(DEC), A thin cylindrical lens does not have a unique optical
DEC = + 0 . 4 3 cm, center. Instead, each power meridian cross section has
an optical center (Figure 18.10a). Here, the line con-
DEC = 4.3 mm out.
taining these optical centers is referred to as the
In Figure 18.9a, the MRP has BI prism, so the cylinder axis. For a plus cylindrical lens the axis lies
optical center of the -7.00 D lens must be out from along the thickest part of the lens (Figure 18.10b). For
the MRP. This verifies that the decentration was
a minus cylindrical lens, the axis lies along the thinnest
out. Clinically, decentrations are specified in mm to
the nearest mm, so the clinical answer is 4 out. part of the lens (Figure 18.10c).
A cylindrical lens bends rays only in the direction
EXAMPLE 18.5 of the power meridian. The associated field of prism
A +6.00 D lens for the right eye is decentered powers have their base directions aligned with the
8 mm out relative to the MRP. What prism does the power meridian and perpendicular to the axis. The
decentration give? base directions point toward the axis on a plus cylinder
and away from the axis on a minus cylinder. For
example, a plus cylindrical lens with a horizontal axis
MRP meridian has a vertical power meridian, and bends
rays vertically (up or down) but not horizontally (Fi-
optical gure 18.11).
center
Consider a point B that is a perpendicular distance
h p from the axis of a cylindrical lens of power C
(Figure 18.12). Since the distance h p is aligned with
the power meridian, it can be used in Prentice's rule:
Z=-Ch where h p is in cm. (18.5)

EXAMPLE 18.6
What is the prism power at a point 11 mm vertically
down from the axis of a -3.00 x 180 lens?
From Eq. 18.5,
Z = - ( - 3 . 0 0 D ) ( - l . l cm) = - 3 . 3 ,
or
= 3.3 BD.

EXAMPLE 18.7
b) What is the prism power at a point 20 mm in from
FIGURE 18.9. a. Outward decentration for minus lens. b. the axis of a +4.00 x 90 lens in front of the right
Outward decentration for plus lens. eye?
384 Geometrie, Physical, and Visual Optics

side

optical
center

FIGURE 18.12. Parameters to determine prism power on a


front cylindrical lens.

\r we are considering a point below the axis, the base is


BD & I @ 60 (Figure 18.13). For a point above the
axis, the prism base is BU & O @ 60. These are the
only two possible base directions for the -3.00 x 150
b) left spectacle lens.
Since the base directions are easy to obtain, we can
put Eq. 18.5 in absolute values to obtain the mag-
nitude of the prism power:
|Z| = |Ch p |, where hp is in cm. (18.6)
In Figure 18.12, point B is a distance h from the axis.
The perpendicular distance h p to the axis is given by
hp = h sin ,
where is the acute angle that the directed distance h
FIGURE 18.10. a. Side view of a plus cylindrical lens. b. Front
view of a plus cylinder lens. c. Minus cylinder lens. makes with the axis. From Eq. 18.6,
|Z| = |C h sin \, where h is in cm.
(18.7)
From Eq. 18.5,
Z = -(+4.00 D)(+2.0 cm) = - 8 , EXAMPLE 18.8
Zelmo Corkscrew's left spectacle Rx is -5.00 x 60.
or When Zelmo reads, his line of sight passes through
= 8 BO. a point 12 mm below the axis. What is the prism
power at that point?
We can figure out the prism base directions on a The base direction is BD & O @ 150 (Figure
cylindrical lens by considering the location of the axis.
For example, on a -2.00 x 150 lens, the base direc-
tions lie along the 60 meridian and point away from -2.00x150
the axis. When the lens is in front of the left eye and

+ 2.00x180
FIGURE 18.13. Prism base directions for a - 2 . 0 0 x 1 5 0 in
FIGURE 18.11. Prism base directions on a +2.00 x 180. front of the left eye.
Prism Properties of Lenses 385

FIGURE 18.14. For left lens, prism at point 12 mm down


from the axis of a 5.00 x 60 lens. For right lens, prism at a
point 6 mm in from the axis of a -4.00 x 140.

18.14). For a 60 degree axis, and a vertical h, the or BD & I @ 135. In particular, decentration of the
angle is 30. From Eq. 18.7, -4.00 x 45 cannot give a BU prism (which is really a
|Z| = |(-5.00D)(1.2cm)(sin30)|, BU @ 90).
|Z| = (5.00D)(1.2)(0.5), The only cylindrical lenses that can give BU or BD
by decentration are those with a vertical power meri-
= (5.00 0)(0.6) = 3.
dian (i.e., a 180 axis meridian). The only cylindrical
So lenses that can give BI or BO are those with a
= 3 BD & O @ 150. horizontal power meridian (i.e., a 90 axis meridian).

EXAMPLE 18.9 EXAMPLE 18.10


Zelmo's right spectacle lens is a -4.00 x 140. What A -4.50 x 125 lens for the right eye is decentered
is the prism at a point 6 mm in from the axis? 5 mm in relative to the MRP. What is the prism
From Figure 18.14, the base direction is BU & I power at the MRP (and hence the prism in the
@ 50, and the angle is 40. From Eq. 18.7, Rx)?
Z = (-4.00 D)(0.6 cm)(sin 40), From Figure 18.15, the prism base direction at
the MRP is BD & O @ 35. From Eq. 18.7,
=1.5
|Z| = |(-4.50D)(0.5cm)(sin55)|,
So || = 1.8
=1.5 BU & I @ 50.
So
=1.8 BD & O @ 35.
When a cylindrical lens is decentered relative to the
MRP, the prism base directions at the MRP are In the clinic, the horizontal prism testing is done
aligned with the power meridian of the cylinder. This separately from the vertical prism testing. Therefore,
severely limits the prism base directions that can be the Rx is usually written in terms of the horizontal and
obtained by decentration. Consider an Rx for the right vertical prism components. A specific procedure for
eye of -4.00 x 45 combined with 1 BU. Decentration dealing with the horizontal and vertical components is
of the cylindrical lens can give either BU & O @ 135 given in Sections 18.10 and 18.11.

18.4 Rotational Deviations by a


Cylindrical Lens

Consider an emmtrope viewing a distant line through


a cylindrical lens. Not only might the distant line
appear blurred (Chapter 15), but it also might appear
to be rotated. We can use the prism power in the
cylinder to qualitatively explain the apparent rotation.
Figure 18.16a shows an ocular view of a distant
FIGURE 18.15. A -4.25 x 125 decentered in. vertical line viewed through a -3.00 x 135. The prism
386 Geometrie, Physical, and Visual Optics

distant
vertical i

distant NK/
horizontal
line

W
plus
ocular view b) ocular view c) v cylinder
axis

FIGURE 18.16. a. Ocular view of apparent rotation of a vertical line by a minus cylinder, b.
Ocular view of apparent rotation of a horizontal line by the same cylinder, c. Plus cylinder lens
giving same rotation as a.

at point A has its base direction up and to the left rotational deviations. Therefore, the plus cylinder
along the power meridian, so a point viewed through must give the same rotational deviations as the minus
A would appear to be at A', which is down and to the cylinder. In Figure 18.16c, the prism base at point F is
right along the power meridian (i.e., toward the down and to the left along the power meridian, and a
apex). There is zero prism power at B, so a point point viewed through F appears to be at F', which is
viewed through B would not appear deviated. The up and to the right along the power meridian (i.e.,
prism at C has its base down and to the right, so an toward the apex). There is zero prism at G, so a point
object viewed through C would appear to be at C viewed through G does not appear deviated. The
which is up and to the left. The apparent position of prism at H is up and to the right, so a point viewed
the vertical line is found by connecting A', B, and C . through H appears at FT, which is down and to the
The line appears to be rotated clockwise. left. The apparent line is found by connecting F', G,
Figure 18.16b shows a distant horizontal line viewed and FT. Indeed, the vertical line appears to be rotated
through the same minus cylinder lens. The prism clockwise (as in Figure 18.16a). You can repeat the
at point D has its base down and to the right along the argument for the horizontal line.
power meridian, so a point viewed through D would Note that if the above plus cylinder originally had a
appear to be at D', which is up and to the left along vertical axis and then was rotated counterclockwise,
the power meridian (i.e., in the direction of the apex). the distant vertical line would appear to rotate clock-
Again, there is zero prism at B, so a point viewed wise (against rotation). Now consider the transposed
through B does not appear deviated. At E, the prism minus cylinder (Figure 18.16a). If it originally had a
base is up and to the left, so a point viewed through E vertical axis and was then rotated clockwise, the dis-
appears at E', which is down and to the right. The tant vertical line appears to also rotate clockwise (with
apparent position of the vertical line is found by rotation). Hence, a general rule: against rotations
connecting D', B, and E\ The line appears to be occur around a plus cylinder axis, and with rotations
rotated counterclockwise. occur around a minus cylinder axis.
The fact that the two lines appear to be rotated in By transposition any lens with cylindrical power has
opposite directions is called the scissors effect. Any both a plus and a minus cylinder axis meridian. So
lens with cylinder power gives a scissors effect both rotations occur simultaneously. In other words,
whenever the object lines are not aligned with the the lines appear to open and close like a scissors.
principal meridians. When the object lines are aligned
with the principal meridians, no rotational deviation is
observed. 18.5 Rotational Deviations and
By transposition, the minus cylinder in Figures Spectacle Magnification
18.16a and 18.16b is equivalent to a spherical lens
combined with a plus cylinder, the axis of which is 90 The rotational deviations produced by cylindrical
from that of the minus cylinder. Since the sphere has power can also be explained by spectacle magnifica-
rotational symmetry, it does not contribute to any tion. Consider a -3.00 x 135 lens in front of an eye.
Prism Properties of Lenses 387

/ \ 18.6 Motions
no
4
ys loss \ In a spherocylindrical correcting lens for a compound
< > hyperopic astigmat, both principal meridians have plus
power; while in the correcting lens for a compound
myopic astigmat, both principal meridians have minus
object \/
power. We can use the with and against rotations of
b) the previous section to find the principal meridians.
Then we can use another type of movement (with and
against motion) to determine if the power in a princi-
power cross
pal meridian is plus or minus.
Consider a vertical principal meridian. An observer
can determine whether the power in that meridian is
image plus or minus by holding the lens close to his or her
eye, and moving the lens up or down. A real object
observed through the lens appears to move in the
object same direction when the lens is diverging (a
phenomenon called with motion) and in the opposite
direction when the lens is converging (against motion).
The latter motions assume that the lens is held close to
the eye. In particular, if converging light emerges from
the lens, it is assumed that the light is still converging
when incident on the eye. In other words, a physically
/ real image is not formed between the lens and the
observer's eye.
The prism power in a lens provides a simple way to
understand the motions. For a vertical principal meri-
dian with minus power, the prism base directions point
object H' away from the optical center. When the person looks
through a point above the optical center, he or she is
d) e) looking through a base up prism, and the object point
FIGURE 18.17. a. Ocular view of spectacle magnification for appears to be deviated down. When the lens is raised
3.00 x 135. b. Square object aligned with principal meridians of so that the person is looking through the optical
the lens. c. Superposition of image and object, d. Horizontal and center, the prism power is zero and the object is not
vertical diagonals in the object, e. Scissors rotation of diagonal in
the image.
deviated. When the lens is further raised so the person
is looking through a point below the optical center,
the prism is base down, and the object appears to be
The principal meridians of the lens are 45 degrees deviated up. So as the lens was raised, the apparent
away from the horizontal and vertical. In the power object position went from down to up, or went with
meridian, the spectacle magnification power factor the motion of the lens.
gives a percent loss (Figure 18.17a). In the axis meri- For plus power in the vertical meridian, the base
dian, the spectacle magnification power factor gives directions point toward the optical center, and a simi-
neither a loss nor a gain. lar analysis gives against motion provided the eye is
Figure 18.7b shows a distant square with its sides immediately behind the lens. A piano lens (zero diop-
aligned with the principal meridians of the 3.00 x tric power) gives no motion.
135. The square is then imaged as the rectangle (solid The with and against motions form the basis of
lines) shown in Figure 18.17c (ocular view). Figure hand neutralization. In hand neutralization, a known
18.17d shows the diagonals on the object square trial lens is combined with an unknown ophthalmic
(dashed). For the object square, the diagonals are lens. When the resultant power is plus, the combina-
horizontal and vertical. The rotational orientation of tion gives against motion. When the resulting power is
the diagonals in the image is found by drawing the minus, the combination gives with motion. When the
diagonals in on the image rectangle as shown by the resulting power is zero, the combination gives no
solid lines in Figure 18.17e. The diagonals are scis- motion. The power of a thin unknown lens is then
sored in the image in exactly the same manner pre- equal in magnitude but opposite in sign to that of the
dicted by the prism argument. trial lens that neutralizes the motion.
388 Geometrie, Physical, and Visual Optics

Thick lenses provide a slight complication in the 2 BU while along a line 2 cm above the axis the prism
analysis. When the first lens is neutralized by the power is 4 BU (Figure 18.18b). A set of iso-prism
second lens, it is the back vertex power of the first lens lines would run parallel to the axis.
that is equal in magnitude but opposite in sign to the We can make a Fresnel equivalent to this lens by
neutralizing power of the second lens. again using prism sections. The sections all run
Ophthalmic lenses usually have a bent form (con- horizontal, but this time each section has the prism
cave back surface), and a trial lens placed against the power given by Prentice's rule for the cylinder (Figure
back gives an air gap. To avoid errors introduced by 18.18c). Figure 18.18d shows front and side views of
the air gap, the trial lens is placed against the front of the Fresnel cylindrical lens.
the unknown ophthalmic lens. This means that the A Fresnel cylinder always has its sections laid out
unknown lens parameter determined by hand neutrali- parallel to the axis. For a plus cylinder, the bases point
zation is the neutralizing power (or front vertex) as toward the axis. The prism power is again given by
opposed to the back vertex power (which is the Prentice's rule for the cylinder.
parameter specified in the Rx). A Fresnel spherical lens can be made in a similar
manner. On a spherical lens, the iso-prism curves are
spherical and centered on the optical center. There-
fore, the prism sections must be annular sections with
18.7 Fresnel Lenses their prism power again given by Prentice's rule.
Figure 18.19 shows front and side views for a plus
Figure 18.18a shows a -2.00x180 cylindrical lens. spherical Fresnel lens.
The prism base directions all point away from the axis. One advantage of a Fresnel lens is that they can be
Along a line 1 cm above the axis, the prism power is made with a large diameter. Augustin-Jean Fresnel
originally invented this type of lens for light houses
where a very large high plus collimating lens was
needed. The weight and thickness in a regular spheri-
2 cal lens was prohibitive, but the Fresnel lenses solve
this problem. A common use of a high plus large
diameter Fresnel lens is in the overhead projector.
Fresnel lenses are also used in car and truck head-
lights.
Plastic membrane Fresnel lenses have ophthalmic
uses for temporary corrections. They can also offer
a) b) cosmetic advantages for high power lenses such as a
person needing a -22.00 D spectacle correction.

FIGURE 18.18. a. - 2 . 0 0 x 1 8 0 lens. b. Isoprism lines, c. FIGURE 18.19. Front and side views of Fresnel spherical
Fresnel lens. d. Fresnel lensfront and side views. lenses.
Prism Properties of Lenses 389

18.8 Prism Effectivity with Lenses


u|tanc| + | D E C | ( A + l )
Figure 18.20 shows a corrected hyperope viewing an |tan e| =
off-axis object point through a decentered spherical
lens. Measured at the MRP, the object point makes an
|tanc|
1 + |DEC|(,v n + -
angle c with the straight ahead line. The prism at the ' ' \u\ v
MRP produces a deviation angle d, and the eye |tan e| =
rotates through an angle e around its center of rota- 1
V
tion. The distance from the spectacle plane to the
center of rotation is C rot . Since \u\ = -u, U = 1/w, and V= llv,
From the triangle involving the object, and the , |tanc| + |DEC|(-U + V)
MRP, we obtain |tane|=J ' '
i-crotv
O - |DEC|
|tan c| = [tancl + lDEClP
|tan e| =
i - crrttv
or
Dimensionally, DEC is in meters in the above equa-
0 = | n | | t a n c | + |DEC|.
tion. We can multiply both sides by 100 to obtain
From the triangle involving the center of rotation of _ llQ0tancl + lDEC(incm)|P
the eye and the spectacle lens' image point, we obtain 100 tan el
i-c r o t v
III + |DEC| A prism diopter value for the angle c is
|tan e| =
v - Crnt Z c = 100 tan c.
Since Just as for Z e , when the object point is in from the
straight ahead direction, a BO direction is assigned to
IH = |mO| = R 0, Z c . Similarly, when the object point is up from the
forward direction, a BD direction is assigned to Z c . In
it follows that other words, the base direction assigned is opposite
the actual deviation, just as the virtual image formed
|i| = o(l M H tanc l + lDECl)> by the prism appears deviated opposite the base direc-
tion (i.e., in the direction of the apex). Then
or
|z | + |z|
|l| = u|tanc| + p-jlDECl. |zl = i -c cv '
Then In this derivation, the prism Z makes the object look
even farther away from the straight ahead line, so Z c
u|tanc|+ |-r |DEC| + |DEC| and Z add. When the prism makes the object point
|tan e| = appear to be closer to the straight ahead line (or even
v - Cmt past it on the opposite side), then Z c and Z subtract.

/- t

\erV \

\
[DEC M R P
^ ^y^x 1 I DEC

o.e.
>i ^^^S^K
FIGURE 18.20. Prism effectivity (e^d)
for a decentered lens and an object point
that subtends angle c at the MRP.
390 Geometrie, Physical, and Visual Optics

In general, Z c and Z act like vectors, so we can The effective prism power decreases as Michael
remove the absolute value signs if we consider the sum approaches the bolt.
in the numerator as a vector sum:
EXAMPLE 18.12a
8 8
=^ < > Chuck Creampuff wears an Rx of -13.00 D com-
bined with 6 BD at a vertex distance of 12.5 mm.
Equation 18.8 is the general (paraxial) prism effectivi- How much does Chuck's eye turn when viewing a
distant cream pie that is straight ahead?
ty equation.
The center of rotation of the eye is typically
For an object point straight ahead, 13.5 mm behind the eye, so
z c = o, C r o t = 12.5 mm + 13.5 = 26 mm = 0.026 m.
and Equation 18.8 simplifies to For the distant pie, U = 0, and V = P . Then Eq.
18.8 gives
Z
z = i-c r f t t v Z =
6
l-(0.026)(-13.00D)'
When there is only a prism in the spectacle plane,
V = U , and the above equation is identical to Eq.
17.19. When there is both dioptric power and prism 6
Z = = 4.5 BD.
power, then V = P + U, and the denominator depends 1.34
on the dioptric power. Even though the prism power is 6 BD, the eye has
only to turn through an angle corresponding to 4.5
Examples for Zc=0 BD. This correlates with the fact that the spectacle
magnification for the minus lens is less than one. In
other words, the minus lens minifies.
EXAMPLE 18.11a
Michael Bolteater wears an Rx of +13.00 D com-
bined with 6 BD at a vertex distance of 12.5 mm. EXAMPLE 18.12b
How much does Michael's eye turn when viewing a Chuck walks toward the cream pie until it is 25 cm
distant bolt that is straight ahead? from the spectacle plane. How much does Chuck's
The center of rotation of the eye is typically eye turn?
13.5 mm behind the eye, so Here U = - 4 D , and V = -13.00 D + (-4.00
D) = -17.00 D. Then Eq. 18.8 gives
C rot = 12.5 mm + 13.5 = 26 mm = 0.026 m.
6ABD
For the distant bolt, U = 0, and V = P . Then Eq. Z =
18.8 gives l-(0.026)(-17.00 D ) '
6ABD or
l - ( 0 . 0 2 6 ) ( + 13.00D)' 6ABD
Z = = 4.2 BD.
or 1.44
6 Again, the effective prism power decreases as the
Z = = 9.1 BD.
0.66 distance between the object and the lens decreases.
Even though the prism power is only 6 BD, the
eye turns through an angle corresponding to 9.1 .
This correlates with the fact that a plus lens has a Examples for Z = 0
spectacle magnification that is greater than one. In
other words, the plus lens magnifies. EXAMPLE 18.13a
Caryn Smith has an Rx of +8.00 D for each eye.
EXAMPLE 18.11b The vertex distance is 10 mm. The distance be-
Suppose Michael Bolteater walks toward the bolt tween the centers of rotation of Caryn's left and
until it is 25 cm from the spectacle plane. Then how right eyes is 64 mm. Caryn holds a book 40 cm from
much must his eye turn? the spectacle plane. How much must Caryn turn
Here U = - 4 D , and V = +13.00 D + (-4.00 her left eye to view a centrally located letter that is
D ) = +9.00 D. Equation 18.8 gives 40 cm from the spectacle plane?
From Figure 18.21a,
6*BD
l-(0.026)(+9.00)' Z c = 100 tan c = 100 ^ = 8.0 BO.
40
A
BD
z = 60.77 = 7.8*BD.
Since Caryn has a plus lens, the rays bend toward
the optical center as shown in Figure 18.3a. From
Prism Properties of Lenses 391

the figure we expect Z e > Z c . For the 40 cm object


distance.
V = (+8.00 D) + (-2.50 D)
= +5.50D,
C r o t = 10 mm + 13.5 mm
64mm = 23.5 mm = 0.0235 m.
3.2cm J Since there was not any prism in the Rx, Z = 0, and
from Eq. 18.8,
8
Z =
a) l-(0.0235)(+5.50D)'
8ABO = 9 2 A B Q
e VZ b U
0.87 '
Suppose Caryn can accommodate enough to see
the letter without her lenses. Then how much must
she turn her eye?
The angle c is still measured from the spectacle
plane, but now P = 0. Therefore, V = U = -2.50 D,
64mm and
BO
Z =
l-(0.0235)(-2.50D)'
_8^BO
Ze
~ 1.059 ~ 7 6 B -
Thus, without her glasses Caryn turns her eye
Crot = 23.5mm
through an angle corresponding to 7.6 BO.
Note from Figure 18.21b, with no lens in place,
we can find the tangent of e directly:
3.2 cm
tane : = 0.076.
40 c m + 2.35 cm
So
Z = 100 tan e = 7.6 BO.
The increase with the plus lenses correlates with
the percent gain in the spectacle magnification pro-
vided by the plus lens.

EXAMPLE 18.13b
Rebecca Ratchet has an Rx of -8.00 D for each
23.5 eye. The vertex distance is 10 mm. The distance
between the centers of rotation of Rebecca's left
and right eyes is 64 mm. Rebecca holds a book
40 cm from the spectacle plane. How much must
Rebecca turn her left eye to view a centrally lo-
cated letter that is 40 cm from the spectacle plane?
From Figure 18.21c,

Z cc = 100 tan c = 100 ^ = 8.0 BO.


40
Also note from Figure 18.21c that we expect Z e <
Z c . For the 40 cm object distance,
V = (-8.00 D) + (-2.50 D)
= - 1 0 . 5 0 D,
FIGURE 18.21. Prism effectivity or convergence demands, a.
Plus lenses, b. No lenses, c. Minus lenses, d. Base in prisms. C rot = 10 mm + 13.5 mm
= 23.5 mm = 0.0235 m.
392 Geometrie, Physical, and Visual Optics

Since there was not any prism in the Rx, Z = 0, and 18.9 Iso-thickness Curves
from Eq. 18.8,
8 The iso-thickness curves for a spherical lens are circu-
lar and centered on the optical center. The iso-thick-
l - ( 0 . 0 2 3 5 ) ( - 1 0 . 5 0 D) '
ness curves for a cylindrical lens are straight lines that
8 are parallel to the axis. Now let us consider the
= 6.5 .
1.24 iso-thickness curves for spherocylindrical lenses.
You may show that without her spectacles Rebecca For a + 5 . 0 0 0 - 3 . 0 0 x 9 0 lens, the power in the
would have had to turn each eye 7.6. The de- vertical meridian is +5.00 D, and the power in the
creased demand is due to the smaller virtual image horizontal meridian is +2.00 D. The thickest point on
formed by the minus lens. the lens is at the optical center. The thickness de-
creases for points progressively farther away from the
Examples for Non-Zero Zc and Z optical center. The decrease in thickness is slower
along the +2.00 D horizontal meridian than along the
EXAMPLE 18.14a +5.00 D vertical meridian. Therefore, the iso-thick-
Suppose Caryn Smith in Example 18.13a has 3 BI ness curves are asymmetric. For a +5.00 O - 3 . 0 0 x 90
prism in her Rx. If all other parameters are the lens, the paraxial iso-thickness curves turn out to be
same, how much must Caryn turn her eye to see horizontally oriented ellipses (Figure 18.22a). At a
the centrally located letter? point on the lens, the prism base direction is directed
From Example 18.13a, inward (since the optical center is the thickest point).
Z c = 8 BO, C rot = 0.0235 mm, and However, the base directions in general do not point
V=+5.50D. directly toward the optical center. Instead, they are
perpendicular to the iso-thickness curve at that point.
The 3 BI prism makes the centrally located letter
The iso-thickness curves for a spherocylindrical lens
appear closer to the straight ahead position. This
shows up mathematically in the numerator of Eq. are ellipses as long as the power in both principal
(18.7), meridians has the same sign (i.e., both plus or both
minus). The ellipses have their major axis aligned with
8 + 3
the meridian in which the power is least in magnitude.
l-(0.0235)(+5.50D)'
For example, the power cross for a - 2 . 5 0 O - 1 . 5 0 x
5 90 has - 2 . 5 0 D vertically and - 4 . 0 0 D horizontally. In
Z = = 5.7 . this case, the optical center is the thinnest point on the
0.87

EXAMPLE 18.14b
Suppose Caryn's prism is 3 BO (instead of the 3 + 5.00D.
BI above). If all other parameters are the same,
how much must Caryn turn her eye to see the
centrally located letter?
From Example 18.13a,
Z c = 8 BO, C rot = 0.0235 mm, and + 2.00D.
V=+5.50D.
The 3 BO prism makes the centrally located letter
appear farther from the straight ahead position.
Again this affects the numerator of Eq. 18.8,
8 + 3 -2.50D.
Z =
l-(0.0235)(+5.50D)'

Z = 1 ^ = 12.6 BO.
0.87
-4.00D.
EXAMPLE 18.14c
What prism would Caryn need to make the cent-
rally located letter look straight ahead?
From the above examples Z c was 8 BO. An 8
BI prism will make the object appear straight
ahead (Figure 18.21d). In Eq. 18.8, the numerator FIGURE 18.22. Iso-thickness curves and prism base directions.
would be zero, and thus, Z would be zero. a. +5.00C-3.00 x 90. b. -2.50 O-1.50 X 90.
Prism Properties of Lenses 393

lens, and the thickness increases slowest in the vertical and


meridian. The iso-thickness ellipses are vertically
oriented (Figure 18.22b). For the minus power in both =50.8.
principal meridians, the prism base directions point So the two meridians are
out toward the edges.
90 + 50.8=140.8,
When the power in the principal meridians have
opposite signs, then the iso-thickness curves are not and
elliptical. Consider a + 3 . 0 0 0 - 5 . 0 0 x 9 0 lens. In the
9 0 - 5 0 . 8 = 39.2.
vertical meridian, the power is +3.00 D, and the
thickness is a maximum at the optical center. In the The iso-thickness curves are straight lines along the
horizontal meridian, the power is -2.00 D and the 39.2 and 140.8 meridians.
thickness is a minimum at the optical center. So starting The curves in between those straight lines are
at the optical center, the lens gets thinner vertically and hyperbolas, which are asymptotic to the straight lines
thicker horizontally. For such a lens, there is some (Figure 18.23a). The hyperbolas are symmetric rela-
off-axis meridian for which the thickness does not tive to each principal meridian. The minus power gives
change. Since the curvature varies like the sine-squared one set of hyperbolas and the plus power gives the
law, the meridians for which the thickness does not other set of hyperbolas.
change are found by setting Pk equal to zero in Eq. At any point on the lens, the prism base directions
16.1, i.e., are perpendicular to the iso-thickness curves. For the
0 = S + C sin2 , hyperbolas associated with the minus power, the prism
base directions point toward the edges (which are
where is the angle with the cylinder axis. For the thicker). For the hyperbolas associated with the plus
above lens, power, the prism base directions point away from the
edges (which are thinner). Along the straight lines, the
0 =+3.00 + (-5.00) sin2 .
prism base directions point toward the nearest minus
Then power meridian (which is thicker).
2 , -3.00 D n _ For a spherocylindrical lens, the iso-thickness cur-
Sin
= ^5=060> ves are either circles, ellipses, hyperbolas, or straight
lines. Given the power cross of the lens, it is relatively
or easy to make a quick estimate of the iso-thickness
sin = 0.77, curves.

+ 3.00D

-2.00D.
-0.25D

+ 3.00D.

+ 3.00D + 3.00D

+ 1.50D
FIGURE 18.23. Iso-thickness
curves, a. +3.000-5.00x90.
b. + 3 . 0 0 0 - 3 . 2 5 x 9 0 . c. +3.00
O -3.00 x 90 or +3.00 x 180. d.
+3.00O-1.50X90.
394 Geometrie, Physical, and Visual Optics

Suppose the cylinder power is changed to -3.25 D urements (Z x ) are made separately from the vertical
(i.e., the lens parameters are +3.00 0 - 3 . 2 5 x 9 0 ) . prism measurements (Z y ).
The power is +3.00 D in the vertical meridian and Consider a thin spherocylindrical lens with parame-
-0.25 D in the horizontal meridian, and the iso-thick- ters S c C x . We can describe an incident ray by the
ness curves are still hyperbolic. The meridians for position parameters hx and hy (measured from the
which Pk is zero are: optical axis) and the directional parameters qx and qy
(Figure 18.24). A similar description exists for the
0 = + 3 . 0 0 + (-3.25) sin2 </>,
emerging ray. One of the central results of Chapter 16
or is that the incident and emerging ray parameters are
related by the matrix equation
sin2 </>-^ =0.92.
1 0 -P -pt\
So /q;\
=74. 0 1 -P f -Pv
The meridians with straight line iso-thickness curves h: 0 0 1 0
are
90 + 74 = 164, \h y / \o 0 0 0 / \hy/
90 - 74 = 16. We can write the above equation more concisely as:
Figure 18.23b shows the resulting hyperbolas. A' = RA,
When the cylinder power is changed to -3.00 D,
the lens parameters are +3.00 O-3.00 x 90. By trans- where R is a 4 x 4 refraction matrix, A is the 4 x 1
position, the lens can also be described as a +3.00 x incident ray matrix, and A is the 4 x 1 emerging ray
180 cylindrical lens, and the iso-thickness curves are matrix (Equation 16.21). As discussed in Chapter 16,
straight horizontal lines (Figure 18.23c). Note how the we can partition the 4 x 4 refraction matrix into four
hyperbolas smoothly change into the horizontal lines. 2 x 2 matrices:
Suppose the cylinder power is now decreased to
-1.50 D (i.e., the lens parameters are +3.00 C R
-1.50x90). Now the previous straight lines change 1
smoothly into the elliptical iso-thickness curves (Fig-
ure 18.23d). As the cylinder power is changed to zero, where P is the 2 x 2 dioptric power matrix
the iso-thickness curves change into circles.
The prism base direction at a particular point on a /P v
lens is always perpendicular to the iso-thickness curve P=
at that point, and points in the direction of the thicker ,P. P
part of the lens (or away from the thinner part). So
when one can sketch the iso-thickness curves, one can
make an estimate of the prism base direction at that
point.
If is helpful to note that on a centered lens with a \,
round shape, the minus cylinder axis meridian always y
has the thinnest edge and the plus cylinder axis meri-
dian always has the thickest edge. An easy way to
remember this is to just think of elementary minus and
(hx'.hy')
it
plus cylinder lenses (Figures 18.10b and 18.10c). y\ y
A 1
/ '
k/
/ 1
/ J /

18.10 Prism Components in Spherocylindrical / / ^

w
/ X
Lenses 7
x^q*

The human visual system is neurologically wired to


(hx,hy) /
treat horizontal and vertical eye movements different-
ly. Hence, in the clinic, the horizontal prism meas- FIGURE 18.24. Horizontal and vertical ray parameters.
Prism Properties of Lenses 395

the elements of which are given by Let us first check the iso-thickness curves to
2 develop some intuition about the answer. Since one
Px = S + C sin 0, power is plus and the other is minus, the iso-
Py = S + C cos2 , thickness curves are hyperbolas. The methods of
P t = - C sin cos 0. the previous section show that the straight line
boundaries are 45 away from the principal meri-
The refraction matrix multiplications for the ray angles dians. The iso-thickness curves are shown in Figure
are equivalent to the following matrix equations: 18.25. From these curves, the prism base compo-
nents at the reading center are up and in.
For the calculations, we first need the dioptric
power matrix:
/-1.32D -2.70 D^
P =
The horizontal dx and vertical dy components of the
deviation are given by \-2.70D +1.32 Dy
Then
dx = q x - q x , -1.32 D -2.70 D \ / - 0 . 3 cm
and
\Zyl \-2.70D +1.32 D / \ - 0 . 9 c r J
dy = q ; - q y
The matrix multiplications give
From the above three equations,

x
tz-\ /+2.8 \ /-2.8 ^

CM::
where hx and hy are in meters. We can multiply both
-0.4 / V+0.4
The horizontal component is 2.8 while the
vertical component is 0.4 BU. The in and up
sides by 100, and note that for paraxial rays directions agree with the iso-thickness curves.

/Zx\ /100d x \ EXAMPLE 18.16


Henritta Germ wears a -2.00 O-5.00 x 50 in front
\ z j \ioodJ' of each eye. When Henritta reads she looks
through each lens at a point 11 mm down and 4 mm
Then for h and hv in cm, in from the optical center. What prism components
A y
occur at that point? What is the difference in the
/Z\ /P v R\/h\ right and left vertical prism components?
(18.9) For each eye,
\Pt P
y/\hyy
-4.93 D +2.46 D
or in matrix form P=
+2.46 D -4.07 D
Z = -Ph, (18.10)
where
+ 3.00D

3.00D.
Z=i h
^J' \hj
Eq. (18.10) is a matrix form of Prentice's rule. It has
the advantage of being valid for any cylinder axis 0.

w
- horizontal
The signs used are given by the coordinate system in
Figure 18.2.

EXAMPLE 18.15
-3.00D.
Myron Delshrimpsky wears a +3.00C-6.00 x 122
in front of his left eye. When Myron reads, he
+ 3.00D.
looks through the lens at a point 9 mm down and
3 mm in from the optical center of the lens. What FIGURE 18.25. Iso-thickness curves for a + 3 . 0 0 C - 6 . 0 0 X
are the prism components at this point? 122. Prism base direction at indicated positions.
396 Geometrie, Physical, and Visual Optics

For the right eye: For the horizontal component Z x , the torsional
component Pt supplies the coupling with the vertical
/Z\ /-4.93D +2.46 D V + 0 . 4 cm meridian. Similarly, for the vertical component Z y ,
the torsional component P t supplies the coupling to the
\Zy/ 1+2.46 D -4.07D/\-l.lcm horizontal meridian.
The matrix multiplications give For spherical lenses, or for spherocylindrical lenses
with axis 90 or 180, the torsional component Pt is zero,
/z*\ /-47\ /+ 4 7 * in which case the horizontal and vertical meridians are
uncoupled and
Uy/ \+5.5/ \-5.5 Zx = - P x h x , when Pt = 0,
or the horizontal component is 4.7 , and the while
vertical component is 5.5 BD.
The iso-thickness curves are elliptical, and the Z
y = -P y h y , when Pt = 0.
base in and down components are consistent with Thus, for the case of a zero off-axis element P t , we
the curves (Figure 18.26). only need the vertical distance h y to compute the
For the left eye:
vertical component, and only the horizontal distance
/Z\ /-4.93D +2.46 D \ / - 0 . 4 cm h x to compute the horizontal component.

\Zy/ 1+2.46 D -4.07 D / \ - 1 . 1 cm EXAMPLE 18.17


Angela Landkey wears a +9.00 D spherical lens.
The matrix multiplications give When Angela reads, she looks through a point
5 mm in and 12 mm down from the optical center of
'-0.7\ /+0.7 her lens. What is the vertical component of the
prism at the reading center?
+3.5Z -3.5' In this case Pt is zero and Py is +9.00 D. So
or the horizontal component is 0.7 BO and the Z y = - ( + 9 . 0 0 D ) ( - 1 . 2 c m ) = +10.8 ,
vertical component is 3.5 BD. or the vertical component is 10.8 BU. Note that
The iso-thickness curves in Figure 18.26 predict since Pt equals zero, we only needed the vertical
that the prism at the reading center should be down distance hy in this computation.
and out, which agrees with the above calculations.
The difference in the vertical components is 5.5
minus 3.5 or 2. Typically, a vertical prism differ- EXAMPLE 18.18
ence of 2 or more is enough to break fusion, so Scott Bushsnake wears a + 2 . 0 0 0 - 8 . 0 0 x 1 2 0
that the person sees double. Thus, Henritta might spectacle Rx (left eye). What is the prism at a point
experience double vision while reading. If Henritta 12 mm below the optical center? Also, what is the
does not experience double vision, she might ex- vertical prism at Scott's reading center that is
perience asthenopic symptoms, such as itching, 12 mm down and 5 mm in from the optical center?
burning, tearing, and/or fatigue. The dioptric power matrix is

-4.00 D -3.46 D
The explicit matrix multiplications, Eq. 18.9 give P=
Zx = - P x h x - P t h y , -3.46 D 0.00 D
and The curious phenomenon here is that Py is zero,
Zy=-Pthx-Pyhy. but Pt is not. From Prentice's rule,

-2.00D.
-4.00 D -3.46 D 0
-7.00D. -2.00D.
-7.00D -3.46 D 1.2 cnW

4.2Z -4.2'

0 0
So, for the left eye, the prism at a point 12 mm
below the optical center is 4.2 BI.
FIGURE 18.26. Iso-thickness curves and prism base direction The iso-thickness curves for this lens are hy-
at indicated positions on 2.00 O5.00 x 50 lenses. perbolic (Figure 18.27). The straight line boun-
Prism Properties of Lenses 397

+ 2.00D. vertical away from the axis as they move from the MRP to the
reading center. In Figure 18.28a, the eyes move to-
6.00D. ward the axis, so the prism components are relatively
small. In Figure 18.28b, the eyes move away from the
axis, so the prism components are larger. In Figure
18.28c, the axes are the same, but the right reading
center is away from the axis while the left reading
center is towards the axis. In the latter case, there may
be vertical imbalance problems even when the Rx is
-6.00 thick the same for the two eyes.
+ 2.00 thin
EXAMPLE 18.19
FIGURE 18.27. Iso-thickness curves for a lens with a vertical Laura Beans has an Rx of -1.00C-5.00 x 60 for
dioptric power curvature component of zero. Prism base direc- each eye. Laura's reading center is 8 mm down and
tions at indicated positions. 5 mm in from the optical center. Compare the
vertical prism powers at the reading center.
The dioptric power matrix is
daries are 30 on either side of the axis (i.e., at 90
and 150). At 12 mm below the optical center, the -4.75 D +2.16 D
iso-thickness curve is a straight vertical line with P=
the thicker part on the side toward the nose. This .+2.16D -2.25 D/
confirms that the prism at that point is BI.
At the reading center: For the right eye,
/Z\ /-4.75D +2.16D\/+0.5cm^
-4.00 D -3.46 D \ / - 0 . 5 cm
\Zy/ V+2.16D -2.25 D/\-0.8cmy
-3.46 D O -1.2 cm

+6.2Z -6.2*

+ 1.7' -1.7'
At the reading center, the horizontal prism is 6.2
BI and the vertical prism is 1.7 BD. Note that the
vertical prism at the reading center occurs in spite
of the fact that Py is zero. Figure 18.27 shows the
down and in prism direction at the reading center.

At the reading center, the difference in vertical


prism powers is called the vertical prism imbalance.
When considering potential vertical prism imbalance
problems for spherocylindrical lenses, it is important
to keep in mind whether the eyes move toward or

FIGURE 18.28. a. Reading centers toward the axis. b. Reading centers away from the axis. c.
Same axis gives one reading center away and one toward the axis.
398 Geometrie, Physical, and Visual Optics

Even though the lenses are identical, the vertical In Section 18.3, we discussed that an arbitrary
prism at the right reading center is 2.9 BD, while prism cannot be obtained by decentering a cylindrical
the vertical prism at the left reading center is 0.7 lens (S = 0). For such a cylindrical lens, det P is zero,
BD. The vertical prism imbalance is 2.2 (i.e., and the inverse matrix does not exist. For spherocylin-
2.9-0.7). drical lenses with a non-zero determinant, this is not a
problem.

EXAMPLE 18.20
18.11 Decentration with Spherocylindrical In Example 18.18, Scott Bushsnake had a
Lenses - f 2 . 0 0 C - 8 . 0 0 x 120 spectacle Rx (left eye). Sup-
pose Scott's Rx also calls for a 3 BU prism. What
In some spherocylindrical Rx's, horizontal a n d / o r ver- decentration (relative to the MRP) will give the
tical prisms can be supplied by decentration. For a prism?
spherocylindrical lens, the matrix version of Prentice's The dioptric power matrix is
rule is
-4.00 D -3.46 D
Z=-Ph. P=
-3.46 D 0.00 Dy
Let DEC be a 2 x 1 decentration matrix where
The determinant is
/DECV\ d = (-4.00 D)(0.00 D) - (-3.46 D)(-3.46 D),
DEC=| d= -12.00 D 2 .
^DEC
The inverse matrix is
Then, as in the derivation of Eq. 18.4, 0.00 D +3.46 D \
1
DEC = - h ,
12.00 D z +3.46 D -4.00 D /
and
The prism matrix is
Z = PDEC.

We can multiply both sides from the left by the inverse


of P to obtain
DEC = P Z. (18.11) Then
DEC = P Z,
The inverse of a 2 x 2 matrix is found by exchanging
or
the diagonal elements, changing the sign of the off-
diagonal elements, and dividing by the determinant d. 0.00 +3.46V 0
Given the dioptric matrix P, where
-12.00 +3.46 - 4 . 0 0 / \ +3
s
IP
1
PX
X t
P = + 10.39
P PI
\ t x
y/ 12.00 -12.00
then
DEC. -0.87 cm
1/ p
y -p
d DEC, + 1.0cm
\ -P. P.
For clinical purposes, the decentrations are round-
where
ed off to the nearest millimeter. Relative to the
rf=PxPy-PtPt. MRP, the decentration that gives the prism is
8.7 mm in and 10 mm up. Figure 18.29 shows the
From Eq. 16.38, the determinant also equals the iso-thickness curves for the decentered lens. The
products of the powers in the two principal meridians, prism at the MRP is BU. Note that even though
or only vertical prism was called for in the Rx, the
decentration has both horizontal and vertical com-
d = S(S + C). ponents.
Prism Properties of Lenses 399

As an aside, it is interesting to note the behavior of


a dioptric power matrix for a 90 clockwise rotation of
the lens. With this rotation, the diagonal elements Px
and Py change positions, and the off-diagonal elements
change sign. Thus, the inverse matrix for a lens with
parameters S c C x could be calculated by comput-
ing the matrix for the lens S o C x (0 90) and then
dividing that matrix by the determinant [which equals
S (S + C)].

FIGURE 18.29. Lens that has been decentered up and in


relative to the MRP. Problems

1. George Jones wears a - 8 . 0 0 D spectacle correc-


EXAMPLE 18.21 tion. George looks through the lenses at a point
Ruby Rose has a right Rx of - 7 . 0 0 0 - 3 . 0 0 x 3 5 7 mm above the optical center. What is the prism
combined with 5 BI combined with 2A BD. What power at this point on the lenses?
decentration relative to the MRP gives the prism? 2. Sally Smith wears +5.00 D spectacles. When Sally
The dioptric power matrix is reads she converges her eyes and looks down so
/-7.99D + 1.41 D \ that her line of sight passes through the lens at a
point 8 mm and 3 mm in from the optical center.
1+1.41 D -9.01 D / What is the prism power at this point on the lens?
3. Peter Smith wears +7.00 D spectacles. Peter looks
The determinant is 6 mm up and in with his right eye along a line 30
d = (-7.99 D)(-9.01 D) - ( + 1.41 D)( + 1.41 D), with the horizontal. What is the prism at the point
d= +70.00 D 2 . on the lens that Peter looks through?
4. Michael O'Rielly has an Rx of O D - 4 . 0 0 3 BI;
The inverse matrix is
OS - 5 . 0 0 0 3 BI. What decentration will result
/-9.01D -1.41 D \ in the proper Rx?
5. Patricia O'Rielly has an Rx of O D +6.00 0 4
= 2
+70.00 D ( - 1 . 4 1 D -7.99D/ BU. What decentration will result in the proper
The prism matrix is Rx?
6. A - 4 . 5 0 D lens is decentered 8 mm in. What is

OC)
the prism in the Rx?
7. A +3.00 D lens is decentered 4 mm up. What is
the prism in the Rx?
8. Maria Corlione wears an OS spectacle lens of
Then
- 2 . 0 0 x 40.
DEC = P Z,
or a. When Maria looks through the lens at a point
10 mm vertically down from the axis, what is
^DEC\ /-9.01 -1.41\/+5\ the prism power of the lens at that point?
= b. When Maria looks 10 mm horizontally in from
DECJ +70.00|_ 1 4 1 _ 7 . 9 9 _ 2 ) ' the axis, what is the prism power at that point?
c. When Maria looks 10 mm down and 3 mm in,
DEC\ /-42.25\ what is the prism power at that point?

= 9. a. Can a - 8 . 0 0 x 90 O D lens be decentered to
DECJ Tmolv +893j' give 3 BI in the Rx? If so, how much?
b. Can a - 8 . 0 0 x 180 O D lens be decentered to
give 3 BI in the Rx? If so, how much?
DEC\ / - 0 . 6 0 cm \ c. Can a - 8 . 0 0 x 150 O D lens be decentered to
give 3 BI in the Rx? If so, how much?
DEcJ 1+0.13 cm/' 10. Claudine Longneck puts on a pair of prisms with
Relative to the MRP, the decentration that gives the power O D 10 BI, OS 10 BI. Assume the
the prism is 6 mm out and 1 mm up. You may center of rotation is 25 mm behind the spectacle
check the iso-thickness curves for consistency. plane.
400 Geometrie, Physical, and Visual Optics

a. When Claudine looks at a distant object, how 13. a. Compare the answer for problem no. 12 to the
much and in what direction does her eye turn? answer than one would get for Briget wearing
b. When Claudine looks at a book 20 cm in front only +15.50 D lenses and nothing else (i.e., no
of the prisms, how much and in what direction prism).
does her eye turn? b. Compare the answers above to the answers
11. Briget Barbot has grown old and had cataract that one would obtain prior to cataract surgery
surgery on both eyes. Following surgery Briget's (i.e., assume Briget wore no correction prior
Rx is +13.006 . When Briget looks straight to surgery).
through her lenses at a distant object, how much 14. What are the horizontal and vertical prism compo-
and in what direction must she turn her eye? nents at a point 12 mm down and 5 mm in from
(Assume the center of rotation is 27 mm behind the optical center on a -2.00 O6.00 x 56 correc-
the spectacle lens.) tion for the right eye?
12. For near vision Briget Barbot wears +15.50 D 15. What are the horizontal and vertical prism compo-
lenses combined with 6 BI. The +15.50 D lenses, nents at a point 9 mm down and 4 mm in from the
which have +2.50 D more plus than her distance optical center on a +2.00 0 - 6 . 0 0 x 1 2 5 correc-
vision lenses, enable Briget to read a book 40 cm tion for the left eye?
in front of the lenses. Assume the horizontal 16. What horizontal and vertical decentration compo-
distance between the center of rotation of Briget's nents would provide the prism in an Rx of
eyes is 62 mm, and that the center or rotation is -7.00 O-4.00X 1524 BI 2 BU for the left
27 mm behind the spectacle lens. How much must eye?
she converge her eyes to read a letter that is 17. What horizontal and vertical decentration would
centrally located (i.e., 31mm horizontally from provide the prism in an Rx of +5.00 O-3.00 x
each eye) in the book that is 40 cm from her 54 4 BU for the right eye?
lenses?
CHAPTER NINETEEN

Stops and
Related Effects

19.1 Field of View any lenses behind it. When there are no lenses behind
the field stop, it serves as the exit port. (Some authors
The field of view of an optical system is the extent of use the words entrance and exit window instead of
the object plane that is imaged by the system. For a entrance and exit port.)
person standing in the middle of a room, the size of Figure 19.1 shows a thin lens followed by an iris
the window limits the person's field of view of some diaphragm. The diaphragm is the aperture stop, and
distant mountains. As the person walks toward the the lens is the field stop. As a result, the lens serves as
window, his or her field of view increases. The limiting both the entrance port and the exit port. The dia-
effect of the window is eliminated once the person phragm is the exit pupil, and the image of the dia-
sticks his or her eyes past the plane of the window. phragm formed by the lens is the entrance pupil.
Thus, both the size of the window and the person's Figure 19.2 again shows a thin lens followed by an
distance from the window influence the field of view. iris diaphragm. This time the diaphragm is the field
A system needs at least two elements to limit the stop, and the lens is the aperture stop. The lens is then
field of view. The window and the person's eye consti- the entrance pupil and the exit pupil. The diaphragm
tute the two elements in the above example. For a is the exit port, and the image of the diaphragm
single lens camera, the size of the film acts as the formed by the lens is the entrance port.
second element. In a multilens system, two of the Figure 19.3 shows a system consisting of two lenses
elements tend to act together to limit the field of view. with an aperture in between. The aperture is the field
One of these elements is the aperture stop of the stop, and the front lens is the aperture stop. The front
system (defined in Chapter 14). The other is called the lens is then the entrance pupil, and the image of the
field stop. More formally, the field stop is the aperture front lens formed by the back lens is the exit pupil.
or lens rim that together with the aperture stop, is The image of the aperture formed by the front lens is
most effective at limiting the field of view of the the entrance port, and the image of the aperture
system. In the case of the person looking out the formed by the back lens is the exit port.
window, the pupil of the person's eye is the aperture Figure 19.4a shows the entrance port and entrance
stop and the window is the field stop. pupil of a two-element system. For rays originating at
The entrance and exit pupils are the object and the axial object point, the entrance pupil specifies the
image space images of the aperture stop (Chapter 14). size of the cone that ultimately passes through the
Similarly, the entrance and exit ports are the object system. Around the axial object point are a set of
and image space images of the field stop. In particular, paraxial off-axis points from which a cone of rays of
the entrance port is the image of the field stop formed the same size passes through the system. In the image
by any lenses in front of it. When there are no lenses plane, the conjugate set of image points have the same
in front of the field stop, it serves as the entrance port. illumination. This set of object points is called the field
The exit port is the image of the field stop formed by of uniform or maximum illumination. The top bound-

401
402 Geometrie, Physical, and Visual Optics

Q.
O o _
CO
Q) SS-
3
(1) 1
- I -
Q.

T
<?
>
4? if
FIGURE 19.1. Iris diaphragm is the aperture stop,
and the lens is the field stop.

entrance ^
port

1
T
/

I
I
image of
diaphragm FIGURE 19.2. The iris diaphragm is the
by the lens field stop, and the lens is the aperture stop.

Q.
O 8 CO O
CO Q.
O
c

to
r 1 1
CD
_L_ >
-
Q. L

#1 T #2|"

FIGURE 19.3. Two lenses with aperture in be-


t tween. Lens # 1 is the aperture stop, and the aper-
image of ture is the field stop.
2 by 1
Stops and Related Effects 403

entrance
port entrance
pupil

FIGURE 19.4. a. Bounding rays for the maximum


or uniform field of view. b. Illumination fall-off in the
image plane, c. Bounding rays for the field of at least
one half illumination, d. Bounding rays for the total
field of view. e. Superposition of the bounding rays.
404 Geometrie, Physical, and Visual Optics

Ft
^F^Fj

FIGURE 19.4. {Cont'd).

ing ray for the field of uniform or maximum illumina- uniform or maximum illumination are shown in Figure
tion passes through the top edges of both the en- 19.5a. With the help of the dashed line parallel to the
trance pupil and entrance port. The bottom boundary optical axis, the field of uniform or maximum illumina-
ray passes through the bottom edges of both the tion F m is given by
entrance pupil and entrance port.
tan(F m /2) = ( a - b ) / d , (19.1)
In general, the set of image points does not end
abruptly, but rather the illumination falls off to zero in where a is the half diameter of the entrance port, b is
a more gradual manner (Figure 19.4b). The bounding the half diameter of the entrance pupil, and d is the
points for the total field of view are the points for distance between the entrance port and the entrance
which only one ray gets through the system. This ray pupil. (The term half diameter is used to avoid confu-
(or its extension) passes through the opposite edges of sion with radius of curvature.)
the entrance port and the entrance pupil (for example, From the triangles in Figure 19.5b, the angular field
the top of the entrance port and the bottom of the of at least one half illumination F h is given by
entrance pupil, as shown in Figure 19.4d).
tan(F h /2) = a/d. (19.2)
The field of at least one half illumination consists of
the set of object points for which the illumination at With the help of the dashed line parallel to the axis,
the image plane is at least 50% of the maximum. The the triangles in Figure 19.5c give the angular total field
bounding ray for the field of at least one half illumina- F t as
tion is the chief ray that passes through the edges of tan(F t /2) = (a + b)/d. (19.3)
the entrance port (Figure 19.4c). (Remember that
In order to find the aperture stop, we usually find
chief rays or their extensions pass through the center
the entrance pupil first. Similarly, to find the field
of the entrance pupil.)
stop, we usually find the entrance port first. A
Figure 19.4e shows the three fields together. In
straightforward approach is to first find all the en-
practice, the total field of view is essentially useless trance pupil candidates and determine the entrance
because no significant amount of light gets through the pupil (as in Chapter 14). Next, calculate the angles
system from the bounding points. On the other hand, that the remaining candidates subtend at the entrance
the field of uniform or maximum illumination may be pupil. From Eq. 19.2, the entrance pupil candidate
too restrictive. Therefore, by convention, the field of that subtends the smallest angle at the entrance pupil
at least one half illumination is used to specify the field is the entrance port. The conjugate object is the field
of view. In other words, field of view without any stop.
specifying adjectives means field of at least one half
illumination. EXAMPLE 19.1
The field of view is specified either as a linear Determine F m , F h , and Ft for the system in Exam-
distance in the object plane (the field is 12.3 m at a ple 14.1.
100 m object distance), or as an angular measure (the In Example 14.1, a +12.00 D lens with a 4 cm
field of view is 7). The bounding rays for the field of diameter is located 8 cm in front of an aperture of
Stops and Related Effects 405

FIGURE 19.5. Geometry to calculate the angular


fields, a. Maximum, b. One half illumination, c. Total.

diameter 2 cm. As determined in the example, the (F m /2) = 6.6,


entrance pupil candidates were the lens and the F m =13.2.
image of the aperture formed by the lens. The
image was virtual, 200 cm from the lens, and had a From Eq. 19.2,
diameter of 50 cm (Figure 14.5a). As shown in
Example 14.1, the lens is the entrance pupil. tan(F h /2) = 25 cm/200 cm = 0.125,
The lens and the virtual image were the only (F h /2) = 7.1,
two candidates for the entrance pupil. Since the F h = 14.2.
lens is the entrance pupil, the virtual image (the
only other entrance pupil candidate) is the entrance From Eq. 19.3,
port. Consequently, the conjugate object (the aper-
ture) is the field stop. tan(F t /2) = (25 cm + 2 cm)/200 cm = 0.135,
From Eq. 19.1, (F t /2) = 7.7,
tan(F m /2) = (25 cm - 2 cm)/200 cm = 0.1153, F=15.4.
406 Geometrie, Physical, and Visual Optics

In the above example, the three fields are clearly EXAMPLE 19.2
different. When the illumination fall off at the image A +20.00 D lens has a diameter of 3 cm. An
plane is evident, we say the system exhibits vignetting. aperture of diameter 1 cm is located 3 cm to the
(In literature, the word vignette means a short work right of the lens. What is the field of view of this
that depicts something subtly and delicately. Another system for an object 100 cm to the left of the lens?
The entrance pupil candidates are the lens itself
definition is a subtle charming scene.) In photography, and the image of the aperture formed by the
vignetting causes the picture to gradually blend into +20.00 D lens. The latter image is virtual, located
the background, giving a nice artistic effect (as in 7.5 cm to the right of the lens, and 2.5 cm in
some wedding pictures). However, optical systems are diameter. The subtended angles at the axial object
usually designed to avoid vignetting by effectively point are given by Figure 19.6a, and
spreading the field of uniform or maximum illumina-
tan w = 1.5 cm/100 cm = 0.015,
tion over the desired area, and then sharply cutting off
the rest. For the above example, this could be done by and
adding another stop at the image plane to cut off the tan q = 1.25 cm/107.5 cm = 0.012.
object space field at 13.2. (The new stop then takes The angle q subtended by the virtual image is
over as the system's field stop). smaller than the angle w subtended by the lens, so

image of
aperture
by the lens

FIGURE 19.6. a. +20.00 D lens 3 cm in


front of an aperture. For an object plane
100 cm from the lens, the image of the aper-
object
plane 7.5cm ture by the lens is the entrance pupil (q < w).
b. Bounding ray for the field of view.
b)
Stops and Related Effects 407

the virtual image is the entrance pupil. It follows 300 cm from the circular opening, serves as the
that the aperture is the aperture stop. By default, object for the - 2 . 0 0 D lens (i.e., the image for
the other entrance pupil candidate (the lens) is the system 1 serves as the object for system 2). Then
entrance port, and since it is not an image, the lens u = -300 cm,
is the field stop.
The field of view without any other adjectives U=-0.33D,
means the field of at least one half illumination. V = P + U = -2.00 D + (-0.33 D) = - 2 . 3 3 D,
From Eq. 19.2 and Figure 19.6b,
v = -42.9 cm.
tan(F h /2) = (1.5 cm/7.5 cm) = 0.200,
F h = 22.6. The system's entrance pupil is virtual and 42.9 cm
behind the circular opening. From Eq. 19.2,
EXAMPLE 19.3a tan(F h /2) = 10 cm/42.9 cm = 0.23,
Consider a one-eyed (monocular) person looking at
(F h /2) = 13.1,
the distant mountains through a 20 cm diameter
circular opening. The entrance pupil for the system and
is the entance pupil of the person's eye. The en- F h = 26.2.
trance port for the system is the circular opening
(which is also the field stop). When the person's The angle of 26.2 is slightly outside the paraxial
entrance pupil is located 300 cm from the circular zone. However, in order to compare the field of
opening, what is the field of view for a static eye? view results to paraxial spectacle magnification
From Eq. 19.2, theory, let us use the small angle approximation for
tan(F h /2) = 10 cm/300 cm = 0.033, the tangent to obtain the value 26.7 for F h . Then,
Examples 19.3a and 19.3b show that the -2.00 D
F h = 3.8. Fresnel lens increases the field of view by a factor
of 7 (i.e., 26.7/3.8).
EXAMPLE 19.3b The spectacle magnification (Eq. 14.3) for a
Suppose a -2.00 D Fresnel lens (20 cm diameter) is -2.00 D lens located 300 cm in front of the person
now mounted in the circular opening. How does is
that change the field of view?
The simple ray diagram in Figure 19.7 shows 1
M spec 1 - d P '
that the field of view is increased by the addition of
the minus lens. In this case, the circular opening
remains the field stop (and entrance port), and the Mspec
l-(3m)(-2.00D)
person's pupil remains the aperture stop. However,
the system's entrance pupil is now the image of the In effect, the Fresnel lens minifies by a factor of 7,
person's pupil formed by the cornea and the and at the same time increases the field of view by
-2.00 D Fresnel lens. The cornea's image, which is a factor of 7.

- ^ e , s ^
entrance pupil
of static eye
Fh/2-
-M-
^ y

FIGURE 19.7. Increase in field of view by minus


lens.
408 Geometrie, Physical, and Visual Optics

19.2 Keplerian Telescope

t
Fh
Figure 19.9 shows a 7X Keplerian telescope consisting
of a +3.00 D objective lens with a 4 cm diameter and a
+21.00 D ocular lens with a 2 cm diameter. The length
is 38.1cm. The candidates for the entrance pupil are
the +3.00 D objective lens and the image A of the
+21.00 D ocular lens formed by the +3.00 D objective
lens. The image A is real and located 266.67 cm in
front of (to the left of) the +3.00 D. The lateral
magnification is - 7 (note the angular magnification of
the telescope is also 7), so the image has a diameter of
14 cm.
FIGURE 19.8. Qualitative relations between field of view and The entrance pupil is the candidate that limits the
magnification. amount of light that comes from the axial object point.
For objects at a finite distance, we simply determine
which candidate subtends the smallest angle. For the
telescope with the object at optical infinity, the axial
The above example shows that minifiers tend to bundles are collimated (rays are parallel). In this case,
increase the field of view. Fresnel lenses as described the smaller candidate determines the size of the bun-
are sometimes used on the backs of vans or motor- dle that passes through the system. (Alternatively, we
homes to increase the field of view. Similarly, convex might say that in the limit of the object going to
(diverging) mirrors are frequently used to increase the optical infinity, the smaller candidate subtends the
field of view. smaller solid angle.) For the 7X telescope, the
Just as minifiers tend to increase the field of view, +3.00 D objective lens is smaller than the image A,
magnifiers tend to decrease the field. Figure 19.8 and so is the entrance pupil. Since the lens is not an
shows a hypothetical relationship between the field of image, it is also the aperture stop. The image A is
view F h and the magnification. The solid line repre- then the entrance port (it is the only candidate), and
sents the most efficient systems. For these systems, as the +21.00 D ocular lens is the field stop.
the magnification is increased, the field of view de-
From Eq. 19.2, the field of view is
creases (points A to B). However, optical systems can
always be designed with smaller fields (points A to C). tan(F h /2) = 7 cm/266.67 cm = 0.026,
Therefore, if one starts with an inefficient system, one (F h /2) = 1.5,
can increase both the field and the magnification until
the boundary for maximally efficient systems is F h = 3.0.
reached (point C to B). For the telescope, the exit port is the +21.00 D

FIGURE 19.9. Keplerian telescope. Objec-


tive lens is the aperture stop and entrance
pupil.
Stops and Related Effects 409

ocular lens, and the exit pupil is the image B of the image C is larger than the +3.00 D objective lens, so
objective lens formed by the ocular lens. This image is the +3.00 D lens remains the entrance pupil.
real and 5.44 cm behind (to the right of) the +21.00 D The candidates for the entrance port are the image
lens. The lateral magnification turns out to be 1/7 C of the ocular lens, and the image D of the field lens
(note the angular magnification of the telescope is 7), at optical infinity. The nodal ray determines the angle
and so the exit pupil has a diameter of 0.57 cm. w that the image D subtends at the objective lens.
The real exit pupil on a Keplerian telescope is easy Since the field lens is the object, the nodal ray is given
to find and measure. One simply illuminates the objec- by
tive lens so that the light travels backwards through
the telescope. Since the exit pupil is real, a sharply tan w = 1.5 cm/33.33 cm = 0.045.
specified circle of light, called the Rams den circle, The image C subtends an angle at the entrance pupil
shows up on a screen at the exit pupil position. of
When a person looks through the telescope with
his/her eye directly behind the ocular lens (say a tan q = 7 cm/238.9 cm = 0.029.
10 mm vertex distance), the pupil of his/her own eye The angle q is smaller than the angle w, so the image
acts like a stop causing a smaller field than the 3.0. As C is the entrance port.
the person backs away from the telescope, the ob- The effect of the +2.50 D field lens is to change the
served field of view increases. The largest observed entrance port from 266.67 cm in to 238.9 cm without
field (3.0) occurs when the entrance pupil of the changing its size. The angle q calculated directly above
person's eye is in the vicinity of the exit pupil of the equals F h /2, so
telescope. If the person continues to back away, the
observed field of view decreases. F h /2 = 1.7,
Thus, the location of the exit pupil is an important and
consideration in the design of telescopes and other
optical instruments. The distance from the ocular lens F h = 3.4,
to the exit pupil is called the eye relief (Figure 19.9). which is a 13.3% increase in the field of view of the
The 5.44 cm eye relief of the above 7X Keplerian telescope.
telescope is probably too long unless it is a rifle scope. The exit pupil is now the image of the objective
The Keplerian telescope forms an internal image. lens formed by the +2.50 D field lens and the
We can increase the field of view of the telescope (and +21.00 D ocular lens. This image is real and located
shorten the eye relief) by placing a plus field lens at 4.9 cm behind (to the right of) the ocular lens. The
the internal image position. The field lens contributes total lateral magnification remains - 1 / 7 , so the exit
a lateral magnification of +1 (object distance zero, pupil size remains 0.57 cm. Thus, the field lens shor-
image distance zero), and so does not change the tened the eye relief to 4.9 cm, while keeping the exit
magnification of the telescope. pupil size the same.
One way to explain the effect of the field lens is by Higher powered field lenses continue to increase
looking at its prismatic effect on the rays. Without the field of view until the field lens itself becomes the field
field lens, the rays in Figure 19.10a would continue stop. To see when the field lens becomes the field
straight (marked 1 and 2), miss the ocular lens, and stop, consider the nodal ray from the field lens (Figure
thus the object point would be outside the field of 19.10c). As the field lens power is increased, the
view of the telescope. The field lens bends these rays image C of the ocular lens moves closer to the objec-
so that they now pass through the ocular lens (1' and tive without a change in size. Ultimately (at position h
2'), and the object point is then inside the field of view in Figure 19.10c), the image C subtends the same
of the telescope. angle at the entrance pupil as the nodal ray for the
An alternate way to explain the field increase is to field lens. If the field lens power is increased further,
look at what the field lens does to the entrance port the image C moves closer, and subtends a larger angle
position. Assume the field lens is a +2.50 D lens that than the nodal ray for the field lens. In the latter case,
is 3 cm in diameter. The location is 33.33 cm behind the image D of the field lens becomes the entrance
the +3.00 D lens. In object space, the field lens is port and the field lens becomes the field stop. The
imaged at optical infinity, so it cannot be the entrance lowest powered field lens that leaves the image C at
pupil. The image C of the ocular lens formed by the position h gives the largest increase.
field lens and the objective lens is real, and located The image distance for position h is found from the
238.9 cm in front (to the left) of the +3.00 D objective triangle in Figure 19.10c
lens (Figure 19.10b). The total lateral magnification is
again - 7 , so the image C has a 14 cm diameter. The tan w = 7cm/y,
410 Geometrie, Physical, and Visual Optics

without field lens

image of
ocular lens
by the two lenses
(objective & field)
in front of it

b)

FIGURE 19.10. Keplerian telescope with a


_ Ji \ re- field lens. a. Without field lens, rays 1 and 2
miss the ocular lens. b. Image C is the entrance
position of image C
for different field lens powers port (q < w). c. Position changes for image C as
a function of field lens power. When the image

nr
J C is to the right of position h, it is no longer the
entrance port. In the latter case, the field lens
serves as the field stop.
v
Stops and Related Effects 411

or For the general argument, let us consider an aper-


v = 7 cm/tan w. ture located a distance tu behind the ocular lens of the
telescope. The image of this aperture formed by the
As calculated above telescope is located at distance v from the objective
tan w = 0.045, lens of the telescope. From Eq. 12.32, the object-
image matrix has to have the form:
so
v = 7 cm/0.045 = 155.5 cm. m -P.
Sio ~~ (19.6)
We can now work backwards to determine the power
of the field lens that gives an image distance of m 0
155.5 cm. The answer is +10.00 D. In this case,
where Pe is the equivalent power of the afocal tele-
F h /2 = w = 2.6, scope and m is the total lateral magnification.
With the help of S b , we can construct the object-
so image matrix as:
F h = 5.2.
Sio = TSbT,
The 5.2 is a 73% increase in field over the 3.0
where Ttt is the translation matrix for the distance tu
without the field lens. The eye relief with the
and is translation matrix for the distance v. The
+10.00 D field lens is 3.2 cm.
translation matrix has its usual form, and the equation
for the object-image matrix becomes

\ 1
19.3 Magnification Equality for 1
Afocal Systems $io

M /
In the above argument, the size of the image C of the
ocular lens did not change as the field lens power was The three matrix multiplications result in
increased. In fact, the lateral magnification m in the
above argument was always numerically equal to the

angular magnification M for the telescope. Since later-
al magnification and angular magnification are differ-
I -p \ e
Sio M
ent entities, this relation is not obvious. In a way, the
numerical equality is an accident that occurs for afocal \ ^ _ P e t
" + S 2 1 + m t
" M
" y P
e /
systems. We can use the system matrix approach of (19.7)
Chapter 12 to prove the numerical relationship.
Equation 13.23 gives the system matrix for an By comparing the second row, second column ele-
afocal telescope: ments from Eqs. 19.6 and 19.7, we obtain
'M m = M-i;Pe.

S,= (19.4) Since Pe is zero for the afocal telescope,

M m = M,
which completes the proof.
where M is the magnification of the telescope. The
The numerical equality holds even when the aper-
equivalent power Pe is zero for an afocal telescope, but
ture is placed against the ocular lens (tM = zero), in
for instructional purposes, we will carry Pe along for a
which case we can consider the ocular lens (with its
while.
size specified by the aperture) as the field stop.
When light travels backwards through the tele-
A similar proof holds for the size of the exit pupil.
scope, the magnification becomes 1/M, and the back-
There the objective lens is the object, and the light is
wards system matrix is
going through the telescope in the forward direction.
We would then use the system matrix St instead of S b .
M Pe In this case the result would be that the lateral magni-
s = l (19.5) fication m is equal to the reciprocal of the angular
M magnification M (i.e., m = l/M). For example, the
412 Geometrie, Physical, and Visual Optics

exit pupil size of the above 7 X Keplerian telescopes is Galilean telescope, the person's field of view steadily
one seventh the size of the objective lens. decreases.
For a two-lens telescope, we can glean another One way to explain the difference is in terms of the
item of information from the above object-image ma- exit pupil location. The exit pupil for the Keplerian
trices. By setting Pe and tM equal to zero and compar- telescope is a real image located behind the telescope.
ing the second row, first column elements in Eqs. 19.6 The person's field of view is maximum when the
and 19.9, we obtain person's entrance pupil is at the exit pupil of the
telescope (static eye). For the Galilean telescope, the
= -Ms 2 1 .
image of the objective lens by the ocular is a virtual
For a two lens afocal telescope, s21 is just the distance image located inside the telescope. The closest the
d between the lenses, or observer can get to the virtual image is to ram his/her
eye against the ocular lens. The person's field of view
v = -Md (two-lens afocal telescope).
is a maximum at that point, and as the person backs
In Section 19.2, the two-lens 7X Kepler telescope has away from the telescope, the person's field decreases.
a length of 38.1 cm, and the entrance port distance did We can quantitatively compare the two situations
equal 7 times 38.1 or 266.67 cm. by using an aperture to simulate the person's entrance
pupil. When we place this aperture at the Keplerian
EXAMPLE 19.4 telescope's exit pupil position, our previous results are
A 9X Keplerian telescope has an objective lens unchanged provided the aperture is the same size as
with a 36 mm diameter and an ocular lens with a the telescope's exit pupil. For the Galilean telescope,
5 mm diameter. When the entrance pupil is the the aperture simulating the person's entrance pupil
objective lens, and the ocular lens is the exit port, becomes the aperture stop for the system. Thus, the
what is the size of the entrance port and the exit entrance pupil is the image of the aperture by the
pupil? telescope, and the objective lens of the telescope
The ocular lens is then the field stop, and the becomes the field stop (and also the entrance port).
entrance port is 9 times its size, or 45 mm. (Actual-
ly, M = - 9 , which means the entrance port is Consider a 7X Galilean telescope consisting of a
inverted.) +3.00 D objective lens with a 4 cm diameter, and a
The objective is the aperture stop, and the exit -21.00 D ocular lens with a 2 cm diameter. The length
pupil is 1/9 its size, or 4 mm. is 28.59 cm. Consider an aperture of diameter 5.7 mm
representing a person's entrance pupil. The latter
EXAMPLE 19.5
aperture is placed 2 cm from the ocular lens of the
Common binoculars are Keplerian telescopes with telescope. The entrance pupil candidates are the ob-
a Porro prism erecting system. What is the magnifi- jective lens, the image of the ocular lens by the
cation of a pair with a 48 mm diameter objective objective, and the image of the aperture by both the
and a Ramsden circle that measures 6 mm? ocular and the objective.
The objective lens is the entrance pupil and the The image A of the ocular by the objective is
aperture stop, and the Ramsden circle is the exit virtual, and located 200 cm behind the objective (to
pupil. Here m equals 1/M, or the right in Figure 19.11). The lateral magnification is
+7 (as expected from the previous section), and so the
M -m
i, virtual image size is 14 cm. The image B of the
From the lateral magnification definition, aperture by both the ocular lens and the objective lens
is virtual, and located 298 cm from the objective lens.
_ object size _ objective lens size The lateral magnification is again +7, and B has size
image size exit pupil size ' of 39.9 mm. The image B is the smallest of the en-
or trance pupil candidates. In the limit of an object at
optical infinity, the image B is the entrance pupil and
the aperture representing the eye's pupil is the aper-
ture stop.
At the center of the entrance pupil, the tangent of
the angle subtended by the objective lens equals 2 cm
19.4 The Galilean Telescope divided by 298 cm, and the angle itself is 0.38. At the
center of the entrance pupil, the tangent of the angle
As a person backs away from a Keplerian telescope, subtended by the image B equals 7 cm divided by
the person's field of view increases to a maximum and 98 cm, and the angle itself is 4.. So the entrance port
then starts to decrease. As a person backs away from a is the objective lens. (Since the objective lens is not an
Stops and Related Effects 413

8 I IB
O iS ^ _1_
t I
Y "
1

1 T

r-
l FIGURE 19.11. Galilean telescope followed by an
1
aperture that is simulating the entrance pupil of the
observer's eye.

image, it is the field stop). From Eq. 19.2, the field of the top and bottom points of the slide.) The dashed
at least one half illumination F h is 0.77. rays that leave the top of the source are shown along
The 0.77 field for the 7 X Galilean telescope com- with the solid rays that leave the bottom. The source
pares to the 3.0 field for the 7 X Keplerian telescope. rays that pass through the slide outside the points e
The Galilean telescope has no internal image, so a miss the lens. This means that on the projector screen
field lens cannot be added. The 7X Keplerian tele- the points outside e' have zero illumination, or they
scope in Section 19.2 has an internal image, and a are outside the total field of view. In effect, the source
+10.00 D field lens increased the Keplerian tele- functions as a stop in the system.
scope's field to 5.2. Clearly, the Keplerian telescope The vertical dashed line from k to k' is the image of
has a field of view advantage over the Galilean tele- the source by the objective lens. The image kk' is the
scope. This is why common binoculars are Keplerian exit pupil of the system. So the source is the entrance
telescopes with a prism erecting system as opposed to pupil (and hence the aperture stop). The objective
a simpler Galilean telescope. lens is the entrance port (and hence the field stop).
Despite the field restrictions, the Galilean telescope One way to increase the field of view (so that the
does have some advantages over the Keplerian types entire slide is imaged) is to increase the size of the
for low vision patients. The low vision telescope is source so that it subtends a larger angle at the slide.
worn mounted in a spectacle frame. As with any However, this frequently is not practical. The pre-
spectacle frame, bumps and wear can cause the frames ferred method is to use a condenser lens, which pro-
to change position relative to the eye. In a Keplerian vides an apparent source (an image of the source),
telescope with its exit pupil at the eye, such a bump which subtends a larger angle at the slide. The con-
may cause the exit pupil to no longer be coincident denser is a large diameter short focal length lens (or
with the eye's entrance pupil, in which case the person lens system).
cannot see through the telescope at all. On the other Figure 19.12b shows a condenser added to the
hand, the Galilean telescope is not as sensitive to system of Figure 19.12a. The dashed rays leave the top
frame position changes, and consequently, is function- of the source, pass through the condenser, the slide,
ally easier to use. The trade-off then is a larger field the objective lens, the point k' on the image of the
(Keplerian) vs ease of use (Galilean). source, and then on to the projector screen. The solid
rays leave the bottom of the source, pass through the
condenser, the slide, the objective lens, the point k on
the image of the source, and then on to the projector
19.5 Condensers screen. All points on the slide are now included in the
total field of view. The image kk' of the source is still
Transparent slides are frequently the objects of regard the exit pupil of the system, but now it is much larger
in optical systems (e.g., microscopes, projectors, etc.). and very close to the objective lens.
The slides need to be illuminated by a primary source. The fact that there are two images in the system,
Figure 19.12a represents a source, slide, and an objec- the image of the source and the image of the slide, is
tive lens that projects a large real image on a distant sometimes confusing. The dashed rays from the top of
screen. The image size shown is correct for the size of the source are easy to follow through the two lenses to
the slide. (If you wish, you may draw in the rays from the conjugate image point k'. Similarly, the solid rays
414 Geometrie, Physical, and Visual Optics

source

FIGURE 19.12. a. Projector system with the source


functioning as a stop. b. Use of a condenser lens to
increase the field of view.

from the bottom of the source are easy to follow to the field stop are determined by the procedure listed
conjugate image point k. Now consider the top point above, and then the total field is calculated from Eq.
on the slide. Two rays diverge from there: one dashed 19.3. However, because of vignetting, the predicted
and one solid. These rays pass through the objective bounding rays for the total field of view may not pass
lens, are converged, and rejoin each other at the through all the other apertures. An easy way to check
bottom image point on the projector screen (i.e., the for vignetting is to trace the rays through the system.
image is inverted). In particular, note how the same For multilens systems there are very efficient com-
set of rays (solid and dashed) gives the two different putational methods to determine the aperture stop and
images. field stop. One of these methods utilizes the system
matrix approach. Another, the so-called point-slope
method, is a modification of the system matrix ap-
proach. (Those interested in pursuing this are referred
19.6 Multilens Systems and Some Typical to other optics or lens design books.)
Aperture Stops In order to determine the field stop, one first has to
know the aperture stop. Consider some instruments in
In multilens systems, one has to watch out for vignet- which the observer's eye is the last element in the
ting. For example, assume that the aperture stop and system. In the previous sections, we saw that the
Stops and Related Effects 415

II \ 19.7 Film Exposure and the f-Number


^ Z"_"error
--
The exposure is the amount of luminous flux (light
energy) delivered to a film. On a simple level, the
exposure E equals the illuminance I, on the film times
the time t that the film is exposed, or
object ^y I mis-focus
plane ^* E = I,*. (19.8)
T Equation 19.8 is known as the reciprocity law, since
a)
it says we can obtain the same exposure with a high
illuminance and a short exposure time vs a low illumi-
nance and a long exposure time. (This law works well
over the linear region of the film response but breaks
down at the extremes.)
Obviously, the illuminance on the film plane de-
pends on the amount of light let into the system,
which in turn depends on the area of the entrance
pupil. The area of the entrance pupil is proportional to
the square of the diameter h of the entrance pupil, so
I,h 2 . (19.9)
For a fixed entrance pupil, the illuminance on the
FIGURE 19.13. a. Size error due to misfocus. b. Telecentric film plane varies inversely with the image area. In
system to eliminate the error. other words, for a small image, the illuminance is
higher (the energy is more concentrated); while for a
large image the illuminance is smaller (i.e., the energy
is spread out more). The image area is directly propor-
aperture stop for the Keplerian telescope eye system tional to the square of the image size I. (Think of a
was typically the objective lens, while the aperture piece of square film that has a size I x I.) The size I
stop for the Galilean telescope eye system was typical- equals the image distance v times the tangent of the
ly the pupil of the observer's eye. For a compound nodal angle. Therefore, the image area is directly
microscope, the aperture stop is typically the objective proportional to the square of , and
lens, and the same is true for an ophthalmic slit lamp.
For a simple magnifier, the pupil of the observer's eye T 1 (19.10)
is the aperture stop, and the same is true for a V

binocular indirect ophthalmoscope. However, for a From Eqs. 19.9 and 19.10,
direct ophthalmoscope, the apparent source serves as
the entrance pupil, so the light source serves as the (19.11)
aperture stop. For a lensometer, the nosepiece aper-
ture is the aperture stop. For distant objects, the image distance is effectively
A telecentric stop is an aperture stop that is located equal to the reduced secondary equivalent focal length
at the focal point of an optical system. Telecentric f, so
stops are widely used in measurement systems. The
telecentric stop minimizes the measurement error T h2
made by a slight defocus of the system. (A slight
defocus can occur because of the depth of focus, which The ratio of the reduced secondary equivalent focal
is discussed later in this chapter.) In Figure 19.13a, the length to the entrance pupil diameter is called the
stop is at the lens. The chief ray is then the nodal ray f-number:
for the lens, and the place where it intersects the scale
is the reading, which has some error in it. In Figure f-number : f (19.12)
19.13b, the aperture stop is at the secondary focal h'
point of the lens. The incident chief ray is now parallel Then the illuminance on the film plane is inversely
to the axis and, hence, gives no error if slightly proportional to the f-number squared:
defocused. Aperture stops can also be placed at the
primary focal point (telecentric in object space). I, a 1 /(f-number)2.
416 Geometrie, Physical, and Visual Optics

The f-number is usually preceded by the symbols, EXAMPLE 19.6


il. For example, an f-number of 11 is written as f/11, A 35 mm camera (the 35 mm is the film width), has
an f-number of 5.6 is written as f/5.6, and an f-number a 55 mm focal length lens (P = +18.18 D). What is
of 1.8 is written as f/1.8. The larger the aperture, the the entrance pupil size for f-numbers of f/1.8, f/11,
and f/22?
smaller the f-number. Fast camera systems (i.e., short From Eq. 19.12, the entrance pupil diameter for
exposure times) have small f-numbers (e.g., f/1.8, the f/1.8 is
f/2.8). The smaller the aperture, the larger the f- f 55 mm
number. Slow camera systems (longer exposure times) h= r = 1 0 = 30.6 mm.
have larger f-numbers (e.g., f/11, f/16). Instamatics f-number 1.8
and older single lens cameras are f/11 systems. The Similarly, f/11 gives 5 mm and f/22 gives 2.5 mm.
better the camera lens system, the wider the aperture
can be, and so the smaller minimum the f-number. EXAMPLE 19.7
Camera lens systems rated f/1.8 are now common. A +8.00 D trial lens has a diameter of 36 mm.
What is the f-number of the lens?
For a fixed exposure E, the reciprocity law gives The focal length of the lens is 125 mm. From
the exposure time t: Eq. 19.12,
125 mm
t= f-number = =3.5.
V 36 mm
Since the illuminance is inversely proportional to the So f/3.5 is the rating of the lens.
f-number squared, we then have
EXAMPLE 19.8
E*t (f-number)2. A +60.00 D human eye has a typical 3 mm en-
trance pupil. What is the f-number of the eye?
Thus, the exposure time t is proportional to the f- The (reduced) focal length is 16.67 mm. Then
number squared. Suppose for f/2, the exposure time is
1 unit. Then 16.67 mm
f-number : 5.6.
3 mm
E = l / ( 2 ) 2 = 0.25.
So f/5.6 is a typical rating for the eye. (Actually,
the human eye functions from about f/2 when dark
The f-number setting that doubles t for the same E is adapted to f/8 in high illuminations.)
2 = 0.25 (f-number)2,
The f-number system enables one to change lens
or systems on a camera, and yet still use the same type of
f-number = 2/0.25 = f/2.8. film. No matter what the lens system, if it is set for
f/8, then the exposure time is the same. Thus, for the
Similarly, the f-number that doubles the exposure time same object, a +10.00 D lens with a +12.5 mm ent-
again (i.e., t = 4) is f/4. A table of t vs the f-number is rance pupil diameter has the same exposure time as a
+20.00 D lens with a 6.25 mm opening. For illumina-
tion purposes, we simply describe both lenses as f/8.
t f-Number
The f-number system is designed for distant ob-
1 2 jects, and works well as long as the image distance is
2 2.8 approximately equal to the secondary focal length. In
4 4 close-up, or macrophotography, the image size is
8 5.6
16 8 about the same as the object size ( m = - l ) . This
32 11.3 occurs when the image distance is twice the secondary
64 16 focal length, or v = 2f. The f-number then needs to be
128 22.6 modified, and is designated f-effective, where
f-effective = . (19.13)
h
Then from Eq. 19.11,
The above settings on a camera are called the f stop 1
settings. On a modern camera, there may be some half
or third stop settings. A typical real sequence is f/1.8, (f-effective)"
f/2.8, f/4, f/5.6, f/8, f/11, f/16, and f/22. Diffraction For v = 2f,
effects due to the small aperture start to become 2f
noticeable at f/22. f-effective = = 2 (f-number).
Stops and Related Effects 417

Similar equations can be worked out for any image or Then


object distance. For example, for v = 1.5 f, f-effective b -
is 1.5 times the f-number.

or
EXAMPLE 19.9
The lighting conditions and the film in a camera
require an f/8 setting. A bellows is used to extend hi; i;d i; '
the lens away from the film plane so that the image and
size is equal to the object size. What setting should
now be made on the camera lens? :Vd-V,
The image size equals the object size when hi;
v = 2f. So the setting is where Vd and V are the corresponding dioptric values.
f-effective = 2 (f-number) = (2)(8) = 16. Now
So f/16 is the setting. (Actually, macrophotography Vd = P + U d ,
may have more complications than this, but for
these you are referred to the photographic lit- and
erature.) V = P + U,
where U = 1/w and U d = l/w d . By subtraction,

19.8 Depth of Field


v d -v=u d -u.
Thus, the distal part of the depth of field expressed
dioptrically is:
Figure 19.14 shows a fixed image screen at a distance i;
b
behind a thin lens of power P. The conjugate object u
point distance is u. Normally, there is a range of <-u=hlr
object distances (wd for distal to up for proximal) for Similarly, we can show that the proximal part of the
which the image on the screen does not have any the depth of field is
detectable blur. (For ud, the rays are dashed, and the
image distance is vd. For wp, the rays are solid, and p
hi;
the image distance is vp.) The object range, ud to wp, is
Therefore, as expressed dioptrically, the distal part of
called the depth of field.
the depth of field equals the proximal part. Let d stand
The diameter of the lens is h, and the tolerable blur
circle diameter is b. (In other words, blur circles of for each part, or
diameter up to b do not result in detectable blur.) The
depth of field clearly depends on b. From the similar (19.14)
hi;
triangles in Figure 19.14, We can then express the dioptric depth of field as plus
v - v, or minus d. (The total dioptric depth is sometimes
expressed as 2d.)

->-f-

FIGURE 19.14. Depth of field for tolerable


ud blur circle size b.
418 Geometrie, Physical, and Visual Optics

Equation 19.14 shows that for a tolerable blur to


circle size b and a fixed image distance v, we can
-2.00 D - 0 . 3 8 D = - 2 . 3 8 D.
increase the depth of field by decreasing the aperture
size h. This is consistent with the pinhole effect. The linear range is from
100
= -61.7cm,
-1.62 D
EXAMPLE 19.10a
A +7.00 D lens is set at f/3.5, and focused for an to
object 50 cm from the lens. The tolerable blur 100
= - 4 2 . 0 cm.
circle diameter is 1 mm. What is the dioptric and -2.38 D
linear depth of field? Again, the linear depth extends out further on the
The focal length for the +7.00 D lens is 143 mm. distal side (11.7 cm) than it does on the proximal
From Eq. 19.11, the aperture size h for the f/3.5 side (8 cm).
lens is
h = 143 mm/3.5 = 40.9 mm. EXAMPLE 19.11
For the object at 50 cm, the image distance is Consider a +60.00 D lens-screen model for the
20 cm. Then from Eq. 19.14, human eye. The typical diameter of a cone in the
fovea of the human eye is 3 (microns). Assume
1.0 mm that the eye cannot detect blur until the blur circles
d= 0.12D.
(40.9mm)(0.20m) are at least the size of the cone. What is the
dioptric depth of field of the eye when the pupil
The dioptric depth of field is 0.12 D. The dioptric diameter is 3 mm?
range is From Eq. 19.14,
-2.00 D + 0.12 D = - 1 . 8 8 D, 3xl0"6m
d =
to (3xl0~ m)(16.67xl0~3m)'
3

-2.00 D - 0.12 D = - 2 . 1 2 D. </ = 0.06D.


The linear depth of field ranges from So the depth of field is 0.06 D.
100 The typical clinical depth of field of an eye is
= -53.1 cm, 0.25 D, which corresponds to a blur circle of
-1.88D
diameter 12.5 /xm.
to
100
= -47.1 cm.
-2.12D
Note that relative to the 50 cm object point, the 19.9 Object Distance Effect on Depth
linear depth extends out farther on the distal side of Field
(3.1 cm) than it does on the proximal side (2.9 cm).
For distant objects, the image distance v equals the
secondary focal length f of the lens. Then from Eq.
EXAMPLE 19.10b
19.14,
The +7.00 D lens is stopped down to f/11. What is
the depth of field? A - b
From Eqs. 19.12 and 19.14,
bf _ b(f-number) or in terms of the f-number,
d= (19.15)
hvi vf dd = b (f-number)P 2 . (19.17)
Thus, the dioptric depth of field is directly propor-
tional to the f-number, and we can set up the ratio: The above two equations show that the short focal
length (high P) systems give larger depths of fields.
d2 (f-number)2 For example, cameras with shorter focal lengths have
(19.16)
d, (f-number) j ' larger depths for the same f/number.
Then Newborn infants typically have about +2.75 D of
d2 _ 11 hyperopia, and an accommodation system that is not
0.12D 3.5' yet well developed. A typical equivalent dioptric
power for newborn human infants is +89.6 D. Thus,
d2 = 0.38 D. compared to the +60.00 D of an adult eye, the infant
The dioptric range is from eye has a larger depth of field, which helps the infant
-2.00 D + 0.38 D = -1.62 D, while the accommodation system is developing.
Stops and Related Effects 419

EXAMPLE 19.12 For distant real objects, we clearly maximize the


A 35 mm camera has a 55 mm focal length lens. depth of field by focusing the system on the hyperfocal
When the lens is set at f/5.6, the depth of field distance.
ranges from 3 to 10 m (fairly distant). When the For near objects, we note that
lens is replaced with a 110 mm lens, set at f/8, what
is the dioptric depth of field?
The dioptric depth of field for the initial lens is i = V = P + U,
0.10 D to 0.33 D. So the camera is focused halfway v
in between (0.22 D) and the depth of field d is plus where U is the object vergence. Therefore, the diop-
or minus 0.12 D. tric depth of field is
We can use Eq. 19.17 and take ratios,
d = ( P + U),
d2 b(8)(+9.09D) 2
0.12 D " b(5.6)( + 18.18D) 2 '
or
or

d = dd(l+), (19.18)
Then where dd is the distance depth, f is the secondary focal
</2 = (0.12D)(1.43)(0.25), length, and u is the object distance.
and When the object distance is four times the focal
length (u = -4f), Eq. 19.18 gives
d2 = 0.04D.
The focal length change from 55 mm to 110 mm d=^dd.
supplied a decrease by a factor of 0.25, while the
stop change supplied an increase by a factor of For an object at the symmetry point, m = - l and
1.43. u = -2f. The dioptric depth of field is then
The dioptric range is 0.22 D 0.04 D, which
gives 0.26 D to 0.18 D. The linear range is then d = da(\ + 4^)=\da. (19.19)
3.8 m to 5.6 m.
Technically, when the system is focused for the hy-
The hyperfocal distance is the distance at which a perfocal distance
camera is focused so that the distal limit of the depth
d=di\+^-\ (19.20)
of field is at optical infinity. Actually, the image \ uh/
distance v differs slightly from f when the system is
However, for most systems, the focal length is much
focused for the hyperfocal distance. However, if we
smaller than the hyperfocal distance, and the approxi-
neglect this small difference, then the hyperfocal dis-
mation used above is
tance equals the reciprocal of dd. When the system is
focused for the hyperfocal distance, the dioptric range d~dd.
is dd dd, or zero to 2dd. For a system focused on the
hyperfocal distance, the linear range is from infinity to EXAMPLE 19.14
one half the hyperfocal distance. An f/5.6 50 mm camera lens has a hyperfocal
distance of 5 m. What is the maximum linear depth
of field for distance objects, and what is the linear
EXAMPLE 19.13 depth of field for an object at the symmetry point
An eye has a depth of field of 0.25 D. What is the (macrophotography), assuming the same tolerable
depth of field of the eye when focused on optical blur?
infinity? What is the depth of field for the eye when The distant depth of field is to a good approxi-
focused on the hyperfocal distance? mation 0.20 D. The maximum depth for distant
The hyperfocal distance is 1/0.25 D or 4 m. For objects occurs when the camera is focused for the
the eye focused on optical inifinity, the dioptric 5 m hyperfocal distance, and is approximately in-
range is 00.25D, or +0.25 D to -0.25 D. For finity to 2.5 m (i.e., dioptric range 0.2D0.2D).
real objects, the depth is from 0 to -0.25 D, and The percent difference relative to the exact
the linear depth is from optical infinity to 4 m. method, Eq. 19.20, is 1%. For the object at the
For the eye focused on the 4 m hyperfocal dis- symmetry point, Eq. 19.19 gives
tance, the dioptric range is -0.25 D 0.25 D, or
zero to -0.5 D. The linear range is from infinity to j - 2 D
^ ~
2m. d=-r-=0.1D.
420 Geometrie, Physical, and Visual Optics

ft 15 10 7 z> "window" showing distance Figure 19.15, the camera is set at f/11 and focused for
for which camera is focused 2.8 m. From the scale, the depth ranges from about 5
m 5 3 2

16 8 4 ^ u - depth of field scale


to 1.8 m. For f/4, the depth would range from about
4 8 16
3.5 to 2.4 m.
16 11 8 5.6 4 2.8 1.8 rE-f/stop

FIGURE 19.15. Depth of field scale as a function of f/stop.


19.10 Depth of Focus
The symmetry point object distance is 10 cm, and
the vergence incident on the lens from that position Figure 19.16 shows a fixed object in front of a thin lens
is -10.00 D. The dioptric range for the depth of of power P. Given a tolerable blur circle size b , there
field is then - 9 . 9 D to -10.1 D, and the linear is a range of image distances (v to vd) over which a
range is screen can be moved before any blur is detected. This
100 range is the depth of focus.
= 10.1cm, We can again set up similar triangles and work out
-9.9 D
100 equations for the depth of focus. The single thin lens
= 9.9 cm. results show that for distant objects and small depths
-10.1 D
(the usual case), the dioptric depth of focus equals the
Note that while the distant linear range was infinity dioptric depth of field to a good approximation, e.g.,
to 2.5 m, the near range was only 9.9 cm to
depth of field = depth of focus = 0.25 D.
10.1cm. This places severe focusing requirements
on macrophotography.
EXAMPLE 19.15
For distant objects, what is the linear depth of
When humans look from far to near, the eyes focus for the lens in the preceding example?
converge, accommodate, and the pupil constricts. This From the preceding example, the distant diop-
is the so-called near triad. The constriction of the tric depth of field was 0.2 D. So for distant ob-
pupil counteracts some of the severe focusing require- jects, the dioptric depth of focus is also 0.2 D.
The lens has a power of +20.00 D, so the diop-
ments for near objects. Another counteracting effect is
tric range for the depth of focus is + 19.8D to
the size of the tolerable blur for near objects. Since +20.2 D. The linear range is from
the lateral magnification is larger for the near object,
the image is larger, and the tolerable blur may also be 1000
= 50.5 mm,
larger. 19.8 D
Many good cameras have a depth of field scale to
marked on the rings around the lens. To set the depth 1000
(assuming one is not limited by the film speed), one = 49.5 mm.
20.2 D
could proceed as follows: first focus the camera (the This means that the camera film has to be flat
best focus is obtained at a wide aperture, for example, within a tolerance of plus or minus 0.5 mm, and is
f/1.8), then set the f-number to give the desired the reason cameras have pressure plates to flatten
depth, and finally set the correct exposure time. In the film.

FIGURE 19.16. Depth of focus for tolerable blur circle


vd size b.
Stops and Related Effects 421

In a general system, the depth of focus and the the percent increase in the field of view? Also
depth of field depend on the exit pupil and entrance specify the new exit pupil location.
pupil sizes. However, the thin lens theory still shows 5. An 8.5 X Galilean telescope has a +3.00 D objec-
the salient features. tive lens with a 4 cm diameter and a 25.50 D
ocular lens with a 2 cm diameter. The observer's
entrance pupil is simulated by an aperture of
diameter 4 mm located 12 mm behind the ocular
Problems lens. Find the entrance pupil, entrance port, and
angular field of view. Why cannot a field lens be
1. A window of width 100 cm is located 400 cm in added to a Galilean telescope?
front of the pinhole of a pinhole camera. If the 6. A pair of binoculars has a +4.00 D f/5.6 objective
film size is not a problem, what is the angular field lens and a 6 mm diameter Ramsden circle. What is
of view of the pinhole camera? the magnification?
2. A +4.00 D lens with a 4 cm diameter is 5 cm in 7. An f/11 lens has a 4.6 mm diameter. What is the
front of a small aperture. For a real object 100 cm diameter of an f/2.8 lens of the same power?
from the lens, the aperture is the aperture stop. 8. If the correct exposure time is 1 /120 second for an
Locate the entrance pupil, the entrance port, and f/5.6 setting, what is the exposure time for an f/16
the linear field of view. setting?
3. A +5.00 D lens is located 6 cm in front of a 9. In dim light, a human entrance pupil opens up to
-2.00 D lens. Each lens has a 3.0 cm diameter. an 8 mm diameter. For a +60.00 D eye, what is
Find the entrance pupil, entrance port, and linear the f-number?
field of view for a real object 50 cm in front of the 10. An optical system focused for 40 cm has a depth
first lens. Then switch the two lenses around and of field that ranges from 33.33 cm to 50.00 cm.
again find the linear field of view. What is the dioptric depth of field? If the lens in
4. An 8.5 X Keplerian telescope has a +3.00 D ob- the system is +12.00 D f/8, what is the tolerable
jective lens with a 4 cm diameter and a +25.50 D blur circle size?
ocular lens with a 2 cm diameter. Specify the 11. A +16.00 D f/16 lens has a 1.00D depth of
entrance pupil, entrance port, and the angular field. What is the depth of field for the +16.00 D
field of view. Also specify the exit pupil. A lens at an f/1.8 setting?
+4.00 D field lens with a 3 cm diameter is now 12. A camera has a 5 m hyperfocal distance. What is
added to the telescope. Specify the change in the the camera's depth of field when focused for the
entrance port and angular field of view. What is hyperfocal distance?
CHAPTER TWENTY

Aberrations

20.1 Detriments to Image Quality wavefronts that are not quite spherical. This means
that perfect point images are no longer formed, and
In a good quality image, both the gross and the fine extended images are degraded in quality. Besides
detail are sharp and clear. In a poor quality image, the degradation, image distortion may occur for nonparax-
fine detail is blurred out, even though the gross detail ial paths.
may still be resolvable. For example, we can easily Diffraction is the ability of waves to bend around
recognize a person (gross detail) in a slightly blurred corners. Even if all the aberrations are controlled, the
photograph, but we cannot see (resolve) the wrinkles ability of an optical system to image fine detail is still
(fine detail) by the person's eyes. The use of a mag- limited by diffraction. Diffraction is discussed in Chap-
nifying system to examine the blurred photograph only ter 22.
magnifies the blur, and confirms the fact that the fine
detail is degraded or lost.
Besides a misfocus (blur), another source of image
degradation is stray light caused by extraneous light 20.2 Dispersion and Chromatic Aberration
sources or by scatter in the system. Sources of scatter
include scratches, dirt, or inhomogenities in the media. When white light is incident on a thick prism, the light
For example, a cataract in the crystalline lens degrades is dispersed into a rainbowlike spectrum (Figure
the retinal image by scattering light. 20.1a). The geometric reason for the dispersion is that
Even for properly focused systems free of scatter, the refractive index of the material is slightly different
there are optical sources of image degradation. These for each incident wavelength of light.
sources are referred to as the aberrations. The aberra- Figure 20.1b shows the index of refraction of oph-
tions can be divided into chromatic aberration and the thalmic crown glass as a function of the incident
so-called monochromatic aberrations. wavelength. One way to obtain such data is to use an
Chromatic aberration is due to the fact that the ophthalmic crown prism and measure the minimum
refractive index of a material varies slightly with the deviation angle for each incident wavelength. Then we
wavelength. Thus, the refracting power of that materi- can use the minimum deviation equation (Eq. 17.7) to
al varies with the wavelength, and consequently, a solve for the index of refraction for each wavelength.
single lens does not focus all wavelengths at the same From Figure 20.1a, the shorter wavelengths (violets
position. and blues) are bent more than the longer wavelengths
The theoretical reason for the monochromatic ab- (reds). This occurs because the refractive index is
errations of spherical systems is that the paraxial greater for the shorter wavelengths (Figure 20.1b).
equations, including V = P + U and m = U/V, break Thus, the violet and blue wavelengths have a larger
down as the light paths leave the thread-like paraxial refractive index than the reds. The human eye is not
region. Consequently, the nonparaxial light paths give as sensitive to violet as it is to blue, so frequently in

423
424 Geometrie, Physical, and Visual Optics

white

white
&>
* z,een
v
'oiet
a)

nl
1.531 4-
FIGURE 20.3. Rainbow rays.

1.521 +
drop near its edge, is reflected once inside the rain-
E u- E Q E drop, and then exits the drop. The dispersion occurs at
the two refractions (entering and exiting). The color
that appears on top of the primary rainbow is found by
extending the outgoing rays back toward the virtual
image position. This shows that red appears on top of
FIGURE 20.1. a. Dispersion by a thick prism, b. Refractive the rainbow, which is opposite to the BD prism case.
index vs incident wavelength. Similarly, the "fire" (play of colors) in a diamond is
due to dispersion. The dispersion occurs at the refrac-
tions when the light enters and exits the diamond.
dealing with chromatic aberration, we label the short
wavelength zone as blue.
For a thin prism of refractive index n in air,
d = (n-l)A, (20.1) 20.3 Longitudinal and Lateral
Chromatic Aberration
where d is the deviation angle and A is the apex angle.
It is obvious from the equation that the higher indices Consider a convex lens with red and blue wavelengths
give the larger deviations. Figure 20.2 shows a white incident from a distant axial point source. The convex
light bulb located in front of a base down (BD) prism. lens is thickest at the optical center. We can crudely
The prism forms a virtual image that appears to be represent the convex lens as two prisms base to base
deviated toward the apex. Since the prism bends the (Figure 20.4a). Since the prisms deviate blue light the
shorter wavelengths down the most, a blue blur figure most, we see that the blue point image is formed
appears on the top of the white virtual image. Similar- before the red point image. For the lens, the linear
ly, a red blur fringe appears on the bottom of the separation of the red and blue point images is a
white virtual image. These colored fringes are chrom- longitudinal (or axial) manifestation of chromatic
atic aberration due to the dispersion. aberration. Experimentally, this corresponds to mov-
While dispersion is the cause of chromatic aberra- ing the image screen between the clear image posi-
tion, it is also the cause of much aesthetic joy. For tions.
example, dispersion is the cause of the colors of a If we place an image screen at the position of the
rainbow. Figure 20.3 shows the ray path for light that red image, then the red point image is surrounded by
makes the primary rainbow. The ray enters the rain- a blue blur circle (Figure 20.4b). The blur circle is a
lateral (or transverse) manifestation of chromatic aber-
ration. Similarly, when the image screen is placed at
blue the position of the blue image, the blue point image is
v i r t u a l
^O^-_
,mage
red^^"
surrounded by a red blur circle.
Figure 20.5 shows the chief ray for an off-axis
bulb object in front of a system consisting of a separated
aperture and lens. Due to the longitudinal chromatic
FIGURE 20.2. Chromatic blur due to a prism. aberration, either the red or the blue image will be
Aberrations 425

a)

FIGURE 20.4. a. Base to base prisms pro-


vide intuition on the origin of longitudinal
chromatic aberration for a plus lens. b. Lon-
gitudinal and lateral chromatic aberration.

slightly blurred. When the aperture is small enough, 20.4 Refractive Efficiency
the depth of focus may be large enough that the blur is
not noticeable. However, the chief ray also specifies In order to quantify our discussion of chromatic aber-
the size of the blurred image. Consequently, on the ration for refraction, we need to use specific wave-
screen, the blue image is smaller in size than the red lengths. The conventional (vacuum) wavelengths used,
image. The image size difference is also a lateral their colors, their sources, and their Fraunhofer de-
manifestation of chromatic aberration, and may be signations are presented in Table 20.1.
noticeable even when the blur is not. This lateral
manifestation is sometimes referred to as a chromatic
difference in magnification. TABLE 20.1
The human visual system is more sensitive to later- Reference vacuum wavelengths for chromatic aberration
al manifestations of chromatic aberration than it is to
longitudinal manifestations. A system that has a zero Wavelength Fraunhofer
(or minimal) longitudinal chromatic aberration may (nm) Color Source designation
have a noticeable lateral chromatic aberration. We can
design systems that have minimal lateral chromatic 656 Red Hydrogen C
589 Orangish- Sodium D
aberration by leaving some longitudinal chromatic yellow
aberration. 489 Aqua blue Hydrogen F
Systems that depend on front surface mirrors are
free of chromatic aberration. This is because chro-
matic aberration depends on the refractive index,
while the basic law of reflection states that the angle of The emission spectrum for an excited hydrogen gas
reflection equals the angle of incidence independent of consists of a series of discrete wavelengths including
the refractive indices. Thus, the image formation by a the C and F lines. Thus, we can easily isolate the C
front surface mirror does not depend on the refractive and F wavelengths and use them for measurements.
index. An excited sodium source emits the characteristic

1 J k ^

L^^- blue "~

^ ^ T , f FIGURE 20.5. Chief ray and size manifestation of


longitudinal chromatic aberration.
426 Geometrie, Physical, and Visual Optics

orangish-yellow wavelength of sodium. (The orangish- are the refractive efficiency v times 10. For ophthalmic
yellow interstate and parking lot lights are sodium crown glass, n D is 1.523, v is 58.6, and the code is then
sources.) The sodium D emission actually consists of 523-586. The codes for some common transparent
two wavelengths very close together (a so-called doub- materials are presented in Table 20.2.
let line). The doublet wavelengths are at 589.0 nm and
589.6 nm. For most uses, the fact that the sodium D
line is a doublet does not matter, and the D line is
treated as a single wavelength at 589.3 nm. However, TABLE 20.2
for precision work, other wavelengths such as the Refractive parameters for some transparent materials
helium d line at 586 nm (a singlet line) are used.
By convention, the refractive index used to com- Material Code
pute the geometric optics refracting powers is n D the
average index for the sodium D doublet (589.3 nm). Water 334-556
Alcohol 363-606
This corresponds to using a sodium source to measure CR-39 Hard Resin Plastic 495-580
the refracting powers of lenses or prisms. Thus, the Ophthalmic Crown Glass 523-586
+4.00 D mark on a trial lens means that the lens is Polycarbonate 580-300
+4.00 D for sodium D light. Similarly, the 6 mark on Dense Flint Glass 617-366
a trial prism means the prism is 6 for sodium D light. Dense Barium Crown Glass 617-549
Highlite Glass 701-310
For the C, D, and F wavelengths, the refractive
indices of ophthalmic crown are as follows: n F = 1.530,
nD = 1.523, and n c = 1.521.
The refracting power for a single spherical air-glass
interface of radius r for each of the three conventional For transparent materials, the refractive efficiency
wavelengths is v ranges from about 20 to 80. From Eq. 20.4, the
higher the refractive efficiency, the lower the
nn-l chromatic aberration. For a +16.00 D ophthalmic
PD =
crown surface, the chromatic aberration is
nF-l +16.00 D
PF = CA = = +0.27D,
58.6
while for a +16.00 D highlite surface,
r + 16.00 D
Of these, PD is the conventional geometric optics CA = = +0.52D.
31.0
dioptric power. The dioptric chromatic aberration CA
is defined as Figure 20.6 is a plot of n D vs v for various types of
glass. The majority of types are in the shaded area.
CA = P F - P C . (20.2) The tendency is for higher index materials to have
lower refractive efficiencies. However, as a result of
It follows from the above four equations that
considerable research, there are modern glass types
n F nr that have a high index with a good refractive efficien-
CA = P n
nn-l
The term on the right is a characteristic of the material
only, and is called the dispersive power . Its reciproc-
al v is more commonly used and is variously called the
refractive efficiency, the nu-value, and V-value, the
Abbe number, or the constringence:
1
(20.3)

Then
P
(20.4) 20 30 40 50 60 70 80

A six number code is used to classify transparent


materials. The first three numbers are the three digits FIGURE 20.6. Refractive index for sodium D vs refractive
behind the decimal in n n . The second three numbers efficiency for various types of glass.
Aberrations 427

cy. (These glasses fall between the bowed curve and gives
the shaded area). Barium crown glass is one such type.
For comparison, dense barium crown has a code of CA = ^ = +1.08D,
617-549, while dense flint glass has a code of 617-366.
with blue focusing before red. According to this
For a 10.00 D barium crown surface, the chromatic
model, the eye has just over 1.00 D of positive
aberration is
chromatic aberration (Figure 20.4b and 20.7a). In
view of the 0.25 D clinical significance, the amount of
CA-^55-HU.D. chromatic aberration in the eye is surprising. A person
while for a - 1 0 . 0 0 D flint glass surface corrected from sodium D light would be myopic for
blue light and slightly hyperopic for red light. The
-10.00 D
CA = = -0.27D. chromatic aberration of the eye is the basis for the
36.6 red-green test commonly used in clinical refraction.
For a thin lens, the total chromatic aberration is the
sum of that for each surface. Then
=
^Alens t-Aj + C A 2 ,
20.5 The Achromatic Doublet
or
Since some materials have different refractive efficien-
CAlcns=^ + ^ , cies, it is possible to combine a positive and a negative
lens of different materials into an achromatic doublet
where P t and P 2 are the front and back surface powers.
(Figure 20.8a). In the thin lens design of an achro-
Since the lens power is the sum of the surface powers
matic doublet, the total power of the doublet is the
it follows that
sum of the components,
P.
^A-l^ne (20.5) Pt = p, + P 2 ; (20.6)
A plus lens converges the blue more than the red, while the zero chromatic aberration condition is
while a minus lens diverges the blue more than the
red. Thus, a minus lens tends to correct the chromatic 0 = Pi + ^, (20.7)
aberration of a plus lens. For positive systems that
consist of both plus and minus lens, the resulting where Px and P 2 are the respective powers of the
chromatic aberration is called positive when the F components and and v2 are the respective refractive
(blue) image is formed closer to the system than the C efficiencies. Once the materials are specified, we can
(red) image (Figure 20.7a), and negative when the C solve Eqs. 20.6 and 20.7 for the unknown powers Fl
(red) image is formed closer than the F (blue) image and P 2 of the components.
(Figure 20.7b). The longitudinal chromatic aberration
is zero when the F and C images are focused in the EXAMPLE 20.1
same place. Design a +10.00 D achromatic doublet using oph-
thalmic crown glass (523-586) and dense flint glass
For chromatic aberration, a 60 D single spherical (617-366).
refracting interface between air and water (334-556) From the condition for zero chromatic aber-
is a pretty good model of the human eye. This model ration,
Pi , P 2
0:
58.6 36.6'
-= + Then
blue red 58.6
Pi = - 3 6 * P 2 = -1->P 2 .
a)
positive CA The power condition is

+ 10.00 0 = ^ + P 2 .
-* + By substituting for Pl, we obtain
red blue
+ 10.00 D = - 1 . 6 0 P2 + P 2 ,
b)
negative CA +10.00 D = - 0 . 6 0 P,

FIGURE 20.7. Longitudinal chromatic aberration for a system. or


a. Positive, b. Negative. P2 = -16.64 D,
428 Geometrie, Physical, and Visual Optics

dense flint

ophthalmic
crown

a)

highlite

cx ophthalmic FIGURE 20.8. a. Achromatic doublet, b. Exagge-


' crown ration of focus differences, c. Achromatic prism.

Since remaining chromatic aberration (which is small) is


^-.^ referred to as the secondary spectrum.
One straightforward way to focus three wave-
we can substitute to obtain lengths at one position is to use a triplet lens (three
P t = -1.60(-16.95 D) = +26.64 D. components with three different materials). The equa-
tions involved consist of three equations in three
So a +10.00 D doublet lens made of a +26.64 D
unknowns. The resulting lens is called an apochro-
ophthalmic crown component and a -16.64 D flint
glass component has zero chromatic aberration (Fi- matic lens. Actually, a triplet lens that does better
gure 20.8a). than focusing three wavelengths in one position can be
We can check the arithmetic in the above exam- designed by considering the exact form of the index of
ple by noting that the sum of +26.64 D and refraction as a function of wavelength. Such lenses are
-16.64 D is +10.00 D, and by noting that referred to as superachromats.

+26.64 D -16.64 D
C A , _ = ^ 58.6
^ - + 36.6 '
CA,. +0.46 D + (-0.46 D) = 0. 20.6 Achromatic Prisms

For the achromatic doublet, the component with Equation 20.1 gives the deviation for a thin prism of
the higher power (in magnitude) has to also have the index n and apex angle A. The prism diopter rating is
higher refractive efficiency, while the component with given by
the lower power (in magnitude) has to have the lower Z = 1 0 0 tan d. (20.8)
refractive efficiency.
In the achromatic doublet, the zero chromatic aber- When d in radians is small
ration condition (CA = 0) specifies that the dioptric Z~100d,
powers are equal for the hydrogen C (red) and F
(blue) wavelengths. That does not mean that the or
dioptric power for sodium D light is equal to the Z~100(n-1)A.
dioptric power for the C and F wavelengths. The
The prism diopter values are specified in terms of the
achromatic doublet focuses the C and F light at the
sodium D wavelengths, or
same position, but does not necessarily focus D light
there (Figure 20.8b shows an exaggerated view). The 100(nD-l)A. (20.9)
Aberrations 429

We can write similar expressions for the prism diopter where the minus sign comes from the required BD.
ratings for hydrogen C (red) and F (blue) light. Then We can substitute for Z, to obtain
the chromatic aberration of the prism (in prism diop- - 6 = 1.89Z2 + Z 2
ters) is defined by
- 6 = -0.89 Z 2 .
CA Z = Z F - Z C . (20.10) So
It follows that +6.74*
The plus sign means the prism is BU. Then
CA = ^ (20.11)
Z, = --1.89 Z 2 =-12.74*
where v is the refractive efficiency of the material. The
So a 12.74 BD ophthalmic crown prism combined
chromatic aberration of two thin prisms in contact is with a 6.74 BU highlite prism constitutes a 6 BD
then achromatic doublet (Figure 20.8c).
We can check the arithmetic by first noting that
C A z = ^ + =^. (20.12) 12.74 BD combined with 6.74 BU is indeed
equivalent to a 6 BD. Second, the chromatic
We can use Eq. 20.12 and prism additivity to design aberration for this doublet is
achromatic prisms just as we design achromatic lenses. 12.74 6.74
CAZ =
EXAMPLE 20.2
58.6 + 31.0 '
Design a 6 BD achromatic prism using ophthalmic CA. -0.22* + 0.22 = 0.
crown (523-586) and highlite (701-310) glass.
The zero chromatic aberration condition is
Z1 Z2
58.6 + 31.0' 20.7 Thick Lenses and Chromatic Aberration
or
Achromatic doublets of high dioptric power are actu-
Z = Z = - 1 89 Z ally thick lenses. A thick lens system that has zero
31 0 longitudinal chromatic aberration might still have
The power condition is some lateral chromatic aberration. The reason is that
-6A = Z!+Z2, the Gaussian principal planes for the hydrogen C light

FIGURE 20.9. a. System with lateral


chromatic aberration even though longitudi-
nal chromatic aberration is zero. b. System
with minimized lateral chromatic aberration.
430 Geometrie, Physical, and Visual Optics

(red) can occur at positions different from the princi- The equivalent power for two thin lenses in air sepa-
pal planes for the hydrogen F light (blue). Figure rated by a distance t is
20.9a shows the secondary principal planes and the
Pe = P, + P2 tP,P 2 .
image space nodal rays for this situation. The image
screen has clear red and blue images (zero longitudinal Then the derivative is
chromatic aberration), but the image sizes are dif- dP, dP, , dP, -2 t d(P.P 2 )
ferent. dn dn dn
dn
Figure 20.9b shows a system in which the lateral
chromatic aberration is minimized by leaving some The chain rule for differentiation gives
longitudinal chromatic aberration. The blue image is dP^ dP dP, dR dP,
focused slightly in front of the screen position, while dn dn dn dn ^2 ^1 dn '
the red image is focused slightly behind the screen
position. The system is designed so that the red and or
blue chief rays coincide. Therefore, at the screen dP^ P, tP P tP,P2
position, the red and blue image sizes (as specified by dn (n-1) +' ( - 1 ) ~ ( = ) (n-1)'
the chief rays) are equal. The condition for the minimum is

^ = 0
dn '
20.8 Separated Lenses of the and this occurs when
Same Material 2tP,P2=P,+P2.
High magnification telescopes and microscopes are not We can solve for t to obtain
possible unless special care is taken to control the P
aberrations. In the ocular lens, it is much more im- t=l/2 ^^
P,P2 '
portant to control the lateral chromatic aberration as
opposed to the longitudinal chromatic aberration. This
can be done with two separated lenses of the same
material by minimizing the equivalent power with The reciprocal of the power is the secondary focal
respect to variations in the refractive index. Since the length of the lens, so
magnification depends on the equivalent power, this
procedure minimizes the variations in image size (la- (20.15)
t = ^ [ ( f 2 ) 1 + (f2)2]
teral chromatic aberration) rather than variations in
image location (longitudinal chromatic aberration). Equation 20.15 says that for the two lenses, the lateral
(For comparison purposes, you might think of min- chromatic aberration is minimized when the separa-
imizing longitudinal chromatic aberration by min- tion t equals one half the sum of the secondary focal
imizing the appropriate vertex power as opposed to lengths.
the equivalent power.) The Huygens eyepiece is a classic design that meets
The standard calculus procedure for finding an the above criterion. The two lenses in the Huygens
extreme (maximum or minimum) is to set the deriva- eyepiece are called the field lens and eye lens (Figure
tive equal to zero. For a single spherical refracting 20.10). The power of the eye lens is twice that of the
interface, field lens.
P = (n-l)/r = (n-l)R,
where R is the curvature. Then the derivative with EXAMPLE 20.3a
respect to the index is A Huygens eyepiece consists of a +20.00 D oph-
thalmic crown field lens and a +40.00 D ophthalmic
crown eye lens. What is the separation between the
lenses, and the magnification of the eyepiece?
We can rewrite this as The separation is equal to one half of the sum of
the secondary focal lengths or
dP^R(n-l)= P
(20.13)
dn (n 1) (n-1) t = - (50.0 mm + 25.0 mm) = 37.5 mm.
For a thin lens of power Pt it follows that
The magnification of the eyepiece (for plane waves
dPt _ P, leaving) is equal to the equivalent power divided by
(20.14)
dn (n-1)' 4. Here
Aberrations 431

virtual , e v e

Huygens Eyepiece FIGURE 20.10. Huygens eyepiece.

Pe = (+20.00 D) + (+40.00 D) So the primary focal point is 16.67 mm behind the


- [(0.0375 m)(+20.00 D)(+40.00 D)], first lens. Hence, the field lens is placed so that the
light converging from the objective lens would form
Pe = +60.00 D - 30.00 D = +30.00 D. its image 16.67mm behind the field lens (i.e., the
Therefore, object for the field lens is virtual).
M = P e /4 = 7.5X. We can get an intuitive feel for how the
Huygens eyepiece works by following the rays in
EXAMPLE 20.3b Figure 20.10. The blue ray that is parallel between
For plane waves leaving, where must the Huygens the lenses bends down at the eye lens. Relative to it
eyepiece be located relative to the image formed by the red ray would be traveling slightly upward
the objective lens? between the lenses (i.e., because of chromatic
We want the image formed by the objective lens aberration the red ray did not bend as much as the
to occur at the primary focal point of the Huygens blue ray at the field lens). Hence, the red ray
eyepiece. We can find the primary focal point from strikes the eye lens at a point farther from the
the neutralizing power. Since optical center than the blue ray. Prentice's rule for
the eye lens gives
Z=-Ph.
1 - tP 2 '
The dioptric power P is smaller for the red ray than
+30.00 D for the blue. However, this is offset by the larger h
P = 1 - (0.0375 m)(+40.00 D) ' value for the red ray. The net effect is for the red
and blue rays to come out parallel to each other
+30.00 D (which gives zero lateral chromatic aberration).
P = = - 6 0 . 0 0 D.
-0.5
An alternative approach is to use the system
matrix:
20.9 The Monochromatic
S = R 2 TR 1 ,
Wavefront Aberration
or
40.00 D
" ( 1.0 1 0.0375 m 1/
Consider a centered spherical system illuminated by
monochromatic light. In Chapter 7, we derived the
1 -20.00 D vergence equation, V = P + U , with the help of the
vO 1 sagittal approximation. While the vergence equation
works well in the threadlike paraxial region, the light
/-0.500 -30.00 D \ that passes through nonparaxial parts of the system

V 0.0375 m 0.250 ) ' does not behave as predicted by the paraxial equa-
tions.
The determinant check gives +1.00. Then The box in Figure 20.11a represents the centered
Pe = + 3 0 . 0 0 D, spherical system with incident monochromatic light
and from an axial point source. The dashed curve repre-
Pn = P e / S l l = +30.00/(-0.500) = -60.00 D. sents the spherical wavefront predicted by the paraxial
equations. This spherical wavefront is also the wave-
The front focal length is front required to give a perfect point image (neglect-
ing diffraction). The solid curve represents the actual
or wavefront leaving the system. In the paraxial region,
1000 mm/m there is essentially no difference between the pre-
ff=" = +16.67 mm. dicted and the actual wavefront. However, as a func-
-60.00 D
432 Geometrie, Physical, and Visual Optics

/ >^spherical

axial object
Vpoint

axial object
y/point
paraxial focus

FIGURE 20.11. Spherical aberration, a.


Wavefront. b. Rays.

tion of the distance from the axis, the actual wavefront light that makes a large angle with the axis. Thus,
differs more and more from the spherical curve. In these systems are nonparaxial, and unless corrected
general, the deviation of the actual wavefront from the have an inherent image degradation problem due to
perfect sphere is referred to as the wavefront aberra- the wavefront aberration. Precision fast or wide-field
tion. For an axial point source, the wavefront aberra- optical systems give an actual exiting wavefront that is
tion has only one component, which is called spherical closer to a sphere than that of other systems. Con-
aberration. sequently, precision systems have a better image
We can also describe spherical aberration in terms quality.
of the rays. The rays in Figure 20.11b correspond to It is important to note that the design of precision
the actual wavefront in Figure 20.11a. The paraxial systems is object distance-dependent. A system that is
rays focus at the desired image point, but each set of designed to minimize the wavefront aberration for an
nonparaxial (or peripheral) rays focus progressively object at optical infinity has some wavefront aberra-
farther from the paraxial image point. These peripher- tion for a near object, and vice versa.
al rays cause a circular blur patch to occur around the
paraxial image point. For an extended object, these
peripheral rays cause image degradation.
Fast optical systems are those with short exposure 20.10 The Seidel or Third
times. One way to make systems faster is to increase Order Aberrations
the aperture size. As the aperture size increases, part
of the light path becomes nonparaxial. Consequently, For an arbitrary system, there are no closed form
unless corrected, fast systems have an inherent image algebraic equations that describe the exact actual
degradation problem due to spherical aberration. wavefront. Therefore, how can we predict what an
For an off-axis object point and non-paraxial light optical system will do? In other words, how are preci-
paths, the actual wavefront leaving the system also sion optical systems designed?
differs from the spherical wavefront required for a One tool is to numerically determine the exact
perfect image. For the off-axis case, the symmetry that actual wavefront for a given system. We can do this by
occurs for the axial case is broken, and the actual calculating exact ray traces through the system and
exiting wavefront may even contain a toricity factor then using the rays to reconstruct the wavefront. An
(e.g., the wavefront shape might be toric, ellipsoid, exact ray trace involves the use of Snell's law,
hyperboloid, or some other awful shape). Thus, for
n1 sin 0; = n2 sin 0r, (20.16)
off-axis point sources, the wavefront aberration has
spherical aberration plus additional components. at each refracting surface. The exact ray trace is
Systems that have a wide field of view deal with contrasted to a paraxial ray trace that involves the
Aberrations 433

small angle approximation, We could proceed further with the approximations


^-^. (20.17) and keep the fifth order term. This would generate
another set of aberrations, the fifth order aberrations.
However, the use of numerical calculations without For ray angles less than 35 degrees, the fifth order
any theory is analogous to looking for a needle in a aberrations are usually an order of magnitude smaller
haystack. The theory provides a guide for where to than the third order aberrations. Here, we discuss only
look. In the absence of exact equations, we can im- the third order aberrations, or Seidel aberrations.
prove the theory with better approximations. For For systems with spherical surfaces, the five Seidel
small angles (expressed in radians), the sine function is aberrations are interrelated. This complicates lens de-
approximated by sign work because the design that minimizes one of
the five frequently makes one or more of the others
03 05 1
sin 0 ~ 0 - ^- + - jj- + (20.18) worse. (In addition, the designs are object distance-
dependent.)
We could start with the paraxial version of Snell's law
and derive the vergence equation V= P + U. The next
higher order approximation would involve keeping the
third order terms for the sine function. Then Snell's 20.11 Spherical Aberration
law becomes
For a centered spherical system, spherical aberration
e,-|[)=n2(er-||). (20.19) is the only wavefront aberration component that oc-
curs for both on- and off-axis point sources. Figure
When we use the third order terms, we find that 20.12 shows spherical aberration for a distant axial
the difference between the actual wavefront and the monochromatic point source located in front of a
perfect sphere can be described by the sum of five convex single spherical refracting interface. After re-
terms. While these terms are all interrelated, we tend fraction, the paraxial rays intersect at the on-axis
to think of them as five separate aberrations, each of paraxial image position. The peripheral rays obey the
which contributes to the wavefront aberration. The exact form of Snell's law, and they intersect at a point
five terms are called spherical aberration, coma, radial closer to the interface than the paraxial focus.
astigmatism, curvature of field, and distortion. Of Both surfaces of a spherical lens contribute to the
these, spherical aberration is present for both on- and total spherical aberration of the lens. For plus thin
off-axis object points. The other four are present only lenses, the peripheral rays focus closer to the interface
for off-axis object points. than the paraxial rays (as in Figure 20.11b). The
The five aberrations that result from the third order longitudinal manifestation of spherical aberration is
term were first described by Ludwig von Seidel in the the difference between the image location for the
1850s, and so they are sometimes referred to as the peripheral rays vs the image location for the paraxial
Seidel aberrations. They are also referred to as the rays. For a screen at the paraxial image position, the
primary aberrations, the third order aberrations, or as lateral manifestation of spherical aberration is the
the aberrations of form since they depend on the exact circular blur patch due to the peripheral rays. The
form (biconvex, equiconvex, meniscus-convex) of the deviation of the rays from the paraxial focus varies
lens. like the square of the aperture size h. This means that

FIGURE 20.12. Spherical aberration for a


convex SSRI.
434 Geometrie, Physical, and Visual Optics

paraxial focus

paraxial focus

FIGURE 20.13. Shape dependence of spherical aber-


ration. a. Light incident on concave side of a lens. b.
Light incident on the convex side of the same lens.

when the aperture size is doubled, the longitudinal the convex side forward (minimum deviation) is less
spherical aberration quadruples. (The h squared de- than the spherical aberration for the same lens with
pendence can be derived by keeping the next term in the concave side forward (maximum deviation). Thus,
the sagittal approximation (Eq. 7.5), that was used to the same meniscus lens when turned around gives
derive V = P + U.) differing amounts of spherical aberration.
Consider an extended object in front of a thin For light traveling right, Figure 20.14 shows a plot
biconvex lens next to an aperture. When the aperture of longitudinal spherical aberration as a function of
is small, only the paraxial region is exposed, and a lens form for a fixed aperture size h and a distant
good quality image is formed. As the aperture is object. The worst spherical aberration occurs for the
opened up, the paraxial image is degraded by the stray meniscus-convex lens with the concave side forward.
light coming from the peripheral rays. For a large The spherical aberration never goes to zero, but there
aperture setting, we can actually obtain a better image is a form (biconvex with the front side curved more
by moving the screen forward from the paraxial image than the back), which gives the minimum spherical
position to a position that would correspond to a circle aberration. Figure 20.14 shows that the specific lens
of least confusion due to the spherical aberration. form is important in lens design.
When we use an annular aperture that lets through
only a narrow region of peripheral rays (i.e., blocks
the paraxial rays), then the image screen must be
moved even further forward to the image position for
the peripheral rays. The quality of this image is poorer
than that of the paraxial image.
For a single lens, spherical aberration is sensitive to
the exact shape or form of the lens. This is under-
standable in terms of Prentice's rule and thick prism
theory. Consider the meniscus-convex lens in Figure
20.13a. The concave side of the lens is facing the
incident direction. Let us consider the peripheral sec-
tion of the lens as a thick prism (large apex angle).
The ray is incident at an angle on the apex side of the
normal to the first surface. This results in a prism
deviation that is close to maximum. Now what hap-
pens when we turn the lens around so that the convex
side faces the incident direction? Then the ray ap-
proaches the surface at an angle on the base side of
the normal (Figure 20.13b). The resulting prism devia- FIGURE 20.14. Spherical aberration and coma as a function of
tion is close to minimum. The spherical aberration for lens shape..
Aberrations 435

FIGURE 20.15. Aspheric refracting interface to correct spherical


aberration.

paraxial peripheral /
f c us
focus o , /

\
\

FIGURE 20.16. Spherical aberration by a diverging lens.

Another method to minimize (or eliminate) spheri- the (virtual) focus for the peripheral rays is closer to
cal aberration is to use aspheric surfaces. Given an the lens than the paraxial focus.
axial set of conjugate points, there is some surface for In a system consisting of plus and minus lenses, the
which every ray leaving the object point passes after peripheral rays are diverged too much by the minus
refraction through the image point. These surfaces are lenses and converged too much by the plus lenses.
referred to as Cartesian ovals. They are usually not Therefore, the spherical aberration from the minus
spherical. Instead they tend to flatten out away from lenses works to correct the spherical aberration of the
the axis (Figure 20.15). (It is interesting to note that plus lenses. For a system that forms a real image, the
the human cornea flattens out away from the axis.) spherical aberration is called positive or undercorrec-
For a spherical surface, we can then think of spherical ted when the peripheral rays focus closer than the
aberration as occurring because the surface is not the paraxial rays, and negative or overcorrected when
proper Cartesian oval for that object distance. peripheral rays focus farther than the paraxial rays
Figure 20.16 shows the effects of spherical aberra- (Figure 20.17). The spherical aberration is corrected
tion for a point source in front of a minus spherical or zero when the peripheral rays focus in the same
lens. Again, the peripheral rays are bent too much and place as the paraxial rays.

peripheral paraxial
focus 1
focus

positive spherical aberration

>+ paraxial
focus peripheral
focus

FIGURE 20.17. Positive vs negative spherical aber-


negative spherical aberration ration.
436 Geometrie, Physical, and Visual Optics

-doublet (1 Ox)

20 15 10 5
longitudinal spherical aberration

FIGURE 20.18. Longitudinal spherical aberration as a


function of distance from the optical center. Plots for single
lens and doublet lens.

Figure 20.18 shows a plot of longitudinal spherical tered on the chief ray. The tangential plane contains
aberration as a function of aperture size h for a single the chief ray and the optical axis, while the sagittal
lens compared to that for a doublet lens of the same plane contains the chief ray and is perpendicular to the
back vertex power. The doublet lens consists of a tangential plane. Consider a point source located ver-
positive component of one material and a negative tically below the optical axis. In this case, the tangen-
component of a different material. The horizontal tial plane is the vertical plane (Figure 20.20). Relative
displacement of the doublet curve has been multiplied to the perfect sphere, an exiting wavefront exhibiting
by 10 in order to make it visible. The performance of coma is curved too much on one side of the tangential
the doublet is clearly far superior to that of the single plane (perhaps the top side), and not enough on the
lens. Below h values of 14, the doublet undercorrects other (perhaps the bottom side). The tangential rays
the spherical aberration, and then overcorrects it for h through the middle of the lens form an image point at
values greater than 14. By adjusting the thicknesses A. The intermediate set of rays converge at B (show-
and surface curves, doublets can be designed to con- ing the difference in magnification), and the marginal
trol both spherical and chromatic aberration. rays converge at C. The corresponding positions in the
Clearly, spherical aberration can be corrected by coma flare pattern are shown on the right.
keeping the apertures small. However, this does not The rays in the tangential plane stay in that plane.
work for fast or wide field systems. For these, lens For the vertical tangential plane, Figure 20.21a shows
forms, doublets, systems of separated lenses, or as- a series of cross-sectional views of the tangential rays
pheric surfaces can be used to control spherical aber- at various positions behind the lens.
ration for a given object distance. As a function of meridian, the wavefront curvature
has a cosine variation, and matches the perfect sphere

20.12 Coma coma A


(negative) ~"2,^,f\
Unlike spherical aberration, coma occurs only for
off-axis points. For an off-axis point source, the
spherical symmetry is broken resulting in a comet-
shaped illumination patch (Figure 20.19). The asym-
metric character of coma makes it very detrimental to
image quality.
Spherical aberration can be defined as the variation
of focus with the distance h that the rays are off-axis.
Coma can be defined as the variation of lateral magni-
fication with the distance h. Coma and spherical aber-
ration are intimately related, and yet a system may be
free of one and still have the other.
*">^ point source
For off-axis points, we need to identify two refer-
ence planes, the tangential and sagittal planes, cen- FIGURE 20.19. Coma for an off axis point source.
Aberrations 437

FIGURE 20.20. Wavefront,


rays, and illumination patch for

in the sagittal meridian. However, because of the creasing and twisting effect before folding up at the
relative torsion of the wavefront in the sagittal meri- coma flare pattern (Figure 20.21c). The net effect of
dian, the sagittal rays are skew rays that do not stay in the ray behavior is that the rays passing through an
a single plane. The surface defined by the outgoing annular region of the lens form a circle at the image
sagittal rays resembles a sheet of paper with a progres- plane (Figure 20.21d). As the radius of the annular
sively increasing crease in the middle. The sagittal rays zone increases, the circles formed by the rays increase
fold up at the coma flare pattern position. Figure in diameter and become progressively decentered. The
20.21b shows a series of cross-sectional views of the cumulative effect of the nonconcentric circles is the
sagittal rays. coma flare pattern.
Rays in the other meridians exhibit an asymmetric When the coma tail is closer to the axis than the


a) 1

A
3 3
h
C

4 4

FIGURE 20.21. Coma. a. Tangential plane, b. Sagittal


plane, c. Arbitrary meridian, d. Annular zone.
438 Geometrie, Physical, and Visual Optics

head, as in Figure 20.19, the coma is labeled negative. ration and coma. Aplanatic points satisfy a sine condi-
Positive coma has its head closer to the axis than the tion called the Abbe sine condition (after Ernst Abbe,
tail. Like spherical aberration, coma is strongly depen- 1840-1905). Figure 20.22a shows an axial ray diagram
dent on lens form. In fact, coma can be positive, for a set of aplanatic points for a converging spherical
negative, or even zero depending on the form of the interface of radius r separating media of index n and
lens. Figure 20.14 shows a plot of coma (together with n' (where n > n'). The object is located r/n to the left
spherical aberration) as a function of lens form. The of the center of curvature C, and the image point is
form of the lens that gives zero coma is close to the located a distance r times n to the left of C. The ray
form that gives minimum spherical aberration. leaving the object point makes an angle a with the
Also, like spherical aberration, coma depends on axis, while the ray leaving the image point makes an
the square of the aperture size. When the aperture is angle a' with the axis. Figure 20.22b shows the off-axis
stopped down, the tail of the coma pattern is elimi- diagram (nodal ray) for an object of size h and an
nated leaving only the head. image of size h'. We can use the two diagrams, Snell's
Spherical aberration is independent of the location law, and the trigonometric law of sines to derive the
of the aperture stop. When spherical aberration is equation
corrected, then coma is also independent of the aper- hn sin a = h'n' sin a', (20.20)
ture stop location. However, when both spherical
aberration and coma are present, coma is dependent which for the aplanatic points is valid for any ray
on the location of the aperture stop in the system. angles a and a'. For conjugate points that are not
Some aperture stop locations result in positive coma, aplanatic, the above sine condition holds only for
some in negative coma, and one eliminates coma paraxial rays and fails for peripheral rays (large a
leaving only the spherical aberration. values). We can write the sine condition in terms of
the lateral magnification
h' n'sina'
m= = : (20.21)
20.13 Aplanatic Systems h n sin a
Since for aplanatic points, the condition holds for any
The aplanatic points of a system are conjugate object angle a, we see that the lateral magnification is a
and image points that are free of both spherical aber- constant for all aperture sizes h.

n > n'

'IL-rr-THP
virtual ji
image object point

a)

FIGURE 20.22. Aplanatic points, a. Axial ray. b.


Nodal ray.
Aberrations 439

20.14 Radial Astigmatism

Consider a centered spherical system with an off-axis


object point. Coma showed that the symmetry of the
exiting wavefront is broken. Another effect of the
broken symmetry is that the exiting wavefront has a
I slidT toric-type component that results in the Seidel aberra-
FIGURE 20.23. Oil immersion microscope objective. tion called radial astigmatism. Because of radial astig-
matism, the centered spherical system produces a
conoid of Sturm complete, for small apertures, with
two line images. For larger apertures, one of the line
As the object point moves off-axis, coma is the first images is curved. Radial astigmatism is sometimes
aberration to appear. Hence, spherical aberration (al- referred to as marginal or oblique astigmatism.
ready present on-axis) and coma are the two mono- In a small enough region, any smoothly varying
chromatic aberrations most important to control in the asymmetric surface (including the actual wavefront)
design of high magnification microscopes. We can take can be approximated by a toric surface. Therefore, it
advantage of the aplanatic points of a converging is not surprising that radial astigmatism appears as one
spherical surface provided we can put the object in the of the Seidel aberrations.
higher index medium ( n > n ' ) . This is the principal Radial astigmatism is analyzed in terms of the
used in the oil immersion microscope objective for tangential and sagittal planes already defined in the
high magnification. discussion of coma. An important factor is that the
The oil immersion microscope objective consists of tangential and sagittal planes vary with the meridional
a series of lenses, of which the first lens is a hemis- location of the object point. Figure 20.20 showed an
phere with a piano front surface. An oil drop with the object point vertically below the axis for a thin lens.
same refractive index as the hemisphere is placed on The vertical plane contains the optical axis and the
the object slide so that the object is effectively in the chief ray, and so is the tangential plane. The sagittal
higher index medium (Figure 20.23). The spherical plane contains the chief ray and is perpendicular to the
side of the hemisphere has a radius r such that when tangential plane. The horizontal meridian contains the
the microscope is in correct focus, the object in the oil intersection of the sagittal plane and the lens.
is at the aplanatic point (i.e., r/n from C). Now consider an object point that is horizontally
In paraxial optics, the Gaussian principal planes displaced from the optical axis. Here the chief ray and
enable us to conceptually treat a whole system in the the optical axis lie in a horizontal plane. So in this
same manner as a single spherical refracting interface. case, the tangential plane is horizontal and the vertical
For nonparaxial rays, the intersection surfaces become meridian contains the intersection of the sagittal plane
curved surfaces. For the system to be free of spherical and the lens. If the object point is rotated in a circle
aberration and coma, the total lateral magnification around the optical axis, the tangential and sagittal
must be a constant independent of h. This occurs planes rotate with the object point. In particular, an
when the Abbe sine condition (Eq. 20.20) relates the oblique object point gives oblique tangential and sagit-
object and image space parameters (Figure 20.24). To tal planes.
do this, the principal surfaces for the system must be Consider a collimated bundle of light that makes a
curved in a special way. For a distant object, the narrow path through the center of a thin lens of index
secondary principal curved surface must be a sphere n and power P. (This case is classically referred to as
centered on the secondary focal point. The Abbe sine oblique central refraction.) When the distant point
condition is surprising in that it relates an off-axis source subtends an angle with the optical axis of the
aberration, coma, to the ray angles of conjugate axial lens, the result of radial astigmatism is an apparent
points. differential increase in the magnitude of the refracting

IF*
FIGURE 20.24. Abbe sine condition for a system.
440 Geometrie, Physical, and Visual Optics

powers in the tangential and sagittal meridians. To the The power gain is frequently noticed by people (espe-
third order, the effective powers Pt and Ps in the cially myopes) who need a slightly stronger spectacle
tangential and sagittal meridians are, respectively, correction. They discover that when they tilt their
given by spectacles, as in Figure 20.25, they can see better.
Pt = P[(2n + sin2 ) I {In cos2 )], (20.22)
EXAMPLE 20.4
and Professor Walter Nutty wears a spherical correction
Ps = P[l + (sin24>/2n)]. (20.23) of -5.50 D. The professor views a distant straight
ahead object point and tilts his lenses 19 about a
Radial astigmatism (RA), in diopters, is the differ- horizontal meridian (Figure 20.25). From the sec-
ence between the tangential or sagittal powers, or ond order equations, what spherocylindrical lens is
equivalent to the tilted sphere? Also, what is the
RA = P t - P s . (20.24) effective power gain?
It follows from the above three equations that In this case, the tangential plane, defined by the
chief ray and the optical axis, is the vertical meri-
RA = P tan2 . (20.25) dian. There are 180 in radians, or 57.3 per
radian. So 19 equals 0.332 radians. Then from Eq.
As third order equations, Eqs. 20.22 through 20.25 are 20.26,
accurate out to angles of 35 degrees (and perhaps
some beyond depending on the precision needed). 4(0.332):
When n is 1.5, and the angles involved are less than P. = - 5 .50D[1 + ]
20 degrees, we can express in radians and use the Pt = -5.50D (1.147)=-6.31 D.
small angle approximations to simplify the above ex-
pressions. To the second order, Equation 20.27 gives
(0.332)2]

.-++ (20.26)
D[,1 +
P =-5.50D 3
Ps = -5.50 D(1.037) = -5.70 D.
J'

and then
-4] (20.27) The power cross for the tilted lens is then

-6.31 D
RA = Pt ?\ (20.28)
From the second order expressions, both Pt and Ps -5.70 D
increase as a function of , with Pt always increasing
faster than Ps. The difference RA increases with the
square of the off-axis angle . Near the axis, is small
and squared is even smaller. Therefore, the first The equivalent spherocylindrical lens parameters
noticeable off-axis effect is usually coma instead of are 5.70 O0.61 x 180. The power gain is best
radial astigmatism. represented by the spherical equivalent of the
When is not large, the most noticeable effect of above result or
radial astigmatism may be the power gain in both the Se = -5.70 D + (-0.61 D/2) = -6.01 D.
sagittal and tangential meridians (as opposed to the
So the professor gained an extra -0.51 D by tilting
difference, RA). The power gain is best represented his spectacles.
by the dioptric location of the circle of least confusion.
For a spherocylindrical lens, the spherical equivalent
corresponds to the dioptric location of the circle of
least confusion. When we apply the spherical equiva-
lent equations to radial astigmatism, we obtain
Se = (Ps + P t )/2, (20.29)
or
R A
o T.
(20.30)
S =P + .
Then from the second order equations, we obtain

Se = p[l + ^ - ] . (20.31) FIGURE 20.25. Power gain by tilted lenses.


Aberrations 441

With a calculator, it is just as easy to compute Ps In plus cylinder form, Ps equals the sphere power,
from the more accurate Eq. 20.23, and RA from the and so the spherocylindrical parameters are
more accurate Eq. 20.25. Then +8.75 O+3.12 x 90. In this case, the tangential line
image (i.e., the line image formed by the tangential
Pt = Ps + RA. (20.32) meridian) is vertical, and
For Example 20.4, the more accurate results are 100
-5.69 0 - 0 . 6 5 x 1 8 0 . = +8.4 cm.
' +11.87D
EXAMPLE 20.5 The sagittal line image is horizontal, and
Two distant monochromatic point sources are 100 A
horizontally side by side. The optical axis of a ^ = T8J5D=11-4cm-
+8.00 D spherical glass (n = 1.50) lens passes
through one of the sources. The other subtends a The dioptric location for the circle of least confu-
32 angle at the optical center of the lens. Give the sion is
effective spherocylindrical parameters of the lens,
and the line image locations for oblique central Se = +8.75 + ^ = +10.31D.
refraction of the light coming from the off-axis As compared to the +8.00 D power of the lens, the
source. + 10.31 D corresponds to a +2.31 D gain.
In this case, the tangential plane (which contains
the chief ray and the optical axis) is the horizontal
plane. The intersection of the sagittal plane and the People typically look straight ahead at a distant
lens falls on the vertical meridian. Then from Eq. object, and drop their line of sight when reading
20.23, (Figure 20.26a). For a spectacle lens with no tilt, this
results in the presence of radial astigmatism while
reading. However, spectacle frames usually have a
PB = + 8 .OOD[I + ^ ] ,
pantoscopic tilt (Figure 20.26b). The pantoscopic tilt is
Ps = +8.00 D(1.094) = +8.75 D. equivalent to a rotation around a horizontal meridian.
From Eq. 20.25, Besides a cosmetic improvement, the pantoscopic tilt
minimizes the radial astigmatism while reading (at the
RA=+8.00Dtan 2 32, expense of leaving some radial astigmatism at dis-
RA= +8.00 D(0.391) = +3.12 D. tance). The typical pantoscopic tilt is about 7.5.
Equation 20.32 gives the tangential power Some spectacle frames also have a face-form tilt,
Pt = Ps + RA, which is equivalent to rotating the lens around a
vertical meridian (Figure 20.27). As a result of the
Pt = +8.75 D + 3.12 D = +11.87 D. tilts, strong prescriptions may behave differently in the
The effective power cross is spectacle frame than during the visual exam when the
tilts are not present.
+8.75 D When a plus spherical lens is tilted around a meridian
and the effective spherocylindrical parameters S c C x
are expressed in plus cylinder form, then S is the sagittal
+11.87D
power Ps and is the meridian around which the tilt
occurred. When a minus spherical lens is tilted around a

straight ahead

,*-*< pantoscopic tilt side view

straight ahead
0
*e' FIGURE 20.26. Lines of sight for distance and reading, a. No
b) tilt. b. Pantoscopic tilt.
442 Geometrie, Physical, and Visual Optics

FIGURE 20.27. Face form tilt.

meridian and the effective spherocylindrical parameters scopic tilt. What are the effective spherocylindrical
S C C x are expressed in minus cylinder form, then parameters of the lens for an object point located
again S is the sagittal power Ps and is the meridian straight ahead?
around which the tilt occurred. We can use the standard Figure 20.26b shows the geometry. Here the
transposition relations to go back and forth between the tangential plane is the vertical plane. The horizon-
plus and minus cylinder expressions. tal meridian contains the intersection of the sagittal
plane and the lens. Then from Eq. 20.23,

EXAMPLE 20.6 Ps = +15.75 D(l + 0.017) = +16.02 D.


A -9.50 D highlite (n = 1.71) lens has a 14 degree From Eq. 20.25,
face-form tilt. What are the effective spherocylin-
drical parameters of the lens for an object point R A = +15.75 D tan 2 13,
located straight ahead? RA = +15.75 D(0.053) = +0.84 D.
Figure 20.27 shows the geometry. Here the
So the effective spherocylindrical parameters are
tangential plane is the horizontal plane. The verti-
+ 16.02O+0.84X 180. By transposition, the effec-
cal meridian contains the intersection of the sagittal
tive parameters are +16.86 O -0.84 x 90.
plane and the lens. Then from Eq. 20.23,

'.'>[' + w] Consider an on-axis point source in front of a plus


lens-aperture combination, and assume negligible
Ps = -9.50 D(l + 0.017) = -9.66 D. spherical aberration. When the lens is rotated, both
From Eq. 20.25, coma and radial astigmatism are present. Thus, at
each line image position, a coma flare is detectable
RA=-9.50Dtan214,
(Figure 20.28a). By decreasing the aperture size, the
RA = -9.50 D(0.062) = -0.59 D. coma flare can be eliminated without changing the
The effective spherocylindrical parameters are position of the line images. In other words, the inter-
-9.66 0 - 0 . 5 9 x 9 0 . val of Sturm caused by radial astigmatism is un-
For completeness, Eq. 20.32 gives the effective changed when the aperture is made smaller. (This
tangential power contrasts with the behavior of longitudinal spherical
Pt = -9.66 D + (-0.59 D) = -10.25 D.
The effective power cross is then

.-9.66D

10.25 D
a)

EXAMPLE 20.7 b)
Carol Strawberry married Tom Jam. Now Mrs.
Strawberry-Jam is aphakic and wears a +15.75 D FIGURE 20.28. a. Coma flare at position of radial astigmatism
CR-39 (n = 1.49) spectacle lens with a 13 panto- line image, b. Curved tangential image.
Aberrations 443

object tangential image


a)

B
\
AI A

tangential image sagittal image FIGURE 20.29. a. Spoked wheel object together
object
with tangential and sagittal image, b. Four points
b) together with their tangential and sagittal images.

aberration.) If we continue to decrease the aperture ellipse). The shallower form is the starting point for
size, we run into the pinhole effect. Here the length of the design of corrected curve spectacle lenses. Note
the line images decreases, and, consequently, the blur that when near objects are taken into account, the
due to radial astigmatism decreases. designs are flatter than for distant objects only. Also
Again, consider an object point vertically below the note that for high plus lenses (high hyperopes or
optical axis. The tangential meridian is the vertical aphakics), there is not a spherical lens that gives zero
meridian. As we rotate the object point in a circle radial astigmatism. This has led to the use of aspheric
around the axis, the tangential and sagittal meridians lenses for high plus corrections.
rotate with the point. As a result, the tangential line In systems with plus and minus lenses, the radial
image is really a curved image (part of a circle). Figure astigmatism from the minus lens acts to correct the
20.28b shows the tangential image for an object point
vertically below the axis. The sagittal line image is
always perpendicular to the tangential line image. For
a large spoked wheel target, the effect is that the
wheel rim is clear in the tangential image plane, and
the spokes are clear in the sagittal image plane (Figure
20.29a). Figure 20.29b isolates three object points
(marked A, B, and C) on the wheel, and shows the
corresponding tangential and sagittal images.
For oblique central refraction, radial astigmatism is
not very sensitive to the form of the lens. However, + 20 +
radial astigmatism through a peripheral part of the
lens, as shown in Figure 20.30a, is dependent on the
form of the lens. Figure 20.30b shows a plot of the
form of a spherical thin lens that gives zero radial
astigmatism for one off-axis angle. The plot is for front
surface power vs back vertex power. The curves turn
out to be elliptical and are referred to as the Tscher-
ning ellipses. The inside (smaller) ellipse is for a
distant object, while the outside (larger) ellipse is for a
near object. i i

For a distant object, lens forms that give zero radial Back Vertex Power
astigmatism exist for back vertex powers ranging from
about - 2 2 D to + 8 D . For most back vertex powers, b)
there are two forms that work: a steep form (top of FIGURE 20.30. a. Radial astigmatism through peripheral por-
the ellipse) and a shallower form (bottom of the tion of the lens. b. Tscherning ellipses for distance and near.
444 Geometrie, Physical, and Visual Optics

radial astigmatism from the plus lenses. Therefore, 20 cm from the center of curvature C. Since the
besides lens form, we can use doublets and systems of central rays in each pencil pass through C, the surface
separated plus and minus lenses to control radial specified by all the image points is a sphere of radius
astigmatism. 20 cm centered on C. If we repeat the same argument
for parallel rays in the glass incident on the same
interface, we would again get a spherical image sur-
face, except that now its radius would be 30cm (i.e.,
20.15 Curvature of Field the secondary focal length for this case is 20 cm, and
then we add 10 cm for the interface's radius of cur-
If the previous Seidel aberrations are corrected, then vature).
an off-axis monochromatic point source would give a A lens has two spherical surfaces, and each contri-
perfect point image (neglecting diffraction). However, butes to the curvature of field. For a distant object the
for a specified object plane, the conjugate image curvature K of the image surface for a thin lens in air
points usually do not lie in a plane. This is the Seidel is given to second order by
aberration called curvature of field. In the presence of
curvature of field, an extended image on a flat screen =-,
gets progressively blurred as a function of distance n
from the optical axis. where P is the power of the lens and n is the refractive
We can get an intuitive feeling for curvature of field index. From the equation, the higher the power P (in
by considering a +5.00 D single spherical refracting magnitude) the greater the image surface curvature K.
interface between air and glass (n = 1.50). Figure For the same power, higher index lenses have less field
20.31 shows three different narrow pencils of rays curvature than lower index lenses.
coming from the distant object. In each pencil, the For a distant object, the curvature K of the image
central ray points toward the interface's center of surface for a series of separated thin lenses in air is
curvature C and so does not bend. Consider the pencil given by
traveling straight ahead. The rays are refracted and
form an image at the secondary focal point. The = . (20.33)
secondary focal length is j n
j
Equation 20.33 is called the Petzval equation (after
Josef Petzval, 1807-1891, who first discovered it).
Positive lenses introduce an inward curvature in the
So f2 is 30 cm from the vertex of the interface, or field (concave relative to the back vertex of the sys-
20 cm from C. tem), and minus lenses introduce an outward curva-
The interface has exactly the same geometry for the ture in the field. The image surface is flat when K
other two pencils. The central ray in each pencil equals zero (the Petzval condition). Petzval used this
passes straight through without bending, and the other condition to design the best portrait lens system of his
rays converge on the central ray to form a real point time.
image at a distance of +30.00 cm from the point where It is interesting to note that for a plus system, a
the pencil's central ray intersects the interface, or minus lens introduced very close to the image plane

curved
image surface

F2

FIGURE 20.31. Curvature of field for a convex


single refracting interface.
Aberrations 445

makes a full contribution to the Petzval condition for a the radial astigmatism equations of the last section.)
flatter field, and yet makes very little contribution to There is a mathematical relation that says the tangen-
the other aberrations. Another point of interest is that tial image surface is always farther from the Petzval
two separated lenses of the same material, one with a surface than the sagittal image plane. However, when
power of +P and the other with a power - P , give a radial astigmatism is present, the best image surface is
flat field (K = 0). For example, a +10.00 D ophthal- specified by the circles of least confusion in the con-
mic crown lens 5 cm in front of a 10.00 D ophthalmic oids and not by the Petzval surface.
crown lens has a flat field, a +5.00 D equivalent Figure 20.32a shows the image surfaces for a dis-
power, and a +10.00 D back vertex power. tant object in front of an equiconvex lens. The tangen-
Curvature of field and radial astigmatism are inti- tial image surface is curved more than the sagittal
mately related, although we can have one without the image surface. Since these surfaces are rotationally
other. In the absence of radial astigmatism, the best symmetric about the optical axis, this diagram is re-
image surface is called the Petzval surface. In the ferred to as the teacup (tangential) and saucer (sagit-
presence of radial astigmatism and curvature of field, tal) diagram. From the mathematical relationship, the
the tangential image plane and the sagittal image Petzval surface is then the plate under the teacup and
plane are both curved. (The curvature of the tangen- saucer.
tial and sagittal image planes was already implicit in For zero radial astigmatism, the T, S, and P sur-

FIGURE 20.32. Tangential, sagittal, and Petzval sur-


faces. a. Equiconvex lens. b. Zero radial astigmatism, c.
Zero curvature of field, d. Anastigmat projector system.
446 Geometrie, Physical, and Visual Optics

faces all coincide (Figure 20.32b). We say that such a 20.34a shows an object in front of a thin plus lens with
system is corrected for radial astigmatism but not for an aperture at the secondary focal point (a telecentric
curvature of field. In the presence of radial astigmat- stop). The chief ray for the peripheral object point is
ism, we say that curvature of field is corrected when incident parallel to the axis. Thus, the object distance
the surface specified by the circles of least confusion is for the chief ray equals the paraxial object distance.
flat (Figure 20.32c). The image distance for the chief ray is the diagonal
Figure 20.32d shows the tangential and sagittal distance g, which is greater than the paraxial image
image surfaces for an anastigmat projector system. distance v. The lateral magnification mper for the
The anastigmat has a flat field and zero radial astig- off-axis point is equal to g/w, and
matism at one off-axis angle a. The radial astigmatism
between a and the axis is minimal; however, beyond g > V
alpha it quickly gets bad. (In actual practice, better u u
results are obtained with the node slightly in front of where m is the paraxial lateral magnification. Since the
the image plane.) peripheral magnification is greater than the paraxial
Radial astigmatism and curvature of field are the magnification, the outer part of the image is magnified
two most important aberrations to control in the more, resulting in pincushion distortion.
design of corrected curve spectacle lenses. (The third The same argument can be repeated for a stop
most important is lateral chromatic aberration.) When located in front of the lens at the primary focal point
curvature of field is taken into consideration as well as (Figure 20.34b). In that case, the chief ray emerges
radial astigmatism, the lenses tend to be flatter than parallel to the axis, and the peripheral image distance
those given by the Tscherning ellipse. is equal to the paraxial image distance. The diagonal
object distance for the chief ray is greater than the
paraxial object distance, so the peripheral lateral mag-
nification is less than the paraxial lateral magnifica-
tion, resulting in barrel distortion. When the thin lens
20.16 Distortion is by itself (no other aperture), there is no distortion.
Similarly, a pinhole by itself forms a distortion-free
Distortion is the last Seidel aberration. Distortion image, and, consequently, pinhole cameras with long
differs from the other aberrations in that it does not exposure times are sometimes used for wide field
cause any blur or loss of resolution. Distortion is due photography of stationary objects such as buildings or
to the fact that the lateral magnification for off-axis landscapes.
points depends on the exact image and object dis- When we view an object through a lens, our own
tances. Distortion has the result that any straight line pupil acts as the aperture stop of the system. An
in the object plane is imaged as a straight line only object viewed through a minus lens shows barrel
when the object line passes through the optical axis. distortion. In Figure 20.34c, the aperture behind the
All other lines are curved (Figure 20.33). lens simulates the pupil of the observer's eye. Here
For a grid target in front of a spherical lens, there both the peripheral image and object distances are
are two types of distortion: pincushion (in which the greater in magnitude than the corresponding paraxial
lines are convex relative to the axial point) and barrel distances. However, the object distance difference is
(in which the lines are concave relative to the axial greater than the image distance difference, so the
point). Pincushion distortion results when the per- peripheral lateral magnification is less than the paraxi-
ipheral lateral magnification is greater (in magnitude) al magnification, and the result is barrel distortion.
than the paraxial lateral magnification. Barrel distor- One way to correct distortion is to design a sym-
tion results when the peripheral lateral magnification metric system so that the barrel distortion introduced
is less than the paraxial lateral magnification. by one part of the system is counteracted by the
In general, distortion depends on lens form, and is pincushion distortion from the other part. A simple
extremely sensitive to the aperture location. Figure example is the orthoscopic lens system in which an
aperture is located midway between two lenses of the
same power (Figure 20.34d).
Distortion in asymmetric systems can be minimized
by making the system satisfy a geometric tangent
condition. However, the wider the field of a lens, the
harder it is to correct for distortion. Photographs
taken with the very wide field "fish eye" lenses show
FIGURE 20.33. Barrel and pincushion distortion. distortion.
Aberrations 447

/ s>
y\ >

\V F >
object - * 7 _

< ^ ^ image
V
a) u
>f

> K

1
FIGURE 20.34. a. Chief ray for pincushion
T distortion, b. Chief ray for barrel distortion, c.
Chief ray for aperture simulating observer's
4 'Y \1 entrance pupil, d. Orthoscopic system.

20.17 Lens Design about f/16. Petzval's portrait lens, also about 1840,
was the first designed lens. It could be used at f/3.5
Obviously, all the aberrations occur simultaneously. For provided the field of view was not large (Figure
large angles, fifth and larger order aberrations also come 20.35b).
into play. Furthermore, the design that improves one of As mentioned, a symmetric lens system is free of
the aberrations frequently makes another worse. There- distortion. In fact, a perfectly symmetrical lens system
fore, lens design is a complicated process. As an is also free of lateral chromatic aberration and free of
example, let us briefly consider how camera lens design coma for one zone. In 1866, Dallmeyer developed a
has progressed. symmetric system called the rapid rectilinear lens,
Figure 20.35a shows a simple meniscus lens design which could be used at f/8 with a field as wide as 50
developed by Wollaston about 1812. Wollaston did not (Figure 20.35c). A number of modern systems, such as
have any coherent theory to guide him. Instead, he the Zeiss planar, are symmetric systems operating as
tried different lenses and different forms until he fast as f/1.4 (Figure 20.35d).
empirically found one that worked well. Around 1840, According to third order theory, a designer has to
Chevalier replaced Wollaston's meniscus lens with an control eight variables: the system power, the five
achromatic doublet. Chevalier's lens could be used at Seidel aberrations, and longitudinal and lateral
448 Geometrie, Physical, and Visual Optics

-^ + >+

a)

-*> + *- +

-> +-

FIGURE 20.35. Examples of camera


lens system design, a. Wollaston. b.
Petzval portrait lens. c. Rapid re-
ctilinear. d. Zeiss planar, e. Cooke trip-
let. / . Zeiss Tessar.

chromatic aberration. Hence, a good lens system has ple element systems. One product of computer aided
to have at least eight free variables in the design. The design is modern zoom television cameras. These
Cooke triplet lens, first designed by Taylor in 1893, is zoom systems have as many as 20 elements in them.
an elegant design that uses just eight variables (Figure You might consider the image quality the next time
20.35e). These include the three lens materials, the you watch your favorite TV program.
form of each lens, and the two spacings between the
lenses. The original Cooke triplet could operate at f/4.
Modern variations of the Cooke triplet, such as the
Zeiss Tessar, can operate as fast as f/1.9 (Figure 20.18 Spherocylindrical Lenses
20.35f).
In the 1930s lens design work got a major boost The previous sections applied to systems of spherical
from the introduction of antireflecting coatings. The lenses. Ophthalmic optics is complicated by the use of
antireflecting coatings enabled people to use more spherocylindrical lenses with toric surfaces to correct
air-spaced lenses without having to worry about inter- astigmatism in the human eye. Such spherocylindrical
nal reflections. The very recent development of new lenses have all the previous aberrations plus some
high index glass types has enabled people to achieve additional coma-type aberrations that are not present
the same quality systems with fewer lenses, e.g., a in spherical systems.
Tessar style vs a Planar style. Finally, the growing As mentioned, the two most important aberrations
power and availability of computers has enabled peo- to control in spectacle lens design are radial astigmat-
ple to quickly calculate exact ray traces through multi- ism and curvature of field. The oblique central refrac-
Aberrations 449

tion equations are easily modified for a spherocylindri- Chapter 10 where we used the paraxial approximation
cal lens tilted (or effectively tilted) around a principal to derive the power equations of a mirror. Of the
meridian. The actual spherocylindrical power in the Seidel aberrations for a mirror, distortion is particular-
sagittal meridian is used to compute the effective ly easy to observe in a convex mirror (such as those
sagittal power Ps, while the actual spherocylindrical used for observation in stores).
power in the tangential meridian is used to compute The Cartesian oval was the surface that for a
the effective tangential power P t . conjugate object and image point gives no spherical
aberration. For a mirror, the Cartesian ovals are the
EXAMPLE 20.8 second order conic surfaces (ellipses, parabolas, and
A +14.000-6.00x90 lens (n = 1.5) has an 0.2 hyberbolas). For an ellipse, the conjugate points free
radian pantoscopic tilt (i.e., 11.5). To second of spherical aberration occur at the two focii of the
order, what are the effective spherocylindrical ellipse (both near points). For a parabola, the conju-
parameters for viewing an object straight ahead? gate points free of spherical aberration are optical
The power cross for the lens is infinity and the focus of the parabola. (One practical
application is the use of parabolic mirrors in auto-
+ 14.( mobile headlights.)

Problems

The tangential meridian is the vertical meridian. 1. What is the chromatic aberration for a +7.00 D
From Eq. 20.26, ophthalmic crown (523-586) lens? Compare this to
the chromatic aberration for a +7.00 D polycarbo-
Pt = + 14.00D[l + ] , nate (580-300) lens.
2. What is the chromatic aberration for a -6.00 D
Pt = +14.00 0(1.053) = +14.75 D. CR-39 (495-580) lens? Compare this to the
The sagittal meridian is the horizontal meridian. chromatic aberration for a -6.00 D highlite (701-
From Eq. 20.27, 310) lens.
3. When a human eye is emmetropic for sodium D
Ps=+8.00D[l ^ f ] , light, is it myopic or hyperopic for blue light (F
+
line)? For red light (C line)?
Ps = +8.00 D(1.013) = +8.11 D. 4. Design a +7.00 D achromatic doublet using oph-
thalmic crown glass (523-586) and highlite glass
The effective power cross is
(701-310).
5. Design a -6.00 D achromatic doublet using poly-
+ 14. carbonate (580-300) and CR-39 (495-580).
6. Design a 5 BU achromatic prism using ophthalmic
crown (523-586) and dense flint glass (617-366).
7. In view of radial astigmatism, what are the effective
spherocylindrical parameters for a +7.00 D lens
with a 12 pantoscopic tilt? (The object point is
The effective spherocylindrical parameters are straight ahead.) In terms of a spherical equivalent,
+ 14.75 0 - 6 . 6 4 x 9 0 . did the lens gain or lose power?
8. In view of radial astigmatism, what are the effective
spherocylindrical parameters for a +9.00 D lens
that is rotated 9 around a vertical axis? (The
20.19 Mirrors object point is straight ahead.) In terms of a spheri-
cal equivalent, did the lens gain or lose power?
Front surface mirrors are free of chromatic aberration. 9. A -8.00 O - 4 . 0 0 x 180 lens has a 10 pantoscopic
However, front surface mirrors do suffer from the tilt. To second order, what are the effective sphero-
monochromatic aberrations of form, including the cylindrical parameters for viewing an object straight
Seidel aberrations. The first clue to this occurred in ahead?
CHAPTER
TWENTY-ONE
Waves and
Superposition

21.1 Introduction many other parts of the electromagnetic spectrum.


From long to short wavelengths some of these are:
In the eighteenth century, people thought that elec- radio waves, television waves, radar waves, mi-
tricity, magnetism, and light were independent phys- crowaves, infrared radiation, visible light, ultraviolet
ical entities, but in 1820, Hans Christian Oersted radiation, x-rays, and gamma rays.
discovered that an electrical current induces a mag-
netic field. (Specifically, an electrical current causes a
compass needle to turn). In 1830, Michael Faraday
added to the relationship by discovering that a chang- 21.2 Basic Wave Properties
ing magnetic field induces an electric field. Today we
know that electricity and magnetism are different In order to discuss the wave properties of light, let us
manifestations of the same basic force called the elec- first review some elementary wave concepts. In gener-
tromagnetic force. al, a wave is a disturbance or change that propagates
Faraday also made another curious discovery. He through the available space or medium. Three charac-
found out that a magnetic field rotates the polarization teristics of low energy wave motion are: energy is
plane of a light beam that is propagating in a medium. transmitted, the medium is not transmitted, and an
This was the first apparent connection between light oscillation with a return to equilibrium is involved.
and the electromagnetic force. Subsequently, in 1886 In a high energy wave propagation, the return to
James Clerk Maxwell summarized the then known equilibrium does not necessarily occur. For example,
electromagnetic phenomena in four equations. With tapping on your desk with your finger generates low
great physical intuition, Maxwell decided that a term energy sound waves traveling through the desk. Ignit-
was missing in one of the equations. After introducing ing a stick of dynamite on the desk initially sends very
the missing term, he derived the fact that accelerating high energy sound waves, i.e., shock waves, through
electrically charged particles should radiate energy, the desk. On about the third shock wave through, the
and this energy propagates through space according to material begins to tear apart, and there is no longer a
wave equations. Furthermore, the calculated propaga- return to equilibrium. In regards to light, typical
tion speed of the electromagnetic waves turned out to everyday light waves are low energy waves. High
be the speed of light. So Maxwell proposed that energy light waves can be generated by modern lasers,
electromagnetic radiation exists and that light is part but those effects are outside the goals of this book.
of the electromagnetic radiation spectrum. For an example of the three characteristics of low
A few years later, Henrich Hertz generated the first energy wave motion, consider a pulse propagating
man-made electromagnetic waves. Hertz waves were down a rope (Figure 21.1). Point X on the rope,
in the region of the electromagnetic spectrum that we which for conceptual purposes can be marked with red
now call radio waves. Today we are familiar with dye, is at the equilibrium position in nos. 1 and 2.

451
452 Geometrie, Physical, and Visual Optics

When the pulse reaches point X, that part of the rope


r~^. gains energy and moves up (nos. 3 through 5). As the
pulse moves on, point X loses energy and returns to
3 the equilibrium position (nos. 5 through 9). Eventual-
ly, point X is back in its original position (nos. 10
4
through 12).
5 We refer to the pulse as a transverse wave. A trans-
6 verse wave is one in which the oscillation is perpen-
dicular to the direction of propagation. The pulse
7
oscillation was vertical (up and down), while the prop-
8 agation direction was horizontal (to the right).
9 Another type of wave is a longitudinal wave in
which the direction of oscillation is in the same plane
|10 as the direction of propagation. Figure 21.2 shows a
11 stretched slinky. When the slinky is pushed in (right)
12 and then pulled back (left), a longitudinal pulse prop-
agates to the right along the slinky. A point on the
FIGURE 21.1. Transverse pulse traveling down a rope.
slinky would first be at equilibrium, then oscillate right
and left as the pulse passes through.
Sound waves are longitudinal waves. Figure 21.3
shows a sound wave generated in a tube. A compres-
sion pulse traveling right is generated by the piston.
The compression is followed by a rarefaction (a less
dense region). The rarefaction is then followed by
another compression. A stationary air molecule oscil-
lates left and right as each pulse passes through.
When a stone is dropped into a pond, a series of
ripples or transverse water waves move outward on
the surface (Figure 21.4). Each water molecule vib-
FIGURE 21.2. Longitudinal waves in a slinky. rates up and down as the ripples pass through. Light
waves are transverse waves similar to the ripples
except that they propagate in three dimensions
(whereas the ripples only propagate along the two-
auiiiiiiiiiiiiiii mum dimensional surface of the water).
Each double-headed arrow in Figure 21.5a repre-
--^lllllllllllllllllllllllllllllllll sents a possible vibration state for a three-dimensional
transverse wave that is propagating perpendicular to
=HIII1IIIIIII1IIIIIIII1I the plane of the paper. One vibration state is vertical,
one is horizontal, and two are oblique. We can resolve
-III I I a vibration state into two perpendicular components
just as we can resolve a vector into two perpendicular
components. Figure 21.5b shows the horizontal and
FIGURE 21.3. Longitudinal sound waves. vertical components of an oblique vibration state. The

FIGURE 21.4. Top and side views of trans-


Top view verse waves generated by a point source.
Waves and Superposition 453

"7!
/
/
/ y components
/
>- vector /
/
L_
a) Vibration state Components

b) and ^-

FIGURE 21.5. a. Some polarization states for transverse waves propagating perpendicular to
the plane of the paper, b. Resolution of polarized light into perpendicular components.

phenomenon of separating a transverse wave into y = Asinp, (21.1)


components is called polarization. Light waves are
transverse waves and can be polarized. Sound waves where y is the displacement, A is the amplitude or
are longitudinal waves and cannot be polarized. maximum displacement, and p is called the phase. For
a harmonic wave, Figure 21.6b shows a plot of the
displacement y as a function of the phase p.
A harmonic wave results when the restoring force
21.3 Harmonic Waves (the force that causes the return to equilibrium) is a
linear function of the displacement y, i.e.,
When a pulse is regularly repeated, a periodic wave is F = ky,
formed. Figure 21.6a shows some periodic waveforms.
A harmonic wave is a periodic wave that can be where F is the restoring force and k is the constant of
mathematically described by a single sine function: proportionality. In low energy situations, many restor-

LrLTLn-TLrL
Spatial
period

oVAWXA
a)
U
Spatial
period

FIGURE 21.6. a. Periodic nonhar-


monic waves, b. Harmonic wave.
454 Geometrie, Physical, and Visual Optics

ing forces satisfy this condition. Therefore, harmonic


waves are ubiquitous.
In dealing with harmonic waves, it is frequently
convenient to talk about the crests or the troughs.
However, we can use the phase p to pick out any point
on the wave that we wish to discuss. From Eq. 21.1, FIGURE 21.7. Double exposure of a harmonic wave moving
the phase of a crest is 90, i.e., right.
y = Asin90= +A.
(In radians, the phase of the crest is 77/2.) The sine of wave. During this time, each water molecule makes
270 is - 1 . Therefore, one complete up-down cycle or vibration. The time
y = Asin270= - A , required for a complete vibration is called the (tem-
poral) period T of the wave. The number of cycles per
and we see that 270 is the phase of a trough. The unit time is called the (temporal) frequency f of the
phase difference between a crest and a trough is 180 wave, and
(i.e., 270-90).
f=l/T. (21.2)
Each of the other parts of the wave has a specific
phase value. For example, consider a wave with an For example, a harmonic water wave with a period of
amplitude of 4 units. At a phase of 30 the displace- \ second (s), has a frequency of 4 cycles/s.
ment is: Since the distance the wave moves during a period
y = 4sin30 = 4(+0.5) = +2. T is just one wavelength, we have the following
relation:
Zero displacement occurs when the phase is 0, 180, distance = (speed)(time),
or 360 degrees. The phase difference between a zero
and the nearest crest or trough is 90. or
A wavefront is a curve on a wave connecting points = .
of equal phase. It is easy to think of the crests (phase
90) as wavefronts (Figure 21.4). However, we could We can solve for the speed v, and substitute from Eq.
connect all points on the wave with phases of 63 (or 21.2 to obtain
any other value), and use these curves as the wave- Af = v. (21.3)
fronts.
For a periodic wave, the spatial period is the length Equation 21.3 is a very important relation between the
of each repeating cycle. The spatial period of a har- wavelength, frequency, and speed of a wave. In par-
monic wave is called a wavelength, and is represented ticular, for a fixed speed, Eq. 21.3 says that the
by the greek letter (lambda). In Figure 21.6b, the wavelength and frequency are inversely related.
distance between A and B equals the wavelength, as
does the distance between B and C. Similarly, the
EXAMPLE 21.1
distance between two adjacent crests equals the wave- A harmonic wave is traveling 10 cm/s and has a
length. In fact, the distance between any two points temporal frequency of 4 cycles/s. What is the
that have a 360 degree phase difference equals one wavelength? From Eq. 21.3,
wavelength. So one wavelength corresponds to a
phase difference of 360 degrees. Half of a wavelength = (10cm/s)/(4 cycles/s) = 2.5 cm/cycle.
corresponds to a phase difference of 180 degrees. Usually we just write
For conceptual simplicity, let us consider a water = 2.5 cm.
wave or ripple propagating to the right. As the wave
moves right, each water molecule on the surface vib-
rates up and down. Figure 21.7 shows a double expo-
sure picture of the wave. The solid curve indicates the 21.4 Intensity
wave at time t1? while the dashed curve indicates the
wave at the later time t2. The water molecule at A The intensity I of a wave is the energy carried or
rises during the time t2 t l5 while the water molecule delivered by the wave per unit area per unit time, i.e.,
at B falls. for a small area and a small time.
If the double exposure is taken at times tj and t2 I = (energy delivered)/[(area)(time)]. (21.4)
during which the wave traveled one wavelength, the
resulting side view is identical to that of the initial The intensity I of a harmonic wave is proportional to
Waves and Superposition 455

the square of the wave's amplitude: In general wave theory, the word intensity refers to
IocA . 2
(21.5) the energy delivered by the wave per unit area and per
unit time. By international convention, when dealing
Equation 21.5 follows from basic physics. Consider with electromagnetic waves, the word intensity is re-
a water wave. A water molecule on the surface vib- placed by the word irradiance. From Eq. 21.5, the
rates up and down as the wave comes through. The irradiance is proportional to the square of the am-
kinetic energy that the wave delivers to the water plitude of the electromagnetic wave. When the ir-
molecule equals 1/2 m (v vib ) 2 where m is the molecule radiance is evaluated according to its ability to pro-
mass and vvib is the vibration velocity. The vibration duce a brightness response in a human observer, we
velocity equals the derivative of the displacement y refer to it as the illuminance.
with respect to the time t, or
vvib = dy/dt.
From Eq. 21.1, 21.5 Frequency Invariance and
vvib = d(Asinp)/dt, Wavelength Changes
vvib = Ad(sinp)/dt, The visible region of the electromagnetic spectrum is a
vVib = A cos p. narrow band centered around a vacuum wavelength of
550 nm. One nanometer (nm) equals 10~ 9 m. Another
Since the kinetic energy is proportional to the square unit that has been used for wavelength is the angstrom
of the vibration velocity, the kinetic energy delivered (), which equals 10~10m (or 0.1 nm). In older
to the molecule is proportional to (A cos p)2. books, you might also run across the millimicron
The potential energy of a particle displaced a dis- (), which equals the nanometer.
tance y by a wave is proportional to 1/2 ky2 where k is As discussed in Section 1.7, light of a single wave-
the proportionality constant in the restoring force. For length typically stimulates a single color response. The
an incident harmonic wave, the potential energy is color-wavelength relations given in Table 21.1 are
then proportional to (a sin p)2. repeated here.
The total energy delivered to the molecule is the
sum of the kinetic energy plus the potential energy.
Then
TABLE 21.1
I oc (A cos p) 2 + (A sin p)2, Color response vs vacuum wavelength
I oc A2(cos2 p -I- sin2 p).
Vacuum Wavelength
Since cos2 p + sin2 p = 1, it follows that Color Response (nm)
IocA2. Red 780-620
For our purposes, we can choose the units to make the Orange 620-590
Yellow 590-560
proportionality constant equal to 1. Then Green 560-490
I = A2. Blue 490-450
Indigo 450-430
Frequently, only relative intensities are considered, in Violet 430-380
which case
Ire = I 2 /Ii = (A 2 ) 2 /(A 1 ) 2 . (21.6)
For electromagnetic radiation, the unit for fre-
EXAMPLE 21.2 quency is called the hertz (Hz). Visible light is a
A water wave has an amplitude of 2, and a second narrow part of the electromagnetic spectrum centered
water wave has an amplitude of 6. What is the on a frequency of about 5.45 x 1014 Hz. The frequency
relative intensity of the second wave to the first? is a very important parameter because it does not
From Eq. 21.6, change when the light moves from one medium to the
Ire, = (6)2/(2)2 = 36/4 = 9. next. For example, Figure 21.8 shows a system with
plane interfaces between air, water, and glass. The
It is the intensity of a sound wave that is the light incident in air has a frequency of 5.45 x 1014 Hz.
stimulus for perceived loudness, and it is the intensity When the light enters the water its frequency is un-
of a light wave that is the stimulus for perceived changed, and similarly when the light enters the glass
brightness. its frequency is unchanged.
456 Geometrie, Physical, and Visual Optics

cones. The light is absorbed in the cones, and the


water (n = 1.33) glass (n = 1.50)
resulting neural signal ultimately produces the red
response. However, when the light is incident on the
/Ay\^/vw\A/ww\A cones, the wavelength is not 633 nm. The wavelength
has changed each time the light enters a new medium.
= 550 nm \=414nm = 367 nm In particular, the wavelength in the vitreous humor is
FIGURE 21.8. Wavelength changes as monochromatic light 633 nm

moves into different media. - = T33^ = 474nm

So for light of vacuum wavelength 633 nm incident on
the eye, the wavelength in the vitreous humor is
The invariance property for the frequency does not 474 nm. In each of the media, this light has a fre-
hold for the wavelength. The velocity v for light in quency of 4.74 x 1014 Hz. Thus, we could say that light
transparent materials is slower than it is in a vacuum, of frequency 4.74 x 1014 Hz is the stimulus for the red
and response. It is more common to say that the incident
c light of (vacuum) wavelength 633 nm is the stimulus
v= for the red response even though the wavelength in
n, the vitreous is 474 nm.
where n is the refractive index of the medium.
Now consider light of frequency f in a medium of
index n. From Eq. 21.3, EXAMPLE 21.3
Light of frequency 6.52 x 1014 Hz is incident on a
fAm = v, human eye. What is the frequency and wavelength
where Am is the wavelength in the medium. Then in the vitreous humor? What is the normal color
response?
f
The frequency is invariant, so it remains 6.52 x
^=n> 1014 Hz in all the media. The vacuum wavelength is
and
A =
fiT or
H 3xl08m/s
6.52 x 1014 Hz
0.460 x 10" m,

The vacuum wavelength equals c/f. Therefore, A = 460xl0~ 9 m = 460nm.


From Table 21.1, the normal color response is
<-i (21.7) blue. The wavelength in the vitreous humor is
In Figure 21.8, the vacuum wavelength is 550 nm, the 460 nm
wavelength in water (n = 1.33) is 414 nm, and the = 344nm.
1.336
wavelength in glass is 367 nm. Remember that the wavelength-color tables are for
Since the wavelength changes whenever the light vacuum wavelengths, not for wavelengths inside
enters a different media, it would be better to identify the eye.
monochromatic light by its frequency, which is in-
variant. However, in many physical optics experi-
ments, it is the wavelength that is directly measured,
and so we tend to identify light by its vacuum wave- 21.6 Superposition
length. (The index of refraction of air is about 1.0003,
so normally we think of the wavelength in air as being The superposition principle states that when two or
equal to the vacuum wavelength. For precision work, more low energy waves are superposed on the same
the difference must be taken into account.) space or medium, the waves travel independently
Under normal circumstances, light of vacuum through each other, and the resultant displacement at
wavelength 633 nm incident on the eye produces a red each position is the algebraic sum of the displacements
response. The frequency of the light is due to each wave.
As an example, consider two pulses traveling in
c = 3 x 10 8 m/s opposite directions down a rope. Figure 21.9a shows
= 4.74 x 1014 Hz.
~ " 633 x 10"9 m the resulting effect as a function of time. Initially, the
The frequency remains the same as the light prop- pulses are traveling toward each other but are not
agates through the cornea, the aqueous humor, the superposed. In exposure no. 2, the waves are super-
crystalline lens, the vitreous humor, and into the posed, and the resultant displacement on the rope is
Waves and Superposition 457

-^V-
^^

.r^i.
-/XL .
7

b)
FIGURE 21.9. a. Superposition of two upward pulses, b.
Superposition of an upward pulse with a downward pulse.

the sum of the individual displacements. Exposure


nos. 3 and 4 are similar. In exposure no. 5, the pulses
have passed each other, and each continues on its
way.
In Figure 21.9b, an upward and downward pulse FIGURE 21.10. a. Constructive interference, b. Destructive
are incident from opposite directions. In exposure nos. interference.
2 to 5, the resultant displacement is the sum of the
individual displacements due to each pulse. In expo-
sure no. 3, one pulse has counteracted the other and simply superposed.) When the component waves are
there is no resultant displacement. However, at the in phase and have the same amplitude, the amplitude
time of exposure no. 3, the points on the rope still of the resultant wave is twice that of either com-
have a non-zero vertical velocity. In other words, for ponent.
exposure no. 3 all the energy carried by the wave is When the crest of one component lines up with the
kinetic, while in the other exposures some of it is trough of the other component, the waves are 180 out
kinetic and some potential. In exposure no. 6, the of phase. In this case, the amplitudes subtract (Figure
pulses have passed each other and each proceeds as if 21.10b), a situation referred to as destructive interfer-
the other pulse had not been present. ence, i.e.,
When two harmonic waves of the same frequency
yr = Aj sin + A 2 sin(p1 + 180).
propagate in the same direction, the resultant dis-
placement is that of a harmonic wave of the same Since
frequency, i.e.,
sin(p! + 180) = -sin p 1?
yr = A{ sin pl + A 2 sin p 2 = A r sin p r ,
we have
The two waves are said to be in phase if their crests
yr = A, sin p 1 + A 2 [ ( - l ) sin p j ,
line up ( = p 2 ). In this case, the resultant amplitude
equals the sum of the two component amplitudes or
(Figure 21.10a), i.e.,
yr = ( A 1 - A 2 ) s i n p 1
yr = Ai sin pj + A 2 sin pl,
So
yr = ( A 1 + A 2 ) s i n p 1 ,
,^,-,), (21.9)
and so
while p r = pj.
,^,+^, (21.8)
For two equal amplitude waves that are 180 out of
while p r = p,. phase, the amplitude of the resultant wave is zero
This is referred to as constructive interference. (Be (complete destructive interference).
careful of the word interference here. Nothing has When two harmonic waves of different frequencies
interfered in a normal sense; the two waves have are superposed, the resultant waveform is periodic but
458 Geometrie, Physical, and Visual Optics

a)

FIGURE 21.11. a. Superposition of harmonic waves


of different frequency, b. Beats from superposition of
two harmonic waves of slightly different frequency.

not harmonic (Figure 21.11a). It turns out that essen- ly through each other. The resultant displacement on
tially any periodic waveform can be synthesized by the water's surface is the algebraic sum of the displace-
adding together a series of harmonic waves of differ- ments from each wave. The circular curves in Figure
ent frequencies (e.g., Fourier series in Chapter 25). In 21.12 represent the crests of each wave. There are
that sense, harmonic waves are a set of fundamental some positions where the two waves always arrive 180
waves. out of phase (e.g., a trough from one always arrives at
When we superpose two harmonic waves of slightly the same time as a crest from the other.) Destructive
different frequencies, we get the waveform shown in interference occurs at those positions resulting in a
Figure 21.11b. In those regions where the waves are node (i.e., no resultant vibration on the water
pretty much in phase (crests lined up), we get a surface). Curves connecting the nodal regions are
constructive interference effect. Since the waves are at actually hyperbolas with the point sources serving as
a slightly different frequency, the two crests drift the focal points of the hyperbolas. The hyperbolic
farther apart until we have a region where a crest and nodal curves become straight lines at points far away
a trough tend to line up. Then we get a destructive from the two sources.
interference effect. The result is the so-called beat In between the destructive interference zones,
phenomenon, where it appears that the energy carried there are regions where a wave from one source
by the waves is clustered into groups or beats. Con- always arrives in phase with a wave from the other
ceptually, the beat phenomenon begins to show how source, and constructive interference occurs. At points
light can have both a wave and a particle character. far away from the sources, the effect of the construc-
tive interference is a resultant wave that travels out-
ward in between the nodal curves.
Consider the waves incident on the plane G, which
21.7 Interference in a Ripple Tank is distant from the two sources (Figure 21.12). The
energy carried by the waves is proportional to the
Ripple tanks are great for developing physical intui- square of the amplitude. The curve in Figure 21.13 is a
tion about wave motion. Consider two vibrators plot of the resultant intensity I as a function of the
hooked together in a ripple tank. A circular wave distance x across the plane G. Because of the interfer-
travels out from each vibrator. According to the ence, the energy carried by the waves (the intensity) is
superposition principle, the waves travel independent- redistributed. At the constructive interference posi-

mmmk FIGURE 21.12. Top view of interference by


two coherent waves.
Waves and Superposition 459

FIGURE 21.13. Intensity plot for in-


terference of two coherent waves (each of
intensity I,).

FIGURE 21.14. Time exposure of standing wave.

tions, the resultant amplitude is twice that of each teger multiple of half wavelength), the two component
component wave, so the energy available is actually waves are always 180 degrees out of phase at the nodal
four times that of each component wave. (Naively, positions. Constructive interference occurs at the an-
this looks like 1 + 1 = 4.) In effect, energy is removed tinode positions, so the amplitude of the standing
from the destructive interference zones and placed in wave is twice that of the components. Each curved
the constructive interference zones. Energy is con- line in Figure 21.14 shows the standing wave at a
served in the process, but it is redistributed. particular time. The envelope for the curved lines
shows the standing wave motion over a time interval.
While energy is available to a detector at the
antinodes, no energy is available at the nodes. Stand-
21.8 Standing Waves ing waves have been set up with visible light. The
original experiment set standing waves up in a photo-
In the ripple tank, the region directly in between the graphic film about 20 wavelengths thick. Exposure
two sources has two component waves propagating in occurred in the antinode regions (A, C, E, and G in
opposite directions. The resultant superposition results Figure 21.15), but there was no exposure at the nodes
in significant effects only when the distance between (B, D, and F). So the film had exposed layers sepa-
the two sources corresponds to an integer multiple of rated by unexposed layers.
half the wavelength. Then a standing wave is formed
by the superposition. The standing wave has stationary
nodes that are separated by stationary vibration re-
gions (the antinodes). 21.9 Coherence
The stationary nodes are formed by destructive
interference between the two component waves, i.e., Whenever two monochromatic light sources vibrate
when the two sources have the right separation (in- together, ripple tank-type interference patterns result.
More precisely, the phase difference between the two
sources must remain constant with time, i.e.,
incident light (Phase of S J - (Phase of S2) = d (at all times).
reflected light
Note that the phase of Sj could abruptly jump, but if
the phase of S2 makes the same jump, then the
H 1 condition is still satisfied. When the condition is satis-
II [ I H I fied, the sources and the resulting waves are said to be
A B C DE F G mutually coherent.
FIGURE 21.15. Standing wave exposure pattern. At the other extreme, suppose the phase difference
460 Geometrie, Physical, and Visual Optics

between the two sources undergoes abrupt changes Suppose the waves are incoherent. Then the phase
with time: difference fluctuates as a function of time. The aver-
age irradiance delivered to a detector is
(Phase of S J - (Phase of S2) = d1 (from 0 to t j ,
(Phase of Si) - (Phase of S2) = d2 (from tx to t 2 ), (,). = (A,) 2 + (A 2 ) 2 + 2A,A2[cos( Pl - p 2 )] avg .
(Phase of S,) - (Phase of S2) = d3 (from t2 to t 3 ),
The cosine is sometimes positive (up to H-1) and
etc. When the phase changes are completely random, sometimes negative (down to - 1 ) . If the phase differ-
then the sources and waves are said to be mutually ences are random and the average is done over a long
incoherent. For incoherent waves, the energy deli- enough period of time, then the cosine term averages
vered to a particular area changes with time. During to zero (i.e., every plus value is ultimately canceled by
one time period there may be destructive interference, a minus value). Then
while during another time period there may be con-
structive interference. (I r ) avg = (A,) 2 + (A 2 ) 2 ,
For two harmonic waves superposed together, the
resultant displacement is the sum of the individual or
displacements (U.vg = I i + l 2 (21-12)
yr = Aj sin + A 2 sin p 2 = A r sin p r . Equation 21.12 says that for incoherent sources, the
average irradiance distribution is just the sum of the
The resulting irradiance (intensity) I r at a point is irradiances of each individual source. This is our
proportional to (Ar)2. It follows from the above that typical everyday experience since our common light
Ir = (A,) 2 + (A 2 ) 2 + 2A 1 A 2 cos( Pl - p 2 ). sources are highly incoherent. Specifically, they usual-
ly change phase much faster than 10~8 seconds, and
The first two terms on the right are just the irradiances the human visual system cannot detect changes faster
that would be achieved by the respective waves in- than about 1/20 second. (In other words, the human
dividually. The third term is called the interference visual system averages over about a 1/20 second
term. period.)
To investigate the interference term, let us first
consider the case of waves in phase. Then p{ equals p 2
and the cosine is +1. Thus, EXAMPLE 21.4
2 2 A wave has an irradiance of 25 units, and another
Ir = (A,) + (A 2 ) +2A 1 A 2 , wave has an irradiance of 9 units, (a) What is the
or resulting (average) irradiance when the waves are
incoherent? (b), What is the resulting irradiance
Ir = (A,+A 2 ) 2 . (21.10) when the waves are coherent and 180 degrees out
of phase? (c), What is the resulting irradiance when
Since I r equals the square A r , the above equation is the waves are coherent and in phase?
just the typical constructive interference result that the For (a), the irradiances add, and
amplitudes add. For the additional condition that (I r ) avg = 25 + 9 = 34.
=
A2 Al9
For (b), the amplitudes subtract. The wave with
Ir = (2A,) 2 =4I,, irradiance 25 has an amplitude of 5, while the wave
with irradiance 9 has an amplitude of 3. Then
where is the irradiance of each component wave by Ar = 5 - 3 = 2,
itself.
When there is a 180 degree phase difference, the and
cosine equals 1. Then Ir = (A r ) 2 = 22 = 4.
For (c), the amplitudes add.
Ir = (A,) 2 + ( A 2 ) 2 - 2 A 1 A 2 , Ar = 5 + 3 = 8,
or and
Ir = ( A , - A 2 ) 2 . (21.11) I r =(A r ) 2 = 82 = 64.
Note that from common experience with inco-
This is the typical destructive interference result in herent sources, we expect a result of 34 units. At
that the amplitudes subtract. When we have the addi- the constructive interference positions, we get 64
tional condition that the amplitudes are equal, the units. At the destructive interference positions, we
resultant irradiance Ir is zero. get 4 units.
Waves and Superposition 461

21.10 Young's Double-Slit Experiment At the constructive interference positions, the crest of
one wave has to arrive at the same time as the crest of
In 1800, the predominant opinion of the scientific the other wave. For this to occur, the path difference
community was that light was not a wave. Certain w must be 0, , 2, 3, etc. The condition for the
known light properties could be explained by the wave maximums is:
theory, but these properties could also be explained by
w = mA, where m = 0 , 1 , 2 , . . . .
the nonwave theories. In 1801, Thomas Young con-
ceived the following experiment. He wanted to de- From the above two equations,
monstrate that light was a wave by generating interfer-
ence effects similar to those in a ripple tank. Young h sin = mA, where m = 0 , 1 , 2, . . . .
wanted to use two pinholes to generate the interfer- (21.13)
ence effects, and he realized the need for mutual The interference maximums are located at the angular
coherence of the light coming through the pinholes. locations given by Eq. 21.13. The central or zero order
He decided to run sunlight through a single pinhole maximum results for m = 0. The maximums given by
and then have that light illuminate the double pinholes the other m values are called the mth order max-
(Figure 21.16). Since the double pinholes were both imums, i.e., m = l gives the first order maximums,
illuminated by the light that went through the single m = 2 gives the second order maximums, etc.
pinhole, Young reasoned that whatever phase changes The screen is a distance t from the slits. Since we
the light incident on the left pinhole underwent would are dealing with a screen far away, we can use the
equal the phase changes of the light incident on the small angle approximation to obtain
right pinhole. Thus, the light coming through the
double pinholes would be mutually coherent. ~ sin 0 ~ tan 0 = x/t, (21.14)
A modification of Young's original experiment was where x is the distance from the center of the screen.
to use slits instead of pinholes. The slits maintain the The linear location x of the mth order maximum is
needed coherence requirements while letting more given by
light through. For monochromatic light, the result of
hx/t = mA,
the double-slit experiment is a series of bright and
dark interference fringes exactly analogous to the or
ripple tank experiment. The dark fringes (unillum-
inated regions) correspond to the nodal regions of the x = mAt/h.
ripple tank (destructive interference). The bright frin- In the small angle region, the maximums are
ges were due to constructive interference. equidistant from each other, and the fringe spacing s
Let us analyze Young's double-slit experiment for a equals the distance x between the central and first
screen far from the double slits. In Figure 21.17, h is order maximum (m = 1), or
the distance between the slits Sl and S2. The wave s = At/h. (21.15)
leaving Sj has to travel a distance w farther than the
wave leaving S2. Since angles with perpendicular sides Destructive interference gives a minimum or dark
are equal, the angle B in the triangle involving w and fringe. Here the path difference must be a half integer
h is equal to 0. Therefore, multiple of A, i.e.,
sin = sinB = w/h. w = (m + 1/2)A, where m = 0 , 1 , 2,

FIGURE 21.16. Young's interference experiment.


462 Geometrie, Physical, and Visual Optics

>x

FIGURE 21.17. Geometry for double-slit in-


terference pattern.

Then Here
h sin = (m + 1/2) where m = 0 , 1 , 2 , . . . . sin 0 = 2(480 x HT 9 m)/0.5 x HT 3 m,
(21.16) sin 0 = 1.92 x 3 .
In the small angle region, the minimums occur halfway From the small angle approximation, 0 is 1.92 x
between each maximum. 10~3 radians (or 0.11 degrees, which is 6.6
minutes).
For a fixed wavelength , the fringes are wider
From Eq. 21.14,
when the slits are closer together (small h). Wide
fringes are easy to see. The fringes are narrow or finer x = t ( 1 . 9 2 x l 0 " 3 ) = 13.4mm.
when the slits are far apart (large h). Fine fringes are The second order blue maximum is not out as far as
harder to see, and may lie below the resolution ability the second order orange maximum. This is consis-
of the human eye. tent with the fact that the fringe spacing increases
For a fixed slit separation (fixed h), the fringes are with increasing wavelength.
wider for a longer wavelength . Hence, red fringes
are spread out farther than blue fringes. Thus, for EXAMPLE 21.6
white light illumination, interference fringes are Monochromatic light is used to produce double-slit
interference fringes with a spacing of 3 mm. The
colored.
screen is 5 m from the slits and the slits are
0.83 mm apart. What is the wavelength?
From Eq. 21.15,
EXAMPLE 21.5a
Light of wavelength 600 nm (orange) is used to A = sh/t,
produce double-slit interference fringes. What is A = (3 x 10~3 m)(0.83 x 10~3 m)/5 m,
the angular and linear location of the second order
maximum when the two slits are 0.5 mm apart and = 0.498 x HT 6 m,
the screen is 7 m away? A = 4 9 8 x l 0 " 9 m = 498nm.
The second order maximum is described by This wavelength is in the green region near the blue
m = 2 in Eq. 22.13. Then border, so it would actually appear as a blue-green,
sin0 = 2A/h, i.e., an aqua blue.
sin 0 = 2(600 x 10~9 m)/0.5 x 10" m,
sin 0 = 2.40 xlO" 3 .
From the small angle approximation, 0 is 2.40 x 21.11 Partial Coherence and Fringe Visibility
10~3 radians. Alternatively, 0 is 0.14 degrees or
8.25 minutes, i.e., there are 60 minutes per degree, Assume that each slit by itself provides the same
so 0.14 degrees = 8.25 minutes. irradiance on the screen (I t = I 2 ). When the light
From Eq. 21.14,
coming through the slits is incoherent, there is no
x = t ( 2 . 4 0 x l 0 " 3 ) = 16.8mm. interference and the irradiance distribution on the
screen in the small angle region is 2 li. When the light
EXAMPLE 21.5b coming through the slits is coherent, interference
For the same slits, what is the location of the fringes appear on the screen with a maximum of 4 I and a
second order maximum for light of 480 nm (blue) minimum of zero (Figure 21.13).
incident on the same slits? Suppose each slit is illuminated with a mixture of
Waves and Superposition 463

60% coherent and 40% incoherent light. When ll During a finite emission time, the atom emits a
equals I 2 , the coherent light contributes zero at the wave packet, which has a finite length (Figure 21.19).
minimums and 4(0.61!) or 2 . 4 ^ at the maximums. Now consider a pair of double slits illuminated by the
The incoherent light adds an irradiance of 0.4(21!) = atom. When the emission time is long enough, the
0.81! to every point on the screen. The net effect is waves that overlap after passing through the slits
that the maximums now have an irradiance of 2.41 j + belong to the same wavepacket, are therefore mutual-
0.8 ll or 3.211,, while the minimums now have an ly coherent, and give fringes of visibility 1. For a
irradiance of 0.81,. As compared to the 100% co- shorter emission time, the wavepackets may only par-
herent case, the interference fringes are still there, but tially overlap, so that interference occurs only part of
the visibility (contrast) is reduced (Figure 21.18). the time. This decreases fringe visibility (partial coher-
The fringe visibility is defined as ence). For still shorter emission times, the same
wavepackets may not overlap at all. This results in no
Visibility = max ^min
(21.17) interference and zero fringe visibility (incoherence).
max min The longitudinal coherence length is essentially the
For 100% coherent light the fringe visibility equals length of the wavepacket emitted during a typical
1.0, while for incoherent light the fringe visibility emission. In a tungsten filament (a hot solid) light
equals zero. For the 60-40 mixture, the fringe visibili- bulb, the typical coherence length is only 3 or 4
ty is wavelengths. For a low pressure gas discharge tube,
the collisions are less frequent, and coherence lengths
Visibility 1 of 5,000 wavelengths (about 2 mm) can result. The
3.21, + 0 . 8 1 ! ' laser (invented in 1960) is an incredible source in
which the coherence length can be as long as 10 n
or wavelengths (about 50 km). The longer the longitudi-
2.4
Visibility = = 0.6. nal coherence length, the easier it is to produce visible
interference fringes.
Note that the fringe visibility equals the percent of For extended sources, there is also a lateral coher-
coherent light. ence effect. Consider a pair of double slits illuminated
The above example shows that partially coherent by two point sources that are close together. Suppose
waves can generate interference fringes, but the vis- that each point source by itself produces interference
ibility or contrast of the fringes is reduced. The per- fringes of good visibility. The resultant irradiance on
cent of coherent light is called the degree of coherence. the screen due to both point sources is the sum of the
The above numerical argument can be repeated for irradiances in the two interference patterns.
any percentage. Thus, we have a general principle. Because the point sources are angularly separated,
The fringe visibility equals the degree of coherence of the central maximums in the two interference patterns
the light coming through the two slits. are not quite lined up. This means that the resultant
One cause of partial coherence (or incoherence) is irradiance pattern has fringes of slightly reduced vis-
the finite emission time of atoms. An isolated atom ibility. If the two point sources are separated further,
emits light for 10"8 s at the most, and then needs to the visibility of the fringes in the resultant pattern
recycle before emitting again. In materials, the emis- decreases still further. The fringe visibility drops to
sions are further shortened by interactions or collisions zero when the two point sources are separated by an
with neighboring atoms. (Hot solids can have emission amount such that the maximum from one interference
times as short as 10~15 s.) pattern lines up with the minimum from the other.

FIGURE 21.18. Reduction in fringe visibility


due to partial coherence.
464 Geometrie, Physical, and Visual Optics

A/WWWVW FIGURE 21.19.


emission time).
Finite length of a wavepacket (due to finite

Now consider an extended source consisting of the lateral coherence length depends on the angular
many points. Suppose the source has a small angular size of extended sources.
size 0S and illuminates a pair of double slits. Then the
light from each point source produces an interference
pattern each slightly misaligned with the others. This
results in a slightly decreased fringe visibility. As the 21.12 Laser Visual Acuity Testing of
angular size of the source gets larger, the different the Visual Pathways
interference patterns are more and more misaligned,
and the interference patterns tend to counteract each Since interference fringes are easy to form with lasers,
other, resulting in decreasing fringe visibility. For an we can use a laser to bypass the optics of the eye and
angular source size 0S and a slit separation h, the fringe test the integrity of the neural part of the human visual
visibility is negligibly small when system (retina, optic nerve, brain). The method is to
0 s >A/h. (21.18) split a laser beam and project two small spots of the
coherent light in the patient's entrance pupil. The light
Thus, interference fringes are visible when 0S is much then diverges from these spots and forms interference
less than /h. Note that the required angular size fringes on the retina. The orientation of the fringes
depends on the slit separation h. can be changed by rotating the spots. Furthermore,
Why couldn't Thomas Young have directly illumi- the fringe spacing can be changed by moving the spots
nated his double slits with light from the sun, and closer together or farther apart. We thus have the
observed interference fringes? The answer is that for tools to ascertain which set of fringes the patient can
Young's slit separation h, the sun's angular size 0S was resolve.
too large. Young solved this by running the sunlight For example, suppose a person with a history of
through a pinhole before illuminating the double slits. retinal disease has cataracts. The surgeon is unsure
The pinhole's angular size was sufficiently small to whether to do cataract surgery since the retina may
give visible interference fringes. Later, Emile Verdet not be functional. Furthermore, the cataracts obscure
made double slits close enough together (h < 0.05 mm) the retina, so it cannot be examined by normal means.
so that he obtained visible interference fringes with To use the laser method to check the visual pathways,
the slits illuminated directly with sunlight. all we need to do is find a small clear enough area in
Verdet's experiment is startling because we normal- the crystalline lens to project the two spots of laser
ly think of the sun as an incoherent source. However, light. Then we can vary the fringe orientation and
it is probably better to think of the sun as having a spacing and get a subjective measure of the potential
small degree of coherence that can be manifested acuity after cataract surgery. Since the interference
under the right conditions. One way to specify those fringes do not depend on a clear image being in focus
conditions is in terms of the h value that first gives on the retina, the patient does not need to be correc-
zero visibility in Eq. 21.18. This h value is sometimes ted during this procedure. A positive result (good
called the lateral coherence length lw. From Eq. 21.18, acuity) would indicate that the patient is a good
candidate for surgery. A negative result is inconclu-
lw = A/0 (21.19) sive, since we may not be sure that the interference
where 0S is the angular size of the source. (Remember fringes actually get to the retina.
that by moving extended sources far enough away, we
can make them act like point sources, i.e., we increase
lw by decreasing 0S.)
In astronomy, the angular size of red dwarf stars 21.13 Other Methods for Producing
has been determined by measuring the lateral coher- Interference
ence length lw of light from the star, and calculating 0S.
(Even more impressive, but outside the scope of this Interference fringes similar to those from the double-
book are the second order coherence measurements, slit experiment can be produced by a number of other
developed by Hanbury-Brown and Twiss, to deter- methods. One way is to use a Fresnel biprism (Figure
mine the angular size of stars smaller than red dwarfs.) 21.20a). When the biprism is illuminated by a point
In summary, the longitudinal coherence length de- source, it forms two virtual image points. In effect, the
pends on the finite emission time of the atoms, while two virtual images function as the two point sources in
Waves and Superposition 465

s-
b)

A/\/k/\/\, FIGURE 21.20. a. Interference fringes by Fresnel biprism. b.


Michelson interferometer.

Young's original interference experiment. The waves the special theory of relativity. (In 1907 Albert
that go through the different parts of the biprism then Michelson became the first American to win the Nobel
overlap and produce the interference fringes. In the Prize in Physics.)
same manner, two slightly tilted mirrors can be used
to generate interference fringes in the reflected light.
The Michelson interferometer is a famous device
that uses interference fringes to make precision meas- 21.14 Speckles and Laser Refraction
urements (Figure 21.20b). Interference fringes are a
natural for precision measurements since a variation of As mentioned, lasers are highly coherent sources.
even half a wavelength (about 280 nm) gives a clear When a diverging laser beam illuminates a diffuse
shift in the interference fringes. The basic ingredients reflecting surface such as a wall, the illuminated region
of a Michelson interferometer are a partially silvered appears to have a series of small black spots in it
mirror (A) together with two fully silvered mirrors (B (similar to a bunch of fly specks). This speckle pattern
and C). The light incident on A is split in two, and is due to interference of the diffusely reflected co-
each part travels to a different mirror. The part that herent light.
goes to B reflects, and travels back to A at which Figure 21.21 shows a highly magnified view of the
partial transmission occurs with the transmitted part diffusely reflecting surface. The waves arriving at posi-
going to the screen. The part that goes to C reflects tion 3 have reflected from different positions on the
and travels back to A at which partial reflection occurs surface, and thus have different path lengths. Depend-
with the reflected part going to the screen. The two ing on the path difference, either destructive interfer-
different paths introduce the phase differences that ence (dark speckles) or constructive interference
lead to the formation of the interference fringes on the (bright speckles) can occur. Similar statements can be
screen. made about any other position, such as 1 or 2.
Michelson used this interferometer to make preci- When an observer views the wall, the interference
sion measurements on the speed of light. In particular, of the diffusely reflected light occurs at the retina, and
he showed that the speed of light is independent of the thus the observer sees the speckles. (A camera records
relative motion of the source and the measurement the speckles on film in the same manner.) Since the
system. In other words, an observer moving towards a speckles are formed by interference, the observer sees
light beam, and an observer moving away from a light the speckles even if the wall is blurred. In fact, we can
beam will both measure the relative speed of the beam consider the speckles on the retina to be conjugate to
as c. This is one of the experimental foundations of speckles in the eye's far point plane.
466 Geometrie, Physical, and Visual Optics

coherent speckle at the retina (position ) is conjugate to the


dark speckle in the far point plane (position 1).
Figure 21.22b shows an uncorrected, unaccommo-
dated hyperope viewing the wall. Again, assume the
diffusely reflected light going into the eye destructively
interferes, giving a dark speckle at . This speckle is
conjugate to the speckle that would have been formed
at position 1. In terms of geometric optics, the real
speckle on the retina () is conjugate to a virtual
speckle at position 1.
Not only are the speckles present, but when an
ametrope moves his or her head back and forth (or up
FIGURE 21.21. Speckle formation due to path length differ- and down), the speckles also appear to move. In fact,
ences from diffuse reflection. they appear to move in opposite directions for a
myope as compared to a hyperope. The speckles do
not appear to move for an emmtrope. It should be
Consider a myope viewing a distant wall (Figure obvious then that the speckle movement can be used
21.22a), and assume that a dark speckle is formed by to determine the refractive error of the observer's eye.
destructive interference in the myope's far point plane This is the basis for laser refraction.
(position 1). The two waves travel independently Let us consider the motion for a myope, hyperope,
through each other, continue on to the eye, and are and emmtrope (Figure 21.23). In each case, we will
recombined at the retina. The path lengths for each start with a dark speckle on the center of the retina.
wave are the same between the far point plane (posi- This speckle is conjugate to a speckle in the far point
tion 1) and the retina (position ), so the waves again plane; then the eye moves up (dashed lines). For the
destructively interfere at the retina. In effect, the dark myope, the speckle in the far point plane does not

incident
coherent
light

myopic eye

coherent light
incident

hyperopic eye
FIGURE 21.22. a. Speckle in myope's far point
plane is conjugate to speckle on the retina, b. Virtual
speckle in hyperope's far point plane is conjugate to
b) the speckle on the retina.
Waves and Superposition 467

EM RM PM
JRM
i t 1 myope
-^-center (shifted)

against
center

center (shifted)
EM RM PM
I RM
t t t ti4 ^ center
with

emmtrope
center (shifted)

EM RM PM

Y no motion - center

FIGURE 21.23. Speckle motion for myope, hy-


perope, and emmtrope.

shift, but now the nodal ray (dashed) for the eye becomes the nodal ray, the speckle remains on the
intercepts the retina at a position above the center. center of the retina, and no speckle motion is per-
The nodal ray defines the new retinal speckle position, ceived.
and it is shifted up relative to the retinal center. So relative to a head movement, the myope per-
However, retinal images are inverted, so an upward ceives against motion, the hyperope perceives with
shift on the retina is perceived as a downward motion. motion, and the emmtrope perceives no motion. The
Hence the eye moved up (EM), the speckle on the higher the degree of ametropia, the faster the per-
retina shifted up (RM), and the perceived speckle ceived motion. The lower the degree of ametropia, the
motion is down (PM). Thus, the myope who moves slower the perceived motion.
his or her head perceives the speckles to move in the Laser refraction is actually performed by illuminat-
opposite direction. ing a very slowly moving surface with laser light. Then
For the hyperope, the solid line shows the rays for the observer does not need to make a head movement
the speckle formed on the center of the retina. The to perceive the motions. A drum that rotates one
dashed lines show the eye shifted up, in which case the revolution per hour works well. Under incoherent
incident top ray becomes the nodal ray for the shifted illumination, the drum appears stationary. Under co-
eye. The nodal ray (dashed) shows that the speckle herent illumination, the speckle motion is perceived
position is now below the center of the retina. So the by an ametrope. A corrected ametrope (or an emm-
speckle moved down on the retina, which is perceived trope) sees no definite with or against motion, al-
as an up motion. Here the speckle appears to move though the speckles may appear to boil around in a
with the motion of the retina. random fashion. (Note that a surface moving one
The bottom drawing is for the emmetropic eye. direction while the patient's head is stationary is equi-
This time when the eye is shifted up so that the top ray valent to the patient moving his or her head the
468 Geometrie, Physical, and Visual Optics

opposite direction while the surface is stationary. that adhere together give the fringes because of trap-
Therefore an uncorrected myope perceives the spec- ped thin layers of air. Some chemical interactions
kles to move opposite to his or her head motion, or produce thin film surface changes, which then result in
with the motion of a moving surface.) colored interference fringes (e.g, the bottoms of some
Laser light is monochromatic, so the terms emme- cooking pans, weathered glass). Some sea shells also
tropic, myopic, and hyperopic refer to the conditions show thin film interference colors because of their thin
for that wavelength. The most common low energy film shell structure.
laser is a helium-neon laser, which puts out a coherent Thin film interference is formed as a result of
monochromatic beam with a 632.8 nm vacuum wave- multiple reflection processes. This is a wave effect;
length. Because of the eye's chromatic aberration, however, we will use rays to simplify the figures. In
blue focuses before red. A person emmetropic for the Figure 21.24, the incident ray hits the thin film at point
middle of the visible spectrum is actually hyperopic for A. Part of the light is reflected (ray 1) and part is
the helium-neon laser beam. transmitted. The transmitted ray travels to point B,
Optometers based on laser refraction have been where part of it is transmitted (ray u) and part is
very useful is studying the subjective accommodative reflected. The reflected part travels to point C, where
response of the eye. A beam splitter is used so that the part of it is transmitted (ray 2) and part reflected. The
person can simultaneously view the object (the accom- part reflected travels to point D where again part is
modative stimulus) and a slowly rotating drum. The transmitted (ray v) and part reflected. The process
laser light is flashed onto the drum for short periods of continues on resulting in the reflected waves 1 to 5 and
time. While the laser is on, the person sees the speckle the transmitted waves u through y.
pattern simultaneously with the accommodative For a thin film, the coherence length of ordinary
stimulus. If the person sees no speckle motion, the light is long enough so that all the overlapping multi-
subjective accommodative response equals the ple reflected waves give interference effects. For a
stimulus. If the person sees speckle motion, then the complete analysis, all the multiple reflected waves
response leads or lags the stimulus. The speckles need to be taken into account, but for our purposes
themselves do not act as an accommodative stimulus; we will consider only the largest amplitude waves
after all, they are formed on the retina independent of namely, 1, 2, u, and v.
the focus. These laser optometers have been a prime If waves 1 and 2 are out of phase, then destructive
tool for testing the accommodative intermediate rest- interference occurs in the reflected light. When this
ing state hypothesis. happens, waves u and v are in phase resulting in
The speckle size is a function of the diffuseness of constructive interference in the transmitted light. As
the surface and the size of the optical system's aper- before, energy is conserved in the process, but it is
ture stop. If the drum is rotating vertically, the speckle redistributed from the reflected light to the trans-
motion is vertical and power is checked in the vertical mitted light.
meridian of the eye. The axis of the drum can be By adjusting the thickness of the thin film, we
rotated so that any other meridian can be checked, could make the destructive interference occur between
and hence laser refraction can be used as a meridional waves u and v, in which case the constructive interfer-
refraction technique. For an off-axis meridian of an ence occurs between waves 1 and 2. Again energy is
astigmat, the torsional component of the dioptric conserved. The energy that is removed from the trans-
power causes the speckle motion to appear in a meri- mitted light is channeled into the reflected light.
dian different from the meridian of rotation. This
complicates the use of laser meridional refraction.

\ 12 3 4 5

21.15 Thin Film Interference

Colorful interference fringes are created by thin films nz


of various materials. These fringes are quite common
because the films are thin enough so that even short
wavepackets give the interference effects. Hence, the
s VVYYB 3
\
fringes show under illumination by sunlight or other
* \v\\ "
common light sources. Some frequently seen examples u v x y
are the color of oil slicks on wet roads and the colored FIGURE 21.24. Multiple reflections responsible for thin film
fringes in soap bubbles. Two clean microscope slides interference.
Waves and Superposition 469

21.16 Phase Changes at Reflection r 12 t 12 A + t 12 r 21 A = 0,


t12A(r12 + r 21 ) = 0,
For thin films, one obvious source of the phase differ-
ences is the over and back path differences across the or
film. A second source is phase changes at the reflec-
tions. To gain some intuition about phase changes at a r,,=
reflection let us consider a pulse propagating along a
rope (Figure 21.25). The rope consists of a lightweight The minus sign indicates that the two reflections differ
section (thin line) connected to a heavyweight section by a phase change of 180 degrees. For dielectrics in
(thick line). Some reflection occurs when the pulse situations not involving total internal reflection, Max-
moves from one section to the other. When the crest is well's equations show that when the light is initially in
initially in the heavy section, the reflected pulse is also the optically less dense (lower n) medium, the re-
a crest (no phase change). However, when the crest is flected light has a 180 degree phase change, whereas
initially in the lightweight section, the reflected pulse when the light is in the optically more dense medium
is a trough (a 180 phase change on reflection). (higher n), the reflected light does not have a phase
change. For example, light in air that reflects at an
G. G. Stokes (1819-1903) used the principle of air-glass interface undergoes a 180 degree phase
optical reversibility to prove that light waves must change (Figure 21.27), while light in the glass that
exhibit reflected phase changes analogous to the rope reflects at a glass-air interface does not change its
pulses. Consider an interface between two dielectric phase.
media of refractive index n{ and n 2 . In Figure 21.26a,
the incident ray represents a wave of amplitude A For completeness, consider the light traveling up
incident on the interface. The amplitude of the re- and to the left in the reversed situation. According to
flected wave is r 12 A. The amplitude of the transmitted the principle of reversibility,
light is t 12 A. *2*2 A + r 1 2 r 1 2 A A ,
In the optically reversed situation, two waves are
incident on the interface (Figure 21.26b). The wave or
r12A is partially reflected, r12(r12A) and partially trans- ti2t2i+(r12)2=l.
mitted, t 12 (r 12 A). The wave t 12 A is partially reflected,
r 21 (t 12 A) and partially transmitted, t 21 (t 12 A). Accord- There are no phase changes with transmission so t12
ing to the reversibility principle, the reversed situation equals t21. Then
must give us back the original wave of amplitude A
traveling in the opposite direction. Hence the two (t 12 ) 2 + (r 12 ) 2 = l.
waves traveling to the bottom left in medium n2 must ^et
cancel each other out (destructive interference). In
terms of amplitudes, T=(t 1 2 ) 2 , and R = (r12)2.

A
FIGURE 21.25. Relation of reflected pulse phase to incident
pulse phase.

r 12A

FIGURE 21.26. a. Reflection and refraction of incident


a) b) light, b. Reflection and refraction of reversed light.
470 Geometrie, Physical, and Visual Optics

180
phase
change

glass

FIGURE 21.27. Phase behavior at reflection.

Then Since
T + R = l,
A =
and we see that T and R are the geometric optics
transmission and reflection factors (no absorption). the condition for a minimum in the reflected light
For near normal incidence, Fresnel's law of reflection becomes
gave 2t = m(A/n 2 ) where m = 0 , 1 , 2, . . .
2
( 2 - ) (100%). where is the vacuum wavelength.
R=
(2 + ) 2 For m = 1, the reflected light has a dark fringe for
2t= A/n2.
For m = 0, the condition for a dark fringe in the
21.17 Soap Films reflected light is
2t = 0.
Consider monochromatic light incident on a soap film
of thickness t in air (Figure 21.24). Here rij and n 3 If there is no film, we obviously do not expect any
equal 1, while n2 is approximately 1.33. At the first reflected light, and the above condition agrees with
reflection (point A), the reflected light changes phase that. However, as the film thickness goes to zero, the
by 180 degrees. The transmitted wave travels to B. A mechanism by which the reflection goes to zero is
percentage of the wave reflects at B and travels back interesting. The air-film reflection gives a 180 degree
to C where part is transmitted (wave 2). For almost phase change, while the interior film-air reflection
normal incidence, wave 2 has traveled over and back does not give a phase change. If the thickness is
or a distance of 2t farther than wave 1. Suppose negligible, these two reflected waves (1 and 2) overlap
and destructively interfere. Gravitational forces cause
2t = mAtf, where m = 0 , 1 , 2, . . . a soap film to thin with time. As t goes to zero, the
and Atf is the wavelength in the thin film. Then the reflected light turns black even before the film breaks,
path difference does not introduce a phase difference thus confirming the above analysis.
between waves 1 and 2. However, wave 1 had a 180 By a similar analysis, constructive interference in
phase change at its reflection whereas wave 2 did not. the reflected light (and hence destructive interference
Thus, wave 1 is 180 degrees out of phase with wave 2, in the transmitted light) occurs when
and they will destructively interfere.
What about the transmitted light? At C the light 2 t = ( m + l), m = 0,1,2,
reflects with no phase change and travels back to D, \ 2 / n9
where part of it is transmitted (wave v). The only
EXAMPLE 21.7a
source of phase differences between waves u and v is
that wave v traveled 2t (over and back) farther than Monochromatic light of 656 nm (red) is incident on
a soap film. What is the smallest non-zero thickness
wave u, but for the thickness condition above, waves u that will give a dark fringe in the reflected light?
and v are in phase, and so they constructively in- Here m = 1, and
terfere.
2t = A/n2,
Because of the interference, there is a minimum or
dark fringe in the reflected light, while there is a
maximum or bright fringe in the transmitted light. 656 nm
More energy is transmitted and less is reflected. t= =2466nm
2(133)
Waves and Superposition 471

EXAMPLE 21.7b
; 2
When white light is incident on the above film,
what color will the reflected light tend to have?
The thickness of 246.6 nm gives a maximum or
bright fringe in the reflected light for a wavelength 1
given by / /c
2(246.6 nm) = (m + 2) T^3 '
B
or n3 = 1.50 \ \
4(246.6 nm)(1.33) = (2m + 1). \ v
For u

m = 0, = 1311.9 nm, FIGURE 21.28. Antireflection thin film.


m = l, = 437.3 nm,
m = 2, = 262.4 nm.
percent reflectance R12 at the n r n 2 interface to equal
Of these wavelengths, 437.3 nm is in the visible
range (indigo or violet). Therefore, a thickness of the percent reflectance R23 at the n2-n3 interface. For
246.6 nm gives destructive interference in the re- normal incidence. Fresnel's law of reflection (Chapter
flected light for 656 nm (red) and constructive inter- 10) gives
ference in the reflected light for 437.3 nm. There
are other wavelengths around 437.3 nm (especially _ (n2-ni)2 1 0 i W
2
the blues) that will also tend to have a constructive ~(7^ '
interference effect. Hence in white light, the re-
flected fringe shifts from the violet toward a blue.
Normally in soap bubbles, the film thickness R23= (n
5^^100%.
+n )3 2
varies with time. For incident white light, the re-
sulting thin film colored interference fringes drift The ideal thin film has an index n2 that makes R12
around with time. If you watch the top of a soap equal to R23. We can set the above two expressions
bubble, these interference fringes turn black or equal to each other, cross multiply, and solve for n2 to
disappear just prior to the bubble breaking. obtain
n2 = VrM^. (21.20)
This is the refractive index for an ideal antireflection
21.18 Antireflection Films
thin film. For example, consider ophthalmic crown
glass (n = 1.523) in air (n = 1.000), the ideal refractive
In lens systems, the amount of light lost by reflection
index is VI.523, or 1.234.
rapidly increases with the number of lenses. For exam-
ple, if 4% is reflected at each surface, the transmission Practical thin films do not necessarily have the ideal
factor T is 0.96 per surface. Consider a system consist- index. A thin film of magnesium fluoride (n = 1.38) is
ing of five separated lenses in air. There are ten widely used as an antireflecting coating for glass len-
surfaces, and the resultant transmission of the primary ses. Some of the practical reasons are good adhesion
wave is (0.96)10 or 0.66. The primary transmitted wave and relatively good stability over time.
is down to 66% even though only 4% is being re- The thickness condition for an antireflection thin
flected at each surface. Furthermore, multiple reflec- film is different from that of a soap film in air because
tions can generate unwanted ghost images in the of different phase changes at the reflections. In Figure
system. These ghost images sometimes cause troubles 21.28, there is a 180 degree phase change for the light
for spectacle corrected ametropes. reflected at position A. There is also a phase change
In the 1930s vacuum technology advanced to the for the light reflected at position B. Therefore, the
point where thin films could be evaporated onto glass only phase difference between waves 1 and 2 is due to
surfaces. This made possible the design of antireflect- the over and back (2t) distance that wave 2 traveled.
ing thin films. The antireflecting thin films work simi- For destructive interference between 1 and 2, we then
lar to the soap bubbles except that the index sequence need
is now different. In Figure 21.28, n 3 , n 2 , and n, are 2t = (m + l/2)A tf , m = 0,1,2,
the refractive indices of the glass, thin film, and initial
For m = 0,
medium, respectively. In order for the antireflecting
film to work with maximum efficiency, we need the 2t = Atf/2,
472 Geometrie, Physical, and Visual Optics

or the red and violet ends of the spectrum. This gives the
t = Atf/4. lens a purplish appearance similar to that of a ripe
plum. (Such a lens is sometimes referred to as a
The minimum thickness of an antireflecting thin film bloomed lens.) For cosmetic reasons, the thickness of
on glass is a quarter of the wavelength in the thin film the antireflection coatings on spectacle lenses is some-
(Atf). Since times adjusted so that the residual reflected light has a
strawish appearance instead of the purple.
A
tf - The thin film thickness that gives destructive inter-
ference in the reflected light gives constructive inter-
we have ference in the transmitted light. In Figure 21.28, con-
sider the wave traveling from position A to B. At B,
t = 4n, (21.21) part is transmitted (wave u) and part is reflected with a
180 phase change. The reflected light travels back to
or position C, and part is reflected without a phase
tn change. The light then travels back across the film,
2=4
and part is transmitted at point D (wave v). Wave v
The product of the thickness and the refractive index, differs from u by a 180 phase change due to the
tn 2 , is called the optical thickness. Thus, we can say reflection at B and by the phase difference due to the
that the reflected light has a minimum when the over and back distance (2t). The condition for de-
optical thickness is a quarter wavelength. structive interference in the reflected light was
2t = (m + l/2)A tf , m = 0,1,2,
EXAMPLE 21.8
An MgFl (n = 1.38) antireflection coating is placed The above condition gives a 180 phase difference due
on a glass lens. What is the minimum thickness of to the path difference. This 180 phase difference adds
the film for incident light of wavelength 552 nm? to the 180 phase difference due to the reflection at B.
552 nm ^ The result is 360 phase difference, which means that
tn2 = - = 138 nm, the two waves (u and v) are in phase and constructive-
ly interfere. The energy that is not reflected (because
138 nm of the destructive interference) is transmitted (because
t = = 100nm. of the constructive interference).
1.38
Thus, an antireflection coated lens has an increased
Figure 21.29 gives the reflectance as a function of transmission. This is important to remember since at a
incident wavelength for uncoated glass, a single layer quick glance an antireflection coated lens can look like
antireflection coating, and multiple layer antireflection a lens with a light absorbing tint. However, the lens
coatings. The single layer coating is not ideal, so the with the absorption tint decreases transmission, whereas
minimum is not zero. The performance is improved by the antireflection coated lens increases transmission.
using layered thin films. Zinc sulfide (n = 2.32) is one This can make a difference in low illumination condi-
of the materials used with MgFl to achieve the layered tions such as night driving.
results. Just as thin films can be antireflecting, we can
When the coating is antireflecting for the middle of design layers of thin films that give constructive inter-
the visible spectrum, some reflection occurs at both ference in the reflected light. Such thin film stacks can
give a much higher percent reflection than silvered
surfaces.

21.19 Interference Filters

Interference filters work on thin film interference prin-


ciples. They are designed to transmit a very narrow
band of visible light (by constructive interference)
while reflecting all other wavelengths. The reflection
of the transmitted wavelength is prohibited by destruc-
FIGURE 21.29. Reflectance as a function of incident wave- tive interference in the reflected light.
length for uncoated and various layer antireflection films. An interference filter consists of a thin film sand-
Waves and Superposition 473

FIGURE 21.30. Transmittance for interference filter


vs absorption filter.

wiched in between highly reflecting surfaces. The high- visible when two clean microscope slides adhere to-
er the reflectance of the surfaces, the narrower the gether trapping some air between them. In optics,
wavelength band of transmitted light. Figure 21.30 such thin film interference is used to test the precision
compares the transmission of a narrow band interfer- of surfaces to within a fraction of a wavelength of the
ence filter to that of an absorption filter. The absortion light. The simplest set-up for spherical surfaces gener-
filter is a broad band filter that also absorbs much of ates circular interference fringes called Newton's rings
the desired light. The interference filter is a high (after Newton who experimentally discovered them
transmittance narrow band filter. An absorption filter even though he did not believe the wave theory.)
heats up under high illumination conditions, whereas In Figure 21.31, the light in the glass incident on A
an interference filter tends to stay cool, since it is is partially transmitted and partially reflected (without
reflecting the unwanted light instead of absorbing it. a phase change). The light in air incident on B is
Consider an interference filter that transmits a partially reflected with a 180 phase change. Since the
wavelength . The thickness of the film that gives reflections introduce a 180 phase difference, waves 1
constructive interference between the multiply reflec- and 2 will be out of phase if the path difference
ted transmitted waves is exactly the correct thickness between them is
for a standing wave of that light to be formed in the
film (between the highly reflecting surfaces). In the 2t = mA, m = 0,1,2, . . .
ideal situation, this is the only wavelength for which where is the wavelength in the air gap. There are
there is a large buildup of energy inside the thin film, then circular dark fringes in the reflected light for
and consequently, the only wavelength that is trans- thicknesses corresponding to air gap thicknesses t,
mitted. Increasing the reflectance of the surfaces helps which corresponds to integer multiples of a half wave-
to finely tune the allowed standing wave, and con- length. Bright fringes occur between the dark fringes.
sequently, the transmission band narrows. When an Note that a dark fringe occurs for t = 0, or for the
interference filter is tilted, the effective thickness contact point. This is not surprising since we do not
changes and the transmitted wavelength decreases. expect reflection from an ophthalmic crown-ophthal-
(The decrease may seem counterintuitive, but a com- mic crown interface. Actually, we already get destruc-
plete geometric analysis shows that it is correct.) tive interference for a non-zero air gap provided the
Precision interference filters are made with multi- gap thickness is considerably smaller than a quarter
thin film layers. Multilayer broad band interference wavelength.
filters also exist. For example, we can design interfer- Flaws of a quarter wavelength shift a dark fringe to
ence filters that refect visible light, but transmit in-
frared (i.e., a cold mirror), or vice versa (a hot
mirror). The Fabry-Perot interferometer is a precision
interferometer that works much like an interference
filter. ^<Z^ sees ((()

21.20 Newton's Rings

Thin air layers can also generate thin film interference.


One example is the thin film interference fringes FIGURE 21.31. Newton's rings.
474 Geometrie, Physical, and Visual Optics

a bright fringe and are thus clearly visible by inspec- spacing on a screen 300 cm from the slits? Same
tion of Newton's rings. Similar thin film interference question for the slits 0.5 mm apart. Same question
methods are used to calibrate precision manufacturing for slits 0.5 nm apart.
equipment. 6. Light from an arc source passes through an inter-
ference filter and is incident on two narrow slits
0.06 cm apart. The interference pattern is formed
on a screen 250 cm away. The interference fringe
Problems spacing on the screen is 2.28 mm. What is the
wavelength of the light?
1. A light wave has a frequency of 5.00 x 1014 Hz. 7. For a given wavelength, does the lateral coher-
What is the vacuum wavelength in nm? When this ence length of a source increase or decrease with
light enters glass (n = 1.50), what is the frequency increasing source size?
and wavelength? 8. When light of wavelength 489 nm is incident on
2. A harmonic wave has an amplitude of 3 units. A soap film (n = 1.33) in air, what is the smallest
second harmonic wave has an amplitude of 12 thickness that gives a maximum in the reflected
units. What is the relative intensity of the second light?
wave to the first? 9. What is the ideal index and thickness for an
3. Light in the vitreous humor has a 406 nm wave- antireflecting thin film on a high index glass (n =
length. What is the normal color classification for 1.68) lens in air. Assume the incident wavelength
this light? is 560 nm.
4. A harmonic wave of amplitude 5 units is super- 10. A 90 nm thick MgFl (n = 1.38) antireflecting thin
posed with a harmonic wave of amplitude 7 units. film is placed on a glass lens with a 1.523 index.
What is the resultant intensity when the waves are What incident wavelength is the thin film antire-
incoherent? When the waves are coherent and in flecting for?
phase? When the waves are coherent and 180, 11. A helium-neon laser beam (wavelength 632.8 nm)
out of phase? is incident on double slits of separation 0.8 mm.
5. Sodium D light has a 589.3 nm vacuum wave- How far is the third order maximum in the inter-
length. When this light passes through two narrow ference pattern on a screen located 6 m from the
slits 0.1 mm apart, what is the interference fringe slits?
CHAPTER
TWENTY-TWO
Diffraction

22.1 The Huygens-Fresnel Principle sources for the production of spherical secondary wave-
lets, and at any later time the new wavefront position is
We are very familiar with the ability of sound waves to the envelope (or surface of tangency) to these secondary
travel around corners. You may have even noticed the wavelets.
ability of water waves to propagate around obstacles.
The ability of a wave to propagate around corners is Huygens' principle can be used to derive the geomet-
called diffraction. The importance of diffraction de- ric optics law of refraction (Snell's law), as well as the
pends on the size of the obstacle or aperture relative law of reflection.
to the size of the wavelength. Music and speech Figure 22.1a shows Huygens' principle applied to
wavelengths range from 1.7 cm to 17 m. A door (an plane waves. The points A, B, C, and D each serve as
aperture) is perhaps a meter across, so that for the a point source for a secondary spherical wavelet. After
long wavelengths, the sound waves bend easily around a time t, each of the secondary wavelets have prop-
the doorway. Diffraction is less evident for the short agated downstream. The new position of the wave-
wavelengths. front is specified by the envelope (or surface of
The wavelengths of visible light are so short tangency) of the secondary wavelets. This envelope is
(10~7 m) that at first glance it appears no bending a straight line indicating that the wavefront is still
occurs (i.e., the law of rectilinear propagation). How- plane.
ever, careful observations show that visible light waves Figure 22.1b shows plane waves incident on an
do diffract. For example, look at a distant street light obstacle. High above the obstacle, the envelope is still
at night and squint. The light appears to streak out flat indicating rectilinear propagation. In the lower
from the light bulb, i.e., it has bent around the corners region, the obstacle blocks some of the secondary
of your eyelids. Alternatively, stand in a dark room wavelets. In the region behind the obstacle, the sec-
and look at a distant light bulb in another room. Now ondary wavelets from points above the obstacle give
move until the doorway blocks half of the distant light an envelope that is curved similar to that of a sphere.
bulb. Due to diffraction around the doorway, the light This curved envelope indicates that some light bends
appears to streak out into the umbra region of the around the obstacle and propagates into the geometric
dark room. optics umbra region. This is the diffracted light. Since
There are several approaches to explaining diffrac- the amount of diffracted light (the luminous flux) is
tion. The most intuitive approach is based on a princi- small, any background light may mask it. However,
ple first put forth by Christian Huygens (1629-1695, with a dark background, it is easy to observe.
inventor of the pendulum clock). Huygens' principle The fall-off in the diffracted light is not as smooth
states: as suggested by Figure 22.1b. In fact, the irradiance
varies in a ripple-like manner (Figure 22.1c). Augustin
All points on a wavefront can be considered as point Fresnel (1788-1827) expanded Huygens' principle to

475
476 Geometrie, Physical, and Visual Optics

1
22.2 Single-Slit Diffraction

A Consider monochromatic plane waves (wavelength )


incident on a single slit of width d. Figure 22.2a shows
B a cross-sectional view in the plane perpendicular to the

i
\ slit length. When the aperture is large, the waves
C going through the middle remain plane waves (little or
no diffraction). When the aperture is very small, only
one secondary wavelet comes through, and the result-
ing cross-sectional shape of the envelope is the circular
D shape of that wavelet. The circular shape indicates a
large amount of diffraction (lateral spreading) of the
plane wave light.
initial
plane
J>7
^ ^
at a later time In Figure 22.2a, the slit has a size such that several
wave / secondary wavelets pass through the slit. At the dis-
tant screen, these secondary wavelets superpose giving
a rippled irradiance distribution, which is referred to
as the single-slit diffraction pattern (Figure 22.2b).
The irradiance in the pattern has a central maximum
at P0, and then falls to zero (the first minimum). Past
the first minimum, the irradiance rises to a relatively
small second maximum before again dropping to zero
"*1
(the second minimum). The rippling continues with
each maximum having less magnitude than the prev-
ious maximum.
According to Huygens' principle, each point on the
screen receives secondary wavelets from each point on
the wavefront in the slit. The resultant displacement is
the sum of the displacements of the secondary wave-
b) lets. In the limit of a distant screen, the central point
becomes equidistant from each point on the slit's cross
section so that all the secondary wavelets arrive in
phase, and the resulting constructive interference gives
the large central maximum I 0 . The other points on the
screen are not equidistant from each point on the slit;
therefore, the secondary wavelets arriving at these
points are not all in phase, so the resultant displace-
ment is smaller than that of the central maximum.
Figure 22.3 shows a view of the geometry for the
irradiance at the point P on the distant screen (where
FIGURE 22.1. Huygens principle, a. Plane waves, b. Diffrac-
P subtends an angle with the axis). For a distant
tion. c. Rippled irradiance in the diffraction pattern. screen the points on the line AC are equidistant from
P. Therefore, the relative phase distribution of the
wave along AC is identical to the relative phase
distribution of the wavelets arriving at P. Figure 22.4a
explain the ripple effect. A simplified statement of shows the relative phase distribution along the line AC
Fresnel's addition to Huygens' principle is: for the angle that gives the first minimum. The first
minimum occurs because each upward displacement
For light in the same bundle (i.e., waves from the same along AC is matched by a downward displacement of
point source) the secondary wavelets undergo mutual equal magnitude. When these wavelets are super-
interference. posed, the upward and downward displacements can-
cel each other and give a zero resultant displace-
It is the constructive and destructive interference of ment.
the secondary wavelets that accounts for the ripple The distribution shown in Figure 22.4a cuts across
effects in the diffraction patterns. one complete spatial period of the wave. Consequent-
Diffraction 477

Pt

very far

b) FIGURE 22.2. a. Single-slit diffraction, b. Ir-


->-P radiance in single-slit diffraction pattern.

ly, the length of the line segment BC must equal one or


wavelength, and the first minimum in a single-slit dsin = . (22.1)
diffraction pattern occurs at
For angles greater than the first minimum, there is
B C = . no longer complete destructive interference, and the
resulting irradiance increases for awhile. Figure 22.4b
From the triangle ABC,
shows the distribution along the line BC that results in
AB sin 0 = , the first maximum away from the center (at approxi-

|\

FIGURE 22.3. Geometry of single-slit diffraction.


478 Geometrie, Physical, and Visual Optics

/ ^ ^ ^
d = 10X
a)

b)

w^ a) 15 10 10 15

/mwAwc d = 5X
c)
FIGURE 22.4. Wave displacement (phase distribution) along
the line AC of the previous figure, a. First minimum, b. Next
maximum, c. Second minimum.
b)

mately BC = 3/2 ). Here the downward lobe cancels

n
one of the upward lobes. The first maximum is then
due to the remaining upward lobe. Due to the cancel-
d=X
lations, the resultant displacement at the first max-
imum is significantly smaller in magnitude than the
central maximum.
Figure 22.4c shows the distribution along the line
15 10 5 10 15
A C that occurs when BC equals 2 . Here the two
-
upward lobes are cancelled by the two downward
lobes, and the resultant displacement on the distant FIGURE 22.5. Single-slit diffraction irradiances as the slit
screen is again zero. This is the second minimum width varies.
location, and the angular location is given by
d s i n 0 = 2 A.
Most of the irradiance is spread out between the
In general, the minimum locations are given by
center maximum and the 5.7 angular location for the
d sin 0 = m A, where m = 1, 2, 3, . . . . first minimum.
(22.2) Suppose the slit width is halved (becomes 5 A).
Then the first minimum location is given by
The secondary maximums occur approximately half-
way between the local minimums. 5 A sin 0 = A,
Note that Eq. 22.2 is an equation for the minimums
in a single slit diffraction pattern. The structure of the or
equation is easy to confuse with Eq. 21.13, which is 0=11.5.
the equation for the maximums in the double-slit
Decreasing the aperture size from 10 A to 5 A caused
interference pattern. Be careful not to confuse the two
the diffraction pattern to spread out about twice as
situations.
far.
Since most of the irradiance occurs inside the first
An aperture with a width d equal to one wave-
minimum, we can use Eq. 22.1 to learn about single- length A causes the first minimum to fall at 90.
slit diffraction. Consider a slit with width d equal to 10 Consequently, no ripples are formed, and a monotonie
times the wavelength . Then from Eq. 22.1, fall-off in the irradiance occurs. The three cases dis-
10 A sin 0 = A, cussed are shown in Figure 22.5.

or EXAMPLE 22.1a
sin 0 = 0.10, A slit of width 0.5 mm is illuminated with 633 nm
and red light from a helium-neon laser. What angular
location gives the first minimum in the single-slit
0 = 5.7 diffraction pattern on a distant screen? On a screen
Diffraction 479

10 m from the slit, what is the linear distance Twist the dinner fork about its handle so that the
between the central maximum and the first min- cross-sectional area between the tines decreases. As
imum? soon as the cross-sectional area is small enough, the
From Eq. 22.1, diffraction pattern will be evident. (A potental con-
(0.5 x 10~3 m) sin = 633 x 10~9 m, versation topic for a candle light dinner.)
or
6.33 x l O - 7
Sm=
5X10- '
22.3 Diffraction by a Square Aperture
sin 0 = 1.27 xlO" 3 .
From the small angle approximation, When a single slit is vertical, the waves are diffracted
0 = 1.27x 10"3 radians, horizontally (Figure 22.6a). When the slit is horizon-
tal, the waves are diffracted vertically (Figure 22.6b).
or in minutes When a vertical slit is placed against a horizontal
= (1.27 x 10"3 radians) (57.3 degrees/radian) slit of the same width, the combination forms a square
x(607degree) = 4.35\ aperture. To determine the diffraction pattern for the
Here the amount of diffraction is quite small. For a crossed slits (or square aperture), we can first consider
screen 10 m from the slit, the linear location of the the horizontally diffracted waves leaving the vertical
first minimum away from the central maximum is slit. Each of the horizontally diffracted waves is as-
y = (10 m) tan 0. sociated with one of the maximums in the diffraction
pattern for the vertical slit, and the wave amplitude
From the small angle approximation, decreases rapidly for the higher order maximums.
y = (10 m)0 = (10 m)(1.27 x 10~3 radians), Each of the horizontally diffracted waves then passes
y =1.27 xlO" 2 m =1.27 cm. through the horizontal slit and is diffracted vertically.
The result is a series of vertical diffraction patterns
For a given slit width, the spread in the pattern
depends directly on the wavelength . In fact for centered on each of the horizontal maximums (Figure
small angles, the spread 0 is directly proportional to 22.6c).
the wavelength. Hence long wavelengths (the reds) The total pattern for the crossed slits or square
are diffracted or spread more than the short wave- aperture is dominated by the maximums along the
lengths (blue or violets). horizontal and vertical axes. The other local max-
imums are much fainter due to the amplitude fall-off
EXAMPLE 22.1b with increasing order of the diffracted waves.
For the same slit width, where is the first minimum
for 486 nm (blue) light?
From Eq. 22.1,
(0.5 x 10~3 m) sin 0 = 486 x 10"9 m, 22.4 Diffraction by a Circular Aperture
sin 0 = 9.72 x 10~4.
From the small angle approximation When a circular aperture is illuminated by a mono-
chromatic plane wave, the waves are diffracted equally
0 = 9.72x 10"4 radians, in all directions. The resulting diffraction pattern has a
0 = 9.72 x 10"4 radians (57.3 degrees /radian) central maximum surrounded by a first minimum, and
x(607degree) = 3.34'. then by a series of progressively fainter rings (Figure
As specified by the location for the first minimum, 22.6d). The bright center bounded by the first min-
the blue light is diffracted about 23% less than the imum is known as Airy's disk. (Sir George Airy
red light. (1801-1892) is credited with being the first person to
use spherocylindrical lenses to correct astigmatism,
When white light illuminates a single slit, each and is well known for the Airy integral that occurs in
wavelength is diffracted independently. The result is the mathematical theory of the rainbow.) The boun-
an irradiance distribution with a white center sur- dary for Airy's disk is given by the first minimum,
rounded by colored fringes. The outer part of the which has an angular location of
pattern tends to be reddish since the long wavelengths d sin 0 = 1.22 A. (22.3)
(red) are diffracted the most. You can easily observe a
white light single-slit diffraction pattern by looking As compared to Eq. 22.1, Eq. 22.3 contains the
through the tines of a dinner fork at a match or weird 1.22 number. Perhaps we shouldn't be too
candle. The room illumination should be dim or dark. surprised at this, because we are already familiar with
480 Geometrie, Physical, and Visual Optics

Diffraction Patterns
vertical

Mil I II
slit

a) I - first maximum
- central maximum

horizontal
3 slit
first maximum
-central maximum

D square
aperture

c)
nun m

O circular
aperture FIGURE 22.6. Apertures and resulting diffraction
patterns, a. Vertical slit. b. Horizontal slit. c. Square
d) aperture, d. Circular aperture.

the weird number that appears in the equation for From the small angle approximation,
the area of a circle. The 1.22 comes from the first zero 0 = 1.53 x 10" 3 radians.
of the Jj Bessel function. For those not familiar with
Bessel functions, a more intuitive explanation is given Alternatively,
in the next section. 0 = 0.087 degrees = 5.2'.
Diffraction by a circular aperture behaves as we On the screen 6 m away, the linear location x of the
would expect from single-slit considerations. The first minimum is
smaller the aperture, the larger the diffraction (i.e., x = (6 m) tan 0.
Airy's disk is larger as shown by a larger 0 in Eq.
From the small angle approximation,
22.3). The larger the wavelength the larger the diffrac-
tion, so again red is diffracted more than blue. x = (6 m)(1.53 x 1 0 - 3 radians),
or
EXAMPLE 22.2 x = 9 . 2 x l 0 " 3 m = 9.2mm.
Plane waves of 500 nm are incident on a circular
aperture of diameter 0.4 mm. What is the angular The value x is the radius of Airy's disk, so the
location of the first minimum in the diffraction diameter is 18.4 mm. (Note that the algebraic effect
pattern? What is the diameter of Airy's disk on a of the 1.22 makes the 550 nm circular aperture
screen 6 m behind the aperture? calculations resemble those for 610 nm light inci-
From Eq. 22.3, dent on a single slit.)

(0.4 x NT 3 m) sin 0 = 1.22(500 x 10~9 m), You can observe a white light circular diffraction
(0.4 x 10" 3 m) sin 0 = 610 x 10~9 m, pattern by making a small pinhole in a sheet of
sin 0 = 1.53 x 10"3. aluminum foil. Then stand in a dark or dim room, and
Diffraction 481

look through the pinhole at a small source (e.g., ponding diffraction pattern smoothly changes from
match, candle, a distant light bulb). that of a square aperture to that of a circular aperture.
Figure 22.8 shows the topographical representation of
the irradiance distributions in the corresponding dif-
fraction patterns. In each case the solid lines are the
22.5 Changing from a Square to minimums, the dots are the local maximums, and the
a Circular Aperture dashed lines are the irradiance values at 50% of the
local maximum.
This section provides some intuition about the factor All of the diffraction patterns are dominated by the
of 1.22 that appears in the Eq. 22.3. The method used large central maximum. The minimums in the diffrac-
is to consider the change in the diffraction patterns as tion pattern for the square aperture are intersecting
the aperture is smoothly changed from a square aper- horizontal and vertical lines, while the minimums in
ture of width d to a circular aperture of diameter d. the diffraction pattern for the circular aperture are
The series of apertures in Figure 22.7 shows a circles.
smooth change from a square aperture to a circular We are particularly interested in the first minimum
aperture. The n values listed correspond to aperture surrounding the central maximum. Because of the
boundaries specified by intersecting horizontal and vertical lines, the first min-
imum is not well defined for the square aperture. The
[xn + y n ] 1 / n = l.
n = 20 aperture is basically square except that the
The square aperture is the limit for large n values corners have been rounded in. A comparison of Fig-
(n>). The circular aperture is given by n = 2. As n ures 22.8a and 22.8b shows that the corresponding
is decreased to 2, the aperture smoothly changes to diffraction patterns are also quite similar. The major
circular. change is that while the square pattern had intersect-
As the smooth aperture change is made, the corres- ing horizontal and vertical minimums, the n = 20 pat-

SQUARE

n=3

FIGURE 22.7. Smooth change of a square aperture to a circu-


lar aperture.
482 Geometrie, Physical, and Visual Optics

Square

,
i
v

,y
1
r"
^^
(
\
v
' )
* N


y ~~"*"**<+ ^""s
N
/ N '
/ \

1 i
/
\ *
\ !
* / \ / v
/
\ / \ * *. * / \- J
^^

o;
r^~
'"
^ _ : ^
"N
;


n = 20

s~ \
/ \
\ /

u
/ \
/ \ / \
I
I !
v /

FIGURE 22.8. Diffraction patterns corresponding to


the apertures of Figure 22.7. The four crosses in c. e.
( : ) mark the positions of the horizontal and vertical min-
imum intersection points in the square aperture dif-
fraction pattern.
Diffraction 483

n=6

FIGURE 22.8. (Cont'd).


484 Geometrie, Physical, and Visual Optics

e) FIGURE 22.8. (Cont'd),

tern has minimums that do not quite intersect. In 22.9) is given by Eq. 22.1. In the small angle approxi-
particular, there is a clearly defined first minimum that mation, the angular location is /d. From the Pyth-
surrounds the central maximum. agorean theorem, the distance to the first diagonal
It is now easy to visualize how the pattern will minimum q0 is given by the V2A/d.
change as the apertures are changed from n = 20 to Since the first minimum for the circular aperture
n = 2. The diffraction pattern will circularize. In par- lies approximately halfway between the axial and diag-
ticular, each minimum curve will circularize. To follow onal locations, we have
how this occurs, the four intersection points closest to (1 + V 2 ) / 2 ~ 1 . 2 1 .
the central maximum in the square pattern are marked
on the n = 6 through the n = 2 patterns. It is clear A more exact treatment shows that the 1.21 value is
from these marks that the first minimum is circulariz- actually 1.22.
ing by moving in toward the central maximum along Here we gained some intuition about the 1.22 that
the diagonals. On the other hand, the first minimum is appears in Eq. 21.3 for the first minimum in the
moving out along the horizontal and vertical axes (see
especially the n = 3 to n = 2 step). In fact, the total
inward movement along the diagonal is approximately y0
equal to the total outward movement along the axes.
In other words, the first minimum for the circular
aperture lies approximately halfway between the diag-
1
onal and axial locations of the first minimum for the
square aperture. i
In the diffraction pattern for the square aperture, - * - ( X/d) ^ xo

the angular location for the first minimum along either FIGURE 22.9. Geometry for the diagonal location of the first
the horizontal or vertical axis (x0 and y0 in Figure minimum in the square aperture diffraction pattern.
Diffraction 485

circular aperture diffraction pattern. The method was Figure 22.11b shows cross sections of a circular
to consider the change in the diffraction patterns as aperture with incident plane waves. Immediately be-
the aperture is smoothly changed from square to hind the aperture, the illumination is exactly as pre-
circular. The result was that the first minimum moves dicted by the law of rectilinear propagation (geometric
approximately equal distances in along the diagonals optics). As the screen is moved away from the aper-
and out along the axes. This consideration alone gives ture, we enter the Fresnel or near field zone. Here the
us a factor of 1.21, which is within 1% of the exact illuminance on the screen starts deviating from that
1.22 value. predicted by geometric optics and we start getting the
diffraction ripple effects. These ripples give illumina-
tion in umbra zones together with decreased illumina-
tion in the unobstructed zones. The ripples change
22.6 Fraunhofer vs Fresnel Diffraction character as we move through the Fresnel zone, finally
blending into those of the Fraunhofer or far field
In each of the obstacles and apertures discussed diffraction zone. Once into the Fraunhofer zone, the
above, plane waves were incident and the diffraction ripples no longer change character but maintain the
pattern was viewed on a distant screen (Figure same angular relationship.
22.10a). In these cases, we can consider the diffracted Fresnel diffraction can result in some counterintui-
waves as plane waves (Figure 22.10b), and we refer to tive behavior of light. The famous Poisson spot is an
the diffraction as far field. Far field diffraction is also example. In 1818, Fresnel had presented his mathe-
known as Fraunhofer diffraction. (Joseph von Fraun- matical work on near field diffraction to a committee
hofer (1787-1826) did much work on diffraction grat- in Paris. One of the committee members was the very
ings and optical spectra.) adept mathematician Simeon Poisson. Poisson studied
When the waves incident on the aperture are con- Fresnel's work, and then pointed out that if the work
verging or diverging, or when the distance from the was correct, there should be a maximum (a bright
aperture to the observation screen is small, the diffrac- spot) directly in the middle of the umbra region at a
tion is referred to as near field or Fresnel diffraction. specific distance in the near zone behind a circular
In this case, the curvature of the diffracted waves is obstacle. Poisson felt that this prediction was ridicul-
important (Figure 22.11a). ous and that Fresnel's work was wrong. On hearing
One of the differences between Fresnel and Fraun- Poisson's comments, one of the other committee
hofer diffraction is that the Fresnel patterns can
change significantly as the distance from the aperture
is varied. On the other hand, the Fraunhofer patterns
are not sensitive to distance changes, and thus, we can
express the Fraunhofer pattern by its angular location
as long as the screen is in the far field zone.
short distance-

a)

large
) distance ]

a)

I
Far
Near
< # (Fresnel)
(Fraunhofer)

large
distance
*/,'/.
b)
b) ''/
FIGURE 22.10. a. Geometry for Fraunhofer or far field dif- FIGURE 22.11. a. Geometry for Fresnel or near field diffrac-
fraction. b. Plane wave representation for Fraunhofer diffraction. tion. b. Variation of diffraction patterns.
486 Geometrie, Physical, and Visual Optics

members, Dominic Arago, set up the experiment and 22.7 Diffraction Limit on Resolution
found that the bright spot does exist.
Some people perceive dark floaters in their eyes. In the previous section, we saw that an image point is
These floaters are naively described as shadows cast actually a Fraunhofer diffraction pattern. In other
on the retina by debris, such as blood cells, floating in words, when the light waves pass through the lens,
the vitreous humor. However, these floaters are not they suffer some diffraction effects (usually small).
geometric optics shadows but rather Fresnel diffrac- Thus a perfect optical system (correct focus, all aberra-
tion patterns formed by the debris. (A good way to tions corrected) still does not form a perfect point
observe floaters is to look at a uniformly illuminated image. Instead the light is spread out some due to the
blank field such as a gray sky.) diffraction. This means that diffraction places a limit
Figure 22.12a shows a way to observe Fraunhofer on the ability of the system to form good quality
diffraction in situations where space is limited. The images. In other words, diffraction places a limit on
light from a near point source is collimated by a lens. the ability of the system to transmit perfect informa-
The collimated light (plane waves) passes through the tion about the object. Optical systems that are "per-
diffracting aperture. The diffracted waves are also fect" (correct focus, aberrations sufficiently corrected)
plane but have an angular deviation. A second lens is are referred to as diffraction limited systems.
then used to bring the diffracted plane waves to a The ability of an optical system to resolve two
focus. The Fraunhofer diffraction pattern appears in points is one example of the detrimental (or limiting)
the image plane of the second lens. effects of diffraction. Figure 22.13a shows two point
The two lenses and aperture can be combined into sources (A and B) in front of a lens with a circular
one lens with a finite aperture size (Figure 22.12b), in aperture. The point sources subtend an angle at the
which case the Fraunhofer diffraction pattern appears nodal point of the lens, and so the geometric image
in place of the geometric optics point image. The points A' and B' also subtend the angle at the nodal
geometric optics assumption is that the lens apertures point.
are large and the diffraction spread is small enough Because of the circular aperture, each image point
that the result is effectively a point image. However, if is actually an Airy's disk with its surrounding rings.
the lens aperture is small, or if we are interested in For simplicity we will neglect the surrounding rings
very fine detail, then the difference between a point and just consider Airy's disk.
image and the Fraunhofer diffraction pattern becomes When the aperture is large, there is little diffraction
important. and Airy's disk is small. In this case, the two illumina-
tion patterns are very close to points, and it is easy to
tell from the image plane that two points existed in
object space. As the aperture size is decreased, the
diffraction increases. Now there is a sizable Airy's disk
centered on each image point (A and B'). Initially, A
and B', are far enough apart that the two Airy's disks
do not overlap. Hence we can still tell from the image
plane that there were two point sources in object
space (i.e., we can still resolve the two points).
a) When the aperture is decreased still further, the
diffraction increases and Airy's disks get larger and
begin to overlap (Figure 22.13b). An experienced
observer looking at the image plane can still tell that
there are two point sources in object space. As the
process continues (aperture size decreases, diffraction
increases), the Airy's disks eventually overlap until
they look like one large diffraction pattern. Here even
experienced observers (as well as automated photo-
detection systems) can no longer tell that there were
two points in object space. In this case, we say the two
points are not resolvable.
b) There are several criteria for the resolution limit.
FIGURE 22.12. a. System to observe Fraunhofer diffraction in Rayleigh's criterion is one that is widely used because
a small area. b. Fraunhofer diffraction pattern at the image of its simplicity. (John William Strutt, the Baron
point. Rayleigh, 1824-1919, was very active in optical re-
Diffraction 487

well resolved

b) just resolved

Rayleigh's
criterion

FIGURE 22.13. Two point


resolution, a. Geometry, b. Dif-
fraction patterns, c. Irradiance
distributions.

search.) Rayleigh's criterion states that the two point In diffractions-limited systems, the larger apertures
resolution limit occurs when the first minimum (the give the best resolution. Note, however, that large
boundary of Airy's disk) of one diffraction pattern aperture systems are usually limited by aberrations as
coincides with the center maximum of the other dif- opposed to diffraction. The smaller apertures help
fraction pattern. When the patterns overlap more than with the aberrations but make diffraction worse.
this, the points are not resolvable. When the patterns
overlap less than this, the points are resolvable.
EXAMPLE 22.3
Equation 22.3 gives the angular location for the
Two dots are on a chart 6 m from a pinhole camera
first minimum in the diffraction pattern for a circular with a diameter of 0.5 mm. For light of wavelength
aperture. For the resolution limit, the angles are 550 nm (the middle of the visible spectrum), how
small, and then Rayleigh's criterion states that the just close can the two dots be and still be resolved?
resolvable angle 0r is From Eq. 22.4,
0r = 1.22A/d, (22.4) 0r = 1.22(550 x 10"9 m)/(0.5 x HT 3 m),
where d is the diameter of the circular aperture (or of 0r = 6.71xlO~7/5.Ox 10"4,
the lens rim). Then for two points subtending an angle 0r = 1.34x 10"3 radians,
0 at the lens: there is resolution when 0 > 0r and no (or 4.61').
resolution for 0 < 0 r . Figure 22.13c shows intensity From the rays in Figure 22.13a, the separation
plots for an unresolved, just resolved (Rayleigh limit), of the two points is
and resolved situation. y = tan 0(6 m).
488 Geometrie, Physical, and Visual Optics

From the small angle approximation, 1.22


s = 2 n sin i '
y=0(6m),
y = (1.34xl0~ 3 )(6m), where is the vacuum wavelength, 2i is the angle
y = 8.04x 10"3 m = 8.04 mm. subtended by the objective lens at the object plane,
and n is the object space index (normally air, but can
be oil as in the oil-immersion objective). The term
EXAMPLE 22.4 (n sin i) is called the numerical aperture, and is printed
A +5.00 D thin lens has an overall diameter of
3.2 cm. What angular separation gives the resolu- on microscope objectives together with the relative
tion limit for two distant point objects emitting magnification. Typical microscope objectives have nu-
light of wavelength 589 nm? Also, how many wave- merical apertures ranging from 0.16 up to 1.25. In
lengths apart are the two image points when just special cases, the numerical aperture may get as high
resolvable? as 1.5.
From Eq. 22.4,
0r = 1.22(589 x 10"9 m)/(3.2 x 10~2 m), EXAMPLE 22.6
A diffraction-limited 50 X microscope objective has
0r = 2.25x 10"5 radians, a numerical aperture of 1.22. For 600 nm inco-
(or 4.63 seconds). herent illuminance, how close can two points be
From the rays, the separation of the two image and still be resolved by the microscope?
points is In this case,
y = tan0(f 2 ). 1.22(600 nm)
s= = 300nm.
The secondary focal length f2 is 20 cm. Then from 2(1.22)
the small angle approximation,
y = 0(20 cm), In diffraction-limited systems, the resolution ability
5 clearly depends on the wavelength. Long wavelengths
y = (2.25xl0" )(20cm), (reds) are diffracted the most, while short wavelengths
y = 4.50 x 10~4 cm = 4.50 x 10"6 m. (blues and violets) are diffracted the least. Therefore,
The number of wavelengths is 4.50 x 10"6 m/589 x resolution is better for the short wavelengths. This is
10 9 m, or 7.6 wavelengths. the reason why the ultraviolet microscope has better
resolution than the visible light microscope.
In complicated optical systems, the aperture stop Electrons are quantum particles as opposed to elec-
provides the maximum amount of diffraction. The tromagnetic radiation. However, all quantum particles
image space effects of this diffraction are figured by propagate according to wave equations and so suffer
using the exit pupil. The object space effects are from diffraction. The electron's wavelength is much
figured by using the entrance pupil. For a telescope, less than that of visible light. Thus, the electron
the object space angle for two point resolution is then microscope gives resolution that is orders of mag-
easy to find. nitude better than the ultraviolet microscope.

EXAMPLE 22.5
A 7X Keplerian telescope has a +2.00D f/12.5
objective lens. What is the angular resolution limit 22.8 Diffraction and Acuity
of the telescope for 486 nm light?
The objective lens is the entrance pupil of the At large pupil sizes, the human eye does not form a
telescope. The f-number is the focal length divided perfect point image due to aberrations. However, at
by the diameter of the lens, so the objective lens
diameter is 50 cm/12.5, or 4 cm. Then Eq. 22.4 smaller pupil sizes (2.4 mm) or less, the human eye
gives appears to be a diffraction-limited system. We can use
a simple thin lens and screen model of the human eye
0r = 1.22(486 x 10"9 m)/(4 x 10"2 m), to investigate the effects of diffraction on visual acuity.
0r = 1.48x 10"5 radians, The most familiar acuity system is Snellen acuity
(or 3.06 seconds). (20/20, 20/40, 20/60, etc.) When expressed as a deci-
mal, Snellen acuity is equal to the reciprocal of the
For a microscope, we want to know the minimum smallest resolvable angular detail expressed in min-
distance s between two resolvable points in the object utes, for example, 20/40 = 0.5, which has a reciprocal
plane. From Rayleigh's criterion, we can show that the of 2. Therefore, the smallest letters that a person with
minimum distance s between two resolvable incoher- 20/40 acuity can resolve have detail that subtend an
ent point sources is given by angle of 2'. The 20/20 letters have detail that subtend
Diffraction 489

an angle of , and 20/60 letters have detail that 0r = (6.77 x 10 4 radians)(57.3 degrees/radian)
subtend an angle of 3'. x (607degree),
Snellen acuity is considerably more complicated
than two point resolution. Nevertheless, we can gain 0r = 2.33'.
intuition by using Rayleigh's criterion to estimate Then
Snellen acuity. Consider a thin lens and screen model 1/2.33' = 0.43 = Snellen acuity,
of the human eye with a 2.4 mm pupil size. According
to Rayleigh's criterion, what is the expected visual or
acuity for the middle of the visible spectrum (555 nm)? 0.43 = 20/x,
From Eq. 22.4, the smallest resolvable angle is and
0r = 1.22(555 x 10~9 m)/(2.4 x 10"3 m), x = 20/0.43 = 46.6.
0r = 6.77xlO~ 7 /2.4xlO~ 3 , So in this case, the expected visual acuity through
the pinhole is about 20/50.
0r = 2.82x 10~4 radians. According to this example, a person with 20/20
The conversion to minutes is acuity would lose acuity when looking through the
1 mm pinhole. A person with about 20/50 acuity
0r = (2.82 x 10"4 radians)(57.3 degrees/radian) would show no change when looking through the
x (607degree), pinhole. A person with 20/250 acuity due to optical
blur would gain to 20/50 when using the pinhole.
0r = 0.97'.
Then
1/0.97' = 1.03 = Snellen acuity,
22.9 Diffraction Halos
or
1.03 = 20/x, Consider a random array of many small circular aper-
and tures. If this array is illuminated by plane waves from
a white point source, each aperture will generate an
x = 20/1.03 = 19.4. Airy type diffraction pattern. When the apertures are
So the predicted Snellen acuity is 20/19.4, or essential- small and close together, the diffraction patterns are
ly 20/20, which is about what we would expect. large and overlap. The overlapping diffraction pat-
The pinhole is used clinically to see if poor acuity is terns produce a readily visible halo, namely, a central
due to optical causes (blur) or pathological causes white disk surrounded by circular colored rings (red
(e.g., senile macular degeneration). If the poor acuity outermost, blue or violet innermost). Similar halos
is due to blur, then the pinhole effect minimizes the occur when the diffraction is due to a random array of
blur and increases the acuity. If the poor acuity is not circular obstacles.
due to optical blur, then the pinhole gives no improve- The obstacles do not even have to be opaque.
ment in acuity. Suspended water (n = 1.33) droplets in air (n = 1.00)
What happens when an emmtrope with good acu- can give diffraction halos. When observed through a
ity looks through a pinhole? Here the pinhole intro- light cloud cover around the sun or moon, these
duces diffraction, and the person's acuity may actually diffraction halos are referred to as coronas. The dif-
decrease through the pinhole. Again, we can use a fraction halos are easy to distinguish from ice crystal
thin lens and screen model to predict the acuity halos. Ice crystal halos are due to refraction and
through a pinhole. dispersion by the ice crystals; they have red on the
inside of the rings, whereas diffraction halos have red
on the outside of the rings.
EXAMPLE 22.7 The sometimes brilliant halos seen through fogged
Based on Rayleigh's criterion, what is the expected up car windows are diffraction halos. You can easily
visual acuity through a pinhole with a 1 mm diam- see such halos by breathing on the side of a clear glass,
eter? and then looking through the fogged area at a small
Let us use the middle of the visible spectrum or
the wavelength (555 nm). From Eq. 2.4, the smal- source (e.g., match, penlight, or distant bulb).
lest resolvable angle is When the cornea swells (becomes edematous),
small droplets of fluid form randomly between the
0r = 1.22(555 x 10~9 m)/(1.0 x 10~3 m), stromal fibers. These random droplets produce a dif-
0r = 6.77x 10 -4 radians. fraction halo that the person sees when looking at
The conversion to minutes is lights. Such halos are one of the warning signs of high
490 Geometrie, Physical, and Visual Optics

ocular pressures produced by uncontrolled glaucoma. interference maximums are modulated by the single-
These halos can also be produced by epithelial damage slit diffraction pattern (Figure 22.14c, where the two
due to poorly fitting contact lenses. patterns are superposed). The double-slit separation is
the same for Figures 22.14a and 22.14c; only the slit
width is different. Note that the single-slit minimums
can actually wipe out some of the double-slit max-
22.10 Diffraction Gratings imums (the missing orders).
Let us keep each slit width very narrow, and
Interference is clearly present in the phenomenon of consider the effect of adding more slits. Initially we
diffraction. In fact, the distinction between the two is have the double-slit pattern in which all the interfer-
sometimes murky. We did not worry about diffraction ence maximums have essentially the same irradiance
in our initial discussion of the double-slit experiment (Figure 22.15a). What changes occur in the pattern
because the width of each slit was so small that the when a third slit (at the same spacing) is added?
first diffraction minimums were very far out (Figure The minimums will certainly be affected. For the
22.14a). When the slit width is larger, the diffraction double-slit case, the minimums result when the two
minimums are closer to the central maximum. Figure waves destructively interfere. Consider a double-slit
22.14b shows the right half of a single-slit diffraction minimum at position A. With three slits, three waves
pattern for this case. For this slit width, the double-slit are incident on A. Two of these waves still destruc-

a)

FIGURE 22.14. a. Double-slit interference pattern for nar-


row widths, b. Single-slit diffraction pattern for wider slit. c.
Double-slit interference pattern for wider slits.
Diffraction 491

added (at the same spacing), it produces a third wave


that arrives in phase with the other two. Therefore,
the maximum positions remain at C. The net effect is
that the main maximums still occur at the same posi-
tions, but now there is a little bump in the irradiance
between each main maximum (the relative irradiance
distribution is shown in Figure 22.15b).
Suppose we now go to four slits (at the same
spacing). The fourth wave will arrive in phase at C, so
the location of the major maximums is unchanged.
However, the minimums are again affected. The
three-slit minimums occurred at B where the three
incoming waves resulted in complete destructive inter-
ference. At B, the fourth wave now gives some ir-
radiance. On either side of B, there are positions D
where the superposition of the four waves gives a zero
resultant displacement (destructive interference).
Consider position A, which is halfway between the
major maximums. For the three slits, A was the center
of the little bump; while for two slits, A was a mini-
mum. For four slits, A is again a minimum because
B A B each pair of two waves gives destructive interference
(i.e., the four waves pair off two by two). For four
slits, the result is two little bumps between each major
maximum (Figure 22.15c).
Because of the little bumps, the irradiance spread
around each major maximum gets narrower as the
number of slits increases, i.e., in the two-slit case, the
irradiance falls off slowly around a major maximum,
while in the four-slit case, the irradiance falls off much
quicker.
In general, n equally spaced slits produce major
maximums at the same angular positions as the
D B A B D double-slit maximums:
h sin = m , m = 0 , 1 , 2, 3, . . . ,
where h is the distance between slits. As n is in-
4 creased, each major maximum becomes narrower and
narrower. Between each major maximum are n 2
little bumps. However, the maximum irradiance in
each little bump gets smaller and smaller as n in-
m=2 m=1 m=0 m=1 m=2 creases. For a very large number of slits, the ir-
radiance of the little bumps becomes negligible, and
d) all the light is channeled into very narrow major
FIGURE 22.15. Interference patterns for the same slit spacing.
maximums. A diffraction grating consists of many slits
a. Two slits, b. Three slits, c. Four slits, d. Many slits (a with a very small slit separation. Figure 22.15d shows
diffraction grating). the zero, first, and second order maximums for mono-
chromatic light incident on a diffraction grating.
For the zero order maximum, there is no angular
tively interfere, leaving the third wave to produce separation of incident wavefronts. However, for the
some irradiance at A. On either side of A, there are first and higher order maximums ( m > l ) , the max-
positions B where the resultant displacement of the imum locations are wavelength-dependent. Since the
three waves is zero (the three waves give destructive maximums are very narrow, diffraction gratings be-
interference). At the double-slit maximum positions come a fantastic tool for analyzing the spectrum of
C, the two waves arrive in phase. When the third slit is wavelengths present in a source.
492 Geometrie, Physical, and Visual Optics

it

m=1 m=0 m=1


a)

Central E E
c c
(mixed) in <o
o co FIGURE 22.16. a. Central and first order diffrac-
tion grating maximums for the sodium doublet wave-
lengths. b. Central and first order maximums for a
hydrogen source.

Consider a low pressure sodium source. The source distant car, street, or porch light. The screen acts like
emits the two sodium doublet wavelengths at 589.0 nm a pair of perpendicularly crossed diffraction gratings,
and 589.6 nm. For a small number of slits (or for and the light appears to streak out in a cross-shaped
dispersion by many prisms), these two wavelengths pattern. (Recall the discussion of diffraction by a
overlap. Because of this, people at one time thought square aperture.) The spacing of the screen may not
there was only one sodium wavelength at 589.3 nm be fine enough to resolve the colors in the source's
(the mid-point). Diffraction gratings give narrow spectrum. You can do even better by looking through
enough maximums to resolve the sodium doublet a pair of sheer curtains at the same light. The threads
wavelengths (Figure 22.16a). in the curtains are closer together than those of the
A low pressure hydrogen gas discharge tube emits screen.
only certain specific wavelengths. A diffraction grating Many feathers are natural diffraction gratings and
can be used to resolve the wavelengths emitted. Figure give very nice spectrums. The ribs on a phonograph
22.16b shows the central and first order maximums. record act like a reflection-type diffraction grating
Besides the F (486 nm) and C wavelengths (656 nm), when viewed near grazing incidence. Video disks have
hydrogen also emits two violet wavelengths in the smaller rib spacings and give diffraction grating colors
visible spectrum. even from small angles.
For double slits, the irradiance of the interference Panty hose give brilliant diffraction halos due to the
maximums depended on the relationship between the randomly occurring orientation of the threads. (These
slit separation and the coherence of the incident light. act like many different crossed diffraction gratings).
The coherence must be high for the larger slit separa- Silk gives similar results. (In fact, the colors produced
tions. For smaller slit separations, the coherence can by silk were the motivation for David Rittenhouse's
be lower. For a diffraction grating (many slits), the slit invention of the first diffraction grating in 1785. Later,
separation is automatically very small, so that the in 1816 Joseph Fraunhofer independently invented the
coherence requirements are not particularly high. diffraction grating and made it famous with a series of
Hence, the maximums from the diffraction grating can spectral experiments.)
be seen with ordinary extended light sources. Perhaps In the crystalline lens of the human eye, radial lens
that is why it is called a diffraction grating as opposed fibers form a randomly oriented diffraction grating
to an interference grating. known as Rabl's laminae on the periphery of the lens.
In a normal eye, Rabl's laminae give a dim halo called
the physiologic halo. Note this halo is different from
the pathologic halos due to corneal swelling. You can
22.11 Simple Observations see the physiologic halo by looking at a bright relative-
ly small source at night. You might stand on a
You can easily observe diffraction grating effects. At sidewalk, and use the lights of a car coming down the
night, look through a window or door screen at a street. (When viewing through a stenopeic slit, only
Diffraction 493

part of the physiologie halo is visible, whereas all of a


pathologic halo from the cornea is visible.) The radial
diffraction grating properties of the crystalline lens
also give the radial streaks that people see when
looking at a star or other bright small light source.
These streaks are referred to as the ciliary corona.
FIGURE 22.17. Fresnel zone plates.

22.12 Iridescence in Nature daries. Each higher order diffraction maximum (first
order, second order, third order, etc.) results in a
Diffraction grating effects are surprisingly common in different point image. More recently, Dennis Gabor
nature. The iridescent green on the neck of a male pointed out a modification in which only the zero and
mallard duck is due to the diffraction grating effect of first order diffracted waves are present. In Gabor 's
the feathers. Similar natural diffraction gratings give modification, the zones change transmission according
the iridescent blue wings of the Morpho butterflies, to a sinusoidal variation. For collimated plane waves
and even the beautiful colors of the eye of the pea- incident on Gabor's zone plate, there is a central order
cock's feathers. The metallic-like sheen from a num- diffracted wave, a first order outgoing (diverging)
ber of beetles is due to the diffraction grating structure diffracted wave, and a first order inward (converging)
of their shells. The metallic green reflection from a diffracted wave. The inward diffracted wave forms a
cat's eye at night is due to a layered structure in the real image point at a focal distance +f, while the
cat's retina (a reflection grating). outward diffracted wave forms a virtual image point
Diffraction gratings give maximums at specific an- (at a focal distance - f ) .
gles that depend on the wavelengths. Therefore, a clue For a distant extended object, the plane wave in
to diffraction grating colors in nature are the satura- each bundle are incident at slightly different angles.
tion of the colors (i.e., fairly monochromatic) and a Hence, the first order diffracted waves for each bundle
dependence of the observed color with viewing angle. leave at slightly different angles. The resulting diffrac-
The gemstone opal is another example of a natural tion maximums form extended images (inverted for
diffraction grating. Opal actually has a series of ran- the real image and erect for the virtual image). Since
domly oriented gratings. The fire or play of color in diffraction depends directly on the wavelength, the
the opal comes from the diffraction grating max- image position (i.e., diffraction maximum) is a very
imums. Shells containing layers of calcium carbonate sensitive function of the wavelength. Thus, zone plates
also give a diffraction grating effect. Examples are have severe chromatic aberration in white light. (Re-
pearls and mother-of-pearl. The silvery pearl-like ap- member that the index of refraction varies only slight-
pearance of some fish scales (e.g., the belly scales of ly with wavelength, so refraction results in much less
herring) come from platelike crystals that lie in the chromatic aberration than a zone plate.)
scale and produce a diffraction grating effect. A Fresnel zone plate can be made by printing dark
circular stripes on white paper, and then photograph-
ing it. The resulting photographic slide functions as
the zone plate. It is more complicated to print a Gabor
22.13 Zone Plates zone plate. However, it turns out that Gabor zone
plates can be made optically by recording a set of
Consider monochromatic plane waves incident on a circular interference fringes on appropriate film. The
diffraction grating made of circular "slits." Because of invention of the laser made the optical method easy to
the circular symmetry, the diffracted waves also have use.
circular symmetry. The most startling phenomenon is Figure 22.18 schematically represents the optical
that a sequence of zones (Figure 22.17) can be found method to make a Gabor zone plate. Mirrors A and D
for which the inward diffracted waves are spherical. are partially reflecting and partially transmitting. Mir-
This means that the diffracted waves form a point rors E and F are plane reflectors. An expanded laser
image, but if we can form a point image by diffraction, beam is split in two by A. One part of the beam is
we can also form an extended image by diffraction. In converged by lens B to form a point image at C
that sense a zone plate functions like a lens. (actually a small Airy's disk and ring pattern). The
Fresnel zone plates were the earliest zone plates diverging spherical waves from the point C are then
made. The transmission across a Fresnel zone plate superposed with the plane waves from the other part
changes abruptly from 0% to 100% at the zone boun- of the beam. Because of the phase differences, a set of
494 Geometrie, Physical, and Visual Optics

B awarded the Nobel prize in physics for his discovery of


holography.

iiMiti^M]'
A Gabor zone plate is essentially a hologram of a
point source. Suppose we want to make a hologram of
two point sources separated both laterally and lon-
gitudinally. We could do this by separating the initial
laser beam into three parts (A, B, and C). We con-
verge A to form one of the points, and B to form the
other. Then C is used to form two sets of circular
interference fringes, one from the superposition with
A, and one from the superposition with B (Figure
22.19a). The two sets of circular fringes are recorded
by direct exposure of the film (Figure 22.19b). (Cross
FIGURE 22.18. Optical method to make a Gabor zone plate by modulation between the two sets of fringes can limit
film exposure to an interference pattern. quality of the hologram).
When the hologram is properly reilluminated, it
generates first order diffracted waves that propagate
circular interference fringes are formed by the super- inward and outward. The two first order waves that
position. When the appropriate film G is placed at the are diffracted outward are spherical. One has A as its
superposition position, the interference fringes are center of curvature, and the other has B as its center
recorded on the film. The circular interference fringes of curvature. The zero order wave is plane and as-
on the developed film have the required spacing and sociated with the wave C. In Figure 22.19c, rays 1 and
sinusoidal transmission of a Gabor zone plate. represent the spherical diffracted wave that has A
A Gabor zone plate forms images by diffraction as its center of curvature (virtual image), while rays 2
even with light sources that we usually think of as and 2' represent the spherical diffracted wave that has
incoherent (i.e., very low coherence). On the other B as its center of curvature. An observer looking at
hand, we needed highly coherent light (such as from a the diffracted waves sees the two virtual image points
laser) to optically make a Gabor zone plate. The spatially separated both longitudinally and laterally.
process of optically making a Gabor zone plate, and (The zero order wave C provides the background
then using it to form an image is better known as illumination.)
holography. Consider a real three dimensional object, such as a
small box in front of a cat. When viewed from an
angle the cat is visible behind the box. When viewed
from straight ahead (axially), the cat is hidden by the
22.14 Holography

In a general sense, a hologram is a very complicated


diffraction grating (or zone plate) that when properly
illuminated results in diffracted waves (maximums)
that form real and/or virtual images. The diffracted
waves are actually reconstructions of the waves that
would come from real objects. Once these waves are
reconstructed, an observer viewing them sees the ob-
jects in three dimensions. Another startling property
of the hologram is that each small section of it con-
tains all the information needed to reconstruct the
whole scene. In other words, when a hologram is cut A 1-4
in half and one of the halves is properly illuminated,
the diffracted wave from that half still forms the whole
image. The Greek work for whole is holos, thus the
name hologram. c)
Holography was invented by Dennis Gabor in 1947
FIGURE 22.19. a. Method to make a hologram of two point
as a method to improve electron microscopy. Visible sources, b. Schematic representation of interference fringes on the
light holograms did not become practical until after holographic film. c. Generation of virtual image points by diffrac-
the invention of the laser in 1960. In 1971, Gabor was ted waves from the hologram.
Diffraction 495

box. This relative shift in position of the objects is and the diffusely reflected light superpose and result in
referred to as parallax, and is a monocular depth a complicated interference pattern that is recorded on
perception cue. the film F. The developed slide is the hologram. If we
Figure 22.20a shows a top view of the set-up to examine the slide through a microscope, we see only
make a hologram of the cat box scene. The cat is the fringes of the recorded interference pattern.
represented by the circle C and the box by the square When the hologram is properly illuminated by light
D. The hologram film is F. Mirrors B and E are plane incident at the same angle as the reference beam,
mirrors, and mirror A is partially reflecting and par- it forms two diffracted waves (the first order max-
tially transmitting. An expanded laser beam is split at imums), together with the zero order or central max-
A. The transmitted part of the laser beam is reflected imum. One of the first order maximums is converg-
at B, and illuminates the cat box scene C and D. The ing (and forms a real image), while the other is div-
diffusely reflected light is incident on the film F. (The erging (and is associated with a virtual image).
rays represent the diffusely reflected light from two of Let us consider the diffracted waves associated with
the points in the cat box scene.) The part the laser the virtual image. These diffracted waves are a recon-
beam reflected at A is called the reference beam. It struction of the waves that would have come from the
remains plane, is reflected at E, and is then incident object. Therefore, the observer sees a conjugate virtu-
on the film F. At the film plane the reference beam al image in three dimensions just as if the objects were

FIGURE 22.20. a. Set-up to make


a hologram, b. Set-up to view the
hologram's image.
496 Geometrie, Physical, and Visual Optics

there. For example, by viewing the hologram from a rainbow, although that is not very accurate scientifical-
side angle near ray 2', the cat is behind the box and ly). The different colors are due to the wavelength
therefore not visible. By viewing the hologram from dependence of the diffraction in the vertical direction.
the normal (between rays 2 and ) , we can see the cat The initial rainbow holograms were actually the
behind the box. (In Figure 22.20b, rays 1 and second hologram of a two hologram process. First a
correspond to the diffracted spherical wave that has a typical hologram is made of the scene (e.g., the cat
center of curvature at a point on the cat, while rays 2 and box). This is referred to as the master hologram.
and 2' correspond to the diffracted spherical wave that Then a laser beam is split into a reference beam and
has a center of curvature at a point on the box.) an illuminating beam. The illuminating beam passes
Let us contrast a hologram with a normal photo- through a horizontal slit and then through the master
graphic slide made with a camera. The information hologram and is incident on a film. The reference
recorded in a normal photograph relates only to the beam is also incident on the film, so that an interfer-
amplitude distribution of the waves. The information ence pattern is recorded (a hologram of a hologram).
(interference pattern) recorded on the hologram re- The developed film is the rainbow hologram.
lates to both the amplitude and the phase distribution When the rainbow hologram is reilluminated, the
of the waves. In technical terms, it is the ability of the light diffracted by it shows the three-dimensional
hologram to record phase information, as well as scene (cat and box) as viewed through the slit. (You
amplitude information, that distinguishes it from the can think of the rainbow hologram as forming an
camera's photographic slide. image of the slit, so even when it is illuminated by an
When the hologram is made, each small part of the extended source, the diffracted light passes through
film receives diffusely reflected light from each point the slit's image.) At the primary viewing angle, typical
on the object. Hence, each small part of the hologram room illumination gives a good quality black and white
contains the whole information. Imagine viewing the image. At higher and lower angles, the diffracted
hologram through a circular aperture. The whole waves give the image in a series of different colors.
three-dimensional image is still present, although the (There are now specialized one-step processes for
angular view is restricted. A small enough aperture making rainbow holograms.)
introduces its own diffraction, and this diffraction The holograms discussed above are transmission
introduces blur into the hologram's image. Thus, if the holograms. Reflection holograms can also be made.
aperture is made smaller and smaller (or equivalently Reflection-type rainbow holograms are now on some
a hologram is cut smaller and smaller), the image credit cards (to hinder counterfeiting), and have ap-
eventually blurs out. peared on several covers of National Geographic
The geometry of making a hologram can be varied Magazine. Reflection-type holograms are also avail-
to emphasize either the real or the virtual image. The able as jewelry items. In each case the three-dimen-
angle of the diffracted waves from a hologram de- sional images are visible under ordinary room illumi-
pends directly on the wavelength of the source. There- nation.
fore, holograms tend to have chromatic aberration Holograms are also used in high tech supermarket
problems. Holograms also tend to have image quality checkout lines to help the electro-optical system read
problems when illuminated by extended sources. the bars and stripes of the universal product code
Many of these problems have been overcome in what (UPC). Typically, laser light is passed through a rotat-
we now refer to as second generation holograms. One ing disk of 21 wedge shaped holograms each of which
method of minimizing both the extended source and deflects the beam in a different way. At least one of
chromatic aberration problem is to limit the depth of the beam deflections will reach and be diffusely re-
the hologram's image to a fraction of the viewing dis- flected from the UPC onto a photodetector, which
tance. electronically reads the signal and feeds it to a compu-
The rainbow hologram is an outstanding example ter. The computer has the product's name, size, and
of second generation holography. The rainbow holog- price stored in memory. It gives the price, prints the
ram's image can be viewed with ordinary room illumi- receipt, and updates the store's inventory. (All that
nation. This is achieved by sacrificing vertical parallax for a can of beans!)
(which does not change the three-dimensionality
much, since it is mostly horizontal parallax that we use
as a depth cue). When scanning horizontally, the
person sees the three dimensional image (horizontal Problems
parallax). When scanning vertical, the person does not
perceive vertical parallax, but instead sees the same 1. A single slit has a width of 0.03 mm. For incident
angular view in different colors (hence the name light of wavelength 520 nm, what is the angular
Diffraction 497

location of the first minimum in the single-slit 4. For a distant object plane, what angular separation
diffraction pattern? What is the angular location of in minutes gives the two point resolution limit for a
the first minimum for 620 nm incident light? +20.00 D lens set at f/5.6? Use a 570 nm wave-
2. A circular pinhole has a 0.1 mm diameter. What is length.
the angular location of the first minimum in the 5. Based on Rayleigh's criterion alone, what is the
diffraction pattern when the incident light has a expected acuity for a human eye with a 3 mm pupil
565 nm wavelength? What is the size of Airy's disk and an incident wavelength of 600 nm? (Use a thin
on a screen located 200 cm behind the pinhole? lens and screen model.) What is the expected acu-
Same questions for incident light of 465 nm wave- ity for looking through a pinhole of diameter
length. 0.5 mm?
3. Two dots are on a chart 4 m from a pinhole camera 6. When a diffraction grating has 5,000 lines per cm,
with a 0.2 mm diameter. For light of wavelength what is the angular location of the second order
550 nm, what is the minimum distance between the maximum for incident light with a 543 nm wave-
two dots that gives resolution according to Ray- length? Same question for light of wavelength
leigh's criterion? 533 nm.
CHAPTER
TWENTY-THREE
Scattering,
Dispersion,
and Polarization

23.1 Dipole Radiation refer to the inhomogeneities as particles. (The parti-


cles might be molecules, dust particles, water droplets,
A dipole consists of equal amounts of positive and ice crystals, aerosol pollutants, etc.) Scattering de-
negative electrically charged particles separated by a pends on the size of the particles, the distance be-
distance s. The dipole is electrically neutral, but will tween particles, and the strength of the interaction
emit electromagnetic waves of frequency f when the between the light and the particles. The strength of
charges periodically oscillate toward and away from the interaction depends on the refractive index and the
each other with frequency f. absorption strength of the particles.
Figure 23.1 shows a sequence of electric field cur- In this section, we will confine our discussions to
ves for an oscillating dipole with a vertical axis. As the transparent particles (i.e., weak absorbers), and as-
charges move toward each other, the electric field sume that the incident light is white. We will also
curves form closed loops. Once the charges pass each assume that the particles are far enough apart so that
other, the closed loops propagate away from the the light scattered by different particles is incoherent.
dipole and represent the radiated electromagnetic The total effect of the scattering is then obtained by
wave. Note that no radiation propagates out directly adding the intensities of the light scattered from each
along the dipole axis. particle. (The incoherent scattering condition is met
Atoms or molecules consist of equal numbers of when the spacing between the particles is much grea-
negatively charged electrons and positively charged ter than the coherence length of the light.)
protons. The atoms and molecules are thus electrically The dependence of scattering on particle size has
neutral (as opposed to ions). When electromagnetic two extremes. One extreme consists of particles that
waves are incident on the atoms or molecules, the are much larger than the wavelength of the incident
electrons are driven to oscillate in a direction opposite light. Here scattering is caused by the geometric optics
to the protons. Hence, the incident radiation sets up a processes of reflection and refraction. The scattering
dipole oscillation in the atoms or molecules. The from large particles is strongly peaked in the foward
dipole oscillation causes a reradiation of an elec- direction, and tends to be wavelength-independent
tromagnetic wave of the same wavelength as the inci- (Figure 23.2a). Common examples are clouds, mist,
dent wave. The reradiated light propagates out in the and fog. (Under certain conditions, dispersion pro-
allowed directions, and is a major component of scat- vides an exception to the wavelength independence.
tered light. Rainbows are a vivid example.)
The other extreme consists of particles that are
much smaller than the wavelength of the incident
23.2 Incoherent Scattering light. (Much smaller means less than a tenth of the
wavelength.) Here dipole reradiation is the prime
Section 1.10 pointed out that scattering is due to contributor to the scattered light. The scattering from
inhomogeneities in a medium. For convenience, let us small particles is strongly wavelength-dependent. Re-

499
500 Geometrical, Physical and Visual Optics

low the fourth power dependence, and hence the blue


color of the sky was due to scattering by the air
# molecules.
The two extremes of incoherent scattering are eas-
ily observed on partly cloudy days. The white of the
clouds is due to scattering by the large transparent
water droplets, or ice crystals. The blue of the sky is
due to Rayleigh scattering by the small transparent air
molecules. Since both are made up of transparent
particles, there is no "white" in the clouds, nor "blue"
in the sky. (Something to contemplate while waiting in
a traffic jam!)

23.3 Rayleigh Scattering


Where does the l/ 4 dependence come from?
Rayleigh gave the following dimensional argument.
He pointed out that the amplitude of the scattered
wave is directly proportional to the volume of the
scattering particle. (Volume has dimensions of length
cubed.) Second, he assumed that the scattered wave-
front is spherical. The energy carried by a small part
of the spherical wavefront is inversely proportional to
the surface area of the sphere. (The surface area of a
FIGURE 23.1. Sequence of electricfieldcurves for an oscillat- sphere is 4 times the radius squared.) Since the
ing dipole. amplitude of a wave is proportional to the square root
of the energy carried, it follows that the amplitude of
the spherical scattered wave is inversely proportional
to the radius of the wavefront.
^ " i f ^
The ratio of the amplitude of the scattered wave to
that of the incident wave is dimensionless. Since the
amplitude of the scattered wave is directly proportion-
al to the volume of the particle (length cubed), and
inversely proportional to the radius of the wavefront
(length), there is still an unaccounted factor in the
FIGURE 23.2. a. Scattering by a large particle, b. Scattering denominator that varies as a length squared. The only
by a small particle. other length in this situation is the wavelength of the
light, and so the scattered amplitude must be inversely
proportional to the square of the wavelength. Hence,
lative to large particle scattering, small particle scatter- the scattered intensity (irradiance) is inversely propor-
ing tends to be more isotropic in direction (Figure tional to the fourth power of the wavelength, or
23.2b).
John Tyndall, a contemporary of Lord Rayleigh, ISca,-p (23.1)
was the first person to experimentally confirm that A considerably more detailed mathematical analy-
small particles scatter the short wavelengths (violets sis shows that the l/ 4 dependence is characteristic of
and blues) more than the longer wavelengths. dipole radiation under the conditions that the dipole
Rayleigh showed that theoretically the scattering de- size is much smaller than the wavelength of the light.
pendence depends directly on the fourth power of the
frequency. Since frequency and wavelength vary in- EXAMPLE 23.1
versely, this means that the scattering varies inversely What is the relative Rayleigh scattering intensity
with the fourth power of the wavelength. Rayleigh Irel for hydrogen F light (blue) compared to hydro-
also pointed out that molecular scattering would fol- gen C light (red)?
Scattering, Dispersion, and Polarization 501

white
rea K
> white / \ * * * K \

bluish / 7^-^-1n-T--r^^

FIGURE 23.3. Rayleigh scattering.

FIGURE 23.4. Multiple scattering.

Here
Irel = V I i ,
where I2 and li are the respective scattered inten-
sities. The F wavelength is 486 nm, and the C path. The light initially scattered appears bluish (i.e.,
wavelength is 656 nm. Then from Eq. 23.1, more blues than reds). However, this light loses blues
Irel = (656 nm/486 nm)4 = 3.3. faster than it loses reds as it travels through the
scattering medium. So the light incident on the second
Thus, for Rayleigh scattering, the blue F light is
scattered 3.3 times more than the red C light. Even scattering center is loaded with reds. When the second
more impressive is the fact that 400 nm light is scattering occurs, the relatively deficient blues scatter
scattered 7.2 times more than 656 nm light. more than the relatively abundant reds. Hence, the
scattered light returns toward a balance of the blues
The smoke coming from the end of a cigarette has and reds, and thus appears white. Small grains of
a bluish appearance (especially when viewed against a sugar and salt are colorless by themselves, but larger
dark background). The blue component is due to amounts of sugar and salt appear white because of
Rayleigh scattering by the small smoke particles. multiple scattering. Multiple scattering also contri-
However, the smoke that is exhaled by the smoker butes to the white of clouds, foam, unfilled paper, and
appears white. The exhaled smoke particles are coated the sciera of the eye. Even large amounts of pure air
with water, and are much larger than a wavelength. tend to give a more desaturated blue than small
This gives the wavelength-independent scattering of amounts. Thus, on a clear day, the sky appears bluer
the geometric optics limit. from the top of a high mountain than it does from sea
In a Rayleigh scattering situation, the transmitted level.
light is deficient in the short wavelengths and loaded in Besides multiple scattering, the sky is also desatu-
the long wavelengths. Thus, the transmitted light and rated by dust particles, or other aerosol pollutants.
the scattered light have a complementary appearance Measurements on such skies show a wavelength de-
(Figure 23.3). Sunrises and sunsets appear orange or pendence that varies from l/ 3 to 1/2.
red since the light reaching the observer's eye has The first general formulation of the scattering prob-
traveled a long path through the atmosphere, and has lem was due to Gustave Mie (1868-1957). For small
lost more blues than reds (Figure 1.6). The red color particles, Mie's equations give Rayleigh scattering.
of a sunset is enhanced by the presence of small dust For large particles, Mie's equations give the wave-
particles in the lower atmosphere. (You can simulate length-independent scattering of the geometric optics
atmospheric scattering by placing a few drops of milk limit. With the aid of computers, Mie's equations can
in a clear glass of water, and then illuminating the be numerically solved for any other region (e.g.,
glass with a penlight. The transmitted light will have a particle size of one or two times the wavelength). In
definite orangish appearance, while the scattered light these cases, there is a wavelength dependence, but it
has a faint bluish appearance.) differs from the l/ 4 dependence of Rayleigh scat-
tering.
Diffraction and/or absorption can join together
with Mie scattering to give unusual results. Oil drop-
23.4 Other Incoherent Scattering Effects lets in the atmosphere from large forest fires have on
occasion given scattering that resulted in a bluish
Multiple scattering rapidly decreases any wavelength appearing sun and moon. This is allegedly the origin
dependence. Figure 23.4 shows a double scattering of the phrase once in a blue moon.
502 Geometrical, Physical and Visual Optics

23.5 Coherent Scattering the material. (The mathematical proof of this is


known as the Ewald-Oseen extinction theorem.) The
When the particles are much closer together than the other constructive interference direction is that given
coherence length of the light, constructive and de- by Snell's law, and the resultant scattered wave in this
structive interference occurs between the light waves direction is the refracted wave.
scattered from different particles. Even for ordinary The constructive interference gives a refracted
sources (the sun, light bulbs, etc.) coherent scattering wave that is phase-shifted relative to the incident
occurs when the particles are much closer together wave. Each successive scatterer contributes an addi-
than the wavelength of the light. tional component to the phase shift, thus making it
Coherent scattering is involved with the transmis- appear that the light travels through the medium at a
sion of light through transparent materials that are speed v that is different from c. We refer to the ratio
inhomogeneous on a microscopic level. Figure 23.5 v/c as the refractive index n of the medium. In prin-
shows light (represented by the rays A) in a vacuum ciple, we can derive the refractive index n from the
incident on a regular array of transparent particles in scattering properties in the medium. (Of course, it is
which the spacing is much less than the wavelength of far easier just to use Snell's law to determine it.) Since
the light. Each particle acts as a source for the coher- Rayleigh scattering shows a dependence on the wave-
ent scattered waves. length of the incident light, we would now expect that
The scattered waves propagate outward and over- the refractive index would also show a wavelength
lap. The material acts similar to a diffraction grating in dependence.
that for most directions the resultant superposition of Deep in the material, there is no reflected compo-
the many coherent scattered waves is destructive inter- nent because of the destructive interference of the
ference. However, two directions give constructive scattered waves in the direction of the reflected light.
interference (analogous to the maximums from a dif- However, within a half wavelength of the surface,
fraction grating). One of the constructive interference there are fewer scattering centers with the result that
directions is parallel to the incident light (a forward complete destructive interference does not occur in
scattered component). This component is 180 degrees the reflected direction. In fact, the reflected light
out of phase with the unscattered part of the incident comes from the scattering within a half wavelength
wave, and destructively interferes with it. This inter- zone of the surface. For a transparent material, blue
ference wipes out the unscattered portion of the inci- light scatters more than red light. On the other hand,
dent wave, a very common sense result since we the half wavelength zone for red is larger than that for
usually do not think of the incident wave as existing in blue. The result is that the amount of light reflected
from a transparent material is wavelength-indepen-
dent.

23.6 Intraocular Light Scatter

The corneal stroma consists of collagen fibrils that are


approximately 25 nm in diameter and have an inter-
fibril spacing of 60 nm. These dimensions are much
less than the wavelength of visible light, so coherent
scattering theory applies to the transparent cornea.
Some residual scattering occurs (i.e., the destructive
interference is not complete), and this scattered (or
diffusely reflected) light enables us to examine the
structure of the cornea with a slit lamp.
The corneal endothelium pumps excess water out
LOOQO\0000 of the cornea into the aqueous humor. In cases of
endothelial trauma, the pumping action decreases and
lakes of water begin to form between the stromal
fibers. Once these lakes approach a size of 1/2 the
wavelength, the efficiency of the destructive interfer-
ence in the scattered light declines. This results in
FIGURE 23.5. Coherent scattering. increased scattering, in which case the cornea starts to
Scattering, Dispersion, and Polarization 503

lose its transparency and takes on a greyish appear- angle, and are immediately altered by immersion in a
ance. fluid, whereas Tyndall blues are independent of view-
In the scierai area, the collagen fibers are much ing angle and do not change color rapidly when im-
larger, and the spacing is about a wavelength. Here mersed.
there is little or no coherent scattering, and the sciera Iris color is due to the scattering-selective absorp-
appears whitish due to large particle incoherent scat- tion combination. The human iris is backed by a
tering. In cases of scierai thinning, the fibers get heavily pigmented epithelial layer. The apparent iris
smaller and the sciera takes on a bluish appearance color depends on the amount of pigment that lies in
due to Rayleigh-type scattering. (In fact, if you look the layers anterior to the epithelial layer. When the
carefully, you will see that most normal scieras have a layers in front have little pigment, the eyes are blue.
slight bluish appearance.) The bluish appearance is due to Rayleigh scattering
Another difference between the sciera and the from the fine gossamer stromal fibers in the anterior
cornea is that the collagen fibers in the human cornea layers. When the anterior layers have more pigment,
are laid down in parallel bundles, while those in the the pigment contributes a yellow component. This
sciera are laid down in a random order. However, this yellow coupled with the scattered blue gives green
regularity does not seem to be crucial to transparency. eyes. Still more pigment in the anterior layers results
In 1967 Goldman and Benedek observed that the very in brown eyes. Albino eyes have no pigment. Here the
thick Bowman's zone of a shark cornea has a com- reddish light diffusely reflected (scattered) from the
pletely random arrangement of collagen fibers and yet blood in the iris capillaries coupled with the scattered
still meets the coherent scattering conditions for trans- blue from the stromal fibers to give a pinkish ap-
parency. pearance.
The crystalline lens is not as transparent as the
cornea. The lens has a slight yellow pigment to it, and
also scatters more light. In cataract formation, the
protein in the crystalline lens clumps together and 23.8 Resonance Scattering
causes additional scattering. Thus, a cataract degrades
vision by scattering light. As the cataract grows, the In the previous section, we considered scattering from
amount of scatter increases. Eventually, in order to transparent structures in the presence of a selective
restore useful vision, the cataractous crystalline lens absorber. Here we want to consider scattering from
must be surgically removed. the absorbers themselves.
The healthy retina scatters about as much light as First we need to further consider how the atomic
the cornea. However, when the blood supply is inter- electrons respond to the incident radiation. The
rupted, the retinal integrity is damaged, and edema atomic electrons are bound to the nucleus by attractive
fluid starts collecting in the nerve fiber layer. This electromagnetic forces. For our present purposes, let
increases the retinal light scattering, and that area us conceptually represent the electromagnetic binding
turns milky grey. force by a spring. In this model, we consider the
electrons to have certain natural vibration frequencies
(i.e., the natural vibration frequencies of the spring).
These natural vibration frequencies are called the
23.7 Scattering and Selective Absorption resonance frequencies of the atom.
When an electromagnetic wave is incident on the
In the animal world, Rayleigh scattering and selective atom, it drives the electron into a dipole vibration that
absorption often combine to give a Tyndall blue color. has the same frequency as that of the incident radia-
The basic mechanism is for the scattering particles to tion. When the incident frequency is far from the
appear in front of a layer of dark melanin pigment. resonance frequency f0, the dipole vibration is small
The scattering particles might be collagen fibers, air (Figure 23.6). However, when the incident frequency
vesicles, protein, keratin, or guanine crystals. The f matches the resonance frequency f0, the vibration is
blue in the feathers of a blue jay are caused by greatly amplified.
scattering from air vesicles. Tyndall blues include We can intuitively understand the amplification as
those on baboons, on a turkey's neck, and on many follows. When the incident radiation is at f0, each
reptiles including chameleons. push that it gives the atomic electron matches the
We can distinguish between iridescent blues due to electron's natural vibration direction (just like correct-
diffraction and interference effects (e.g., the Morpho ly pumping a swing). Hence the resulting vibration
butterly) from Tyndall blues (scattering) by the fact becomes very large. On the other hand, when the
that interference colors change rapidly with viewing incident frequency is very different from f0, the result-
504 Geometrical, Physical and Visual Optics

y 1<T>N ,^n ffo

ffo

ffo

FIGURE 23.6. Vibration response of an atomic electron with a


resonance frequency f0 illuminated by light of frequency f.

ing vibration is small, because the resonance amplifica-


tion does not occur. (This is analogous to trying to
pump a swing at the wrong times.) FIGURE 23.7. Absorption strength k for two different damp-
The magnitude of the electron's vibration is limited ing factors.
by the damping factor g. The damping factor depends
on those things that drain off energy from the elec-
tron. These include reradiation, or loss of energy
through atomic collisions (heat). From the differential
equations for a damped oscillator, the square of the Figure 23.7 is a plot of the absorption strengh k as
electron's vibration amplitude is proportional to a function of the incident frequency f. For a small g,
the absorption strength k forms a deep, narrow "hole"
1 centered on f0. For a larger g, the hole is shallower.
( f ; - f T + (gfr Now we are ready to consider resonance scattering.
Consider white light incident on a sodium vapor.
For a medium that is not too dense (few atomic Sodium has a strong resonance (small g) at the sodium
collisions), the damping factor g is small. Then when f D wavelength of 589.3 nm. (Actually, there are two
is far from f0, the first term in the denominator is large overlapping resonancesat 589.0 nm and 589.6 nm
and the vibration amplitude is small. However, when f and 589.3 nm is the average of the overlapping reso-
equals f0, the first term in the denominator vanishes, nances.) For transparent particles (no resonances in
and only the small second term remains. Here the the visible spectrum), we would expect Rayleigh scat-
vibration amplitude is very large (i.e., the resonance tering from the vapor. However, the sodium atoms
amplification). have an amplified vibration response to 589.3 nm, and
For a material, we would expect the absorption hence reradiate very strongly at that wavelength. Con-
strength k to somehow depend on the electron's vibra- sequently, the expected Rayleigh scattering is over-
tion amplitude. We would also expect that k would whelmed by the yellow reradiated light of the sodium
depend on the density of the material. For materials doublet (Figure 23.8). The reradiation of the light at
that are not too dense (e.g., a gas), we can use some the resonance frequency of the atom is referred to as
simplifying approximations together with the elec- resonance scattering (or resonance radiation). As a
tron's vibration amplitude to derive the following ex- consequence, the transmitted light is deficient in the
pression for the absorption strength k: sodium yellow wavelengths.
A good diffraction grating shows that sunlight is
-SEgf missing some very specific wavelengths including the
(23.2)
(fo-fT + (gfr sodium D wavelength. The theory is that these wave-
lengths are emitted by the hot core of the sun, but that
where E is a parameter that depends on the atomic the "cooler" vapors in the sun's outer atmosphere
electron's mass and charge. Note that the denominator undergo resonance scattering. Hence, the transmitted
is the same as that of the square of the electron's light that is incident on earth is deficient in the reso-
vibration amplitude. nance wavelengths of the elements in the vapor.
Scattering, Dispersion, and Polarization 505

white Sodium Vapor white (deficient in 589.3 nm)


>

FIGURE 23.8. Resonance scattering (or reso-


yellow (589.3 nm) nance radiation).

23.9 Dispersion damping factor causes the refractive index curve to


turn down and equal 1 at f0. Above the resonance,
In Section 23.5, we saw that a refracted wave is n decreases some more and then starts to increase
actually made by a coherent scattering process. The again. As f goes to infinity, the refractive index
refracted wave appears to travel through the increases to 1. The behavior of the refractive index
medium at a speed of c/n, where n is the refractive near and above f0 is referred to as anomalous
index of the medium. The refractive index is a dispersion. (The meaning of n < l is discussed in
function of the frequency (or wavelength) of the the next section).
incident light. The fact that the different frequen- Even though the above equations are derived
cies of electromagnetic radiation travel through a under a low density approximation, they actually
material at different speeds is called dispersion. work well for many optical materials. It is clear
Dispersion is responsible for chromatic aberration. from the equations that both absorption and disper-
The higher the refractive index, the slower a sion are tied to the magnitude of the electron's
wave propagates through a medium. We would vibration response to the incoming radiation. In
intuitively expect that the refractive index would that sense, both the absorption behavior and the
depend on the density of the medium. Since a dispersion behavior are determined by the reson-
coherent scattering process makes the refracted ances of a material.
wave, we might also intuitively expect that it would For a material with multiple resonances, the
depend on the vibration response of the electrons in resonance term in Eq. 23.3 is replaced by the sum
the medium. Again, for a material that is not too of the terms for each resonance:
dense (e.g., a gas), we can make some simplifying
SE(ff-f 2 )
approximations, and derive that + (23.4)
2 at f 2 ) 2 + ( g f) 2 '
SE(fg-f )
n= l + (23.3) where m is the number of resonances and fx is the
( f - f 2 ) 2 + (gf)2*
2
i l resonance frequency. Figure 23.10a shows plots
of the index of refraction and the absorption
Note that the denominator of the second term is strength curve for a hypothetical material with
the same as that of the square of the electron's three resonances. Figure 23.10b shows the actual
vibration amplitude. In fact it is the same as the
denominator for the absorption strength k.
Figure 23.9 shows a plot of the refractive index n
vs the incident frequency f for a small non-zero
damping factor g. For f well below the resonance
frequency f0, the refractive index increases slowly
with increasing f. This slow increase is characteristic
of transparent materials and is referred to as nor-
mal dispersion. As f approaches the resonance, the
refractive index starts to increase much faster. For a
zero damping factor, the refractive index n would
go to infinity at the resonance frequency f0. How-
fo f^
ever, the non-zero damping factor prohibits the
refractive index from going to infinity. Instead, the F I G U R E 23.9. Dispersion curve for a single resonance.
506 Geometrical, Physical and Visual Optics

phenomenon (Section 21.6). In a vacuum, both har-


monic waves travel with speed c, and so the beat also
travels with speed c.
However, in a dispersive medium, the two har-
monic waves travel at different phase velocities.
Therefore, the crests of the two waves will not stay
lined up, and the beat is going to propagate at a
different speed than either component. The beat vel-
ocity is slower than that of either component, and is
called the group velocity u. The group velocity u is
related to the phase velocity v by

u = v (dispersive term),
or

u= (dispersive term).

In a vacuum, the dispersive term is zero, and n


f equals 1, so u = v = c. When n is less than 1 (anomal-
b) ous dispersion), the dispersive term is large enough so
FIGURE 23.10. a. Dispersion and absorption curves for a that u stays less than c.
material with three resonances, b. Dispersion curve for crown Light sources have a finite emission time; con-
glass. sequently, wavepackets have a finite length (Figure
21.19). Similar to beats, we can consider the wave-
packets to be made up of a superposition of a number
refractive index (dispersion curve) for clear crown of harmonic waves each of slightly different frequency.
glass. The glass has several strong resonances well In a vacuum all the harmonic components, and con-
below the visible region. The closest resonance to sequently the packet itself, move with the speed c. In
the visible region is the resonance in the ultraviolet. a dispersive medium, each harmonic component
The resonance in the ultraviolet means that normal moves with a different speed. This has two effects:
glass becomes opaque in the ultraviolet. The disper- (1), the packet spreads out as it moves through the
sion in the visible region is due to the nearness of medium; (2), the packet moves slower than the phase
the resonance in the ultraviolet. If that resonance velocity of the harmonic components (exactly analog-
had been further away, the dispersion would be ous to the beat behavior except that more than two
smaller in the visible region. components are involved).
One way to achieve a higher index glass is to The theory of relativity says that energy (informa-
pick a chemical composition that results in the tion) cannot propagate faster than the vacuum speed c
ultraviolet resonance being closer to the visible of light. The packets convey the energy (or the infor-
region. These high index glasses tend to absorb mation), and the relativity limit says that the packet or
some blue (hence, the transmitted light is slightly group velocity must be less than c. For anomalous
yellowish). The high index glasses also tend to have dispersion where n < 1, the group velocity is still less
more dispersion than the lower index glasses. than c because of the dispersive term. (Michelson was
Barium crown glass is an exception in the sense that the first to actually measure a difference between a
it has low dispersion for a high index glass. phase and a group velocity. He found that when the
phase velocity for CS2 is c/1.64, the group velocity is
c/1.76.)

23.10 Group vs Phase Velocity

The speed v, where v = c/n, is the speed at which a 23.11 Polarization Fundamentals
point of fixed phase propagates in a harmonic wave of
frequency f. For example, a crest is a fixed phase of 90 Electromagnetic waves are transverse, and can there-
degrees. When two harmonic waves of slightly differ- fore be polarized. For electromagnetic waves prop-
ent frequencies are superposed, the result is the beat agating perpendicular to the plane of the paper, the
Scattering, Dispersion, and Polarization 507

double-headed arrows in Figure 23.11 show the vib- the square of the sine function. Equation 23.5 is
ation plane of the electrical field component for three known as the law of Malus after Etienne Malus (1775-
different polarization states: vertical, horizontal, and 1812) who discovered polarization of light in 1808.
diagonal. The superposition of two perpendicularly polarized
A dipole with a vertical axis produces vertically waves never gives destructive interference. After all, a
polarized light. However, in our natural environment, vertical displacement cannot cancel a horizontal dis-
light is incident from many atomic and molecular placement. For complete destructive and constructive
dipoles with differing orientations. In addition, the interference to occur, the light that is superposed must
same atomic dipole radiates for at most 10~8 seconds, be identically polarized.
and then may change orientation before radiating In Young's double-slit experiment, the incident
again. Hence, the light in our normal environment is a light was unpolarized. This unpolarized light consisted
completely random mixture of waves vibrating in all of equal amounts of uncorrelated horizontal and verti-
planes. Such light is called unpolarized. In unpolarized cal polarization components. The interference fringes
light, there is a randomly fluctuating phase relation- were produced by the right slit horizontal components
ship between the waves vibrating in different planes. interfering with the left slit horizontal components,
Partially polarized light can be made by mixing polar- and similarly the vertical components interfering with
ized and unpolarized light. each other. This is easily checked by using polarization
The arrow in Figure 23.12 represents the amplitude filters, such as polaroids, over the slits. When the
Am of a wave that is plane-polarized at an angle with filters both transmit only vertically polarized light, the
the horizontal. The irradiance (intensity) Im of the interference fringes still have a visibility of 1. Fringe
wave is proportional to (A m ) 2 . Like a vector, the visibility of 1 also occurs when both filters transmit
polarized light can be resolved into perpendicular only horizontally polarized light. However, when one
components. From Figure 23.12, the horizontal com- filter transmits vertically polarized light and the other
ponent has an amplitude of A m cos 0, while the verti- transmits horizontally polarized light, no interference
cal component has an amplitude of A m sin . The fringes occur (i.e., zero visibility). Thus, for two
oscillations of the horizontal and vertical components waves to be mutually coherent, the waves must have
are correlated. the same polarization state together with a correlated
The irradiance Ix of the horizontal component var- phase relation.
ies like the square of the amplitude: Partial coherence (fringe visibility less than 1) can
occur for polarization states that are neither aligned
Ix = (A m cos0) 2 .
nor perpendicular. Here, the horizontal components
Then of one beam interfere with the horizontal components
of the other beam, and similarly for the vertical
Ix = I m cos 2 0. (23.5)
components. Suppose the diagonally polarized light in
A similar equation for the vertical component contains Figure 23.12 is superposed with horizontally polarized
light. When the two beams have a fixed phase rela-
tion, the horizontal components give constructive and
destructive interference. The vertical component of
the diagonally polarized light has nothing to interfere
with, so it provides the uniform background that
decreases the visibility of the interference fringes.
a) ' b) c)

FIGURE 23.11. Vibration meridians for polarized light, a.


Vertical, b. Horizontal, c. Oblique.
23.12 Polaroid

In a dielectric material, the electrons are bound to the


atoms and/or the molecules. In a metal, the conduc-
tion electrons are free to move throughout the metal.
Because the conduction electrons are free to move,
they respond very strongly to incident light, which is
why metals are such fantastically strong absorbers of
Am cos light.
FIGURE 23.12. Resolution into perpendicular polarization When light is incident normally on a thin metal
components. disk, the vertically polarized component drives the
508 Geometrical, Physical and Visual Optics

electrons to vibrate vertically, while the horizontally ferred to as the analyzer since it is analyzing the plane
polarized component drives the electrons to vibrate polarization properties of the light incident on it.
horizontally. Suppose the metal disk is replaced by a H-sheet starts to lose its effectiveness for the shorter
grid of very fine vertical metal wires. The conduction visible wavelengths (the violets), so when two H-sheet
electrons are free to run up and down the metal wires, polaroids are crossed, a little violet is transmitted. For
and so they still strongly absorb light that is plane- the most part, we will neglect the violet leakage.
polarized vertically. However, the horizontal move- When unpolarized light is incident on an ideal
ment of the conduction electrons is restricted. When polaroid, the transmitted light consists of the 50% that
the wire diameters are small relative to the wavelength is vibrating in the plane of the Polaroid's transmission
of the light, the restriction prevents the absorption of axis. Such an ideal polaroid is referred to as HN-50. In
the horizontal component, and it is transmitted (Fig- reality, there are surface reflections (about 4%) at
ure 23.13). (In other words, the vertical grid "looks" each surface. So we would be down to a 42% trans-
like a solid metal disk to the vertically polarized light, mission for unpolarized light (HN-42). In real
and "looks" transparent to the horizontally polarized polaroids, some absorption also occurs. The widely
light.) used types of H-sheet polaroid are HN-38, HN-32,
A standard oven rack would serve as a wire grid and HN-22.
polarizer for microwaves. A wire grid polarizer has For plane-polarized light incident on an ideal
been constructed for infrared (2,160 wires/mm). polaroid (HN-50), the irradiance of the light trans-
While no wire grid polarizer has been constructed for mitted varies like the law of Malus (Eq. 23.5), pro-
visible light, polaroid works on the same principle. vided is the angle between the polarization plane of
J-sheet polaroid was invented in 1928 by Edwin the incident light and the transmission axis of the
Land, then a 19-year-old undergraduate student at polaroid. When the polarization plane is aligned with
Harvard. Besides polarizing, J-sheet scattered light the transmission axis of an ideal polaroid (0 = 0),
and thus had a hazy appearance. In 1938, Land inven- 100% of the light is transmitted. When the polariza-
ted the widely used H-sheet polaroid. The process tion plane is parallel to the absorption axis, it is
consists of heating a sheet of clear poly vinyl alcohol, perpendicular to the transmission axis. Here, equals
and stretching it in one direction. During the stretch- 90, and no transmission occurs.
ing, the sheet's long hydrocarbon molecules become
aligned. The sheet is then dipped in an ink solution
rich with iodine. The iodine impregnates the plastic EXAMPLE 23.2
and attaches to the straight long chained molecules Vertically polarized light is incident on an ideal
polaroid with an absorption axis rotated 35, from
effectively forming a chain of its own. The iodine's the vertical. What percent of the light is trans-
conduction electrons can move along the iodine chains mitted?
as if they were long thin wires. The transmission axis is perpendicular to the
The absorption axis of the polaroid is aligned with absorption axis, or 55 from the vertical (Figure
the iodine chains. The transmission axis of the 23.14). Before calculating, you might ask whether
polaroid is perpendicular to the iodine chains. Thus, you expect the percent transmission to be high
when the iodine chains are vertical, the transmission (greater than 50%) or low (less than 50%). Since
axis is horizontal. (Warning: many elementary books the absorption axis is closer to the polarization
have simplified picket fence models that give the plane, we should expect a low transmission.
For the ideal polaroid (HN-50), we neglect the
impression that the chains are aligned with the trans- surface reflections. From Eq. 23.5,
mission axis.)
When two perfect polaroids are perpendicularly I = Im cos2(55),
crossed (e.g., one with a horizontal absorption axis I = Im(0.57)2 = Im(0.33).
and one with a vertical absorption axis), no light is So for this case, 33% of the incident light (i.e., Im)
transmitted. The second polaroid is frequently re- is transmitted (in agreement with the expectations).

FIGURE 23.13. a. Unpolarized lightcompo-


nents are not correlated, b. Wire grid polarizer, c.
b) Transmitted component.
Scattering, Dispersion, and Polarization 509

L^s* absorption Malus, Eq. 23.5:


^^ axis I = Imcos2(30),
transmission ^J/^
I = Im(0.75).
axis This light is polarized at an angle of 30 with the
vertical, or 60 degrees with the horizontal. So the
transmission for the last polaroid is
I = Imcos2(60),
I = Im(0.25).
The total transmission for the three polaroid system
FIGURE 23.14. Polaroid with an absorption axis 35 degrees is then
from the vertical. (0.50)(0.75)(0.25) = 0.094,
or 9.4% of the incident light.
Thus, when all three polaroids are present,
EXAMPLE 23.3 9.4% of the incident light is transmitted. When the
Unpolarized light is incident on a two polaroid middle polaroid is removed, leaving only the two
system. The second polaroid has a transmission axis crossed polaroids, no light is transmitted. The 9.4%
rotated 25 relative to the first. Neglecting surface transmission occurs because the middle polaroid
reflections, what is the percent of the original acts to rotate the plane of the polarized light. When
incident light that is transmitted through both the first and last polaroid are perpendicularly cros-
polaroids? sed, the maximum transmission for incident un-
We will assume the polaroids are ideal (HN-50). polarized light through a three polaroid system
Then the first polaroid transmits 50%. We should occurs when the middle polaroid has a transmission
expect a high transmission for the second polaroid axis at 45 to that of the first polaroid (see the
since the incident polarization plane is fairly close problems).
to the transmission axis. Equation (23.5) gives the
transmission of the second as:
I = Imcos2(25),
I = Im(0.82). 23.14 Polarization by Reflection
The total transmission is the product, (0.50)(0.82),
Light can be partially or completely polarized by
or 41%.
specular reflection from dielectrics. Anyone who has
looked through polaroid sunglasses at the sun's reflec-
tion from water has probably noticed this effect. It is
only necessary to tilt the head from side to side, thus
23.13 Polarization Plane Rotation tilting the polaroids, to observe the intensity changes
by Polaroids in the reflected light that is the transmitted through
the polaroids.
An ideal system with two crossed polaroids transmits We can use Figure 23.15 to help understand why
no light. A remarkable property of polaroids is that the reflected light may be partially or completely
when a third polaroid is placed between the two polarized. Unpolarized light in the medium of refrac-
crossed polaroids, transmission can occur. This trans- tive index n, is incident on the flat surface of the
mission is due to rotation of the polarization plane by medium with refractive index n 2 . Figure 23.15 is
the middle polaroid. drawn for the case where the angle between the
refracted ray and the reflected ray is 90. The incident
EXAMPLE 23.4 unpolarized light can be resolved into two perpendicu-
Consider a three polaroid system in which the first lar components: one oscillating perpendicular to the
polaroid has a vertical transmission axis and the last paper (represented by the dots) and the other oscillat-
polaroid has a horizontal transmission axis. Assume ing in the plane shown by the double-headed arrows.
that the transmission axis of the middle polaroid The light represented by the dots sets up dipole
makes an angle of 30 with the vertical. For inci- vibrations in the surface zone of the second medium.
dent unpolarized light, what percent of the original
light is transmitted through the system? The axis of these dipoles is perpendicular to the plane
We will assume HN-50 (ideal) polaroids, so the of the paper, and so the dipoles reradiate maximally in
first polaroid transmits 50% of the light, and that that plane. From a coherent scattering viewpoint, both
light is polarized vertically. We can find the trans- the reflected and refracted waves are then generated
mission of the second polaroid from the law of by constructive interference of the reradiated waves.
510 Geometrical, Physical and Visual Optics

FIGURE 23.15. Geometry for polarization by reflection.

The light represented by the double-headed arrows Equation 23.6 is referred to as Brewster's law after Sir
sets up dipole vibrations that are in the same plane as David Brewster who discovered it empirically. (Brew-
the arrows. The axes of these dipoles are parallel to ster also invented the kaleidoscope). Equation 23.6
the direction of the reflected light. Since dipoles do specifies that for dielectrics Brewster's angle depends
not radiate along their axes, no light is reradiated in only on the ratio of the refractive indices of the two
the reflected direction. Hence, no reflection occurs for materials. In particular, for a material of index n2 in
this component of the incident light. air (nl = 1), Brewster's angle can be used to determine
Thus, when the reflected direction is perpendicular the refractive index of the material.
to the refracted direction, the light that is specularly
reflected is plane-polarized in the plane of the surface. EXAMPLE 23.5
The incident angle at which this occurs is referred to Unpolarized light in air is incident on ophthalmic
as Brewster's angle. We can derive an equation for crown glass (n = 1.523). At what incident angle will
Brewster's angle by starting with Snell's Law: the light reflected be completely polarized?
n1 sin 0B = n2 sin 02. From Eq. 23.6,
tan 0B = 1.523/1,
Because the angle of reflection equals the angle of
or
incidence, and because of the 90 angle between the
reflected and the refracted rays, we have 0B = 56.7.
18O = 0B + 9O+0 2 , Similarly, Brewster's angle for an air-water (n =
1.333) surface is 53.1 degrees.
or
02 = 9O-0 B . The above discussion shows that when unpolarized
Then light is incident at Brewster's angle on a horizontal
dielectric plane (e.g., glass or water), the light specu-
nl sin 0B = n2 sin(90 - 0B). larly reflected is plane-polarized horizontally. When
Since cos 0 = sin(90 - 0), Snell's law becomes vertically polarized light is incident at Brewster's angle
on a horizontal dielectric plane, no light is specularly
nl sin 0B = n2 cos 0B, reflected.
or Polaroid sunglasses with a vertical transmission axis
sin 0B _ n2 do not transmit the horizontally polarized light re-
flected at Brewster's angle from flat horizontal sur-
cos 0B n ! ' faces. Such sunglasses reduce the glare that arises
and from specular reflection from water surfaces or roads.
They do not reduce glare arising from reflections from
tan0B=. (23.6)
a vertical surface, such as office building windows.
Scattering, Dispersion, and Polarization 511

In cases where the reflected light is highly polar- 23.16 Birefringence


ized, the refracted light tends to be slightly polarized.
If this process is repeated many times (e.g., a stack of When one views objects through a properly oriented,
slightly separated microscope slides), the refracted transparent, polished calcite crystal, the image is dou-
light will eventually also become plane-polarized. bled (Figure 23.17). This phenomenon is known as
While specular reflection tends to polarize light, birefringence, or double refraction. The light as-
diffuse reflection does not. In fact, polarized light that sociated with each image is plane-polarized perpen-
is diffusely reflected is depolarized. dicular to that of the other.
The underlying reason for the appearance of the
double images is that the calcite crystal is anisotropic,
which means the material has a different structure in
23.15 Polarization by Scattering different directions. The wire grid polarizer is aniso-
tropic, and so are most crystal lattices.
Scattered light can become wholly or partially polar- We will not consider the exact structure of the
ized for essentially the same reasons that apply to calcite lattice. Instead, Figure 23.18 shows a simple
specularly reflected light. Figure 23.16 shows unpolar- spring model of an anisotropic lattice with a horizontal
ized light incident on a scattering center. The dots and a vertical axis. The horizontal set of springs
represent the component vibrating perpendicular to represents a different molecular force than the vertical
the plane of the paper. This light sets up a dipole set of springs. Since the forces are different, the
vibration in the scattering center, and the dipole re- natural or resonant frequencies differ for vertically vs
radiates maximally in the plane of the paper (including horizontally polarized light. The refractive index and
points A, B, and C). The vertical double-headed the absorption strength for the vertically polarized
arrows represent the other polarization component of light is determined by the vertical resonances. The
the incident light. This component sets up a dipole refractive index and the absorption strength for the
vibration with a vertical axis. Since dipoles do not horizontally polarized light is determined by horizon-
radiate along their axes, no reradiation occurs in the tal resonances.
vertical direction (point B). Thus, the scattered light Figure 23.19 shows dispersion curves for both the
reaching point B is plane-polarized perpendicular to vertical direction (i.e., the refractive index nv vs fre-
the plane of the paper. At other points (e.g., A and quency f) and the horizontal direction (nH vs f). The
C), the scattered light is partially polarized. vertical resonance occurs at frequency f2, while the
As a consequence, skylight at 90 to the sun is horizontal resonance occurs at frequency f3. For inci-
polarized. As the angle varies from 90, the polariza- dent frequency fx, the anisotropic medium is transpar-
tion decreases. The polarization of skylight is notice- ent for both the horizontally and vertically polarized
able when wearing polaroid sunglasses. (Bee's eyes light. However, the refractive indices are different for
have built-in polarization detectors, and they appear the two polarizations. This difference in refractive
to use the polarization information in their dance that
tells other bees where to look for honey sources.)
Multiply scattered light is depolarized. Thus, while
the blue sky tends to be polarized, the white clouds
are not.
r "7
/ M M P p KK
/
L.
incident unpolarized light FIGURE 23.17. Double images of birefringence.
k k k k k

WW

FIGURE 23.18. Representation of forces in an anisotropic ma-


FIGURE 23.16. Geometry for polarization by scattering. terial.
512 Geometrical, Physical and Visual Optics

and that these images are polarized perpendicular to


each other.
The refractive indices can be specified independent
of crystal orientation as that for the o-ray, n 0 , and that
for the e-ray, n e . Table 23.1 shows some refractive
indices for sodium D light (589.3 nm).

TABLE 23.1
Refractive indices for some birfringent materials

Calcite 1.658 1.486


Quartz 1.544 1.553
Sodium Nitrate 1.585 1.337
Ice 1.309 1.313

There is an orientation for the calcite crystal for


which the o-ray and the e-ray are not separated,
FIGURE 23.19. Dispersion curves for the vertical and horizon-
although each still experiences its own refractive
tal components.
index. The light path along this orientation is called
the optical axis of the crystal. Crystals, such as calcite,
which have one optical axis are called uniaxial and
index leads to the birefringence or double refraction have two principal indices of refraction. More compli-
phenomenon. cated crystals, such as mica, have two optical axes
Figure 23.20 shows unpolarized light incident nor- (biaxial) and three principal indices of refraction. An
mal to a calcite crystal. The polarization component isotropic medium has only one index of refraction.
perpendicular to the plane of the paper (the dots) For the frequency f2 in Figure 23.19, the horizontal
behaves as expected, and the associated ray is called light is transmitted, while the vertical light experiences
the ordinary or o-ray. The polarization component a resonance and is absorbed. A crystal that transmits
represented by the double-headed arrows does not one component and absorbs the other is called di-
behave as expected. (In a sense, this component acts chroic. J-sheet polaroid was made with dichroic her-
as though the crystal has an effective normal that is pathite crystals.
different from that of the crystal's surface.) The ray Before polaroid was invented, the most common
associated with this component is labeled the extra- polarizers were prisms made from birfringent materi-
ordinary or e-ray. Figure 23.20 shows that the e-ray als. These included the Nicol, Glan-Thompson, and
and the o-ray give the double images of Figure 23.17, Wollaston prisms.

e-ray

>
o-ray

FIGURE 23.20. Extraordinary and ordinary


rays for a calcite crystal.
Scattering, Dispersion, and Polarization 513

23.17 Circular Polarization in a circle. For light coming out of the plane of the
paper, Figure 23.23 shows the rotation directions for
Unpolarized light can be considered as composed of right vs left circularly polarized light.
equal amounts of perpendicularly polarized light Circularly polarized light consists of equal amounts
where the phase relations between the two compo- of plane-polarized components where the components
nents randomly fluctuate. Plane-polarized light can are correlated and have a 90 phase difference. Other
also be regarded as composed of two perpendicular phase retardation values result in elliptical polariza-
components, but in this case the phases of the two tion. When elliptically polarized light is incident on an
components are correlated. Figure 23.21 shows the electrically charged particle, it drives the particle to
horizontal and vertical components for light that is rotate in an elliptical pattern. Actually, plane- and
plane-polarized at 45 to the vertical. As the time circularly polarized light are special cases of elliptically
sequence advances from tj to t 5 , the components polarized light.
oscillate in phase. When plane-polarized light is inci- Figure 23.24 shows the phase retardations needed
dent on an electrically charged particle, it drives the to change the plane-polarized light on the left to each
particle to oscillate in the plane of the polarization. of the other polarization states. A phase retardation of
By using anisotropic materials, we can retard the either 90 or 270 gives circular polarization. A phase
phase of one component relative to the other. Figure retardation of 180 gives plane polarization perpen-
23.22 shows the effect of a 90 phase retardation. At dicular to the original plane. A 360 phase retardation
t l5 the vertical component is maximally displaced, gives the same polarization plane. Other phase re-
while the horizontal component has a zero displace- tardation values give elliptically polarized light. The
ment. At t 2 , the vertical displacement is decreasing, difference between the listed values gives the phase
while the horizontal displacement is increasing. At t 3 , retardations needed to go from one state to another.
the horizontal component has reached maximum dis- For example, each circular polarization state differs
placement while the vertical component is zero. As a from a plane-polarized state by a 90 phase retar-
function of time, the vector sum of the horizontal and dation.
vertical displacements gives a rotating arrow. In other The above analysis shows how circularly polarized
words, the electric field vector of the electromagnetic light can be made from two perpendicular plane-
wave is now circularly rotating as the light propagates polarized components. A similar analysis shows that
through space. We call this light circularly polarized. equal amounts of correlated right and left circularly
When circularly polarized light is incident on an elec- polarized light can be combined to make plane polar-
trically charged particle, it drives the particle to rotate ized light.

FIGURE 23.21. Motion picture sequence for the horizontal and


vertical (correlated) components of linear polarized light.

t
~* S l \
ti t2 t3 U t5 te t7 te to

/ \ FIGURE 23.22. Motion picture sequence of a phase-


\ \ / retarded component resulting in circular polarization.
514 Geometrical, Physical and Visual Optics

with a 45 transmission axis followed by a polaroid


with a 135 transmission axis does not transmit light.
Suppose a fullwave plate for sodium D light (589 nm)
has a horizontal and vertical axis and is placed be-
tween the two crossed polaroids. The 589 nm light
leaving the first polaroid is plane-polarized, and the
plate produces a 360 phase retardation, which gives
the same polarization plane (Figure 23.24). The
FIGURE 23.23. Right and left circular polarization directions for
589 nm light is not transmitted by the second polaroid
light coming out of the plane of the paper. (the analyzer).
However, the fullwave plate gives a 360 phase
retardation only for 589 nm. For incident wavelengths
different from 589 nm, the actual phase retardation

/O ^ O
produced by the fullwave plate differs from 360. For
the other phase retardation values, the light coming
through the plate is elliptically polarized, and some of
270
90 180 it is transmitted by the second polaroid. Since we are
FIGURE 23.24. Phase retardations that give various polarization dealing with little or no transmission for some wave-
states. lengths, and some significant transmission for more
distant wavelengths, the transmitted light is colored. A
589 nm fullwave plate between crossed polaroids gives
23.18 Phase Retarders a beautiful blue color. By adjusting the plate thick-
ness, other colors may be obtained. These colors
In anisotropic materials with a horizontal and vertical resemble those obtained from thin film interference,
axes, the phase retardation between the horizontal and hence are called interference colors.
and vertical polarization components is introduced For the appropriate incident wavelength, a half-
because the wavelength of the light in the medium is wave plate placed between the two crossed polaroids
different for each component (e.g., / vs / ). flips the plane of polarization 90 resulting in 100%
The amount of phase retardation depends on these transmission at the second (ideal) polaroid. Wave-
interior wavelengths plus the thickness of the material. lengths for which the plate is not quite a halfwave still
Figure 23.25 shows horizontal and vertical components have a high transmission at the second polaroid, so for
of wavelength incident on a halfwave plate. The incident white light the output remains white.
plate thickness is equal to 1 interior wavelength for the Mica and sheets of stretched polyvinyl alcohol are
horizontally polarized light and 1.5 times the interior often used to make phase retarders. Common cel-
wavelength for the vertically polarized light. The exit- lophane acts as a phase retarder (because of the
ing components both have wavelength , but now they stretching during its manufacture). Some isotropic
are 180 out of phase. A quarterwave plate gives a 90 transparent materials, such as plastic, can become
phase retardation, and a fullwave plate gives a 360 anisotropic under stress. When these materials are
phase retardation. placed between two crossed polaroids, the differing
If we neglect the violet leakage, then a polaroid areas of stress produce differing amounts of phase

FIGURE 23.25. A halfwave plate.


Scattering, Dispersion, and Polarization 515

retardation, and the result is a colored map of the


stress areas. This effect is called photoelasticity. The
stress patterns in heat-tempered glass can also be seen
through crossed polaroids. (Sometimes a person wear-
ing polaroid sunglasses can see the stress patterns in
tempered window glass. In this case, the scattering of
the skylight serves as the polarizer, and the person's
polaroids serve as the crossed polaroid analyzer.)
It was originally believed that heat-tempered spec-
tacles lenses that gave a uniform Maltese cross stress
pattern were more impact-resistant than lenses that
gave an irregular stress pattern. However, later re-
search showed that there is little or no correlation
between the uniformity of the stress pattern and the
impact resistance. Chemical-tempered lenses do not
show a stress pattern, and yet they have better impact
FIGURE 23.27. a. Right circular polarizers back to back
resistance than heat-treated lenses. (transmission), b. Left and right circular polarizer back to back
(no transmission).

23.19 Circular Polarizers plate, and their effect is to flip the polarization plane
by 90, resulting in 100% transmission at the second
We can make a right circular polarizer by placing a polarizer. Note that the right circularly polarized light
polaroid with a transmission axis at 135 in front of a from the first polaroid and phase plate is highly trans-
quarterwave plate with a horizontal and vertical axis mitted when incident on the backwards right circular
(Figure 23.26a). If the quarterwave plate is rotated 90 polarizer (i.e., the second combination).
clockwise, the other polarization component is re- If the first quarterwave plate is turned 90, the first
tarded and the combination is then a left circular combination is a left circular polarizer (Figure
polarizer (Figure 23.26b). The circular polarizer is 23.27b). Here the two quarterwave plates are perpen-
definitely directionally dependent. Unpolarized light dicularly crossed. One cancels the retardation of the
passing backwards through the circular polarizer re- other (zero net retardation), and they leave the polari-
mains unpolarized as it passes through the quarter- zation plane unchanged, resulting in no transmission
wave plate and then is plane-polarized by the at the second polaroid. In this case, the left circularly
polaroid. polarized light from the first combination is not trans-
Consider two identical right circular polarizers in mitted through the backwards right circular polarizer.
which the quarterwave plates have a horizontal and Similarly, a backwards left circular polarizer trans-
vertical axis. Now rotate the second one 180 around a mits left circular polarized light, but does not transmit
vertical axis so that it is backward relative to the first right circularly polarized light. Thus, backwards circu-
(Figure 23.27a). In this case, the two polaroids have lar polarizers serve as analyzers for circular polarized
perpendicularly crossed transmission axes, while the light.
two quarterwave plates are still identically aligned. While linear polaroids eliminate glare due to specu-
For light incident on the combination, the first lar reflection in the vicinity of Brewster's angle, they
polaroid plane polarizes the light. Then the two do not have the same property for light reflected at
aligned quarter wave plates act like a single halfwave normal incidence. For normal incidence, we can use
the polarizing-analyzing property of circular polarizers
to reduce glare from reflection. Consider a video
display terminal (VDT) with glare problems. When a
right circular polarizer is placed over the VDT, the
right circular polarized light incident on the VDT is
specularly reflected as left circularly polarized light.
This left circular polarized light cannot pass backwards
through the right circular polarizer, so the glare is
a) \As b) \ J y removed.
FIGURE 23.26. Polaroid and quarterwave plate that make a In a similar manner, circular polarizers have been
circular polarizer, a. Right, b. Left. used to control corneal reflex glare in direct ophthal-
516 Geometrical, Physical and Visual Optics

moscopy, and to improve the view of the corneal must make sure the screen does not depolarize the
endothelium in slit-lamp biomicroscopy. light.
Polarizers can also be used to present the slightly
dissimilar views that serve as the stimulus for stereop-
sis. If the brain's stereopsis system is working, two flat
23.20 Optical Activity and Induced Effects slightly dissimilar targets result in a single three-
dimensional perceived image. Systems that test this
Certain materials, particularly organic ones, rotate the include the Wirt four dot system and the stereofly.
plane of polarization of light that passes through them.
This phenomenon is known as optical activity, and is
important in organic chemistry and biochemistry.
23.22 Ocular Birefringence and
(You can investigate optical activity by placing com-
Haidinger's Brush
mon corn syrup between crossed polaroids.)
In various circumstances, electric or magnetic fields
Both the cornea and the crystalline lens are birfrin-
applied to materials also rotate the plane of polariza-
gent; however, there doesn't seem to be any clinical
tion. These are called induced effects. Induced effects
advantage to this birefringence. The macula area of
are important in information storage, communication,
the retina also has some birefringence. This results in
and other electro-optical applications. Many cal-
the entoptic phenomenon called Haidinger's brush,
culators have liquid crystal displays that use cross
which is clinically useful.
polaroids and induced effects to make the numbers
Suppose one stares through a polaroid at the blue
visible.
sky. When the polaroid is suddenly rotated clockwise
90, a yellowish brush appears momentarily and then
fades. When the polaroid is continuously rotated at
the right speed, a rotating yellowish brush is seen
23.21 Binocular Vision and Polarized Light against the blue sky background. (The rotating brush
resembles an airplane propeller.)
Polarized light is used clinically for studying binocular Haidinger's brush is due to the radially arranged
vision. Consider the letters END, where the E is nerve fiber layer of Henle, which lies over the rods
covered with a vertically transmitting polaroid and the and cones of the macular region. The nerve fibers are
D is covered with a horizontally transmitting polaroid. birfringent and act like a yellow radial polarizing
Now consider a person wearing a horizontally trans- filter. When the polarization plane is switched, the
mitting polaroid over the left eye, and a vertically brush is visible, but then quickly fades as does any
transmitting polaroid over the right eye. The person stabilized retinal image.
sees the N with both eyes (50% transmission). With A Haidinger's brush apparatus consists of a blue
the left eye, the person sees ND. With the right eye, filter and a rotating linear polaroid placed over a light
the person sees EN. We can use this arrangement to bulb. It is used to detect, diagnose, and train am-
tell if the person is suppressing, and if so which eye is blyopia due to eccentric fixation. In eccentric fixation,
being suppressed. For normals, we can also use this acuity is reduced because the person fixates with a part
arrangement to compare clarity of the left and right of the peripheral retina instead of the macula. When a
images. normal observer uses the Haidinger's brush apparatus,
We can use a vectographic polaroid to actually he or she sees the brush centered on the fixation
superimpose letters that are perpendicularly polarized. target. When an eccentric fixator uses the apparatus,
A vectograph consists of two perpendicularly stretched he or she sees the brush centered on a point different
sheets of poly vinyl alcohol. As such, there are no than the target point.
conduction electrons, and so, no polarizing properties.
When iodine ink is used to print a letter, for example
an O, on one side, the polaroid iodine chains are
formed, and the O acts like a polarized O. The iodine Problems
ink can then be used to mark a different letter, say an
X, on the other side, in which case the X is polarized 1. What is the relative Rayleigh scattering intensity
perpendicular to the O. When a person looks through for 600 nm light compared to 500 nm light?
a polaroid at that spot on the vectograph, either the X 2. Horizontally polarized light is incident on an ideal
or the O is seen depending on the Polaroid's orienta- polaroid in which the transmission axis is rotated
tion. Vectographic slides are available for both far and 30 degrees from the horizontal. What percent of
near visual testing. When projecting the slide, one the incident light is transmitted?
Scattering, Dispersion, and Polarization 517

3. Unpolarized light is incident on an ideal polaroid 1.701), what incident angle results in 100% polari-
in which the transmission axis is rotated 60 de- zation in the reflected light?
grees from the horizontal. What percent of the 7. For light in ophthalmic crown glass (n = 1.523)
incident light is transmitted? incident on a glass-air interface, what incident
4. Unpolarized light is incident on a two polaroid angle results in 100% polarization in the reflected
system. The transmission axis of the second light?
polaroid is rotated 80 degrees from that of the 8. What is Brewster's angle for polycarbonate (n =
first. If the polaroids are ideal, what percent of the 1.58)?
incident light is transmitted by the two polaroid 9. What does the statement, "There is no blue in the
systems? sky nor white in the clouds" mean?
5. In a three polaroid system, the first and last 10. What is the physical explanation for Mikala
polaroid have perpendicular transmission axes, Creampuffs green eyes (i.e., iris color)?
while the middle polaroid has a transmission axis 11. In terms of coherent scattering theory, why does
rotated 45 degrees from the first. For ideal an edematous cornea lose transparency?
polaroids, what percent of the incident light is 12. Explain why light reflected from a dielectric sur-
transmitted through the three polaroid system? face at Brewster's angle is polarized.
6. For light in air incident on highlite glass (n =
CHAPTER
TWENTY-FOUR

Emission,
Absorption, and
Photons

24.1 Blackbody Radiation suspended ball, which in turn absorbs and reradiates.
Let I be the irradiance (intensity) of the radiation in
A blackbody is a material that absorbs most of the the cavity. Let ai be the absorptance of the ball and
incident light. By conservation of energy, the fraction the radiation emitted by the ball. When equilib-
a of the incident light absorbed (the absorptance), the rium is achieved, the suspended ball emits as much as
fraction r of the incident light reflected (the reflect- it absorbs, or
ance), and the fraction t of the incident light trans- E^aJ.
mitted (the transmittance) must add to 1: Now consider a second suspended ball in the cavity.
r + t + a=l. (24.1) The irradiance is still the same, and for equilibrium
Scattered light is included in the reflectance or trans- E 2 = a 2 I.
mittance depending on its direction of travel. The From the above two equations,
absorptance a depends on the surface characteristics as
well as the interior absorption strength k of the ma- (24.2)
terial. In particular, surfaces that diffusely reflect usu-
ally absorb more than smooth surfaces of the same The above relation, known as Kirchhoff s law,
material. shows that the amount of emittance is directly propor-
Metals have a high interior absorption strength k tional to the absorptance a. A ball that does not
but also a high reflectance r. Therefore, the absorp- absorb much (small a) also does not emit much (small
tance a of a polished metal surface is small. On the E). A ball that absorbs a lot (large a) also emits a lot
other hand, solid carbon appears black because of its (large E). A famous demonstration of Kirchhoff s law
high absorptance and low reflectance. is to coat a pyrex rod with lamp-black, and then heat
You know from everyday experience that hot ob- the rod in a flame. When the rod is withdrawn the part
jects radiate light. The electromagnetic radiation from coated with lamp-black glows a brighter red than the
a hot object is called thermal radiation. The spectral uncoated part.
distribution of thermal radiation is largely a function A perfect blackbody (a = 1) is the strongest emit-
of the temperature of the source. We can investigate ter. In an absorbing material, a cavity with a small
this temperature dependence by considering the ther- opening approaches an ideal blackbody (Figure 24.2).
mal radiation from an ideal blackbody. When light is incident on the hole, it passes into the
To understand the emission properties of a black- cavity and reflects back and forth from the interior
body, consider a ball suspended by an insulating string cavity walls until it is absorbed. Very little light comes
in a closed cavity (Figure 24.1). When the cavity is back out of the small opening. Hence, the opening
heated to a certain temperature, the walls emit elec- appears to be an ideal black spot. (The pupil of your
tromagnetic radiation. This radiation is incident on the eye appears black for similar reasons.)

519
520 Geometrie, Physical, and Visual Optics

FIGURE 24.3. Blackbody irradiance distributions as a function


of frequency for several different temperatures.

FIGURE 24.1. A ball suspended in a cavity. One characteristic of blackbody radiation is that
the peak wavelength Am is inversely proportional to
the temperature T of the blackbody. This is known as
Wien's displacement law:
constant (24.3)

The constant equals 2.90 x 10 nm-K, where K is de-


grees Kelvin.

EXAMPLE 24.1
FIGURE 24.2. A cavity that serves as a blackbody radiator. What is the peak wavelength for a blackbody at
5,800 K (the sun)?
From Wien's displacement law,
When the material is heated up, the surfaces
radiate, and the cavity fills with electromagnetic radia- 2.90 x 106 nm-K
= = 500nm.
tion. Because there are many radiating points on the 5800 K
surface of the cavity, the radiation gets very concen- Another characteristic of blackbody radiation is
trated in the cavity, and begins to leak out of the that the total irradiance It is directly proportional to
opening. The concentration of the radiation in the the fourth power of the temperature:
cavity makes the emittance at the cavity hole much
higher than that of any other point on the outer It = (constant)T4. (24.4)
surface. This is known as the Stefan-Boltzman law. The strong
For two temperatures, Figure 24.3 shows the ir- temperature dependence results in a dramatic increase
radiance I of the electromagnetic radiation emitted in irradiance as T increases.
from the hole as a function of the frequency f. For an Because of the above two laws, care must be taken in
ideal cavity, this spectrum is independent of the judging apparent colors. A dim light may have a peak
specific absorbing material. For low temperatures, wavelength in the red. When a rheostat is turned to
most of the radiation is in the infrared. You can feel increase the current in a filament, the filament gets
this radiated heat with your hand, even though you hotter and emits more light (Stefan-Boltzman). How-
cannot see it. As the temperature is increased, the ever, the peak wavelength emitted also shifts up (Wien
emittance increases and moves toward the higher fre- displacement), so the incident light is bluer than before.
quencies (shorter wavelengths). At about 800 K, some
radiation becomes visible, and we say the object is red
hot. As the temperature is increased still more, the
trend continues (white hot at 6,000 K and blue hot at
10,000 K). The sun's spectrum resembles that of a 24.2 Photons
blackbody at 5,800 K. The spectrum of a candle flame
resembles that of a blackbody at 1,800 K. A hot Classical electromagnetic wave theory correctly de-
tungsten filament, called a greybody, produces a spec- scribes a wide range of everyday physical phenomena,
trum with the shape of a blackbody spectrum at but it breaks down at the atomic level. The blackbody
3,600 K, but the tungsten filament's irradiance has a radiation spectrum was one of the early breakdowns.
magnitude of about one third of the blackbody ir- In classical electromagnetic wave theory, the black-
radiance. body spectrum is analyzed by examining the standing
Emission, Absorption, and Photons 521

wave modes set up in the cavity. The classical analysis naively associate with waves. Other quantum particles,
correctly fits the low frequency part of the blackbody such as electrons, protons, neutrons, pions, etc., also
spectrum, but goes haywire at the higher frequencies exhibit interference and diffraction effects. (Some
(Figure 24.4). As a function of increasing frequency, have felt that the term wavicles is a better name than
the experimental curve peaks and turns down, while quantum particles.)
the classical prediction continues to grow. This is For monochromatic electromagnetic radiation, the
known as the ultraviolet catastrophe. (In fact, accord- energy E carried by a single photon is directly propor-
ing to the ever increasing function predicted by the tional to the frequency f:
classical theory, matter would radiate away all its
energy and collapse. The estimated time for this to E = hf, (24.5)
occur is about 17 seconds.) where h is Planck's constant, which equals 6.63 x
Another breakdown in classical electromagnetic 10~ J-s (Joule-seconds). Equation 24.5 was first in-
wave theory involved the photoelectric effect. In the troduced by Max Planck in 1901 to explain the black-
photoelectric effect, electromagnetic radiation of high body radiation spectrum. However, Planck interpreted
enough frequency incident on a metal results in the the equation as saying that while the radiation could
ejection of electrons. Classical theory predicts that have a continuous energy, the electrons in the black-
electrons should be ejected, and that the mechanism is body walls could only emit the radiation in units of hf.
a "sunbathing" mechanism. In particular, classical In 1905, Albert Einstein explained the photoelec-
electromagnetic theory says that the metal needs to tric effect by saying that the radiation itself consisted
sunbathe for awhile before the electrons come out (a of particles, each with energy hf. The radiation can
time lag), that the effect should be frequency-indepen- consist of any integer number of particles, and hence
dent, and that the maximum kinetic energy of the the total radiated energy is an integer multiple of hf.
ejected electrons should be dependent on the ir- On an energy level diagram, a free electron with a
radiance of the light. The experimental results are that kinetic energy of magnitude s is assigned a -I-s value.
the effect is very frequency-dependent (it does not A free electron with zero kinetic energy is assigned a
occur if the frequency is not high enough), that the zero value. An electron that is bound by an energy of
maximum kinetic energy of the ejected electrons is magnitude w is assigned a - w value (Figure 24.5).
independent of the irradiance, and that when the According to Einstein, electrons bound in a metal can
frequency is high enough the electrons are ejected be freed by absorbing a single photon with enough
immediately (no time lag). energy to overcome the binding energy (hf>w).
Both of the above breakdowns are solved by the When the photon does not have enough energy to free
quantum theory of light. The quantum theory correct- the electron (hf<w), the absorbed energy is dissi-
ly describes electromagnetic radiation interactions at pated as heat. Thus, the photoelectric effect is ex-
the atomic level, as well as encompassing and includ- tremely frequency-dependent.
ing the classical electromagnetic wave theory at the Since the electron is freed by absorbing one photon
macroscopic level. A central feature of the quantum of high enough energy, the kinetic energy of that
theory of light is that the energy carried by elec- electron is the difference between the photon's energy
tromagnetic radiation is not continuous, but comes in (hf) and the binding energy w of the electron. This
packets called photons. kinetic energy is independent of the irradiance. A
While photons appear to be particles, they also higher irradiance means more electrons come out, but
exhibit the interference and diffraction effects that we the maximum kinetic energy is still the same. Finally,

theoretical

+s

FIGURE 24.4. The ultraviolet catastrophe of classical elec-


tromagnetic wave theory. FIGURE 24.5. Energy level diagram for a metallic electron.
522 Geometrie, Physical, and Visual Optics

when an incident photon has a high enough energy, an are referred to as rays rather than waves. The single
electron absorbs it and is immediately ejected (no time photon energy for these x-rays is greater than
lag). Einstein's photon concept explained both the 12,390 eV.
blackbody spectrum as well as the photoelectric effect. Photons are not electrically charged but are emit-
(It is interesting that Einstein received the Nobel prize ted and absorbed by electrically charged particles.
for his work on the photoelectric effect as opposed to Photons, together with neutrons, have a zero rest
the theory of relativity.) mass, but photons never occur at rest. Photons are
Planck's constant is so small that the energy carried always moving with the speed of light. Attempts to
by a visible light photon is minute. Normal amounts of stop a photon (i.e., bring it to rest) simply result in its
visible light consist of an incredible number of absorption.
photons. A 60 watt light bulb puts out about 1019 A moving photon has a mass equivalent m that is
photons per second. Under these conditions, millions found from Eq. 24.5 together with Einstein's relativity
of photons can be added or subtracted without visual equation
notice.
E = mc ,
Since
Af = c, where E is the photon energy, m is the mass equival-
ent, and c is the speed of light. The result is
we can rewrite Eq. 24.5 as
m = (hf)/c2.
he
E= (24.6) Since a photon has a mass equivalent, it carries mo-
mentum as well as energy. The momentum p of a
The energy carried by a single photon is frequently particle is given by
expressed in terms of electron volts (eV). An electron
volt is the amount of energy that an electron gains p = mv,
while moving through a potential difference of 1 volt. where m is the mass and v is the velocity. For a pho-
Note that we are not saying that a photon is like an ton in a vacuum, v = c, and we can use the mass
electron; we are only using the electron volt as a equivalent for m to obtain
measure of energy. In other words, a 2eV photon
carries energy equal to that gained by an electron p = (hf)/c,
while moving through a potential difference of 2 volts.
In convenient units, or
p = E/c. (24.8)
he = 1239 nm-eV.
From Af = c, it follows that p = h/A.
Therefore, Eq. 24.6 becomes Besides energy and momentum, a photon also car-
1239 nm-eV ries angular momentum. A naive picture is that the
E= (24.7) photon is spinning as it moves through space. In the
quantum theory, photons are spin 1 particles (as op-
posed to electrons, which are spin 1/2 particles, or
EXAMPLE 24.2 pions, which are spin zero particles). It is the photon
What is the single photon energy for the C (red) spin that is responsible for the polarization properties
line of the hydrogen spectrum (A = 656nm)? of light.
From Eq. 24.7,
1239 nm-eV
E= 1.89 eV.
656 nm
What is the single photon energy for violet light 24.3 Bohr Model of the Atom
of vacuum wavelength 413 nm?
From Eq. 24.7, Besides thermal radiation, light is also emitted when a
1239 nm-eV high voltage is placed across a tube of gas. Unlike the
= 3.00eV. continuous blackbody spectrum, the spectrum from a
413 nm
low pressure gas discharge tube consists of a series of
Note that the shorter the wavelength, the higher the discrete wavelengths called a line spectrum. The line
frequency, and the more energetic each photon is. spectra are resolved by use of prisms or diffraction
Diagnostic x-rays have a wavelength of less than gratings (Figure 22.16). Line spectra were first ex-
0.1 nm. Consequently, their wave properties are not plained by the Bohr model of the atom.
particularly noticeable, which is historically why they According to the Bohr model of the atom, the
Emission, Absorption, and Photons 523

electrons that orbit the atomic nucleus exist only in emitted when an electron in hydrogen jumps from
discrete energy levels. An electron can jump to a E 3 to E j ?
lower energy level by emitting a photon (Figure The energy levels are farther apart than in
24.6a). Similarly, an electron can jump to a higher Example 24.3a so the photon has more energy and
level by absorbing a photon (Figure 24.6b). the emitted light has a higher frequency (lower
In the Bohr model, the allowed electron energy wavelength).
From the Bohr model,
levels in the hydrogen atom are given by
E 3 = -13.6 eV/(4)2 = -0.85 eV.
-13.6 eV
Em = m = 0,1,2,. (24.9) Then the energy E of the emitted photon is
(m + l ) 2 !
= 3 - 1 ,
The lowest energy level, called the ground state, is
given by E 0 = -13.6 eV (Figure 24.7). This means that = -0.85 eV- (-3.4 eV) = 2.55 eV.
an incident 13.6 eV photon has enough energy to free From Eq. 24.7,
the ground state electron, thus ionizing the atom. In = 1239 eV-nm/2.55 eV= 486 nm.
the absence of incident radiation, the ground state is These photons constitute the light in the F line of
the stable state of the atom. The higher states, called the hydrogen spectra.
excited states, are not stable. The first excited state
occurs at = -13.6 eV/22 = -3.4 eV. Similarly, the Figure 24.7 shows the possible downward trans-
second excited state occurs at E 2 = -13.6 eV/32 = itions for hydrogen. Those that end in the ground state
-1.51 eV. (E 0 ) are referred to as the Lyman series. Those that
An electron stays in an excited state for 10 end in the first excited state Ej are referred to as the
seconds or less, and then jumps to a lower level Balmer series, and those that end in the second ex-
emitting a photon in the process. This is called sponta- cited state E 2 are referred to as the Paschen series.
neous emission. The energy of the emitted photon Table 24.1 gives the series, the initial and final states,
equals the difference in the energy levels. An electron and the vacuum wavelengths of the emitted photons.
can jump to a higher energy level by absorbing a The jumps to the ground state are all large, and these
photon that has an energy equal to that of the jump. transitions give ultraviolet photons. The visible wave-
lengths come from the Balmer series. The Paschen
EXAMPLE 24.3a series has smaller energy jumps and gives infrared
What is the energy and wavelength of a photon photons.
emitted when an electron in hydrogen jumps from Because of the energy level structure, atoms absorb
E2 to E,? only those wavelengths that they can emit. Therefore,
= 2 - 1 , when white light is incident on hydrogen with elec-
= -1.51 eV- (-3.4 eV) = 1.89 eV. trons in the first excited state E 1? the photons ab-
sorbed are those with the three Balmer series wave-
The emitted photon has an energy of 1.89 eV. From lengths, and the electrons end up in the appropriate
Eq. 24.7,
higher states. When the hydrogen gas is thick enough,
= 1239 eV-nm/1.89 eV= 656 nm. the three wavelengths of the Balmer series are defic-
These photons constitute the light in the C line of ient or missing from the transmitted radiation. Fraun-
the hydrogen spectra. hofer discovered the deficient or missing lines in the
sun's spectrum and labeled them by letters. The defici-
EXAMPLE 24.3b ent C and F lines are due to absorption by ''cooler"
What is the energy and wavelength of a photon hydrogen in the sun's outer atmosphere.

E, E 1

Eo FIGURE 24.6. Atomic electron energy levels and


Eo
allowed transitions, a. Single photon emission, b.
a) b) Single photon absorption.
524 Geometrie, Physical, and Visual Optics

0. free electron (no Kinetic engergy)


1 E L
p ~ EE4
-1. "
I I I ! ^-3
L2
-2. r Mill
Pascnen
-3. Y Y ?YY Series
1 1 1 Series
Balmer 11 Q
-4.
-5.
-6.
-7.
-8.
-9.
LU -10.
iM ^ ' ^M
-11.
-12.
-13.

-14. U Lyrria n,3eries "-0


FIGURE 24.7. Transitions giving the hydrogen spectra.

T A B L E 24.1 a certain wavelength have a low transition rate be-


Hydrogen energy level transitions and emitted tween the appropriate energy levels. In materials,
wavelengths transitions can also occur nonradiatively, in which case
the energy differences show up as heat loss or gain.
Series Initial Final Wavelength (nm) When the pressure is increased in a gas discharge
tube, the atoms bump into each other more, the
Lyman 95 transitions are disrupted, and the line spectra broaden
97
103
into band spectra. If the pressure is increased still
122 more, the band spectra broaden into continuous spec-
Balmer 434 tra. Part of this broadening is related to the formation
486 of molecules. Molecules provide many additional sets
656 of energy levels for the electrons with the resulting
Paschen 1,282
1,875 increase in possible wavelengths that can be emitted
and absorbed. In particular, organic dyes provide
additional energy levels with very high transition rates.
Liquids and solids have either a band spectrum or a
When an atomic electron is in a particular state, the continuous spectrum. The band absorption spectra are
electrical charge density of the atom is constant with responsible for the color of many substances.
time. When the atom makes a transition between
states of energy E m and E k , the electrical charge
density oscillates with a frequency fmk where
24.4 Fluorescence and Phosphorescence
f = (24.10)
One sign of the contemporary world is the abundance
The frequencies fmk are the resonance frequencies of of fluorescent colors. These colors stand out sharply
the atom. (In the simple spring model of Chapter 23, because the luminous flux emitted at the fluorescent
the resonances appeared as the natural vibration fre- wavelengths may be far greater than the luminous flux
quencies of the spring.) incident at those wavelengths. In most cases, the
The transition rate between the different energy energy that drives the fluorescent radiation comes
levels varies and is determined by the structure of the from an incident higher frequency radiation (such as
individual atom. Atoms that emit strongly at a certain ultraviolet).
wavelength have a high transition rate between the Figure 24.8 shows the ray diagram for a fluorescent
appropriate energy levels. Atoms that emit weakly at substance. Incident high frequency radiation causes
Emission, Absorption, and Photons 525

longer wavelength emission of the material making it


appear whiter under incandescent light. Many de-
tergent whiteners contain a similar dye.
Sodium fluorescein is a fluorescent dye widely used
in contact lens evaluations. The dye is not absorbed by
lower frequency intact corneal epithelial cells because it is a water-
Eo soluble molecule that will not penetrate the lipid
FIGURE 24.8. Energy levels for fluorescence. membranes of the cell. When these membranes are
damaged (perhaps by the contact lens), the dye is
absorbed, and these areas then fluoresce at a wave-
the transition from E 0 to E 2 . Assuming radiative length of about 522 nm when irradiated by near ul-
jumps, the electron then emits a lower frequency traviolet (365 nm to 470 nm). Fluorexon, a higher
photon and jumps down to E l t Finally, the electron molecular weight dye, is used with those hydrophilic
emits another lower frequency photon and jumps back lenses that absorb sodium fluorescein.
to E 0 . In terms of wavelengths, the incident radiation The electron in an atomic excited state has a
has a shorter wavelength than the emitted radiation. certain probability to decay or jump to a lower energy
The higher frequency photon may be ultraviolet, level. Usually these probabilities are such that the
while one or both of the lower frequency photons is jump occurs within 10~8 seconds of excitation. How-
visible. There are cases where illumination by visible ever, there are some excited states, called metastable
blue results in fluorescent radiation in the longer states, which have a very low probability of decay.
visible wavelengths (reds or oranges). The strength of Electrons may stay in the metastable excited states for
the fluorescence depends on the transition rates be- seconds, minutes, or even hours. When the metastable
tween the different states. states are populated by incident radiation, the material
may continue to glow or emit radiation long after the
EXAMPLE 24.4
original source is removed. This is the phenomenon of
A substance has energy levels of E o = -10.0eV, phosphorescence. The distinction between fluoresc-
E1 = -8.0 eV, and E 2 = -5.5 eV. Can this material ence and phosphorescence is a matter of time. The
fluoresce? If so, what incident wavelength is material is classified as phosphorescent if it continues
needed to excite the material? What wavelengths to fluoresce for more than one microsecond after the
are emitted in the fluorescence? exciting radiation is shut off.
For fluorescence, the incident radiation must Clock hands that glow in the dark are a common
have the energy needed for the electron to jump example of phosphorescence. Television and video
from the ground state to E 2 . Therefore, display terminals (VDTs) use short-acting phosphor-
= 2 - 0 , escent materials. These phosphors are excited by an
= -5.5 eV- (-10.0 eV) = +4.5 eV. incident beam of electrons rather than by electromag-
So netic radiation, and they continue to emit visible light
for a short period of time after the electron beam has
= (1239 nm-eV)/4.5 eV= 275 nm (ultraviolet). passed on.
The electron in E2 can decay to El by emitting a Since 1939, fluorescent lights have come into wide
photon of wavelength use. They consist of a gas discharge tube usually
1239 nm-eV containing argon (with possibly some neon) and a
= = 496nm (blue). small amount of mercury. The walls of the tube are
-5.5eV-(-8.0eV)
coated with a phosphor. During the discharge, the
Similarly, the electron in Ex can decay to E 0 by
emitting a photon of wavelength ultraviolet emitted by the mercury (254 nm), excites
the phosphor, which then emits visible light. Figure
1239 nm-eV 24.9 compares the spectrum for a cool white fluores-
-8.0 eV-(-10.0 eV) = 620 nm (orangish-red).
=
cent light (solid line) to the spectrum for a tungsten
Here incident radiation of 275 nm can result in filament bulb (dashed line). The visible lines of the
fluorescence at 496 and 620 nm. mercury spectrum are also passed by the tube and
show up as spikes in the fluorescent spectrum.
Ultraviolet lamps (i.e., black lights) can be used to Fluorescent lights operate at the relatively low tem-
trigger fluorescence. Many minerals fluoresce natural- perature of 40C, and are more energy efficient than
ly. Some white substances, such as paper or shirts, incandescent lights. The spectral distribution of a
contain a fluorescent dye that emits in the blue end of fluorescent light can be varied by changing the phos-
the spectrum. This emission balances the natural phors used.
526 Geometrie, Physical, and Visual Optics

''KJJ
emission is much higher than stimulated emission.
Therefore, stimulated emission was overlooked until
Einstein pointed out its existence in 1916. Stimulated
emission raised the possibility of a process whereby a
large number of coherent photons could be produced.
The modern result is the laser.

1
1 I I 1 H
600 500 400
24.6 Laser Theory
FIGURE 24.9. Emission by a cool white fluorescent bulb (solid
curve) vs a tungsten filament bulb (dashed curve). Lasers are sources of coherent light. The word
LASER is an acronym for Light Amplification by
Stimulated Emission of Radiation. Lasers were pre-
24.5 Stimulated Emission ceded by MASERs (Microwave Amplification by
Stimulated Emission of Radiation). The maser was
According to the quantum theory, all states have some independently invented in the early 1950s by Charles
natural fluctuations. If an excited state happens to Townes of the United States and Alexander
fluctuate too much, a downward jump results with the Prokhorov and Nikolai Basov of Russia. These three
emission of a photon. This is called spontaneous emis- men shared the 1964 Nobel prize for their work on the
sion, and is responsible for the instability of the maser. In 1958, Townes and Arthur Schawlow set
excited states. forth the general conditions needed for a laser. Within
Now consider absorption by an atom that has a two years (1960), the first successful laser was built by
resonance frequency f01 connecting the ground state Ted Maiman of the United States. Arthur Schawlow
E 0 and the first excited state El. When a photon of received the 1981 Nobel Prize in Physics for his work
frequency f is incident on this atom, it perturbs the on lasers and laser optics.
atom, causing the electrical charge density to oscillate In a laser, a photon is used to trigger a stimulated
with the frequency of the photon. When f equals f01, emission resulting in two coherent photons. These two
the perturbation induces the electron to absorb the photons each trigger another stimulated emission re-
photon and jump to the first excited state Ej. sulting in four coherent photons. The four trigger
Now suppose a photon of the same energy is more stimulated emissions resulting in eight coherent
incident on the excited atom in state Ex. This photon photons. If the process continues, an avalanche of
again perturbs the atom causing the electric charge coherent photons is produced (i.e., light ampli-
density to oscillate with the frequency f of the photon, fication).
but this frequency is the resonance frequency f01, and In a medium at a certain temperature, most atomic
in this case induces the electron to emit a photon and electrons are in the ground state E 0 , the next most are
jump down to E 0 . The emitted photon is coherent in the first excited state E 1? and a decreasing number
with and travels in the same direction as the incident are in each successively higher excited state. Suppose
photon. This is called stimulated emission (Figure we try to start an avalanche by stimulating a transition
24.10). from Ej to E 0 . The initial coherent photons might be
Absorption and stimulated emission are related incident on atoms in E l 5 but they might also be
processes, and the corresponding transition prob- incident on atoms in the ground state. When the
abilities are identical in magnitude. Spontaneous emis- atoms are in the ground state, the photon is absorbed
sion is separate and has its own probability. Under and the avalanche is quenched. Therefore, in order to
normal light levels, the probability of spontaneous get an avalanche, we need more atoms in El than in
the ground state. This is called a population inversion.
Population inversions can be obtained by using
metastable excited states, since on an atomic scale the
metastable states live a very long time. The metastable
states are populated by correctly "pumping" the
medium. The pumping consists of supplying energy to
the medium. There are a variety of ways to do this:
ultraviolet radiation, electrical discharge, atom-atom
collisions, chemical reactions, etc.
FIGURE 24.10. Stimulated emission. Maiman's laser was a ruby laser, a three level laser
Emission, Absorption, and Photons 527

24.7 Laser Applications and Types


Laser light is highly monochromatic, coherent, and
. Ei (metastable) collimated. Furthermore, the laser energy output can
pump be incredibly high. Because of the high collimation,
stimulating the laser can deliver this energy to a distant target. In
addition, practically the entire power output can be
photon
concentrated into an extremely small area by focusing
the laser beam.
Eo Many of the laser applications are due to the
coherence (precision measurement by interference,
FIGURE 24.11. Energy diagram for a three-level laser.
holography, spatial filtering). Some are due to the
monochromaticity (fiber optics communication, the
triggering of specific chemical reactions). Others are
pumped by a xenon flashlamp. Figure 24.11 shows the due to the ability to focus the laser light into such a
basics of an energy level diagram for a three level small area (the optical stylus in laser video disk sys-
laser. The pump induces the transition from E 0 to E 2 . tems). Finally, there are the brute force power appli-
The electron then makes a transition (possibly non- cations (welding, cutting hard materials such as
radiative) down to Ej. Since E1 is metastable, the diamond, vaporizing small amounts of material).
electrons tend to stay there, and hence the number of Continuous beam helium-neon lasers are the most
electrons in Ej can be built up until there are more in common low power (milliwatt) lasers. The pump oc-
Ex than in E 0 (the population inversion). curs in the helium, and the energy is then transferred
Eventually, one of the metastable states sponta- to a metastable neon state by collisions between the
neously decays to the ground state and emits a helium and neon atoms (Figure 24.13). The neon lases
photon. This photon is incident on other Ex atoms and at the 632.8 nm transition between two excited states,
stimulates the emission of a second photon that is and then makes other transitions back down to the
coherent with the first. These two are then incident on ground state. The use of the helium makes the process
two more Ex atoms and stimulate the emission of two more efficient, so that a larger amount of laser ir-
more coherent photons. The stimulated emission av- radiance is obtained per unit of input energy.
alanche is now started. Since the pumping radiation A more powerful (watt range) gas laser is the argon
operates at a different wavelength (E0>E2), it does ion laser, which generates laser wavelengths at 488 nm
not compete with the avalanche. and 514.5 nm. There are a number of other powerful
Usually, the lasing medium is placed between two gas (or vapor) lasers, such as the krypton ion laser
reflectors so that the coherent photons travel back and (emission at a number of wavelengths including 530.9,
forth across the medium, thus building the strength of 568.1, and 647.1 nm).
the avalanche. Such a set-up is called a cavity oscil- In terms of power output, the carbon dioxide
lator. In order for the amplification to build in a cavity (C0 2 ) gas laser is in a class by itself. The C 0 2 laser
oscillator, the cavity must be designed so that a stand- actually emits in the infrared (10,600 nm). Even in
ing wave of the coherent photons is set up. The continuous mode, it can emit in the kilowatt range,
increase in number of the coherent photons corre- enough to cut through 2 cm of steel in a matter of
sponds to increasing the amplitude of the standing seconds.
wave. Eventually, either a gate mechanism is opened Excimer lasers contain rare gas halides such as
allowing the coherent light out (a pulsed laser), or a ArF. The word excimer is derived from excited state
small percentage of the light is steadily drained off aimer. These substances are unusual in that they have
through a partially transmitting mirror (a continuous no ground state, and are therefore automatically meta-
beam laser). The light emerging from the cavity oscil- stable. The excited state dimer can decay to the
lator consists of highly collimated plane waves (Figure individual atoms giving off a photon in the process:
24.12). ArF* > Ar + F + photon.

laser beam
laser -A&::S&K. -
FIGURE 24.12. Collimated plane waves of a laser beam.
528 Geometrie, Physical, and Visual Optics

-collision = - metastable
stimulating
coherent photons
photon

pump

"Eo -Eo
Helium Neon FIGURE 24.13. Simplified energy level diagram for a helium-
neon laser.

An incident photon of the correct energy can stimulate enabling us to deliver large amounts of energy to the
this decay, in which case the emitted photon is co- interior of the eye without surgically cutting the eye
herent with the stimulating photon. Excimer lasers are open. Furthermore, this energy can be precisely fo-
exciting because they can lse in the ultraviolet where cused into a small area. The surgical effects of the
each individual photon carries more energy. Pulsed laser beam are due to the wavelength (frequency) and
chemical reaction lasers have delivered pulses as short the amount of energy delivered. The effects are either
as 30 femtoseconds (10 _15 s). Powers as high as 1012 photochemical, ionizing, or thermal.
watts have been obtained during ultrashort bursts. Precise photochemical reactions can be initiated by
The Ruby laser (the first laser) is a solid laser, and intense laser beams of the proper energy. Excimer
is still widely used in pulsed mode at 694.3 nm. The (ultraviolet) lasers can break molecular bonds and
ruby laser requires high pumping powers because the volatize fragmented molecules, thus making very pre-
transitions terminate in the ground state, so the popu- cise incisions. These lasers have been used experimen-
lation inversion is hard to maintain. If the laser transi- tally for corneal surgery.
tion terminates in an excited state, the population A focused NdrYAG laser pulse can deliver enough
inversion is easier to maintain and requires less pump- energy to a small area to ionize (free) the bound
ing power. This is the basis for the neodymium laser, electrons resulting in a pool of free electrons called a
where the neodymium is imbedded in glass or yttrium plasma. Rapid expansion of the plasma creates shock
aluminum garnet (YAG). The Nd:YAG laser can waves that incise the target tissue. Transparent materi-
produce hundreds of watts at the infrared wavelength al can be incised in this way since the result is indepen-
of 1060 nm. dent of the tissue pigmentation. NdiYAG lasers are
Semiconductors, such as light emitting diodes being used for such things as discission of the posterior
(LEDs) can also be made to lse. The semiconductor lens capsule and peripheral iridectomy.
lasers are the smallest lasers (less than 0.5 mm). They Thermal effects result mainly from photovaporiza-
lse in the near infrared, and their power output tion or photocoagulation. A great advantage of photo-
ranges from milliwatts to watts. Their main application vaporization is that it produces a bloodless incision by
is in optical communications, video-disk systems, and sealing the blood vessels. Carbon dioxide (C0 2 ) lasers
other integrated optical-electronic equipment. have been used for photovaporization.
The high powered lasers mentioned above can be The argon ion and krypton ion lasers are used for
used as pumps to generate population inversions in photocoagulation particularly of the retinal region.
other materials. Under these conditions almost any The wavelength used is important since the ocular
material can be made to lse. In particular, liquid dyes contents scatter shorter wavelengths more than longer
can be used to make lasers that can be tuned for wavelengths (similar to Rayleigh scattering) and since
different wavelength emissions. the different ocular pigments have different absorption
strengths for the wavelengths. Therefore, different
tissues can be targeted by using a different wave-
length. For example, the yellow macular pigment
24.8 Laser Surgery xanthophyll absorbs the short wavelengths (less than
510 nm). Therefore, to avoid macular damage in pan-
The laser has revolutionized many surgical procedures retinal photocoagulation in the treatment of diabetic
especially where the eye is involved. The infrared and retinopathy, longer wavelengths are used (argon green
visible laser wavelengths are transmitted by the cornea 514.5 nm, krypton green 530.9 nm, krypton yellow
Emission, Absorption, and Photons 529

568.1 nm). A disadvantage of the greens and yellow is amplitude. In the wave mechanics version of quantum
that they are absorbed by hemoglobin, which can theory, a single particle, including a photon, is repre-
cause unwanted damage in the presence of subretinal sented by a wavepacket (Figure 24.14), and the square
hemorrhages. Krypton red, 647.1 nm, is not absorbed of the packet's wavefunction gives the probability of
by oxygenated hemoglobin, and is safer to use in the finding the photon at a particular position at a particu-
presence of subretinal hemorrhages. Krypton red also lar time. (In other words, the probability of finding
penetrates deeper into the retinal structures. A dis- the energy.)
advantage of krypton red is difficulty in controlling In Figure 24.14a, the photon's wavefunction lies
choriodal bleeding. between A and C. We know the photon is somewhere
between A and C, but we do not know exactly where.
The photon is most likely to be at the positions with a
large wavefunction displacement and less likely to be
24.9 Laser Safety at the small displacement positions.
Suppose the photon represented in Figure 24.14a is
High powered lasers are dangerous to any part of the passing through an aperture that is rigged to close
human body. Low powered lasers, such as the mil- quickly at point B thereby dividing the photon's
liwattt helium neon lasers, are dangerous only to the wavefunction into two parts. However, the photon
eye. The laser beam is well collimated (very direction- itself is not divisible. So does the photon get through
al), and has a small enough beam diameter so that the the aperture or not?
entire undiverged beam can enter the eye. Further- Based on the wavefunction information, we do not
more, the eye focuses the collimated laser light into an know the answer. All we can determine is the prob-
even smaller area on the retina. A retinal burn may ability of whether the photon gets through. Let us
result even before the person can flinch. A foveal burn assume that 40% of the square of the wavefunction
results in severe and irreversible loss in acuity. For this gets through. This means that the photon had a 40%
reason, great care should be exercised in the vicinity chance of getting through and a 60% chance of not
of laser beams. Besides a direct exposure, the laser getting through. If we repeat this experiment many
operator should avoid specular reflections of the laser times, we cannot predict what happens to any one
beam that might cause an exposure from an un- photon, but we do know that statistically the photon
expected direction. Interference filters are available will get through in 40% of the cases and will not get
that protect the eye by reflecting the laser wavelength through in 60%.
while transmitting other wavelengths.

24.11 Uncertainty
24.10 Photons and Probability
The beats shown in Figure 21.11b were the result of
In classical harmonic wave theory, the energy carried the superposition of two harmonic waves of slightly
by the wave is proportional to the square of the wave's different frequencies. The photon wavepackets of Fi-

s\ /I/ 1 (\, \

A
t
B c
1 I
V

FIGURE 24.14. Single photon wave func-


tions. a. Short emission time. b. Longer
b) emission time.
530 Geometrie, Physical, and Visual Optics

gure 24.14 can be made by the superposition of a Heisenberg's uncertainty principle applies to all
number of harmonic waves each of slightly different quantum particles. To understand the effect of a
frequency. Photons with a long emission time have measurement, consider an electron moving with a
wavepackets that resemble harmonic waves, and well-defined momentum p. One way to see objects is
hence only a few harmonic waves need to be super- to shine light on them and then observe the diffusely
posed to generate the photon wavepacket (Figure reflected (scattered) light. So to see the electron, we
24.14b). Thus, photons with long wavepackets have a will shine light on it and then observe the light that the
relatively well-defined frequency and wavelength. electron scatters. However, an electron is so small that
Many different harmonic waves are needed to a major collision results when an incident photon hits
make a short photon wavepacket (Figure 24.14a). So it. Therefore, while the scattered photon gives us
photons with short wavepackets have a large uncer- information on the electron's position, we have now
tainty in wavelength or frequency, i.e., the photon's changed the electron's momentum increasing the un-
frequency is fAf where Af is large. Since the certainty in Ap.
photon's energy is related to its frequency, photons [The Vietnam War protests provide a real world
with a large uncertainty in frequency also have a large analogy to the idea that measurements change the
uncertainty in energy. action. In theory, the television reporters with their
We can generalize these uncertainty relations by cameras were present at the protests only to record
first noting that the uncertainty in frequency is inverse- (measure) the events. Presumably, the protest events
ly proportional to the emission time At: would occur exactly the same whether or not the
television reporters were there. However, the protests
Af~l/At, seemed to gain momentum when the television crews
or showed up. So the measuring device, the television
crews, influenced the event they were suppose to be
AfAt~l. measuring.]
Then
hAfAt~h,
and from Eq. 24.5,
24.12 Diffraction and Single Photons
AEAt~h. (24.11)
The length of a photon wavepacket, Ax in Figure When collimated light is incident on a narrow slit, we
24.14, equals c times At, so Eq. 24.11 becomes obtain a single-slit diffraction pattern on a distant
screen. Now suppose the light level is decreased until
AE[Ax/c]~h. (24.12) only one photon at a time is incident on the slit. As
Then from Eq. 24.8, time passes the individual photons seem to land at
random positions on the screen, but after awhile, the
ApAx~h. (24.13) cumulative effect of the energy delivered by the
photons is equal to the irradiance of the classic single-
Equation 24.13 is the famous Heisenberg uncer-
slit diffraction pattern.
tainty principle, which couples the uncertainties in
momentum and position together. A photon with a The irradiance on the screen is related to the
short wavepacket has a small Ax. Then since A p ~ square of the wavefunction. In the quantum theory,
h/Ax, the uncertainty in momentum is large. Consider the square of the wavefunction tells the probability of
the photons that get through the aperture of Section where each individual photon will hit the screen. The
24.10. These photons have a shorter wavefunction most likely position for the photons to hit is at the
than the incident photons. Thus, closing the aperture central maximum. The least likely position for the
increased the uncertainty in the photon's momentum photon to hit is at the minimum positions given by the
(and consequently in the energy). photon's wavefunction.
Position measurements make Ax smaller, while We can relate the single-slit diffraction to the
momentum measurements make Ap smaller. Heisen- Heisenberg uncertainty principle. We know the posi-
berg^ uncertainty principal says that when we make tion of the photons that pass through the slit to within
one uncertainty smaller the other gets larger. In other an uncertainty Ax equal to the slit width a. According
words, the precise simultaneous knowledge of momen- to Heisenberg's uncertainty principle, the position
tum and position is not possible. Basically, the meas- measurement results in a momentum uncertainty per-
urement of position changes (increases the uncertainty pendicular to the beam of
in) the momentum and vice versa. Ap~h/a.
Emission, Absorption, and Photons 531

For a small , we can use the small angle approxima-


tion for the tangent to obtain the angular deviation
from the straight ahead direction:

=
P '
Then
h/a
=
P
a)
Since the photon's momentum p equals h/, we have
h/a
= ~? ^r~
h/'
or A,*

a
This equation says that the increased uncertainty in
momentum causes the photons to spread out in the VA
angular zone bounded by /, but this equation is just
the small angle version of Eq. 22.1 for the first min-
imum in a single-slit diffraction pattern. So we see that b)
the single slit makes a position measurement on the
FIGURE 24.15. Single photon buildup of double-slit interfer-
photons, and that position measurement changes the ence pattern, a. A few photonsinterference fringes not obvious.
photon's momentum, which results in the angular b. More photonsinterference fringes emerging.
spread that we classically call diffraction.

also noted for his correct prediction of the existence of


antimatter.)
24.13 Interference and Single Photons For the double-slit interference experiment every-
thing seems fine until someone asks, "Which slit does
Consider a double-slit experiment in which only one the photon go through?" The way to experimentally
photon at a time is incident on the double slits. The determine the answer is to rig the slits so that we can
irradiance in the double-slit interference pattern is tell which slit the photon goes through. One possible
related to the square of the wavefunction and defines way to do this is to put a red transmitting filter in front
the probability of where the photon will hit the screen. of one slit and a green transmitting filter in front of the
The photon has a high probability of hitting the screen other slit. Then we would look to see if the photon
at or near the maximum positions, and little or no that hits the screen is green or red. When we do this,
probability of hitting the screen at or near the min- the interference fringes vanish. After all, even from
imum positions. classical wave theory, red and green light have differ-
We do not know exactly where any one photon will ent frequencies and are not mutually coherent.
hit the screen, but we do know the probability. If we Perhaps we can use monochromatic light and
run enough identical photons through the double slits, polarizing filters to tell which slit the photon goes
we eventually end up with double-slit interference through. Let us put a polaroid with a vertical transmis-
fringes. In Figure 24.15, each dot shown represents a sion axis in front of one slit and a polaroid with a
photon hit on the screen. The first frame (a) shows the horizontal transmission axis in front of the other slit.
results after a few photons. The second frame (b) Then we can measure the polarization of the photon
shows the results for more photons. Here the photon incident on the screen to determine which slit the
build up is already beginning to resemble the double- photon went through. When we do this, the interfer-
slit interference fringes. ence fringes again vanish. After all, even from classi-
From the single photon double-slit interference cal wave theory, light plane-polarized perpendicular to
experiments, it is clear that interference is a single each other is not mutually coherent and so no interfer-
photon effect. As Paul Dirac said, "The photon inter- ence occurs.
feres with itself." In fact, this is already evident in the In fact, any experiment rigged to tell what slit the
single photon diffraction experiments. (Paul Dirac is photon goes through destroys the coherence, and
532 Geometrie, Physical, and Visual Optics

therefore, the interference does not occur. Interfer- the mirror to C. In wave theory, we superpose the
ence only occurs when we cannot tell which slit the waves reflecting from each small area of the mirror
photon goes through. In general, interference is a (1,2,3, . . .). The effect of the superposition is that
single particle phenomenon that occurs when there is the waves coming from area 1 destructively interfere
more than one possible path for the particle and we with those from area 2, and those coming from area 3
cannot tell which path the particle took. In order to destructively interfere with those coming from area 4.
make the correct predictions, we need to superpose The only waves that do not destructively interfere are
the possible paths. Clearly, one approach is to use those that reflect from the area around B, the geomet-
wave theory to give the superposition. ric optics reflection point. These paths give waves that
Figure 24.16a shows light from source A incident constructively interfere at C, and so we say that the
on a mirror. After reflecting, the light travels to the light reflected from area B (the geometric optics law
detector at C. Now by the law of reflection, the ray of reflection), but, in fact, this analysis shows that we
that leaves A and gets to C reflects at the position do not really know where the photon is reflected from.
marked B on the mirror. Suppose the light level is cut The result is a superposition of all the possible paths.
until there is only one photon at a time incident on the Suppose we section the mirror, and keep only those
mirror. Based on our discussions so far, we need to parts that would constructively interfere (i.e., areas 1,
consider every possible path of the photon from A to 3, 5, etc.). In particular, we will eliminate area B, the

FIGURE 24.16. a. Contributions of various areas


to single photon reflection, b. Single photon paths for
selected areasa reflection grating.
Emission, Absorption, and Photons 533

geometric optics reflection region (Figure 24.16b). Problems


Even though we have eliminated B, the photon still
has a good probability of arriving at the detector C 1. What is the peak wavelength for a blackbody at
due to the constructive interference from the paths 4,800 K? For a blackbody at 1,000 K?
that we kept, i.e., from wave theory, each small 2. What is the single photon energy for light of
section acts like a point source that sends waves out in wavelength 500 nm? For light of wavelength
all allowed directions. At point C, the waves from the 600 nm?
sections that we kept constructively interfere. So the 3. Describe the emission spectra for a low pressure
sectored mirror is a reflection-type diffraction grating, gas with the following energy levels: -7.50 eV,
and point C is a maximum of the diffraction grating. -5.0 eV, -3.0 eV.
This process somehow occurs even for a single photon. 4. The emission spectra of a low pressure source
includes wavelengths at 510 nm and 610 nm. If the
ground state energy level is at -15.00 eV, list two
different possibilities for the other energy levels.
24.14 Photon Finale How could one experimentally check which
scheme is correct?
In these last photon sections, a discussion has been 5. In terms of an energy level diagram, show why
made of how photons and probability theory are some rocks when irradiated by ultraviolet give off
intricately linked. Second, some single photon prop- visible light. What is this called?
erties are shown, as well as how they merge with 6. What does the word laser stand for?
classical wave theory. There is much, much more to 7. What does the word excimer stand for?
the quantum theory than indicated here, but the rest is 8. For retinal photocoagulation, what factors in-
beyond the scope of this book. fluence the decision to use an argon green vs a
The conceptual interpretation of quantum theory is krypton red laser?
a real challenge to modern man. Nevertheless, the 9. A double-slit interference experiment is changed
general quantum theory is a very successful physical so that we can determine which slit the light goes
theory, the products of which include antimatter, through. How does the change affect the interfer-
quarks, electron microscopes, lasers, nuclear power, ence fringes?
radioactive isotopes for medical applications, and 10. Interference fringes are formed on a screen in a
semiconductors (which have led to solid state elec- double-slit experiment. Then the light level is
tronics and many of the modern high tech devices that decreased to the smallest possible nonzero level.
we enjoy, including personal computers and handheld Describe the resulting illumination pattern on the
calculators). screen, as a function of time.
CHAPTER
TWENTY-FIVE

Spatial Distribution
of Optical
Information

25.1 Spatial Frequency be transferred by the optical system even though the
higher spatial frequencies are wiped out.
There is a variety of detail in the castle scene in Figure To completely judge the quality of an optical sys-
25.1. The finest detail is in the grating on the entrance. tem, we need to know how it images each of the
The portals on the top of the castle walls constitute the spatial frequencies present in the object. The units for
next finest detail. Then in order of decreasing detail, spatial frequency are repetitions per unit length, more
we have the towers and the main blob of the castle commonly called cycles per unit length (e.g., cycles/
itself, which is the grossest detail. mm = mm _ 1 ). The spatial frequency information can
We can classify the detail in terms of the number of also be converted to angular terms such as cycles per
repetitions per unit distance. Across the width of the radian, cycles per degree, or cycles per minute.
castle, the towers repeat three times, and the portals
repeat 13 times. If the gate were the width of the
entire castle, the grating would repeat about 75 times.
The finest detail has the highest repetition rate, while 25.2 Sine vs Square Wave Gratings
the grossest detail has the lowest repetition rate. The
repetition rate is more formally referred to as the A crude way to analyze the information transfer for a
spatial frequency. (Be careful not to confuse the spa- given spatial frequency is to use a chart consisting of
tial frequency with the wavelength or with the tempor- equally spaced black and white bars (Figure 25.2a).
al frequency of the light waves.) The repetition rate of each set of bars (one white and
It turns out that the quality of the information one black) determines the spatial frequency. Figure
transferred from the object plane to the image plane 25.2b shows a plot of the luminance for the bar chart
by an optical system is a function of the spatial as a function of distance across the chart. This lumi-
frequency. For example, consider an uncorrected low nance distribution is referred to as a square wave
myope who looks at the distant castle but cannot distribution, and the bar chart is called a square wave
resolve the detail on the entrance grating because of grating. Figures 25.2c and 25.2d show the bar chart
his blurred retinal image. This myope can still resolve and luminance distribution for a higher spatial fre-
the towers and possibly the portals. A person with quency.
added blur (perhaps an uncorrected medium myope) The method of analysis consists of comparing the
cannot resolve the portals, but can still count the three modulation (also called visibility or contrast) in the
towers. With additional blur (perhaps an uncorrected object to the resulting modulation in the image. The
high myope), the person loses the ability to resolve the modulation M is defined as
three towers, but perhaps can still see the main grey
blob of the castle itself. This example shows that (25.1)
information for the lower spatial frequencies can still

535
536 Geometrie, Physical, and Visual Optics

[UU] ruin RJUl

ui/ LnnJ

FIGURE 25.1. A castle scene with a variety of spatial


frequency components.

where Lmax and Lmin are the maximum and minimum For a proper analysis of information transfer as a
luminances, respectively. Contrast is usually defined in function of spatial frequency in an optical system using
terms of the luminance difference between the object incoherent light, we need to use a grating chart with a
L0 and the background L b ; contrast = (L 0 - L b )/L b . sinusoidal luminance distribution. We refer to the
When we specify that the mean luminance Lm serves sinusoidal chart as a sine wave grating. The solid curve
as the background for the bar chart, then the modula- in Figure 25.3a is the luminance distribution for a sine
tion equals the contrast. wave grating, while the dashed line is the luminance

a)

J
Lmax I
t 1-maxT

FIGURE 25.2. a. A low fre-


Lmin quency square wave grating, b.
Its luminance distribution, c. A
higher frequency square wave
d) grating, d. Its luminance distri-
bution.
Spatial Distribution of Optical Information 537

Lf -max

t-min-l-

b)

FIGURE 25.3. a. Object luminance distribu-


tion (solid) for a sine wave grating vs that of a
square wave grating (dashed) of the same fre-
quency. b. Blurred image luminance distribu-
tion for sinusoidal object distribution of a. c.
Phase-shifted image.

distribution for a square wave grating of the same 25.1 and with a little algebra obtain
frequency and amplitude. (To imagine a sine wave
grating, think of a blurred bar chart.) (25.4)
As a function of the distance x across a sine wave
grating, the luminance is described by For a sine wave grating, the absolute minimum
luminance value is zero, in which case a equals L m .
L0(x) = Lm + a sin(27rfx), (25.2) For those sine wave gratings that have a minimum
where the argument of the sine function is in radians. luminance greater than zero, a is less than L m . Hence,
When the argument is in degrees, the sinusoidal lumi- the modulation M is always less than or equal to 1.
nance distribution is given by We can now factor Lm out of Eq. 25.2 and write the
luminance distribution for an object grating with mod-
L0(x) = Lm + a sin(360fx). (25.3) ulation M0 as
In Eqs. 25.2 and 25.3, f is the spatial frequency of the L0(x) = L J l + M0sin(2*rfx)]. (25.5)
sine wave grating, Lm is the mean luminance, and a is
the amplitude of the deviation from the mean. Consider a misfocused optical system with a sine
For a sine wave grating, wave grating given by Eq. 25.5 as the object. One of
the advantages of sinusoidal luminance distributions is
l^max+ a,
^ and L -a. that for most optical systems the blurred image lumi-
We can then substitute the above equations into Eq. nance distribution L; is also sinusoidal.
538 Geometrie, Physical, and Visual Optics

The spatial frequency f of the image sine wave has an object luminance distribution of
grating is related to the object spatial frequency f by L0 = 10.0 + 8.0 sin(27r3 mm_1x).
f = f/m where m is the lateral magnification of the
When the MTF for 3 cycles/mm is 0.40, what is the
system. Due to the blur, the modulation Mj of the image luminance distribution? (Assume a zero
image grating is less than the modulation M0 of the phase shift).
object grating. Furthermore, if the blur happens to be For unit magnification, the conjugate image dis-
asymmetric, such as would be given by the aberration tribution is a sinusoidal distribution of the same
coma, the image grating is phase shifted through an spatial frequency in which the modulation is de-
angle relative to the object grating. So in general, creased by a factor of 0.40. We can rewrite the
object distribution as
Li(x) = L J 1 + M; sin(27rf x - )]. (25.6)
L0 = 10.0[1 + 0.80 sin(27r3 mm_1x)].
When magnification occurs, f is less than f. When The object modulation M0 is then 0.80, and the
minification occurs, f is greater than f. For a lateral image modulation is
magnification of unit magnitude, f = f.
Usually, the blur is symmetric, in which case there Mj = MTF 0.80 = (0.40)(0.80) = 0.32.
is no phase shift ( = 0). Then The image distribution is
Li(x) = L m [l + Mi sin(27rf'x)]. (25.7) ^ = 10.0[1 + 0.32 8(23 mm_1x)],
or
For unit lateral magnification, Figure 25.3b shows the
blurred image distribution for the sinusoidal object ^ = 10.0 + 3.2 sin(27r3 mm_1x).
distribution given in Figure 25.3a. For completeness,
Figure 25.3c shows the blurred image distribution for a The transfer of information in a perfect optical
90 phase shift. system is limited by diffraction (i.e., a diffraction-
The fact that the blurred image of a sine wave limited system). Diffraction has little effect on low
grating of a single spatial frequency f is a sine wave spatial frequencies. As the spatial frequency is in-
grating of a single spatial frequency f is a special creased, diffraction causes an increasing loss of infor-
property of sine wave gratings. In particular, a blurred mation, so that eventually very fine detail is complete-
square wave grating is not imaged as a square wave ly lost (Rayleigh's criterion in Chapter 22). Fig-
grating. In fact for enough blur, a square wave grating ure 25.4 shows a plot of an MTF as a function of
is imaged as a sine wave grating. For example, the sine increasing spatial frequency f. Because of diffraction,
wave grating of Figure 25.3b could be the blurred the MTF eventually goes to zero at a frequency fc
image of the square wave grating represented by the called the cut-off frequency. All higher spatial fre-
dashed line in Figure 25.3a. quencies are wiped out by the system. In a sense the
cut-off frequency fc corresponds to Rayleigh's criterion
for resolution.
When other sources of blur are present, then the
25.3 The Modulation Transfer Function MTF is less than or equal to the diffraction-limited
MTF at each spatial frequency. Depending on the
For incoherent light, the modulation transfer function amount and source of blur, the system may have
(MTF) of the optical system for the object spatial an effective cut-off fn sooner than the diffraction cut-
frequency f is the ratio off f.

MTF(f) = Mi (25.8)
Mn MTF
where Mj is the sinusoidal image modulation and M0 is
the sinusoidal object modulation. A perfect optical
system preserves the modulation, so the MTF is 1. For
optical systems with blur, aberrations, scattering, etc.,
the image modulation is decreased, so the MTF is less
than 1. When the image is completely blurred out, the
information is lost and the MTF is zero. Hence,
0<MTF<1.

EXAMPLE 25.1 FIGURE 25.4. Modulation transfer function plotted vs spatial


An optical system with unit lateral magnification frequency f.
Spatial Distribution of Optical Information 539

25.4 Periodic Gratings and Linear Systems spatial frequency f can be represented as the sum of a
series of sinusoidals of integer multiples of f. The
Certainly one of the reasons that a sine wave grating is mathematical formulation of this statement is known
special is that most optical systems image it as a sine as Fourier's theorem.
wave grating. However, there is another reason why As a result of Fourier's theorem, we can consider
sine wave gratings are special. An optical system tends the luminance distribution Lp(x) of any periodic grat-
to function as a linear system. This means that if the ing of spatial frequency f as represented by the sum of
response for a sine wave grating Lx of spatial fre- sinusoidal distributions of frequencies which are in-
quency x is A, and the response for a sine wave teger multiples of f, or
grating L2 of spatial frequency f2 is B, then the re-
sponse for the object luminance (hl + L 2 ) is (A + B). Lp(x) = L m [l + M 1 sin(27rfx-0 1 )
Note that for a linear system, a spatial frequency f + M2 sin(27r2fx - 2)
can appear in the image distribution only when the
spatial frequency f (where f = m f ) appears in the + M3sin(27r3fx-</>3) + . . .].
object distribution. For a system with unit magnifica- (25.9)
tion, this means that the only spatial frequencies The lowest frequency (f) component supplies the gross
appearing in the image are those already present in detail and is called the fundamental harmonic. Then
the object. finer and finer details are successively supplied by the
Consider a series of sinusoidal distributions where higher frequency components, i.e., the third harmonic
the lowest spatial frequency is f, and the higher spatial (3f) supplies finer detail than the second harmonic
frequencies are integer multiples of f (i.e., 2f, 3f, 4f, (2f), and so on.
etc.) With an oscilloscope, it is easy to show the Figure 25.5a shows two cycles of a luminance distri-
resultant sum of such a series is a periodic distribution bution for a periodic grating described by the sum of
of frequency f. In fact, any periodic distribution of two sine waves. The lower frequency sinusoidal distri-

FIGURE 25.5. a. Object luminance


distribution, b. Conjugate image
luminance distribution.
540 Geometrie, Physical, and Visual Optics

bution supplies the gross shape (the large slowly vary- or


ing bumps), while the higher frequency sinusoidal Li = 5 + 1 . 5 s i n ^ x ) .
supplies the fine detail (the little bumps).
The general equation for the luminance distribution The image luminance distribution is plotted
of a periodic object grating with zero phase compo- in Figure 25.5b. The low frequency (large
nents is bumps) information is degraded but still present.
The fine detail (small bumps) is blurred out.
Lp(x) = L J 1 + Ml sin(27rfx) + M2 sin(27r2fx)
+ M 3 8(23) + . . .]. (25.10) EXAMPLE 25.3
The MTF for a linear optical system with a
Let us assume that the optical system is linear, has lateral magnification of unit magnitude is
a lateral magnification of unit magnitude, and does not
introduce any phase shifts. Then the image luminance
distribution Lj is given by
MTF
Li(x) = L m [l + N1 sin(27rfx) + N2 sin(27r2fx)
+ N3sin(27r3fx) + . . . ] , (25.11) 1.0
1.0
where the modulation transfer functions for each spa- 0.9
0.8
tial frequency are 0.6
MTF(f) = N 1 /M 1 , 0.3
0.1
MTF(2f) = N 2 /M 2 , 0.0
MTF(3f) = N 3 /M 3 ,
etc.
When the object luminance distribution is given by
EXAMPLE 25.2 L0 = 6[1 + 0.50 sin(27r2x) + 0.35 sin(2<7r4x)
Consider a system with unit lateral magnification
and the following MTF values: + 0.25 8(26)],
what is the image luminance distribution? (Assume
no phase shifts are introduced by the system.)
f MTF
Here the non-zero transfers are:
1.0 N2 = 0.9 M2 = 0.9(0.50) = 0.45,
0.5 N4 = 0.6 M4 = 0.6(0.35) = 0.21,
0.3
0.0
N6 = 0.1 M6 = 0.1(0.25) = 0.025.
Since the optical system is linear, the image distri-
bution is
When the object has the luminance distribution Lj = 6[1 + 0.45 sin(27r2x) + 0.21 sin(27r4x)
of Figure 25.5a described by + 0.025 sin(27r6x)],
L0 = 5 + 3 sin(27rx) + sin(27r3x),
The system transferred the fundamental harmonic
what is the image luminance distribution? (As- well. There was some degradation of the middle spa-
sume no phase shifts are introduced by the tial frequency, and the highest spatial frequency (the
system.) finest detail) was essentially lost. The solid curve in
Here Figure 25.6 shows one cycle of the object luminance
L0 = 5[1 + 0.60 sin(27Tx) + 0.20 sin(2<7r3x)], distribution, while the dashed curve shows one cycle
of the image luminance distribution.
and
We previously said that the blurred image of a
N,= 0.5M,= 0.5(0.60) = 0.30, square wave grating is a sine wave grating. We can
N2 = 0.3 M2 = 0.3(0) = 0, now show this mathematically. A square wave is
N3 = 0 M3 = 0(0.20) = 0. periodic. The Fourier sine series for the square wave
luminance distribution is
Since the optical system is linear, the image
distribution is Lsquare(x) = L J 1 + (C/l) sin(27rfx) + (C/3) sin(27r3fx)
Li=5[l + 0.30sin(27rx)], + (C/5)sin(27r5fx)+ . . . ] ,
Spatial Distribution of Optical Information 541

12 +

Lt

FIGURE 25.6. Object (solid) and


image (dashed) luminance distribu-
tions.

where the constant C equals 4/. The sin(27rfx) term This section pointed out that the luminance distri-
supplies the gross shape. Then each successive term bution for any periodic grating can be represented as
(3f, 5f, etc.) provides more and more fine detail. the sum of a series of sinusoidal luminance distribu-
Figure 25.5a is the sum of the first two terms in the tions. The lower frequency components are respons-
series, and we can see that the gross square wave ible for the gross detail, while the higher frequency
shape is already emerging. The addition of the higher components are responsible for the fine detail. In a
order terms sharpens the fine detail. linear optical system, each sinusoidal component is
For an ideal system (diffraction neglected), the imaged by the system according to the MTF for that
MTF is 1 and a square wave grating is imaged as a spatial frequency. The image then is the sum of the
square wave grating. For a diffraction-limited system, resulting sinusoidal image components.
the MTF falls to zero, and so some high frequency Rayleigh's criterion supplies information equivalent
terms in the sum for L square are missing. If the diffrac- to the cut-off frequency fc. The MTF analysis is much
tion is small and the fundamental frequency f is low, better because it supplies information about the entire
then the lower order terms are not affected much and spatial frequency range from zero to fc. Figure 25.7
the image still looks like a square wave grating. How- shows MTFs for two optical systems (A and B). The
ever, if the diffraction is larger, the fundamental MTF for system B has a lower cut-off frequency than
frequency f is higher, or another source of image the MTF for system A. However, over the middle
degradation is present, then the lower order fre-
quencies are affected and each is transferred by a
different MTF factor. Here the luminance distribution
of the image is no longer a square wave. Finally, when MTF
the diffraction (or other sources of blur) are large
enough so that all the frequencies above the fun-
damental frequency are wiped out, then the image of
the square wave grating becomes
Li(x) = Lm + Mi sin(27rfx) + 0,
where Mt equals the MTF times the constant C. In this
case the square wave grating has been imaged as a sine
wave grating (Figure 25.5b). FIGURE 25.7. MTFs for different systems.
542 Geometrie, Physical, and Visual Optics

spatial frequencies, the MTF for system A is less than


that for system B. Suppose a human observer is to use
these two systems, and the cut-off for the human
observer is at fd, which corresponds to the middle of
the spatial frequencies transferred by the systems. For
all spatial frequencies lower than fd, system B has a
better MTF. Therefore, optical system B is the better
system for the observer to use even though a
Rayleigh's criterion analysis would have said that sys- a)
tem A is the better system. (This example is based on
a dilemma that appeared in the development of televi-
sion systems. The dilemma was that people preferred
one system (B) over another (A) even though accord-
ing to Rayleigh's criterion system A was better.)

25.5 Spread Functions and the MTF


Ideally, the image of a point source is a point image.
However, even in a properly focused system, diffrac-
tion and the aberrations cause the illuminance to
b)
/
spread out around the point image position. The point FIGURE 25.9. a. Line spread function, b. Effect on sinusoidal
spread function is a relative measure of the luminous luminance distribution.
flux at the positions around the ideal point image. In a
diffraction-limited system, the point spread function is
determined by the Fraunhofer diffraction pattern When the gratings of the previous section are
(Airy's disk and its rings). In the presence of aberra- imaged, the point spread function causes the degrada-
tions, misfocus, or scatter, the light is spread even tion in the modulation. Because of the symmetry of
more. the gratings, the blur is in the horizontal meridian, and
The ideal image of a vertical line is a vertical line we can work directly with the line spread function
with zero horizontal width. However, due to the S(g). Given a system with a line spread function, and
system's point spread function, the light in the image a sine wave grating as an object, we want to show
plane is spread horizontally (Figure 25.8). A horizon- mathematically that, for incoherent light, the image is
tal plot S(g) of the relative amounts of luminous flux is also a sine wave grating.
called the line spread function (Figure 25.9a). If the For simplicity, let us assume that the system has a
point spread function has been measured, then the lateral magnification of unit magnitude, and that the
line spread function can be calculated from the point sinusoidal grating has a mean luminance of 1. With
spread function. Alternatively, the line spread func- these assumptions, the object luminance distribution is
tion can be directly measured.
L0(x) = l + M0sin(27rfx). (25.12)
The perfect image distribution is identical to the ob-
line ject distribution. However, because of the line spread
point
function S(g), the illuminance from a number of the
ideal horizontal image points overlaps, as represented
by the overlapping circles in Figure 25.9b. Because the
light is incoherent, the resultant illuminance is the sum
of all the overlapping illuminances.
point spread line spread
To analyze these contributions, first divide the
image plane up into small vertical strips each of
horizontal width Ag. Illuminance equals the luminous
flux incident per unit area. So for the j t h strip, the
W contribution to the illuminance from the line spread
FIGURE 25.8. Point and line objects and resulting spreads in function equals S(g) times Ag.
image plane. The contribution to the illuminance distribution at
Spatial Distribution of Optical Information 543

point x from the strip centered on point Wj equals the We can use Eqs. 25.16 through 25.18 together with
contribution from the line spread function times the Eq. 25.13 and
ideal illuminance in the strip containing wj5 or
sin(a - b) = sin a cos b - cos a sin b,
Six-WjJil + MoSinpTrfWjDAgj. to obtain
Then Lj = 1 + M 0 |F| sin(27rfx - ). (25.19)
gj=x-wj, Equation 25.19 shows that, in the presence of a line
spread function S(g), a sine wave grating is imaged as
in which case we can write the product as
a sine wave grating with a possible phase shift , and a
5( 8 ){1 + 0 8[27(- 8 ;)]} modulation M; where
The line spread function can change as the object is M^MolFl. (25.20)
moved further and further off axis. We will restrict
ourselves to an area, called an isoplanatic patch, over From Eq. 25.8, we see that the modulation transfer
function equals |F|.
which the line spread function does not change. Then
the total distribution at point x is the sum of the So not only have we shown that sine wave gratings
are imaged as sine wave gratings, we also see that the
distributions from each strip.
MTF is directly related to the line spread function
L,(x) = S( g j ){l + M0 sin[27rf(x - gj )]} A gj . S(g) through the sums of Eqs. 25.14 and 25.15.
In the limit as the strip width Agj goes to zero, the
Once gj is large enough, the line spread function S(gj)
sums given in Eqs. 25.14 and 25.15 become the in-
is zero, and the series can be terminated.
tegrals
From the trig relationship
sin(a - b) = sin a cos b - cos a sin b, F s = | o S(g)sin(27rfg)dg, (25.21)

we have
F c = j o S(g)cos(27rfg)dg. (25.22)
MX) = S( gj ){l + M0[sin(277fx) cos(27rfgj)
^ (2) (2 ] )]} The above integrals are called the Fourier sine and
cosine transforms, and we say that the MTF is the
For simplicity, let us pick units so that the sum of the Fourier transform of the line spread function.
spread function is 1. In other words, the area under The phase shift is also a function of the spatial
the line spread function is 1, or frequency f. The MTF and the phase shift are consi-
XS( gj )A gj = l. dered two vectorlike components of the optical trans-
j form function (OTF), i.e.,
Then we can bring the spread function inside the curly OTF=(|F|,0). (25.23)
brackets and consider the sum term by term to obtain
The above representation is the polar coordinate rep-
Lj(x) = 1 + M0FC sin(27rfx) - M0FS cos(27rfx), resentation for the OTF. The rectangular component
(25.13) representation is
where OTF=(Fc,Fs). (25.24)
F.^S^sin^fg^Agj, (25.14) A simplification in the general theory results when
the line spread function is symmetrical. Then the sum
F^Es^cos^Trfg^Agj. (25.15) (or integral) F s is zero, and Eq. 25.18 shows that the
j
phase shift is either zero or 180 (depending on
The above two equations are similar to the sums of
whether F c is positive or negative). Hence, only non-
vector components, so the resultant sum also acts like
symmetrical line spread functions cause a phase shift
a vector. Therefore,
other than 180 in the image grating.
Fc = |F| cos and F s = |F| sin , A 180 phase shift is a phase reversal where the
(25.16) minimums replace the maximums and vice versa (like
the negative of a photograph). So other than a pos-
in which case sible phase reversal, the image grating is shifted rela-
|F| = [F* + F*]1/2, (25.17) tive to the object grating only for a nonsymmetrical
line spread function. Off-axis blur patches, or asym-
and metric aberrations such as coma, cause a nonsymmet-
tan< = F s /F c . (25.18) rical line spread function.
544 Geometrie, Physical, and Visual Optics

,
a) ^ - ^ b)
FIGURE 25.10. Spurious resolution for radial lines.

For a symmetric line spread function, we can sim- MTF(f) = l - ( f / f c ) , (25.28)


plify Eq. 25.19:
where fc is the incoherent cut-off frequency. Accord-
L,(x) = 1 + Mi sin(27rfx), (25.25) ing to Eq. 25.28, the (horizontal) MTF for an inco-
herent diffraction-limited system with a square aper-
where
ture is a straight line that extends from 1 to 0 (Figure
Mj = M 0 F c . (25.26) 25.11). A square aperture system with aberrations,
defocus, scatter, etc., has an MTF that is less than the
In this formulation F c is identified as the MTF:
straight line given by Eq. 25.28.
MTF = FC. (25.27) In image space, the cut-off frequency is given by
What happens optically when F c goes from positive fc = w/(Ai;), (25.29)
to negative (Figure 25.10a)? For a positive F c near
where v is the image distance. For a distant object,
zero, the grating is imaged with a decreased modula-
tion and no phase shift. When F c is zero, no modula- v = f2, and i;/w equals the system's f-number. Then
tion is transferred (the grating is blurred out). For the fc = l/[A(f-number)]. (25.30)
negative F c , the grating reappears but is phase-
reversed. This phenomenon is called spurious resolu- It is often convenient to express the image spatial
tion and is shown in Figure 22.10b for the image of frequency in cycles per mm. Let us take a wavelength
some radial lines. As the radial lines get closer, the of 555.55 nm as representative of the middle of the
spatial frequency increases. Frequently, the first zero visual spectrum. Since 555.55 nm = 555.55 x 10"6 mm,
is taken as the effective cut-off for the useful trans- the cut-off frequency becomes
ferred spatial frequencies. fc = (1800cycles/mm)/(f-number). (25.31)
Let us summarize the crucial inputs into the deriva-
For example, an f/22 system has an image space
tion. We assumed a linear optical system, and that the
cut-off frequency of 81.8 cycles/mm, while an f/1.8
light is incoherent. We also assumed that the point
system has an image space cut-off frequency fc of
spread function does not change over the area that we
1,000 cycles/mm. Because of diffraction, the spread
are considering (called an isoplanatic patch). Finally,
function for the f/22 system is larger than that for the
we assumed that the object was a sine wave grating.
Given these assumptions, and the standard trig-
onometry behavior of cosines and sines, we showed
that the image is a sine wave grating and that the MTF
is related to the sum of cosines and sines times the line
spread function. (In the calculus formulation, the
sums are replaced by integrals, and we say that the
MTF is the Fourier transform of the spread function.)

25.6 MTFs of Various Systems

Consider a diffraction-limited system with a square


aperture of width w. In this case, integral calculus
equations can be used to derive an algebraic equation FIGURE 25.11. DifTraction-limited MTFs for a square and
for the MTF. The result is: circular aperture.
Spatial Distribution of Optical Information 545

f/1.8 system. Clearly, fc decreases when the spread limited behavior at a pupil size of about 1 mm, but at
function increases. 3 mm, the effective cut-off fn is about 200 cycles/mm
(corresponding to a 20/10 visual acuity). The effective
EXAMPLE 25.4 lower cut-off is due either to the aberrations and
Consider an f/4 incoherent system with a square scatter present in the eye, or to the size and spacing of
aperture. What is the MTF for an image space the retinal receptors (cones) in the foveal area.
frequency of 200 cycles /mm? When the blur circle is much larger than the dif-
The cut-off frequency is
fraction pattern, a closed form equation for the MTF
fc = (1800 cycles/mm)/4 = 450 cycles/mm. can be calculated. This equation is referred to as the
Then from Eq. 25.28, geometric M T F and is not accurate for small amounts
of blur. For a square aperture, the horizontal
MTF = 1 - (200/450) = 0.56.
geometric M T F is
We can also use integral calculus to derive an MTF = sin(t)/t, (25.33)
algebraic equation for the incoherent M T F of a dif- where
fraction-limited system with a circular aperture. This t = (27rD)(f/f n ), (25.34)
equation is:
fn is the cut-off for no defocus, and D is the number of
_1 2 1/2 quarter diopters of blur.
MTF= - {cos (f/f c ) [f/fc][l-(f/fc) ] },
77
(25.32) EXAMPLE 25.6
where the arc cosine is in radians and the image space Consider a system with a square aperture and a 200
cut-off frequency is again given by Eq. 25.29. In the cycle/mm cut-off for no defocus. (Except for the
circular case the M T F initially decreases a little faster square pupil, this corresponds to the human eye).
For 50 cycles/mm, what is the geometric MTF for a
than the straight line M T F for the square aperture.
1/4 D error, and a 1/2 D error.
Then the circular M T F gently curves and stretches out For the 1/4 D error,
to the same cut-off Figure 25.11. A degraded system
with a circular aperture has an M T F that is less than t = (2)(50/200) = /2.
that given by Eq. 25.30 at each spatial frequency. Then the geometric MTF is
MTF = sin(7r/2)/(ir/2) = 0.64.
For a the 1/2D error,
EXAMPLE 25.5
Consider an incoherent f/4 system with a circular t = (4)(50/200) = 7,
aperture. What is the MTF for an image space and the geometric MTF is
frequency of 200 cycles/mm? How does this com-
pare to the previous example? MTF = sin(7r)/7T = 0.
The cut-off frequency is the same as the previ- For the same system, what is the geometric MTF
ous example: fc = 450 cycles/mm. So f/fc = for a spatial frequency of 20 cycles/mm and a
200/450 = 0.44. Then 1.00 D defocus?
For the 1.00 D error,
MTF = - {cos _1 (0.44) - [0.44]
t = (8)(20/200) = 0.877.
x[l-(0.44)2]1/2}, Then the geometric MTF is
MTF = - {1.11 - [0.44][0.90]} = 0.45. MTF = sin(0.877)/(0.877) = 0.23.
77
The horizontal MTF for the square aperture was A geometric M T F can also be derived for a circular
0.56. aperture, in which case it has the form J ^ z J / z where
Jj is the first order Bessel function. Some calculated
When blur by misfocus is introduced into a system, geometric MTFs for a schematic eye with a 3 mm pupil
the M T F decreases from the diffraction-limited re- are given in Figure 25.12.
sults. As the blur is increased, the middle of the M T F We can also determine MTFs for prisms. The
dips until an effective cut-off is reached earlier than fc. aberrations of a thick prism introduce blur into the
This agrees with the castle scenario in Section 25.1. image so that the M T F for prisms decreases as the
Optically, a 60.00 D human eye with a 3 mm pupil prism diopter value increases. MTFs also decrease due
acts like a f/5.6 optical system. If the eye were to scatter, as in the case of dirty windshields.
diffraction-limited, the cut-off frequency would be An interesting situation occurs when there is a
about 320 cycles/mm. The eye shows diffraction- central obstruction in the system. A telescope with an
546 Geometrie, Physical, and Visual Optics

actually helps. The corresponding results for an annu-


MTF
lar aperture are shown in Figure 25.13c.
The higher frequencies are also helped by a process
called apodization. In apodization, a gradient tint is
applied around the aperture borders, so that transmit-
tance drops gradually to zero instead of abruptly. In

V/f
.20. .20
100
/* visual
200
r
JO.
f(cycles/mm)- the Fraunhofer diffraction pattern, this has the effect
of removing light from the rings and concentrating it
80 40 20 acuity
10 into the central Airy's disk. The result is a change in
ar the MTF resembling that of the central obstruction.
FIGURE 25.12. Geometric MTFs for a schematic eye with a The MTF drops for the lower spatial frequencies, but
3 mm pupil.
then picks up for the higher spatial frequency. (Apodi-
zation means without the feet, a figurative representa-
tion for Airy's disk without the rings.)
objective mirror has such an obstruction. Consider the
obstruction shown for the square aperture in Figure
25.13a. The corresponding MTF (solid lines in Figure
25.13b) drops to zero much quicker than that for a 25.7 Coupled Systems
clear aperture (dashed lines) and then rises to a value
greater than that for the clear aperture before it drops MTF theory can be applied to any linear system
back to zero again. If the low frequency components whether it is optical, electronic, or even film exposure
are the most important in the object, then the obstruc- and processing. Frequently, two linear systems are
tion degrades the image. If the higher frequency com- used in tandem. For example, for spatial frequency f a
ponents are the most important, then the obstruction camera lens has an MTF relating the object and image

k 1.0

MTF

f >-

MTF

FIGURE 25.13. a. Square aperture with central obstruction.


b. Resulting MTF. c. MTF for circular aperture with central
obstruction.
Spatial Distribution of Optical Information 547

modulation:
L(x)1
Mi = (MTF) 1 M 0 . H(f)l
,4A 2 x 0 2
Now if the film is operating in a linear range, the
process of exposing and developing the film also has
an MTF relating the image modulation to the modula-
/
-2x0
<s
2x0
tion of the developed film.
H
Mfllm = (MTF) 2 M i .
2x0
It is then easy to show that the MTF for the total FIGURE 25.14. A nonperiodic object, L(x), and its spatial
process is the product of the individual MTFs, i.e., frequency components H(f).
from the above two equations
Mfilm = [(MTF) 2 (MTF),]M 0 ,
there is a continuous range of frequencies in the
or higher order of components. These components can
(MTF) tot = (MTF) 2 (MTF),. be computed with calculus methods.
Figure 25.14 shows a nonperiodic object luminance
Suppose, for a given spatial frequency, that a camera distribution that has straight line sides that extend out
lens has an MTF of 0.85, and for that spatial fre- to a distance of 2x0 on either side of the central
quency the film has an MTF of 0.71. Then the total maximum. The maximum luminance value is A2. The
MTF for the developed picture is the product function H(f) represents the continuous range of spa-
(0.85)(0.71) = 0.60. tial frequency components in the object luminance
This can also be applied to the human visual system distribution. The value of H(f) represents the weight-
provided the neural system is functioning linearly. ing of the component of frequency f. In this case, the
Then the total MTF for the visual process is the analytic expression for H(f) is [Asin(277-x0f)/(7rf)]2.
product of the MTF for the optics of the eye and the Given an object, once we know the spatial fre-
MTF for the neural system. The MTF for the eye's quency components we can apply the MTF theory to
optical system is a low pass system, while the MTF for obtain the equation for the image luminance distribu-
the neural system is a high pass system. The product is tion. In that sense, the MTF analysis is theoretically
then peaked at a spatial frequency for which the independent of the exact object.
human vision system is most sensitive.
Note that the product does not apply to coherently
coupled systems. In other words, for two lenses, the
total MTF is not the product of the MTFs for each 25.9 A Nonlinear Example
individual lens. The reason is that the first lens may
have a decreased MTF due to an aberration (e.g., Consider a system with a lateral magnification of unit
positive spherical aberration), while the second lens magnitude. One of the conditions for linearity of this
has a decreased MTF due to the opposite aberration system was that a spatial frequency f appears in the
(e.g., negative spherical aberration). When used to- output if, and only if, the spatial frequency f appeared
gether the aberration of one lens may cancel the in the input. When a system is nonlinear the linear
aberration of the other lens, so that the doublet has a MTF theory does not apply.
better MTF than either lens by itself. Consider a tandem system consisting of the lenses
and the photodetectors (electronic or neural). The
photodetectors have a threshold. When the luminous
flux falls below that threshold, the photodetector does
25.8 General Objects not respond. Suppose the image is a single frequency
sinusoidal grating (Figure 25.15a), and the mean lumi-
Consider a general object such as the castle in Figure nance Lm of the image corresponds to the threshold.
25.1. The castle is not periodic like the grating. Never- Then the detection system responds only to the lumi-
theless, the luminance distribution of the castle scene nance levels above L m , and the detection system
can also be represented by a series of sine and cosine output (Figure 25.15b) consists of a periodic series of
luminance distributions. In the periodic case, the spa- separated bumps. Now the output is no longer de-
tial frequencies of the higher order sinusoidal compo- scribed by the sine wave of frequency f. According to
nents were integer multiples of the fundamental fre- Fourier's theorem, the periodic separated bumps can
quency. For a nonperiodic object, such as the castle, be represented by a sum of sine waves with spatial
548 Geometrie, Physical, and Visual Optics

the aperture size was decreased, but he found that for


a certain aperture size the image abruptly vanished
leaving only a uniformly illuminated background. This
led Abbe to develop some rather different ideas about
the image formation process.
For the compound microscope shown schematically
in Figure 25.16a, a point source S provides the light,
which is then collimated by a condensing lens. The
a)
x ^- collimated plane waves are incident on the microscope
slide. In conventional theory, diverging light leaves

a the slide and is converged by the objective lens to


form the magnified real image I inside the microscope.
Now let us consider Abbe's alternate formulation.
Suppose the object on the microscope slide is a dif-
fraction grating. When the collimated plane waves are
incident, the diffraction grating generates a series of
b) diffracted waves (the zero order along the axis, first
FIGURE 25.15. a. Sinusoidal object, b. Above threshold part order at an angle to the axis, second order at a larger
of the image. angle, etc.). Each diffracted wave is considered a
plane wave traveling at a different angle (Figure
25.16b). The objective lens of the microscope con-
frequencies that are integer multiples of the fun- verges each diffracted wave to form a point image
damental frequency f. Thus, the photodetector (actually an Airy's disk) in the secondary focal plane
threshold has resulted in the introduction of a number of the objective (marked by F 2 ). These point images
of spatial frequencies (2f, 3f, 4f, etc.) that were not are the diffraction grating maximums of Chapter 22.
present in the input. This is an example of a nonlinear The zero order maximum is labeled by 0, the first
response. order maximums by 1, and the second order max-
imums by 2. The light then diverges away from the
point images and the diverging waves overlap at the
internal image plane. All the light originated from the
25.10 The Coherent Transfer Function point source S, and for small aperture sizes the coher-
ence length of the overlapping light is such that inter-
The previous MTF discussion was for incoherent light. ference effects occur. The result is a set of interference
We could repeat the discussion for coherent light, in fringes in the internal image plane I. In fact, these
which case we have to deal with sums of amplitudes interference fringes are the image of the diffraction
rather than sums of intensities. In this case a coherent grating.
transfer function can be derived. One of the charac- Abbe realized that any small object diffracts light,
teristics of a coherent transfer function is that its so that the same theory can be applied, and the
cut-off frequency is one half that of the incoher- internal image is really an interference pattern. Abbe's
ent cut-off frequency fc. Therefore, when looking at theory explains why the image abruptly vanishes as the
the transfer function literature, pay attention to which objective lens aperture is decreased in size. Once the
cut-off frequency is being used. Sometimes the coher- aperture is small enough for the first and higher order
ent cut-off is called fc, in which case the incoherent diffracted waves to be blocked, there are not any
cut-off is at 2f . waves for the zero order wave to interfere with. In this
case there is no interference pattern and hence no
image.
Figure 25.17 shows another ray diagram for the
25.11 Abbe Theory of Image Formation formation of the internal image. We can now interpret
this ray diagram two different ways. One way is the
In the late 1800s Ernst Abbe was employed by the geometric optics method in which we follow the di-
optical firm of Carl Zeiss to improve the optical verging rays from each object point until each bundle
properties of microscopes. In an effort to control the is converged to a point image. The second way is to
aberrations, Abbe worked on a number of designs follow each set of parallel rays leaving the grating at a
with smaller and smaller apertures. Because of diffrac- particular angle to the axis, i.e., three parallel rays are
tion, Abbe expected the image to slowly get worse as traveling upward, three parallel to the axis, and three
Spatial Distribution of Optical Information 549

FIGURE 25.16. a. Internai image formation


by a microscope objective, b. Optics of diffrac-
tion grating as a microscope slide.

FIGURE 25.17. Rays for either the geomet-


ric or diffraction analysis.
550 Geometrie, Physical, and Visual Optics

downward. These parallel rays correspond to the 25.12 Spatial Filtering


plane-diffracted waves, and they converge to form the
series of point images (maximums) in the F 2 plane of Given the Abbe theory of image formation, we can
the objective. The rays from each maximum then manipulate the resulting image by altering what hap-
overlap at the internal image plane (where the inter- pens in the transform plane. Figure 25.18 shows a
ference forms the image). system to perform this alteration. A point source S
Let us consider what additional image quality ef- and condenser C collimate the incident light. (Laser
fects depend on the point images in the F 2 plane of the light is easiest to use.) The object slide is placed at
objective. When the aperture is large enough to pass of lens D. Lenses D and E are separated by the sum of
the first order waves together with the zero order their focal lengths. From a geometric optics viewpoint,
wave, interference occurs in the internal image plane. when an object is placed at F t of lens D, plane waves
However, the interference fringes formed in this case travel between the lenses, and lens E forms a conju-
have rounded edges. In effect, this "image" of the gate image I in its secondary focal plane.
square wave diffraction grating is the sine wave grating Now suppose the object slide is a diffraction grat-
of the same frequency (the fundamental harmonic). ing. Diffracted plane waves leave the slide at various
As the aperture size is increased so that the higher angles, and are converged by lens D to form a series
order diffracted waves are allowed through, the inter- of point images (maximums) at F 2 . The composite of
ference distribution builds into a square wave pattern these point images is the diffraction pattern or Fourier
just like a Fourier series, i.e., each higher order transform of the object. The light diverges from these
diffracted wave sharpens the fine detail. point images and is incident on lens E. Lens E con-
We can further check the Fourier series analogy by verts the light into overlapping plane waves. The
isolating each of the orders of diffracted waves (e.g., interference pattern formed at the secondary focal
the left and right first order diffracted waves, then the plane of lens E is the image I. Lens D took the
left and right second order diffracted waves, etc.). The diffraction grating object and formed the Fourier
interference fringes generated by each order are sine transform (the diffraction pattern) at F 2 , while lens E
wave gratings, which are the Fourier components of takes the transform and forms the image.
the square wave grating. The position in the F 2 plane When the diffraction grating consists of vertical
specifies the spatial frequency, and the intensity of the lines, the Fourier transform (or diffraction pattern) at
maximums specifies the magnitude of the sine wave F 2 consists of a row of horizontal point images (Figure
component at that spatial frequency. This can be 25.19a). When the diffraction grating consists of
verified mathematically by showing that the light dis- horizontal lines, the Fourier transform consists of a
tribution in the F 2 plane is the Fourier transform of row of vertical point images (Figure 25.19b). When
the square wave distribution of the original diffraction the diffraction grating consists of crossed lines, then
grating. each horizontal point image is the center for a vertical
In the MTF sections, we noted that an optical row of point images (or vice versa), and the result is a
system transferred information about each spatial fre- square array of point images (Figure 25.19c).
quency differently. Here we see that the information For the crossed grating, the interference pattern,
about each spatial frequency is in fact physically trans- which is the final image, also consists of crossed lines.
ferred through different positions in the F 2 or Fourier When a horizontal slit is placed in the Fourier trans-
transform plane. form plane (at F 2 in Figure 25.18), the interference

f object slide

FIGURE 25.18. System with an accessible transform


plane (F 2 ).
Spatial Distribution of Optical Information 551

an object slide is placed in the object plane, higher


order diffracted waves appear resulting in higher order
maximums in the F 2 or transform plane. These higher
a) order maximums are unaffected by the mask that
blocks the zero order wave. The light from the higher
order maximums passes through lens E to the image
plane. The resulting interference in the image plane
gives a phase reversed image in which high frequency
information (edges), which were originally very faint,
b) are now strongly evident. Figure 25.20 shows this for
the case of a circular aperture; a is the object grating,
while b is the filtered image grating. This principle is
the basis for dark field microscopy. The phase reversal
is understood by noting that dark areas in the un-
filtered image are due to destructive interference be-
tween the zero order wave and the higher order
c) . waves. When the zero order wave is blocked, this
destructive interference is removed.
FIGURE 25.19. Objects and diffration patterns in the trans- A similar result is achieved in phase contrast mi-
form plane, a. Vertical lines, b. Horizontal lines, c. Perpendicular croscopy. Suppose the object is essentially transparent
lines. but does have a variation in the refractive index. In
the geometric optics image, these areas are still trans-
parent and not visible. However, the phase variations
pattern in the image plane consists only of vertical do diffract a little light out of the zero order wave.
lines, i.e., the object is the crossed grating in Figure Suppose a plane that introduces a phase shift is placed
25.19c, but the horizontal slit passes only the max- in the path of the zero order wave. Then the zero
imums shown in Figure 25.19b, so the image is the order wave is phase-shifted relative to the diffracted
vertical grating shown in Figure 25.19b. In this case, waves. Because of the shift, some destructive and
we say that the image was spatially filtered. When a constructive interference between the diffracted waves
veritcal slit is placed in the transform plane, the and the zero order waves occurs in the image plane.
interference pattern in the image plane consists only of This interference makes the high spatial frequency
horizontal lines. phase information visible. The visibility is further en-
Suppose a square aperture is placed in the trans- hanced when the phase plate is partially absorbing.
form plane and then decreased in size. Each of the (Sometimes photochromic glass is used for the phase
point images in the image plane corresponds to one of plate.) Some soft contact lens surface deposits that are
the terms in the Fourier series (or transform) for the hard to see with regular microscopy are readily visible
image. So if a square aperture is placed in the trans- with the phase contrast technique.
form plane and decreased in size, the square wave
luminance profiles in the image plane lose high fre-
quency components and, hence, start to lose fine
detail. When the square aperture is small enough to
block all second and higher order diffracted waves, the
interference pattern in the image plane is just the
11 i
fundamental sine wave component of the object. In m m M
other words, the horizontal and vertical lines are now
blurred to the point where they have a single sine a)
IIP
wave distribution. When the square aperture is de-
creased further so that only the central maximum
remains, the image abruptly vanishes and we have
only uniform illumination in the image plane.
When no object is present in the object plane, lens
D takes the incident collimated light and forms a point
image in its F 2 plane. This point image is the zero
order maximum. If a mask is made that just blocks
this point image, the image plane is dark. Now when FIGURE 25.20. a. Object grating, b. Spatially filtered image.
552 Geometrie, Physical, and Visual Optics

When laser light (highly coherent) is used as the 25.19c, a diagonal slit is placed in the transform plane
illuminating light, spatial filtering can be performed (F 2 plane in Figure 25.18). Then a diagonal set of
with essentially any type of object slide. In this case, point images is transmitted through the slit. This
the light distribution in the F 2 plane of the objective is results in a perpendicularly diagonal set of interfer-
the Fourier transform of the object. By manipulating ence fringes in the image plane. In this case, we
what happens in the transform plane, we can affect the started with a target that has horizontal and vertical
image. For example, consider a magnified photograph lines, and ended up with a filtered image that has
of a television picture. The horizontal scan lines of the diagonal lines.
television are superimposed on the desired picture.
These scan lines are periodic and so they give specifi-
cally placed vertical maximums in the Fourier trans-
form plane. The rest of the light in the transform
plane is the Fourier transform of the image. A mask Problems
can be made to block only the periodic maximums
corresponding to the scan lines, while allowing all 1. An optical system with unit lateral magnification
other light from the transform plane to pass on. The has an object luminance distribution of
resulting interference pattern is the picture image
without the scan lines. L0 = 13 + 8sin(277-4x).
When a laser beam is diverged, it typically gives a When the MTF for 4 cycles/mm is 0.6, what is the
dirty illumination pattern because of diffraction from image luminance distribution?
each dust particle or finger print. One way to clean the 2. An optical system with unit lateral magnification
beam is to spatially filter it. First, the beam is con- has an object luminance distribution of
verged to a focus. The light diffracted by the dust,
L0 = 15 + 4 sin(277x) + 4 sin(27r3x) + 4 sin(27r5x).
fingerprints, etc. does not go to the point image but
lies at some lateral position in the image plane. If a The MTF for the system is given by 1 - 0.2 f. What
pinhole or small circular aperture is positioned to pass is the image luminance distribution?
only the point image, then the diverging light leaving 3. What is the image space cut-off frequency for
that image gives beautifully clean illumination. 633 nm light incident on an f/5.6 system? For
Spatial filtering techniques can be used to enhance 483 nm light incident?
or depress certain spatial frequencies in an image. In 4. A diffraction-limited system with a square aperture
fact, spatial filtering techniques have even been used has a cut-off frequency of 600 cycles/mm. What is
to color code different spatial frequency information. the MTF for a frequency of 400 cycles/mm?
In short, spatial filtering techniques give us tremend- 5. A diffraction-limited system with a circular aper-
ous power in recapturing hidden information. ture has a cut-off frequency of 600 cycles/mm.
It is possible to obtain artifacts in spatial filtering. What is the MTF for a frequency of 400 cycles/
Suppose for the crossed grating target of Figure mm?
APPENDIX A

Basic Matrix
Algebra

When you go to a grocery store to buy a dozen For a matrix A, we refer to the element in the ith row
oranges, you typically put the oranges in a bag to take and j t h column as a::.
home. The bag makes it easier to transport the
oranges. In a sense, a matrix is a bag of numbers. We
use the matrix (the bag) because it simplifies handling
the numbers and enables us to do operations that are Matrix Addition
difficult to do without matrices.
The individual numbers or elements of the matrix Matrices can be added together if, and only if, they
are identified by a row number and a column number. have the same shape, that is, an M x N matrix can be
Row numbers are written first. Consider a matrix with added to another M x N matrix. For example, a 3 x 2
three row and two columns, a so-called 3 x 2 matrix. matrix can be added to another 3 x 2 matrix. A 3 x 2
matrix cannot be added to a 2 x 2 matrix.
column column In matrix addition, the individual elements are
1 2 respectively added. The technical statement is:
2\ row 1 Let A, B, and C be M x N matrices. Then A + B = C if
and only if
A= row 2
aij+b^Cij,
\i a32/ row 3 for all values of i and j .
For example, when Example: 3 x 2 matrices

/
A= l -1
3
*\
0
/- 31
0 ; B
/4

1
3\

9
A=

\ 2 5/ \ 2 5/ \6 -6/
the individual elements are Then
a =
n 3, a12 = 1;

a21 = 1, a22 = u; C=A+B= 0 9

a^j z, ^32 ^ \8 - 1 /
553
554 Geometrie, Physical, and Visual Optics

In matrix addition, the mathematical properties of if and only if


commutation and association hold.
Cij=2kaikbkj.
A + B = B + A. (commutation)
(A + B) + C = A + (B + C). (association) Example:
3
/
A = -1 0 ; B=
Sealer Multiplication
\ 2 5/
When a single number (a sealer) multiplies a matrix,
then each element in the matrix is multiplied by that Here A is a 3 x 2 matrix, and B is a 2 x 1 matrix. Then
number. the product matrix C, where C = AB, is a 3 x 1 matrix.
Let q be a number (sealer) and A a matrix. Then The corresponding algebraic equations are:
C = qA, C
ll = a
i 1 t*ll + 2^>21 >

if and only if i.e., first row of A and first column of B,


c^qaij, C
21 = a
2lbn + a
22^21'

for all values of i and j . i.e., second row of A and first column of B,
Example: q = 0.5, and A is the matrix used previ- C = a
31 3irjll +
32^21'
ously. Then
i.e., third row of A and first column of B.
/ 3 1\ / 1.5 0.5\ The numerical results are:
C = qA = 0.5 - 1 0 -0.5 0 / (3)(7) + ( l ) ( - 4 ) \ / 17\
\ 2 5 / 2.5/ (-l)(7) + (0)(-4) = - 7

\ (2)(7) + (5)(-4)/ \ - 6 /

Matrix Multiplication Note that for the matrices A and B above, BA is not
defined. We cannot multiple a 2 x 1 matrix from the
Matrix multiplication has some algebraic properties right by a 3 x 2 matrix.
that differ from those of sealers (numbers). The learn-
ing of these special rules is the investment that enables
one to harness the power of matrix algebra. Example: 2 x 2 matrices:
Two matrices can be multiplied together only when
the number of columns of the left matrix equals the 1 -4
number of rows of the right matrix. The product
matrix has the number of rows of the left matrix and 7 3
the number of columns of the right matrix, i.e., an
M x N matrix can be multiplied from the right by an Then the product matrix H, where H = EF, is also a
N x P matrix, and the product matrix is M x P. As an 2 x 2 matrix. The numerical results are:
example, a 3 x 2 matrix can be multiplied from the
right by a 2 x 1 matrix, and the product matrix is a f(2)(-l) + (l)(7) (2)(-4) + (l)(3)N
3 x 1 matrix. A 3 x 2 matrix cannot be multiplied by H = EF =
another 3 x 2 matrix. v(3)(-l) + (5)(7) (3)(-4) + (5)(3)y
Consider an allowed matrix multiplication, C = AB.
The element c^ is obtained by taking the elements in
the il row of A and multiplying each by the cor- or
resonding element in the j t h row of B, and then adding
all the results together. This is algebraically stated as: 5 -5
H
C = AB, 32 3
Basic Matrix Algebra 555

Example: For the matrices E and F above, Let us explicitly verify this for the matrix F above:
/-l -4\/2 W -1 -4W1 0\
J = FE = FU
\ 7 3/\3 51 7 3/\0 \i
/ ( - l ) ( 2 ) + (-4)(3) (-l)(l) + (-4)(5^ The matrix multiplications give
J=
(7)(2) + (3)(3) (7)(l) + (3)(5) y /(-!)(!) +(-4)(0) (-l)(0) + (-4)(l)^
-14 -2 FU =
(7)(l) + (3)(0) (7)(0) + (3)(l)
JH
23 22
or
Note from the above two examples that H T ^ J . 1 -4\
There are some exceptions, but in general matrix FU = | = F.
multiplication does not commute, i.e., 7 3/
AB .
Thus, the order of matrix multiplication is crucial.
There is an associative law for matrix multipli-
cation: The Inverse Matrix
(AB)C = A(BC). Consider a square matrix A. If it exists, the inverse
There is also a distributive law: matrix A - 1 is defined such that
A(B + C) = AB + AC. AA _ 1 =A _ 1 A = U.
For the special case of 2 x 2 matrices, i.e.,

The Unit Matrix


Square matrices have the same number of rows and
columns (e.g., 2 x 2 , 3 x 3 ) . The diagonal elements of the inverse matrix is given by
a square matrix have the same row and column num-
ber (e.g., a n , a22, etc.). The off-diagonal elements -a,
have different row and column numbers (a^ where
i ^ j , as in a12, a 23 ).
Consider square matrices. The unit matrix U has
diagonal elements equal to 1 and off-diagonal ele- where det A stands for the determinant of A. For the
ments equal to 0. 2 x 2 matrix, the determinant equals the product of
The 3 x 3 unit matrix is the diagonal elements minus the product of the off-
diagonal elements.
/
1 0 0 oet A *^i ^ a22 aj 2 a 2 j.
\
0 1 0 Note that for the inverse matrix, the diagonal ele-
u= ments change place, the off-diagonal elements change
\0 0 1/ sign, and all elements are divided by the determinant.
The 2 x 2 unit matrix is
Example:
1 0
U 8 3
0 1 A=
\4 2
For a square matrix A, the unit matrix has the proper-
ty that Then
AU = UA = A. detA = (8)(2)-(4)(3) = +4,
556 Geometrie, Physical, and Visual Optics

2 -3 3. Find C = AB where
A"1 = (1/4)
,-4 8 -2 5
A= | ; B=
0.5 -0.75 3 - 1 / \ 7
A_1 =
-1 2 4. Find C = AB where

-(; -} -c i
You should be able to verify this by explicitly checking
that the product A A - 1 equals the unit matrix U.
If the determinant of a matrix is zero, then the
matrix does not have an inverse and is called a singu-
lar matrix. A matrix that has an inverse is called a 5. Use A and B from the previous problem to find
nonsingular matrix. D = BA. Check if D equals the product C of the
Suppose A and C are known matrices, and previous problem, or not.

AB = C. 6. Find the product E where


We can then find B by multiplying both sides from the
5 3\/l 0
left by A -1 (provided that the inverse exists): E= '
A AB = A *C, -4 2/\0 1
UB = A C,
7. Find the product S where
or
B = A *C. /I 0\/l -4
S='
0.2 l/\0 1

Vectors and Matrices 8. Label the following matrices as singular or non-


singular. Find the inverse matrix for those that are
Vectors can be represented either as column matrices nonsingular.
(i.e., M x 1), or as row matrices (i.e., 1 x M). Con- 3 -4
sider the vector G with a horizontal component -1-2
A=
and a vertical component - 3 . We could represent the -3 5
vector as the 2 x 1 matrix
1 1
+2
G = 1 1
2 0

0 2

Problems 9. Given the following equality, find the matrix B.


(Hint: Use the inverse matrix.)
1. Find C = A + B where

< -J
7 -6
B= 2
> - ( 7
5 4/ \-l
2. Find C = qA where 10. Given the following equality, find the matrix A.
(Hint: Use the inverse matrix and watch the
/3 2\ order.)
A= 1 -6 ; q=-0.4. 9 7\ 12 0

\4 9/ -3 -2/ \5 1
Basic Matrix Algebra 557

11. Find C = AB where Then show that H = PG, where P is a product


matrix that depends on R, R 1 , and A.
/ cosw sinw\ /cosw -sinw\ 13. The trace of a square matrix is defined as the sum
A = ; B = of the diagonal elements. For a 2 x 2 matrix,

(
\ - s i n w cosw/ \sinw cosw/ a a
il i2\
(Hint: sin2 w + cos2 = 1.) Does B have any special
relationship to A? a a
21 22/
12. Let G, G', H, and H' be 2 x 1 matrices (i.e.,
column vectors). Let R and A be 2 x 2 matrices trace A = a n + a22.
where Let C = A + B. Then prove that the traces also
G' = RG, and H' = RH while H' = AG'. add, i.e., trace C = (trace A) + (trace B).
Answers to
Selected Problems

Chapter 1 Chapter 5

(1) 1.67 (1) v = +25.0 cm, real, - 0 . 5 , U 5 f = - 2 . 2 2 D , V5b =


(3) 31.2, toward +5.00 D
(5) 5.32, away (3) v = -25.0 cm, virtual, +2.5, U5f = -20.00 D,
(9) long wavelengths V5b = -3.33D
(5) v = +10.7 cm, real, +0.36, U5f = +2.86 D, V5b =
+ 17.50 D
(7) v = -5.88 cm, virtual, +0.59, U5f = -20.00 D,
V5b = -9.19D
(9) v = -50.0 cm, virtual, -2.5 D, U 5 f = + 4 . 0 0 D ,
Chapter 2 V5b = -1.82D
(11) +7.14 cm, -12.5 cm
(1) -12.5 D, -2.38 D, --0.81 D, losing (13) -37.0 mm
(3) + 12.16 D, -4.81 D (15) 0, 0
(5) +9.61 D (17) virtual, +30.0 cm
(7) -12.09 D, -6.33 D (19) +60.98 cm, -15.24 cm
(9) 12.5 cm, 16.67 cm (21) +10.00 D, -1.25
(H) 5.0 cm (23) u = -18.18 cm
(13) stays clear and gets larger (25) +3.00 D
(27) 13.1mm

Chapter 6
Chapter 3
(1) far point real and 10.26 cm in front of cornea,
(1) virtual, optically real, left near point real and 5.97 cm in front of cornea,
(3) A, virtual image, converging myopia, -9.75 D, -11.42 D
(5) Lens 5, Lens 4 (3) far point virtual and 40.0 cm behind cornea, near
(7) Lens 3, Lens 2 point real and 16.67 cm in front of cornea, hy-
(9) E, D peropia, +2.50 D, +2.41 D
(11) image formed by the spectacle lens, virtual, (5) no, outside her far point
real (7) yes, 1.00 D

559
560 Geometrie, Physical, and Visual Optics

(9) 5.26 D, 6.97 D (5) -5.74 D, -3.09 D


(11) real, 303 cm to 58.8 cm, +3.67 D, infinity to (7) +3.00 D, +33.33 cm from front vertex
73 cm (9) +52.20 D, 9.39 mm
(13) 72.7 (11) +5.16D
(13) virtual and 11.97 cm behind cornea, +8.36 D
(15) +17.30 D
(17) -11.13D, +13.13D, +2.00D
Chapter 7

(la) +52 cm, -0.33, U 5 f = - 1 . 0 5 D , V5b =+3.32 Chapter 10


(le) +136.5 cm, -2.50, U5f=-3.33D, v5b =
+ 1.19D (la) + 16.67 cm, real, -0.33
(le) +14.00 cm, +0.64, U5f=+5.26D, v5b = (lc) +40.91 cm, real, -2.27
+ 17.33 D (le) + 10.00 cm, real, +0.20
(lg) +33.78 cm, -0.35, U5f=-1.08D, v5b = (2b) -6.00 cm, virtual, +0.40
+3.47 D (2d) -30.00 cm, virtual, -2.00
(li) -15.63 cm, + 1.63, U 5 f = - 1 5 . 6 0 D, v 5 b = (2f) -12.50 cm, virtual, -0.25
-4.85 D (3) 50 cm behind the mirror, m = + 1 , +1.00D
(2a) -25.33 cm, +0.33, U5f=-2.22D, v5b = (5) -6.25 D, -16.00 cm, -8.31 D, -16.00 cm
-5.01 D (7) 6.7%, 87.0%
(2c) +152.0 cm, +5.00, U5f=+4.00D, v5b = (9) 2.1%, -3.68mm, virtual, erect, +0.0184
+ 1.03 D (11) +1.65 D
(2e) -9.92 cm, +0.60, U5f = -7.60 D, V5b = -6.70 D (13) +64.64 D, 7.58 mm
(2g) -160.7cm, -5.43, U5f=+3.04D, V5b =
-0.60 D
(3a) apparent depth is 5.58 cm
(3b) apparent depth is 22.78 cm Chapter 11
(5) -20.45 cm, 10.91mm
(7) -1.99, +8.00 D
(9) +13.79 D (1) +9.09 D, +12.50 D, +20.00 D, +15.08 cm, -0.92
(11) 13.8 mm (3) +82.71 cm, -2.22, +6.60 cm from the principal
planes
(13) +5.00 D, +38.60 D
(15) +8.04 D (5) + 10.06 D
(7) +8.85 D, +10.91 D, +6.90 D, the primary princi-
pal plane is 3.20 cm behind the front vertex, the
secondary principal plane is 7.10 cm in front of the
back vertex, the nodal points are +4.79 cm from
Chapter 8 the principal planes
(9) + 11.22 D, +11.05 D, +12.11 D, the primary prin-
(1) 24.6 cm, 188 cm cipal plane is 1.40 mm in front of the front vertex,
(3) -3.39 D, 10 cm 6.55 mm in front of the back vertex
(5) in air 4.14 cm behind the plastic-air interface
(7) reduced system is a +10.00 D lens 4.48 cm in
front of a -4.00 D lens, back vertex power is
+ 14.12 D, neutralizing power is +6.61 D Chapter 12
(9) -50.00 D, +4.00 D
1.45 -2.03 D^
(H) -4.00 D
(13) +4.50 D (1)
VO.lOm 0.55
+2.03 D, +3.68 D, +1.40 D
/0.82 +3.56D\
Chapter 9
(3)
(1) +3.22 D \0.06m 1.48
(3) +5.36 D -3.56 D, -2.41 D, -4.34 D
Answers to Selected Problems 561

(5) -6.69 D, +9.11 D (4e) vertical blur ellipse, CMA, WTR, -2.00
-2.00 x 180
/0.72 -4.40 D^ (6) +0.25 0 - 0 . 5 0 x 1 8 0
(7) (7) - 3 . 0 0 0 - 2 . 0 0 x 6 0
V0.26m -0.20 (9) - 2 . 5 0 0 - 2 . 0 0 x 6 0
+4.40 D, -22.00 D, +6.11 D (11) +2.75 D
(13) +6.00 0 - 1 . 5 0 x 9 0
/-2.300 -2.900 D^ (15) +7.92 0 - 4 . 3 3 x 5 6
(8) (17) vertical 17.6% gain, horizontal 9.9% gain, verti-
, 0.440 m 0.120 cal rectangle
'-1.397 +3.726 D (19) 3.1% gain (135 meridian), 4.3% loss (45 meri-
(12) dian), parallelogram
, 0.231m -1.332

Chapter 13
Chapter 16
(1) 2.5 X, relative distance magnification
(3) 7.0 X (1) +3.68 D, 1.88 D
(5) No, Yes /-3.00 0.00
(7) 6.4 X (3a)
(9) -12.0 X, 54.17 cm \ 0.00 +1.00
(11) 107.14 cm
(13) 4.6 X -2.00 +1.73
(15) 0.43 X, 258.21cm (3c)
+ 1.73 0.00
-1.00 +2.00
(3e)
Chapter 14 +2.00 -1.00

(1) virtual image of the aperture (located 13.33 cm (5a) +1.00O-4.00X 160
behind the lens) (5c) - 3 . 5 0 O - 2 . 0 0 x 180
(3) entrance pupil 7.41 cm behind the +6.50 D lens, (7) -1.04 0 - 7 . 6 8 x 3 3 . 3
exit pupil 4.26 cm in front of the -3.50 D lens (9) -5.75 0 - 2 . 3 7 X 13.1
(5) located 3.15 mm behind the cornea and 3.42 mm (11) +4.63 0 - 2 . 2 6 X 137.3
in size (13) 40 and 130
(7) 9.8% loss (15) 2.00 D
(9) 1.010 (or 1% gain), 0.939 (or 6.1% loss), 0.948 (17) -7.85 D
(or 5.2% loss)
(11) Many lenses will work. One is a CR-39 (n =
1.495) with a +9.87 D front surface and a
5.0 mm central thickness
(13) 1.049 (or 4.9% gain), 0.907 (or 9.3% loss) Chapter 17

(1) 52.9, 32.4 28.1, 28.1, 27.2, total internai


reflection
Chapter 15 (3) 23.2, 19.2
(5) 2.9, 5.0
(1) +50.0 cm, horizontal (7) 4.0
(2) +40.0 cm (horizontal), +18.18 cm (vertical), (9) 1.5 BU
+25.0 cm (11) 5.7 BD & O @ 34.2
(4a) vertical line, SMA, WTR, -2.00 x 180 (13) 3.2 , 5.1 BU, left
(4c) horizontal blur ellipse, CHA, WTR, +4.00O (15) 2.2 BD & O @ 141.3
-2.00 x 180 (17) 9.4
562 Geometrie, Physical, and Visual Optics

Chapter 18 Chapter 22

(1) 5.6 BU (1) 0.99, 1.18


(3) 4.2 BD & O @ 30 (3) 1.34 cm
(5) 6.7 mm up (5) 20/17, 20/100
(7) 1.2 BU
(9a) yes, 3.75 mm out
(9b) no
(9c) no Chapter 23
(11) 9.2
(15) 3.3 , 1.1 BD (1) 2.1
(17) 5.7 mm out, 12.1mm up (3) 50%
(5) 12.5%
(7) 33.3

Chapter 19
Chapter 24
(1) 14.3
(3) +5 D lens, virtual image of the - 2 D formed by (1) 604 nm, 2,900 nm
the +5 D lens, 25 cm, 31 cm (3) 275 nm, 500 nm, 620 nm
(5) Entrance pupil is the virtual image of the 4 mm
aperture formed by the telescope, entrance port
is the +3.00 D objective lens, 0.68, no internal
image. Chapter 25
(7) 18.1mm
(9) f/2.1 (1) 13 + 4.8sin(27r4x)
(11) 0.11 D (3) 282 cycles/mm, 370 cycles/mm
(5) 0.22

Appendix A
Chapter 20 '10 -2
(1)
(1) +0.12 D, +0.23 D 0 1
(3) myopic for blue (F), hyperopic for red (C) 41
(5) polycarbonate +6.43 D, CR-39 -12.43 D
(3)
(7) +7.410-0.31x90 -16
(9) -8.08 O -4.41 x 180
74 - 2 5
(4)
29 -46
37 69
(5)
Chapter 21 34 -9
(1) 600 nm, same frequency, 400 nm 1 -4
(3) 542.4 nm, green (7)
(5) 17.7 mm, 3.54 mm 0.2 0.2
(7) decreases -4.43
(9) 108 nm (9)
(11) 1.42 cm +5.29
Index

Page numbers in italics indicate figures; page numbers followed by t indicate tabular material.

. See Angstrom Absorptance, 519 dioptric power and off-axis


Abbe, Ernst, 438, 548 Absorption, 6 meridians and, 345-347,
Abbe sine condition, 438, 439, 439 reflection and, 6 346, 347
Aberrations, 14, 423-449 selective, 6 Aerial image, 31-32
achromatic doublet and, 427-428, scattering and, 503 Afocal microscopes, 261, 261-263, 262
428 strength of, 6, 7, 504, 504 Afocal systems, 224
achromatic prisms and, 428, Absorption axis, of polaroid, 508 magnification equality for, 411-412
428-429 Absorption filter, 473 Afocal telescopes, 261, 261-263, 262
aplanatic systems and, 438, Accommodation, 4, 88-89, 89 alternate equations for, 263-264
438-439, 439 aging and, 26 Against motion, 387
chromatic, 423 ametrope and, 98 Against rotation, 386
dispersion and, 423-424, 424 amplitude of, 90 Against-the-rule (ATR) astigmatism,
longitudinal and lateral, through contact lenses, 98-99, 99 315, 317, 317
424-425, 425 emmtrope and, 26, 26, 90-91 Aging, accommodation and, 26
monochromatic, 423, 431-432, hyperope and, 27, 91-92, 92 Air
432 maximum, magnifiers and, 255, equidistance for thin lenses in, 46,
positive and negative, 435, 435 255-256 47
thick lenses and, 429, 429-430 through spectacles, 94, 94-95, 95 equivalent thickness and, 140
coma as, 434, 436, 436-438, 437 algebraic approach to, 94, Gauss systems in, 211, 211
curvature of field and, 444, 97-98 glass interface with, 128, 129
444-446, 445 spherocylindrical lenses and, 321, imaging in, reflection and, 172,
detriments to image quality and, 321-322 174, 177, 177-179
423 Accommodative demand, 89 index of refraction of, 10
diffraction and, 423 spectacle, 93-94 reduced systems and. See Reduced
distortion and, 446, 446, 447 Achromatic doublet, 427-428, 428 systems
lens design and, 447-448, 448 Achromatic lens, 428, 429-430 Airy, George, 479
mirrors and, 449 Achromatic prisms, 428, 428-429 Airy's disk, 479-481, 486
radial astigmatism and, 437, Actual angular magnification, 265 Alhazen, 9, 25
439-444, 440-443 Acuity Alpha decay, 360
refractive efficiency and, 425t, diffraction and, 488-489 Ametrope
425-427, 426, 426t, 427 laser testing of visual pathwavs for, axial, 288
Seidel or third order, 432-433 464 corrected
separated lenses of same material Snellen, 488-489 by contact lenses,
and, 430-431, 431 Adaptation, in problem solving, 1 accommodation and, 98
spherical, 432-436, 433-436 Add, 101, 102 retinal image sizes of, 102,
positive and negative, 435, 435 Addition 103, 103-104
spherocylindrical lenses and, matrix, 553-554 spectacle accommodative
448-449 vector, 370-371 demand and, 93-94

563
564 Index

Ametrope (continued) terminology for, 265-266 oblique, 317


mixed, 288 Aniseikonia, 283, 292 perpendicularly crossed cylindrical
range of clear vision for, 101 Anisometropia, 292 lenses and, 300-303, 301-303
refractive, 288 Anisotropie materials, 511 radial, 437, 439-444, 440-443
spectacle magnification and, 281, light propagation through, 14 spectacle magnification and
281, 288, 290, 291 Anomalous dispersion, 505 power factor and, 325
spherical, 27 Anterior chamber angle, viewing, 359 shape factor and, 325-326
system matrices and, 240-241 Antimetropia, 292 spherocylindrical lenses and
telescopes for, 287-288 Anti-node, 459 accommodation and, 321,
Ametropia, spherical, 27 Antireflecting coatings, 448 321-322
Amici prism, 376, 376 Antireflection films, 471, 471-472, 472 effectivity and, 320-321
Amplitude Aperture standard axis notation and, 299,
of accommodation, 90 changing from square to circular, 299-300, 300
of harmonic wave, 453 diffraction and, 481, 481-484, toric surfaces and, 306-308, 307
Analyzer, 508 484-485 transposition and equivalent
Anastigmat projector system, 445, 446 circular, diffraction by, 479-481, combinations and, 304-306
Angle(s) 480 virtual line images and, 312-313,
Brewster's 510 electron, 488 313
critical, 356, 358, 358 numerical, 488 with-the-rule, 315, 317, 317
prisms and, 356 square, diffraction by, 479, 480 Astronomical telescope, 261
deviation, 360-363, 360-364, 379 Aperture stops Atom, Bohr model of, 522-524, 523,
incident, 360-363, 360-364 for Galilean telescope, 412 524, 524t
reduced, 750, 150-151 multilens systems and, 414-415, ATR. See Against-the-rule astigmatism
Angle of incidence, 6 415 Axial ametrope, 288
Angle of reflection, 6 spectacle magnification and, Axis
Angle of refraction, 9 267-268, 268-272, 270-273 absorption, of polaroid, 508
Angstrom (), 455 Apex, of prisms, 355 of cylinder, 295
Angular magnification, 247-266 Apex angle, of prisms, 355 optical, 44, 44, 381
actual, 265 Aphakia, 27 Axis meridian, 295-299, 299
afocal telescopes and, 261, spectacle lens correction for, Axis notation, standard, 299, 299-300,
161-163, 161 87-88, 88 300
alternate equations for, Aplanatic systems, 438, 438-439, 439 Axis refinement, with Jackson cross
263-264 Apodization, 546 cylinder, 348-349, 349-350, 351
apparent, 265 Apparent angular magnification, 265
Badal principle and, 256, 256 Aqueous humor, 4
collimating magnifier and, with flow of, 359 Back focal length, 208
object to eye distance Arago, Dominic, 486 Back vertex power, 148, 155
unchanged, 252, 254, 254-255 Argon ion lasers, 527, 528 plane refracting interfaces and,
compound microscope and, 252, Aristotle, 25 147-149, 147-149
257-259, 258 Aspheric surfaces, 435, 435 Badal's principle, 256, 256
conventional reference distance Asthenopic symptoms, 396 Balmer series, 523
and, 253 Astigmat, equal mixed, 315, 347 Band spectra, 524
converting telescope to microscope Astigmatic bundle, 302 Barrel distortion, 446
and, 264-265 Astigmatic system, 295 Base-apex meridian, 367
equivalent power formulation for, Astigmatism, 27, 295-326 Base down (BD) prism, 367-368
260-261 circle of least confusion and Base in (BI) prism, 355, 357, 367-368
general equation for simple spherical equivalent and, Base out (BO) prism, 367-368
magnifier and, 257 310-312, 311, 312 Base up (BU) prism, 367-368
magnifiers and maximum classification and correction of, Basov, Nikolai, 526
accommodation and, 255, 315, 316-319, 317-319 BD. See Base down prism
255-256 clock dial chart and, 319, 319-320, Beam, 21-23, 22
without physical size change, 248, 320 boundaries of, 22
248-249, 249 collapsing conoid by equalizing from common flashlight, 22
plus lens and cross cylinders and, 303-304, diverging, 22, 22
as collimating magnifier, 251t, 304 lateral magnification and, 76, 76
251-253, 353 cylindrical lenses and, 295-299, from slide projector, 22
as simple magnifier, 250-251 296-298, 302 Beat phenomenon, 458, 506
relative distance magnification and, extended objects and multiple Benham's top, 6
248, 249-250 conoids and, 308-310, 308-310 BI. See Base in prism
relative size magnification and, meridional telescopes and, Biconcave lens, 153, 153
247, 247-248 322-325, 323, 324 Biconvex lens, 153, 153
Index 565

Bifocal lenses, range of clear vision CHA. See Compound hyperopic collapsing by equalizing cross
and, 101-102 astigmat cylinders, 303-304, 304
Binocular vision, 355 Chevalier's lens, 447 multiple, extended objects and,
polarized light and, 516 Chief ray, 273 308-310, 308-310
Birefringence, 511, 511-512, 512, 512t Chromatic aberrations. See toric surfaces and, 306
ocular, Haidiger's brush and, 516 Aberrations, chromatic twisting sheets of light and,
Blackbody, perfect, 519 Chromatic difference in magnification, dioptric power and off-axis
Blackbody radiation, 519-520, 520 425 meridians and, 329-330, 330
Blur, 23 Ciliary corona, 493 Constructive interference, 457
minimization of, image formation Ciliary muscle, 4 Contact lenses
by, 23-26, 24, 25 Circle of least confusion, 301, 348-349 accommodation through, 98-99, 99
Blur circle, 14, 15, 302 spherical equivalent and, 310-312, corrected near point and, 99-100
size of, 23, 23 311, 312 hard, corrections and, 159,
Blur ellipse, 298 Circular polarization, 513, 513, 514 159-160
Blurred image, size of, relative materials and, 515, 515-516 tear, 160, 160-161, 161
magnification and, 272, 273-275, Clear vision, range of. See Range of Context, in problem solving, 1
273-275 clear vision Continuous spectra, 524
BO. See Base out prism Clock dial chart, astigmatism and, 319, Contra-ocular view, 299
Bohr model, of atom, 522-524, 523, 319-320, 320 Contrast, 535-536
524, 524t Clouds, scattering in, 8 Conventional reference distance, 253
Boundary(ies) incoherent, 500 Convergence insufficiency, 355
of beam, 22 CM A. See Compound myopic astigmat Converging bundles, 22, 22-23
focal points and, 67 Coherence, 459-460 Converging lenses. See Lens(es),
Brewster, David, 510 degree of, 463 converging
Brewster's angle, 510 partial, fringe visibility and, 459, Converging light, 23
Brewster's law, 510 462-464, 463, 464 reduced systems and, 141,
BU. See Base up prism Coherence length 141-142, 142
Bundles lateral, 464 Converging mirrors, 167, 168
astigmatic, 302 longitudinal, 463 Converging wavefronts. See
converging, 22, 22-23 Coherent transfer function, 548 Wavefront(s), converging
homocentric, 14 Collagen fibers, coherent scattering Convex mirror, 167
lateral magnification and, 76, 76 and, 503 Convex surfaces, 107, 108
monocentric, 14 Collimated light, 312 convergence and divergence and,
Collimating magnifier 167
object to eye distance unchanged Convolution, 310
and, 252, 254, 254-255 Cooke triplet lens, 448
Calcite crystal, birefringence and, 511, plus lens as, 25It, 251-253, 252 Cornea, 4
511-512, 512, 512t Color coherent scattering and, 502-503
Camera of iris, 503 as source of astigmatism, 315
depth of field and, 417, 417-418 wavelength and, 5t, 5-6 swelling of, diffraction halos and,
object distance effect on, Color perception, 4 489-490
418-420, 420 psychophysiological dependence of, Cornea-to-retina matrix, 240
depth of focus and, 420, 420-421 6 Coronas, 489
film exposure and f-number and, Coma, 434, 436, 436-438, 437 ciliary, 493
415-417 positive and negative, 436, Corrected curve spectacle lenses, 446
Camera obscura, 25 436-438, 437 Coupled systems, 546-547
Carbon dioxide gas lasers, 527, 528 Compound hyperopic astigmat (CHA), Crepuscular rays, 10
Cardinal points, reversibility and, in 315 Critical angle, 356, 358, 358
Gauss system, 216, 216-218, 217 Compound microscopes, 252, 257-259, Crystal
Cartesian ovals, 435 258 anisotropic, light propagation
Cataracts, 27, 423 Compound myopic astigmat (CMA), through, 14
coherent scattering and, 503 315 calcite, birefringence and, 511,
Catoptric power, 176 Concave mirror, 167 511-512, 512, 512t
Cavity oscillator, 527 Concave surfaces, 107, 108 dichroic, 512
Center, optical, 44, 45, 379-383, Condenser(s), 413-414, 414 Crystalline lens, 4
380-383 Condenser lens, 413 coherent scattering and, 503
Centered system, 229 Cones, 4, 418 lacking, 27
Center of rotation, 372-374 Conjugate planes, 204-206 spectacle lens correction for,
Centrads, 371, 371 Conjugate points, 37 87-88, 88
Central refraction, oblique, 439 optical axis and, 44 as source of astigmatism, 315
Central thickness, 279-280 Conoid of Sturm, 302, 304, 305 Cues, 4
566 Index

Curvature, 15 apertures and system matrix and, 336-337


change in media and, 130-131 changing from square to translation matrix and, 333,
dioptric power and off-axis circular aperture and, 481, 333-334
meridians and, 329 481-484, 484-485 vector addition method and,
dipole radiation and, 329, 330 circular, 479-481, 480 345-347, 346, 347
of field, 444, 444-446, 445 square, 479, 480 vergence matrix and, 351-352
of meridian, lens clock to measure, far field, 485 for single spherical refracting
331-333 Fraunhofer versus Fresnel, 485, interfaces, 112, 112-113, 113
of mirror, center of, 171 485-486, 486 Dioptric power curvature component,
Curves, iso-prism, 381, 381 halos and, 489-490 330-333, 331, 332
Cylinder, 295, 296. See also Jackson holography and, 494, 494-496, 495 Dioptric power matrix, dioptric power
cross cylinder Huygens-Fresnel principle and, and off-axis meridians and, 337-340,
axis of, 295 475-476, 476 338, 339
Fresnel, 388 limit on resolution and, 486-488, finding spherocylindrical
Cylindrical lenses. See Jackson cross 487 parameters and, 341, 341-342
cylinder; Lens(es), cylindrical; in nature, 493 trace and determinant of matrix
Lens(es), spherocylindrical near field, 485 and, 340
Cylindrical wavefront, 295 pinhole effect and, 24, 25 Dioptric power torsional component,
single-slit, 476-479, 477, 478 330-333, 331, 332
single photons and, 530-531 Dipole oscillation, 499, 500
zone plates and, 493, 493-494, 494 Dipole radiation, 499, 500
Dallmeyer's rapid rectilinear lens, 447 Diffraction gratings, 490-492, 490-492 Dirac, Paul, 531
Damping factor, 504, 504 observations and, 492-493 Dispersion, 505, 505-506, 506
Dark field microscopy, 551 spatial filtering and, 550-551 anomalous, 505
da Vinci, Leonardo, 25 Diffraction limited systems, 486-488,457 chromatic aberration and, 423-424,
Decay, alpha, 360 modulation transfer function and, 424
Decentration (DEC) 544-545 normal, 505
amount of, 382-383 Diffuse reflection, 7, 7 prisms and, 355
with spherocylindrical lenses, Diopters, 16, 299 Displacement law, of Wien, 520
398-399, 399 of prisms, 368, 368-369, 369 Distance(s)
Depth of field, 417, 417-418 Dioptric depth of field, 417-418 effect on depth of field, 418-420,
dioptric, 417-418 Dioptric power, 57 420
linear, 418 equivalent, 207 hyperfocal, 419
object distance effect on, 418-420, focal length relationships and, object and image, 57-58, 59
420 59-60, 60 reduced, 140
Depth of focus, 420, 420-421 neutralizing, 148, 155 Distance of most distinct vision, 253
Depth perception, 3 off-axis meridians and, 329-352 Distortion, 446, 446, 447
Desaturated green response, 5-6 adding obliquely crossed barrel, 446
Descartes, Ren, 10 spherocylindrical lenses and, pincushion, 446
Descartes' Law, 10 342, 342-344, 343 Diverging beam, 22, 22
Destructive interference, 457, 458 axis refinement with Jackson Diverging lenses. See Lens(es),
Determinant, 340 cross cylinder and, 348-349, diverging
Determinant condition, system matrices 349-350, 351 Diverging mirrors, 167, 168
and, 234-235 curvature and torsion and, Diverging wavefronts. See
Deviations 329, 330 Wavefront(s), diverging
maximum, 360-363, 360-364 dioptric power components Double-slit experiment
minimum, prisms and, 363-365, and, 330-333, 331, 332 interference and single photons
364-365 dioptric power matrix and, and, 531
at normal incidence, prisms and, 337-340, 338, 339 of Young, 461, 461-462, 462
363, 365-366, 366 trace and determinant of, Doublet lens, 436, 436
rotational 340 Double vision, 396
by cylindrical lenses, 385-386, finding spherocylindrical Dove prism, 376, 376
386 parameters from matrix Downstream vergence, 19-21, 19-21,
spectacle magnification and, and, 341, 341-342 95-96, 96
386-387, 387 over-refraction and, 344-345
by thick prism, 360-363, 360-364 refraction matrix and, 334, Eccentric fixation, 516
Deviation angle, 360-363, 360-364, 379 334-336 Effectivity
Diamond "fire," 424 residual refractive error and, astimatism and spherocylindrical
Dichroic crystal, 512 347-348, 348 lenses and, 320-321
Diffraction, 5, 14, 475-496 Sturm's conoid and twisting prism, 372-374, 373
aberrations and, 423 sheets of light and, with lenses, 88, 389, 389-392,
acuity and, 488-489 329-330, 330 391
Index 567

vergence, 88, 96 Extended source, 21 Flat slabs, 139-141, 139-141


Einstein, Albert, 521-522, 526 point, 21-23, 22 ray tracing through, in equi-index
Electric field, 451 Extended target, 308-310 media, 138-139, 139
Electromagnetic equations, of Maxwell, Extra-focal length, primary and Floaters, 486
168 secondary, 162 Fluorescence, 524-525, 525, 526
Electromagnetic force, 451 Extra-ocular muscles, 355 Fluorescent lights, 525
Electromagnetic radiation, 4-5 Extraordinary ray, 512 Fluorexon, 525
speed and wavelength of, 5 Eye, 4, 4. See also Cornea; Retina; f-number, film exposure and, 415-417
Electromagnetic spectrum, 4 Vision Focal length. See also Extra-focal
Electron aperture, 488 ametropic, system matrices and, length
Electron volt, 522 240-241 back, 208
Ellipse(s), Tscherning, 443 Gullstrand's 1 schematic, 239, dioptric power and, 59-60, 60
Elliptical polarization, 513 241-242, 242t equivalent
EM A. See Equal mixed astigmat Gullstrand's 2 schematic, 239, primary, 209
Emission 239-240 secondary, 208
spontaneous, 526 Gullstrand's reduced, 239 front, 209
stimulated, 526, 526 of nautilus, 25 power and, in Gauss system,
Emmtrope, 26, 26 of rattlesnake, 25 208-210, 209, 210
accommodation and, 26, 26, 90-91 real images as objects for, 31-32, secondary, for single spherical
astigmatism and, 313-314 32 refracting interfaces, 115
range of clear vision for, 92, 93 reduced, system matrices and, 239, Focal plane, secondary, front vertex
reflection and, 182-183, 183 242 matrix to, 238, 238-240, 239
speckles and, 466 refracting states of, 26, 26-27, 27 Focal points. See also Primary focal
spectacle magnification and, 288 Eye lens, 430 point; Secondary focal point
thin lens eye model of, 81-82, 82 Eye models, 157 boundary role of, 67
unaccommodated, 26 thin lens. See Thin lens eye models equidistance for thin lens in air,
Emmetropic meridian, 314 Eyepiece, 257 47, 47
Endoscope, 359-360, 360 Huygens, 430, 431 of mirror, 169, 170, 171
Endothelial trauma, 502-503 Eye relief, 409 real, 203, 203-204
Entrance port, 401 reflection and, 169, 169-171, 170
Entrance pupil for single spherical refracting
relative magnification and, 213-21A interfaces, 113-117, 114-117
spectacle magnification and, Fj. See Primary focal point virtual, 204, 204
267-268, 268-272, 270-273 F 2 . See Secondary focal point Focus, depth of, 420, 420-421
Equal mixed astigmat (EMA), 315, 347 Fabry-Perot interferometer, 473 Fog, scattering in, 8
Equiconcave lens, 153, 153 Face-form tilt, 441 Fogging, 319
Equiconvex lens, 153, 153 Faraday, Michael, 451 improper corrections and, 101
Equivalent, spherical, 343 Far field diffraction, 485 of windows, diffraction and, 489
Equivalent air systems. See Reduced Far point, for emmtrope, 82 Fourier,
systems Farsightedness. See Hyperope theorem, 539
Equivalent air thickness, 140 Fata morgana mirage, 10 transform, 543
Equivalent dioptric power, 207 Feathers, as diffraction gratings, 492 Fovea, 4, 355
Equivalent focal length Field Fraunhofer, Joseph von, 485, 492
primary, 209 curvature of, 444, 444-446, 445 Fraunhofer designations, 425, 425t
secondary, 208 depth of, 417, 417-418 Fraunhofer diffraction, Fresnel
Equivalent mirrors, 187-192, 188, object distance effect on, diffraction versus, 485, 485-486, 486
190-192 418-420, 420 Frequency, 454
Equivalent power, 221 electric, 451 resonance, 503-504, 504
for magnification, 260-261 magnetic, 451 spatial, 535, 536
Euler's law, 332 Field lens, 409, 430 Frequency invariance, wavelength
Evidence, in problem solving, 1 Field of at least one half illumination, changes and, 455t, 455-456, 456
Ewald-Oseen theorem, 502 404 Fresnel, Augustin, 475-476
Excimer lasers, 527-528 Field of uniform or maximum Fresnel biprism, 464-465, 465
Exit port, 401 illumination, 401, 404 Fresnel cylinder, 388
Exit pupil Field of view, 401, 402-408, 404-408 Fresnel diffraction, Fraunhofer
for Keplerian telescope, 412 total, 404 diffraction versus, 485, 485-486, 486
relative magnification and, 272, Field stop, 401 Fresnel lenses, prism properties of,
273-275, 273-275 Film exposure, f-number and, 415-417 388, 388
Exploded system, 158 Filter(s), interference, 472-473, 473 Fresnel prisms, 374-375, 375
Extended images, 23, 308-310 Filtering, spatial, 550, 550-552, 551 Fresnel's law, 167
Extended object, multiple conoids and, Fish scales, as diffraction grating, 493 for paraxial light, 167-168
308-310, 308-310 Fixation, eccentric, 516 Fresnel zone plates, 493, 493
568 Index

Fringe visibility, partial coherence and, Grey body, 520 spectacle magnification and, 279,
459, 462-464, 463, 464 GRIN, 13 288, 291
Front focal length, 209 Group velocity, phase velocity versus, thin lens eye model of, 84, 84-85,
Front vertex, matrix to secondary focal 506 85
plane, 238, 238-240, 239 Gullstrand, Allvar, 162, 241 Hyperopic meridian, 315
Front vertex power, 148, 155 Gullstrand's 1 schematic eye, 239, Hz. See Hertz
plane reflecting interfaces and, 241-242, 242t
147-149, 147-149 Gullstrand's 2 schematic eye, 239,
f-stop, 416 239-240 Ice crystal halos, 489
Full wave plate, 514 Gullstrand's reduced eye, 239 Illuminance, 455
Illumination, uniform or maximum,
field of, 401, 404
H-sheet polaroid, 508 Illumination pattern, 298
Gabor, Dennis, 494 Haidiger's brush, ocular birefringence Illusions
Gabor zone plate, 493-494, 494 and, 516 hole-in-the-hand, 3
Galilean telescope. See Telescopes, Half wave plate, 514, 514 moon, 4
Galilean Halos Image(s), 35
Gauss, Carl, 162 diffraction, 489-490 aerial, 31-32
Gauss equation 46, 365 blurred, size of, relative
derivation of, 244 pathological, 493 magnification and, 272, 273-275,
for spectacle magnification, physiologic, 492-493 273-275
284-287, 285, 286 22, 365 erect and inverted relationships
Gauss system, 195-226 Hand neutralization, 387 and, 71
in air, 211, 211 Harmonic, fundamental, 539 extended, 23, 308-310
cardinal points and reversibility Harmonic waves, 452-454, 453-454 formation of, Abbe theory of, 548,
and, 216, 216-218, 217 Heisenberg uncertainty principle, 530 549, 550
conjugacy of principal planes and, Helium-neon laser, continuous beam, ghost, 186, 186
204-206, 205, 206 527, 528 line, 295
Newton's equation and, 218-220, Hertz, Heinrich, 451 mirror, 32, 33
219 Hertz (Hz), 455 movement of, 70, 70-71, 71
nodal points and, 213-216, High energy wave, 451-453, 452, 453 optically real, 35
214-216 Hole-in-the-hand illusion, 3 physically real images versus,
nodal slide and, 220, 220 Holograms, 29, 30, 494, 494-496, 495 30-31, 31
power and focal length relations master, 496 point, 13, 295
and, 208-210, 209, 210 rainbow, 496 Purkinje-Sanson, 187
principal plane and reflection, 496 quality of, detriments to, 423
primary, 200, 200-202, 202 transmission, 496 real, as objects for eye, 31-32, 32
secondary, 195-197, 196-199 Holography, 494, 494-496, 495 real line, 312-313, 313
ray diagrams and, 203, 203-204, second generation, 496 retinal, size of, 102-104, 102-104
204 Homocentric bundle, 14 size of, for distant objects, 16-11,
reduced, 212-213, 213 Homogeneous material, 8 77
SSRI and thin lens as, 218, 218 Horizontal magnification, 322, 323, 325 in two-lens system, 35-37, 36
summary of equations for, 226 Huygens, Christian, 475 virtual, 32-34, 33, 35
symmetry points and, 211-212 Huygens eyepiece, 430, 431 virtual line, 312-313, 313
thick lenses and, 221-224, 222, 223 Huygens-Fresnel principle, 475-476, Image distances, 57-58, 59
examples and, 224-225, 225 476 Image point, 14, 15
vergence equations and, 206, Hyperfocal distance, 419 Image space, 37-39, 38
206-208, 208 Hyperope, 27, 27 for single spherical refracting
Generalized vergence, 21 accommodation and, 27, 91-92, 92 interfaces, 128-129, 129
Geometric optics, 2 corrected, 97-98 Incidence, angle of, 6
Ghost images, 186, 186 by contact lens, near point Incident angle, 360-363, 360-364
Glan-Thompson prism, 512 and, 100 Index of refraction, 8-9, 16
Glass-air interface, object space/image ocular accommodation through of air, 10
space and, 128, 129 spectacles and, 94, 94 Induced effects, 516
Glaucoma, 359 retinal image sizes of, 103, Infants, 418
Gold, absorption by, 6, 7 104, 104 Inferior mirage, 10, 11
Goldman lens, 359 by spectacle lenses, 86-87, 87 Infinity, optical, 15
Gradient index fiber, 359 virtual objects and, 34-35 Infrared light, 5
Grating. See Diffraction grating improper corrections and, 100-101 Intensity, of waves, 454-455
Grating target, 318 range of clear vision for, 92-93, 93 Interference
Green response, desaturated, 5-6 speckles and, 466 constructive, 457
Index 569

destructive, 457, 458 ruby, 526-527, 527, 528 design of, aberrations and,
methods for producing, 464-465, safety and, 529 447-448, 448
465 semiconductors as, 528 different, for same eye, 283-284
in ripple tank, 458, 458-459, 459 spatial filtering and, 552 diverging, 41-43, 43, 44, 153, 153
single photons and, 531, 531-533, ultraviolet, 527-528 common features with
532 for visual acuity testing of visual converging lenses, 54-55, 55
thin film, 468, 468 pathways, 464 optical axis of, 44, 44
Interference filters, 472-473, 473 Laser refraction, speckles and, predictable rays and off-axis
Interferometer 465-468, 466, 467 object points and, 52,
Michaelson, 465, 465 Laser surgery, 528-529 52-54, 54
Fabry-Perot, 473 Laser theory, 526-527, 527 primary focal point for, 46, 46
Interval of Sturm, 302 Lateral coherence length, 464 secondary focal point for, 45,
Inverse matrix, 555-556 Lateral magnification, 60-61, 61 46
Inversion, left-right, inversion and, 186 bundles and beams and, 76, 76 doublet, 436, 436
Investigation, as problem solving image size for distant objects and, equiconcave, 153, 153
technique, 1 76-77, 77 equiconvex, 153, 153
Iridescence, in nature, 493 for mirrors, 183 eye, 430
Iris, 4 radiuscope and, 183-184, 184 eye models and, 157
color of, 503 plane refracting interfaces and, field, 409, 430
Irradiance, 455 137, 137-138, 138 Fresnel, prism properties of, 388,
Iseikonic lens, 283 for single spherical refracting 388
Iso-prism curves, 381, 381 interfaces, 118-119, 119 Goldman, 359
Iso-thickness curves, 392, 392-394, 393 in two-lens system, 71-74 iseikonic, 283
Isotropie materials, 512 unit, 69t, 69-70, 70 Koeppe, 359
rectilinear propagation and, 10 Left-right inversion, reflection and, 186 magnifying, 248
Lens(es). See also Contact lenses; meniscus-concave, 153, 153
Crystalline lens; Multilens systems; meniscus-convex, 153, 153-154
Two-lens systems minus, 88
J-sheet polaroid, 508
achromatic, 428, 429-430 cylindrical, 305
Jackson cross cylinder (JCC), 312
biconcave, 153, 153 Newton's equation and, 161-164,
axis refinement with, dioptric
biconvex, 153, 153 162, 163
power and off-axis meridians
bifocal, range of clear vision and, objective, 257
and, 348-349, 349-350, 351
101-102 obliquely crossed, 337
Chevalier's, 447 ocular, 257
condenser, 413 ophthalmic, motions and, 388
Kepler, Johannes, 9, 25 contact. See Contact lenses optical center of, 44, 45
Keplerian telescopes. See Telescopes, converging, 41, 42-43, 43, 44, 49, planoconcave, 153, 153
Keplerian 50-52, 51, 153, 153 planoconvex, 153, 153
Keratometry, 184-185 common features with plus. See Plus lenses
Kirchhoffs law, 519 diverging lenses, 54-55, 55 prism properties of, 379-399
Knapp's law, 291 predictable rays and off-axis decentration with
Koeppe lens, 359 object points and, 47-49, spherocylindrical lenses and,
K readings, 185 47-50 398-399, 399
Krypton ion lasers, 528 primary focal point for, 46, 46 Fresnel lenses and, 388, 388
secondary focal point for, 45, iso-thickness curves and, 392,
45 392-394, 393
Land, Edwin, 508 Cooke triplet, 148 motions and, 387-388
Large particle scattering, 499 cylindrical. See also Jackson cross Prentice's rule and, 379-381,
LASER (Light Amplification by cylinder; Lens(es), 380, 381
Stimulated Emission of Radiation), spherocylindrical prism components in
526 astigmatism and, 295-299, spherocylindrical lenses and,
Laser(s) 296-298, 302 394-397, 394-398
argon ion, 527, 528 crossed, collapsing conoid by prism effectivity with lenses
carbon dioxide, 527, 528 equalizing, 303-304, 304 and, 389, 389-392, 391
excimer, 527-528 minus, 305 prism-lens combinations and,
helium-neon, continuous beam, perpendicularly crossed, 380, 381-383, 382, 383
527, 528 300-303, 301-303 prisms in cylindrical lenses
krypton ion, 528 plus, 305 and, 383-385, 384, 385
Nd:YAG, 528 prisms in, 383-385, 384, 385 rotational deviations and
neodymium, 528 rotational deviations by, spectacle magnification and,
pulsed, 527 385-386, 386 386-387, 387
570 Index

Lens(es) (continued) bundles and beams and, Linear magnification. See Lateral
rotational deviations by 76, 76 magnification
cylindrical lens and, unit, 69t, 69-70, 70 Linear systems, periodic gratings and,
385-386, 386 object and image distances 539, 539-542, 541
rectilinear, rapid, Dallmeyer's, 447 and, 57-58, 59 Line image, 295
separated, of same material, object movement /image Line spectrum, 522
aberrations and, 430-431, 431 movement and, 70, 70-71, Longitudinal coherence length, 463
shape and power approximations 71 Longitudinal wave, 452, 452
and, spectacle magnification and, power of, 153-154, 154 Low energy wave, 451-453, 452, 453
282-283 primary focal point for, 46, 46 Low vision telescope, 413
shape and power factors and, reversibility and finding the Lyman series, 522
spectacle magnification and, 281, object and, 77-78, 78
281-282 secondary focal point for, 45,
shapes of, 153, 153 45-46, 46
spectacle. See Spectacle(s) spectacle magnification and, . See Millimicron
spherical, working of, 41-43, 42, 275-278, 276 MA. See Mixed astigmat
43 symmetry points and, 68, Macro-photography, 416
spherocylindrical 68-69, 69 Magnetic field, 451
aberrations and, 448-449 two-lens systems and, 71-74, Magnification
accommodation and, 321, 72-74 angular. See Angular magnification
321-322 two lenses in contact and, chromatic difference in, 425
decentration with, 398-399, 74-75 horizontal, 322, 323, 325
399 unit lateral magnification and, lateral. See Lateral magnification
effectivity and, 320-321 69t, 69-70, 70 linear. See Lateral magnification
iso-thickness curves for, vergence equation for, 57, 58 oblique, 325
392-393 tinted, imaging with, 37-39, 38 shape, 280
obliquely crossed, adding, 342, Lens clock shape factor, 325-326
342-344, 343 to measure curvature of meridian spectacle. See Spectacle
prism components in, 394-397, on toric surface, 331-333 magnification
394-398 single spherical refracting interfaces unit, for single spherical refracting
spherical equivalent of, 312 and, 131, 131-132 interfaces, 124-125, 125t
tear, 160, 160-161, 161 Lens effectivity, 88 vertical, 322, 325
thick, 221-224, 222, 223 Lens measure, 131 Magnification equality, for afocal
chromatic aberration and, 429, Lensometer, 256 systems, 411-412
429-430 Lens system, orthoscopic, 446, 447 Magnifier, 408
examples with, 224-225, 225 Light(s) collimating. See Collimating
front and back vertex powers circularly polarized, 513, 513, 514 magnifier
of, 155-156 collimated, 312 simple, 257
motions and, 388 converging, 23 Magnifying lens, 248
piano, magnification properties reduced systems and, 141, Magnifying power, 266
of, 279-280, 280 141-142, 142 Maiman, Ted, 526
thin, 44, 45. See also Thin lens eye diffraction of. See Diffraction Major meridian, 315
models fluorescent, 525 Major Reference Point (MRP), 382
in air, equidistance of focal infrared, 5 Malus, Etienne, 507
points for, 47, 47 monochromatic, 5 Malus, law of, 507
boundary role of focal points paraxial, Fresnel's law for, 167-168 Marginal ray, 273
and, 67 partially polarized, 507 MASER (Microwave Amplification by
bundles, beams, and lateral polarized, 507, 507 Stimulated Emission of Radiation),
magnification and, 76, 76 propagation through anisotropic 526
in different media, 156-157 crystals, 14 Master hologram, 496
dioptric power: focal length quantum theory of, 521 Matrix algebra, 553-556
relationships and, 59-60, 60 speed in a medium, 8-9 Matter, quantum theory of, 5
equations for, 57-78 twisting sheets of, conoid of Sturm Maxwell, James Clerk, 451
erect and inverted and, dioptric power and off-axis Maxwell's electromagnetic equations,
relationships and: single meridians and, 329-330, 330 168
lenses, 71 ultraviolet, 5 Media
as Gauss system, 218, 218 unpolarized, 507 changes in, without change in
image sizes for distant objects visible, 4, 455 curvature, 130-131
and, 76-77, 77 wavelength of. See Wavelength thin lenses in, 156-157
lateral magnification and, Light wave, 451-453, 452, 453 Meniscus-concave lens, 153, 153
60-61, 61 Linear depth of field, 418 Meniscus-convex lens, 153, 153-154
Index 571

Meridian(s) Moon illusion, 4 Nicol prism, 512


axis, 295-299, 299 Motion(s), prism properties of lenses nm. See Nanometer
base-apex, 367 and, 387-388 Nodal plane, for single spherical
emmetropic, 314 MRP. See Major Reference Point refracting interfaces, 125-126,
hyperopic, 315 MTF. See Modulation transfer function 125-126
major, 315 Multilens systems. See also Two-lens Nodal points, 54
principal, 295, 329 systems in Gauss system, 213-216, 214-216
Meridional ray, 333 typical aperture stops and, primary and secondary, 213
Meridional refraction, 352 414-415, 415 for single spherical refracting
Meridional telescopes, astigmatism and, Multiple scattering, 501, 501 interfaces, 117-118, 118
322-325, 323, 324 Multiple spherical surface systems, Nodal rays, 50, 50
Metastable states, 525 144-147, 145-147 reflection and, 170-176, 171-174,
Michelson interferometer, 465, 465 Multiplication 176
Microscope(s) matrix, 554-555 Nodal slide, 220, 220
afocal, 261, 261-263, 262 sealer, 554 Nonhomogeneous material, 8
compound, 252, 257-259, 258 Muscle(s) Nonselective scattering, 8
converting telescope to, 264-265 ciliary, 4 Normal dispersion, 505
oil immersion, 439, 439 extra-ocular, 355 Normal to the surface, 6
optical properties of, 548, 550 Mutually coherent waves, 459-460 Numerical aperture, 488
Microscopy Mutually incoherent waves, 459-460
dark field, 551 Myope, 26, 27
phase contrast, 551 corrected Object(s), 35
Microspherometer, 183-184, 184 accommodation and, 97, 98-99 distant
Mie, Gustave, 501, 501 by contact lenses, 98-99, 100 image size for, 76-77, 77
Millimicron (), 455 near point and, 100 imaging, 129-130, 130
Minifiers, 407, 408 ocular accommodation through extended, multiple conoids and,
Minus lenses, 88 spectacles and, 95, 95 308-310, 308-310
cylindrical, 305, 317, 442 spectacle lenses and, 86 finding, reversibility and, 77-78, 78
Mirage, 10, 11 improper corrections and, 100, 101 movement of, 70, 70-71, 71
fata morgana, 10 near point for, 90, 90 optically real, 35
inferior and superior, 10 range of clear vision for, 90, 90, physically real objects versus,
Mirror(s). See also Reflection 92, 93 29, 29, 30
aberrations and, 449 speckles and, 466 real, reflection and, 171-173, 172,
center of curvature of, 171 spectacle magnification and, 279, 173, 173-174, 174
concave, 167 288, 291 in two-lens system, 35-37, 36
converging, 167, 168 thin lens eye model of, 82-84, 83, virtual, 34, 34-35
convex, 167 84 reflection and, 173, 174,
diverging, 167, 168 Myopic meridian, 314 174-176, 176
equivalent, 187-192, 188, 190-192 Object distances, 57-58, 59
focal point of, 171 Object-image matrix, 237, 237-238
primary, 170 Nanometer (nm), 455 Objective lens, 257
secondary, 169 Nautilus eye, 25 Object points, off-axis, predictable rays
partially silvered, 186-187, 187 Nd:YAG laser, 528 and
plane, 185, 185-186 Near field diffraction, 485 converging lenses and, 47-49,
in two-lens system, 36-37 Near point(s) 47-50
unit lateral magnification for, 183 corrected, 99-100 diverging lenses and, 52,
Mirror images, 32, 33 for myope, 90 52-54, 54
Mixed ametrope, 288 Nearsightedness. See Myope Object space, 37-39, 38
Mixed astigmat (MA), 315 Near triad, 420 for single spherical refracting
Modeling, in problem solving, 1 Negative chromatic aberration, 435, 435 interfaces, 128-129, 129
Modulation, def., 535 Negative coma, 436, 436-438, 437 Oblique astigmatism, 317
Modulation transfer function (MTF), Negative spherical aberration, 435, 435 Oblique central refraction, 439
538, 538 Neodymium lasers, 528 Oblique magnification, 325
spread functions and, 542, Nerve fiber layer of Henle, 516 Ocular accommodation. See
542-544, 544 Neutralization, hand, 387 Accommodation
of various systems, 544, 544-546, Neutralizing dioptric power, 148, 155 Ocular accommodative demand, 89
546 Newborn infants, 418 Ocular birefringence, Haidiger's brush
Monocentric bundle, 14 Newton, Isaac, 162 and, 516
Monochromatic aberrations, 423 Newton's equation, 161-164, 162, 163 Ocular lens, 257
wavefront, 431-432, 432 Gauss system and, 218-220, 219 Ocular view, 320
Monochromatic light, 5 Newton's rings, 473, 473-474 Oersted, Hans Christian, 451
572 Index

Off-axis meridians, dioptric power and. Penumbra, 10 principal. See Principal plane
See Dioptric power, off-axis Percent gain (loss), 277-278 sagittal, 436, 436-438, 437
meridians and Perception, 3-4 symmetry, 68
Off-axis object points. See Object Perfect blackbody, 519 tangential, 436, 436-438, 437
points, off-axis, predictable rays and Period, 454 Plane interface, 135. See also Plane
Oil immersion microscope, 439, 439 Periodic gratings, linear systems and, refracting interfaces
"Old age" vision, 26 539, 539-542, 541 Plane mirrors, 185, 185-186
Opal, as diffraction grating, 493 Periodic wave, 451-453, 452, 453 Plane refracting interfaces, 135-151
Ophthalmometer, 184-185 Petzval, Josef, 444 converging light and reduced
Optical activity, 516 Petzval condition, 444-446 systems and, 141, 141-142, 142
Optical axis, 44, 44, 381 Petzval equation, 444-446 diverging and converging
Optical center, 44, 45, 379-383, Petzval's portrait lens, 447 wavefronts and, 136, 135-137,
380-383 Petzval surface, 445 137
Optical fibers, 359, 359-360 Phase, of harmonic wave, 453 flat slabs and, 139-141, 139-141
Optical infinity, 15 Phase changes, at reflection, 469, ray tracing through, in equi-
Optically isotropic materials, 10 469-470, 470 index media, 138-139, 139
Optically real images, 29-31, 31, 35 Phase contrast microscopy, 551 front and back vertex powers and,
physically real images versus, Phase retarders, 514, 514-515 147-149, 147-149
29-31, 31 Phase velocity, group velocity versus, lateral magnification and, 137,
Optically real objects, 29, 35 506 137-138, 138
physically real objects versus, 29, Phonograph record ribs, as diffraction reduced angles and, 150, 150-151
30 gratings, 492 as special case of spherical
Optical microspherometer, 183-184, Phosphorescence, 524-525, 525, 526 interfaces, 135, 136
184 Photocoagulation, with lasers, 528-529 systems with multiple spherical
Optical path length, 111 Photoelasticity, 515 surfaces and, 144-147, 145-147
Optical systems, precision, 432 Photoelectric effect, 521 vertex neutralization and, 150
Optical thickness, 472 Photography. See Camera Plane waves, 15, 16
Optical tube length, 259 Photons, 5, 520-522, 521, 533 collimated, 527, 527
Optics probability and, 529, 529 Planoconcave lens, 153, 153
geometric, 2 single Planoconvex lens, 153, 153
physical, 2 diffraction and, 530-531 Plate
physiological, 2 interference and, 531, full wave, 514
scope of, 2 531-533, 532 half wave, 514, 514
Optometers, based on laser refraction, wavefunction of, 529, 529 quartex wave, 515, 515
468 Photon spin, 522 Plus lenses, 88
Ordinary ray, 512 Photovaporization, with lasers, 528 as collimating magnifier, 251t,
Orthoscopic lens system, 446, 447 Physically real images, 30, 31 251-253, 252
Oscillation, dipole, 499, 500 optically real images versus, 29-31, cylindrical, 305, 441
Oscillator, cavity, 527 31 Point(s)
Over-minusing, 101 Physically real objects, optically real cardinal, reversibility and, in Gauss
Over-plussing, 101 objects versus, 29, 30 system, 216, 216-218, 217
Over-refraction, dioptric power and off- Physical optics, 2 conjugate, 37
axis meridians and, 344-345 Physiological optics, 2 optical axis and, 44
Physiologic halo, 492-493 focal. See Primary focal point;
Pincushion distortion, 446 Secondary focal point
Pinholes, 23-26, 24, 25 image, 14, 15
Pantoscopic tilt, 441, 441 to determine cause of poor acuity, nodal. See Nodal points
Panty hose, as diffraction gratings, 492 489 object. See Object points, off-axis,
Parallax, 494-495, 496 to distinguish optical defect from predictable rays and
Paraxial light, Fresnel's law for, neural pathology, 25-26 principal
167-168 Planar system, Zeiss, 447 primary, in Gauss system, 200
Paraxial ray, 14 Planck, Max, 521 secondary, in Gauss system,
Paraxial ray trace, zigzag, 195-197, Planck's constant, 521, 522 195
196-199 Plane(s) symmetry, 68, 68-69, 69
Partially polarized light, 507 conjugate, 204-206 in Gauss system, 211-212
Partially silvered mirrors, 186-187, 187 focal, secondary, front vertex Point image, 13, 295
Paschen series, 523 matrix to, 238, 238-240, 239 Point sources, 13, 14
Pathological halo, 493 nodal, for single spherical extended, 21-23, 22
Patterns, in problem solving, 1 refracting interfaces, 125-126, Poisson, Simeon, 485
Pencil, 14, 14 125-126 Poisson spot, 485
Penta prism, 376, 376 polarization, 509 Polarization, 453
Index 573

binocular vision and, 516 in Gauss system, 200, 200-202, 202 Prism-lens combination, 380, 381-383,
circular, 513, 513, 514 Primary source, 2 382, 383
materials and, 515, 515-516 Principal meridians, 295, 329 Prism properties. See Lens(es), prism
elliptical, 513 Principal plane properties of
fundamentals of, 506-507, 507 primary, 285-287 Problem solving techniques, 1-2
optical activity and induced effects in Gauss system, 200, Prokhorov, Alexander, 526
and, 516 200-202, 202 Ptolemy of Alexandria, 9
by reflection, 509-511, 510 secondary, 285-287 Pulsed laser, 527
by scattering, 511, 511 in Gauss system, 195-197, Punctum proximum. See Near point
Polarization plane rotation, by 196-199 Punctum remotum. See Far point
polaroids, 509 Principal point Pupils. See Entrance pupil; Exit pupil
Polarized light, 507, 507 primary, in Gauss system, 200 Purkinje-Sanson images, 187
Polarizers, circular, 515, 515-516 secondary, in Gauss system, 195
Polaroids, 507-509, 508, 509 Principal section, 360
absorption axis of, 508 Prism(s), 355-376 Quantum particles, 521
H-sheet, 508 achromatic, 428, 428-429 Quantum theory of light, 521
J-sheet, 508 Amici, 376, 376 Quantum theory of matter, 5
polarization plane rotation by, 509 apex of, 355 Quartex wave plate, 515, 515
transmission axis of, 508 base down, 367-368
vectographic, 516 base in, 355, 357, 367-368
Polaroid sunglasses, 510, 515 base of, 355 Rabl's laminae, 492
Population inversion, 526 base out, 367-368 Radial astigmatism, 437, 439-444,
Porro prism, 375, 375-376 base up, 367-368 440-443
Position of least confusion, 302 centrads and, 371, 371 Radial lens fibers, 492
Positive chromatic aberration, 435, 435 deviations and Radiation
Positive coma, 436, 436-438, 437 minimum, 363-365, 364-365 blackbody, 519-520, 520
Positive spherical aberration, 435, 435 at normal incidence and, 363, dipole, 499, 500
Power 365-366, 366 electromagnetic, 4-5
astigmatism classification by, 315 thick prisms and, 355, thermal, 519
catoptric, 176 360-363, 360-364 Radiuscope, 183-184, 184
dioptric. See Dioptric power diagnosis of extra-ocular muscle unit lateral magnification and,
equivalent, 221 problems using, 355 183-184, 184
for magnification, 260-261 diopters and, 368, 368-369, 369 Rainbow, 424, 424
focal length and, in Gauss system, dispersion and, 423-424, 424 Rainbow hologram, 496
208-210, 209, 210 Dove, 376, 376 Ramsden circle, 409
magnifying, 266 effectivity of, 372-374, 373 Range of clear vision, 90, 92-93, 93
vertex, front and back, 148, Fresnel, 374-375, 375 for emmtrope, 90-91, 92, 93
155-156 Fresnel BI, 464-465, 465 for hyperope, 91-93, 92, 93
plane refracting interfaces and, Glan-Thompson, 512 for myope, 90, 90, 92, 93
147-149, 147-149 Nicol, 512 through various lenses, 101,
Power cross, 299 ophthalmic base directions and, 101-102
Power factor, spectacle magnification 367-368, 368 Rapid rectilinear lens, 447
and, astigmatism and, 325 Penta, 376, 376 Rattlesnake eye, 25
Power gain, 440, 440 as polarizers, 512 Ray(s), 13, 14, 522. See also
Power meridian, 295-299 Porro, 375, 375-376 Predictable rays
Precision optical systems, 432 reflecting, 375, 375-376, 376 chief, 273
Predictable rays, 54 Risley, 372, 372 crepuscular, 10
converging lenses and off-axis thick, 355, 360-363, 360-364 defined, 13-14
object points and, 47-49, 47-50 deviation by, 360-363, extraordinary, 512
diverging lenses and off-axis object 360-364 marginal, 273
points and, 52, 52-54, 54 thick prism comparisons and, 366, meridional, 333
Prediction, in problem solving, 1 366-367 nodal, 50, 50
Prentice's rule, 231, 379-381, 380, 381 thin, 355, 366, 367, 367 reflection and, 170-176,
Presbyopia, 26 thin prism combinations and, 369, 171-174, 176
Primary extra-focal length, 162 369-371, 370 ordinary, 512
Primary focal point ( F J , 46, 46 total internal reflection and, 356, paraxial, 14
of mirror, 170 357-360, 358-360 zigzag ray trace and, 195-197,
for single spherical refracting Wollaston, 512 196-199
interfaces, 115, 115 Prism-diopter, 368, 368-369, 369 sagittal, 436, 436-438, 437
Primary nodal point, 213 Prism effectivity, with lenses, 389, skew, 333
Primary principal plane, 285-287 389-392, 391 tangential, 436, 436-438, 437
574 Index

Ray diagrams total internal, prisms and, 356, spectacle magnification and,
Gauss system and, 203, 203-204, 357-360, 358-360 386-387, 387
204 unit lateral magnification and Rotational symmetry, 299
reflection and, 170-176, 171-174, radiuscope and, 183-184, 184 Ruby laser, 526-527, 527, 528
176 vergence equations and, 176-177 Rule of thirty, 320
scaling of, 55, 56 Refracting interfaces. See Plane
Ray extension, 54 refracting interfaces; Single spherical
Rayleigh scattering, 8, 500-501, 501 refracting interfaces Sagitta, 108-109, 109
Rayleigh's criterion, 486-488 Reflecting holograms, 496 Sagittal approximation, 108, 109-110
Ray tracing, through flat slab, in equi- Refracting states, of eye, 26, 26-27, 27 Sagittal plane, 436, 436-438, 437
index media, 138-139, 139 Refraction, 9, 9-10, 10 Sagittal ray, 436, 436-438, 437
Reading cap, on telescope, 264-265 angle of, 9 Sealer multiplication, 554
Reading center, 396 central, oblique, 439 Scattering, 8, 8, 423
Real focal points, 203, 203-204 index of, 8-9, 16 coherent, 502, 502
Real images, as objects for eye, 31-32, of air, 10 incoherent, 499-500, 500, 501, 501
32 laser, speckles and, 465-468, 466, intraocular light, 502-503
Real line images, 312-313, 313 467 large particle, 499
Real objects, reflection and, 171-174, law of, for paraxial rays, 14 multiple, 501, 501
172-174 meridional, 352 polarization by, 511, 511
Reciprocity law, film exposure and, over-refraction and, dioptric power Rayleigh, 8, 500-501, 501
415 and off-axis meridians and, resonance, 503-504, 504, 505
Rectilinear propagation, 10, 11 344-345 selective absorption and, 503
Red-green test, 427 Refraction matrix, 230-231, 231 selective and nonselective, 8
Reduced angles, 150, 150-151 dioptric power and off-axis small particle, 499-500
Reduced distance, 140 meridians and, 334, 334-336 Schawlow, Arthur, 526
Reduced eye, system matrices and, Refractive ametrope, 288 Scientific method, problem solving
239, 242 Refractive efficiency, 425t, 425-427, techniques of, 1-2
Reduced systems, 144 426, 426t, 427 Scissors effect, 386
convergent light and, 141, Refractive error, residual, dioptric Scissors movement, 323, 324
141-142, 142 power and off-axis meridians and, Sciera, coherent scattering and, 503
Gauss, 212-213, 213 347-348, 348 Secondary extra-focal length, 162
single spherical refracting interfaces Relative distance magnification, 248, Secondary focal point (F 2 )
and, 142, 142-144, 143 249-250 of mirror, 169
Reduced vergence, 21 Relative size magnification, 247, for single spherical refracting
Reflectance, 519 247-248 interfaces, 113-117, 114
Reflecting prisms, 375, 375-376, 376 Resolution for thin lens, 45, 45-46, 46
Reflection, 6-7, 67, 167-192 diffraction limit on, 486-488, 487 Secondary nodal point, 213
absorption and, 6 spurious, 544 Secondary principal plane, 285-287
angle of, 6 Resonance frequencies, 503-504, 504 in Gauss system, 195-197, 196-199
diffuse, 7, 7 Resonance scattering, 503-504, 504, Secondary source, 2
emmtropes and, 182-183, 183 505 Secondary spectrum, 428
equivalent mirrors and, 187-192, Retina, 2 - 3 , 4, 355 Second generation holography, 496
188, 190-192 matrix of cornea to, 240 Seidel, Ludwig von, 433
focal points and, 169, 169-171, 170 peripheral, 247 Seidel aberrations, 432-433
ghost images and partially silvered Retinal image, size of, 102 Selective absorption, 6
mirrors and, 186, 186-187, 187 with spectacle versus contact scattering and, 503
imaging examples and, 172, 174, lenses, 102-104, 103-104 Selective scattering, 8
177, 177-182, 180 Retinoscopy, streak, 325 Semiconductor lasers, 528
keratometry and, 184-185 Reversibility SHA. See Simple hyperopic astigmat
law of, 6-7, 167 cardinal points and, in Gauss Shadows, 10
left-right inversion and, 186 system, 216, 216-218, 217 Shape factor magnification, astigmatism
mirrors and other interfaces and, finding the object and, 77-78, 78 and, 325-326
167-169, 167-169 principle of, 37, 78 Shape magnification, 280
phase changes at, 469, 469-470, Ripple tank, interference in, 458, Shells, as diffraction gratings, 493
470 458-459, 459 Silk, as diffraction gratings, 492
plane mirrors and, 185, 185-186 Risley prisms, 372, 372 Silver, absorption by, 7
polarization by, 509-511, 510 Rittenhouse, David, 492 Similar triangles, 297, 298
ray diagrams and nodal ray and, Rods, 4 Simple hyperopic astigmat (SHA), 315
170-176, 171-174, 176 Rotation, center of, 372-374 Simple myopic astigmat (SMA), 315
specular, 6-7, 67, 509 Rotational deviations Sine wave gratings, square wave
surface, 6 by cylindrical lenses, 385-386, 386 gratings versus, 535-538, 536, 537
Index 575

Single-slit diffraction, 476-479, 477, 478 coherent transfer function and, 548 band, 524
Single spherical refracting interfaces coupled systems and, 546-547 continuous, 524
(SSRIs), 107-132 general objects and, 547, 547 electromagnetic, 4
changing media without changing modulation transfer function and, line, 522
curvature and, 130-131 538, 538 prisms and, 355
convex/concave terminology and, of various systems and, 544, secondary, 428
107-108, 108 544-546, 546 Specular reflection, 6-7, 67, 509
derivation of vergence equation nonlinear example and, 547-548, Spherical aberration, 432-436, 433-436
and, 110-112, 111 548 positive and negative, 435, 435
dioptric power and, 112, 112-113, periodic gratings and linear systems Spherical ametropia, 27
113 and, 539, 539-542, 541 Spherical equivalent, 343
exploding, 157-159, 158 sine versus square wave gratings circle of least confusion and,
focal points and, 113-117, 114-117 and, 535-538, 536, 537 310-312, 311, 312
as Gauss system, 218, 218 spatial filtering and, 550, 550-552, Spherical interfaces. See also Single
imaging a distant object and, 551 spherical refracting interfaces
129-130, 130 spatial frequency and, 535, 536 plane interfaces as special case of,
imaging examples and, 119-123, spread functions and the 135, 136
119-123 modulation transfer function Spherical surfaces, multiple, systems
lateral magnification and, 118-119, and, 542, 542-544, 544 with, 144-147, 145-147
119 Spatial filtering, 550, 550-552, 551 Spherocylindrical lenses. See Lens(es),
lens clock readings and, 131, Spatial frequency, 535, 536 spherocylindrical
131-132 Speckles, laser refraction and, 465-468, Spherocylindrical parameters, finding
nodal plane and, 125-126, 125-126 466, 467 from dioptric power matrix, 341,
nodal point and, 117-118, 118 Spectacle(s). See also Lens(es) 341-342
object and image space and, accommodation through, algebraic Spontaneous emission, 523, 526
128-129, 129 approach to, 94, 97-98 Spread functions, modulation transfer
reduced systems and, 142, corrected curve, 446 function and, 542, 542-544, 544
142-144, 143 corrections and, 86-88, 86-88 Spurious resolution, 544
reflection and, 167 improper, 100-101 Square wave gratings, sine wave
sagittas and sagittal approximation ocular accommodation through, 94, gratings versus, 535-538, 536, 537
and, 108-110, 109 94-95, 95 SSRIs. See Single spherical refracting
series of, 135, 136 stress pattern and, 515 interfaces
strange cases and, 126-128, 127, Spectacle accommodative demand, Standing waves, 459, 459
128 93-94 Stefan-Boltzman law, 520
symmetry points and, 123-124, 124 Spectacle magnification, 267-293 Stenopeic slit, 329
unit magnification and, 124-125, aperture stop and entrance pupil Step index fiber, 359
125t and, 267-268, 268-272, 270-273 Stereopsis, 516
vergence equations for, 176 approximation of, 278-279 Stigmatic systems, 295
Skew ray, 333 comparison of different lenses for Stimulated emission, 526, 526
Sky, Rayleigh scattering and, 500 same eye and, 283-284 Stokes, G.G., 469
SMA. See Simple myopic astigmat exit pupil and blurred image sizes Streak retinoscopy, 325
Small particle scattering, 499-500 and, 272, 273-275, 273-275 Strutt, John William (Baron Rayleigh),
Smoke, Rayleigh scattering and, 501 Gauss formulation for, 284-287, 486-487
Snell, Willebrord, 10 285, 286 Sturm, J.F.C., 302
Snellen acuity, 488-489 magnification properties of thick Sturm interval, 302
Snell's law, 10, 356 piano lens and, 279-280, 280 Sturm's conoid. See Conoid of Sturm
Soap films, 468, 470-471 power factor and, 281, 281-282 Subsystem matrices, 235, 235-236
Sodium D doublet, 426 astigmatism and, 325 Sun dogs, 365
Sodium fluorescein, 525 relative, 288-292, 289-292, 293t Sunglasses, polaroid, 510, 515
Sound waves, 452, 452 rotational deviations and, 386-387, Superachromats, 428
Space 387 Superior mirage, 10
image, 37-39, 38 shape and power approximations Superposition, 456-458, 457, 458
for single spherical refracting and, 282-283 Surface reflection, 6
interfaces, 128-129, 129 shape factor and, 281, 281-282 Surgery, with lasers, 528-529
mathematical use of, 37 astigmatism and, 325-326 Symmetric system, 446, 447
object, 37-39, 38 summary of equations for, 293 Symmetry, rotational, 299
for single spherical refracting telescopes and, 279 Symmetry planes, 68
interfaces, 128-129, 129 for ametropes, 287-288 Symmetry points, 68, 68-69, 69
Spatial distribution, 535-552 thin lens, 275-278, 276 in Gauss system, 211-212
Abbe theory of image formation Spectacle mounted telescopes, 264 for single spherical refracting
and, 548, 549, 550 Spectrum, 5, 423-424, 424 interfaces, 123-124, 124
576 Index

System matrices, 229-244 through spectacle lenses, 94, Tscherning ellipses, 443
ametropic eyes and, 240-241 94-95, 95, 97-98 Tunneling, 360
derivation of Gauss equation and, corrected near points and, 99-100 Two-lens systems, 71-74, 72-74
244 downstream vergence equation images and, 35-37, 36
determinant condition and, and, 95-96, 96 lateral magnification and, 71-74
234-235 emmtropes and, 81-82, 82 mirrors and, 36-37
dioptric power and off-axis range of clear vision for, objects and, 37-37, 36
meridians and, 336-337 90-91 system matrices and, 236, 236-237
front vertex to secondary focal hyperopes and, 84, 84-85, 85 thin, in contact, 74-75
plane, 238, 238-240, 239 range of clear vision for, Tyndall, John, 500
Gullstrand 1 schematic eye and, 91-92, 92 Tyndall blue, 503
239, 241-242, 242t improper corrections and, 100-101
object-image, 237, 237-238 lens effectivity terminology and, 88
refraction, 230-231, 231 myopes and, 82-84, 83, 84
Ultraviolet catastrophe, 521
subsystem, 235, 235-236 near point and range of clear
Ultraviolet lasers, 527-528
translation, 229, 229-230 vision for, 90, 90
Ultraviolet light, 5
two lens systems and, 236, 236-237 range of clear vision and, 90,
fluorescence and, 525
vertex vergence relations and, 237, 90-93, 92, 93
Umbra, 10
243-244 through various lenses, 101,
Uncertainty, 529, 529-530
101-102
Under-minusing, 101
retinal image sizes and, 102
Under-plussing, 101
spectacle versus contact lenses
Unit lateral magnification, 69t, 69-70,
Tangential plane, 436, 436-438, 437 and, 102-104, 103-104
70
Tangential ray, 436, 436-438, 437 spectacle accommodative demand
Unit magnification, for single spherical
Target and, 93-94
refracting interfaces, 124-125, 125t
extended, 308-310 spectacle lens corrections and,
Unit matrix, 555
grating, 318 86-88, 86-88
Universal product code (UPC), as
Telecentric stop, 415 Third order aberrations, 432-433
hologram, 496
Telescopes Toric surfaces, 306-308, 307
Unpolarized light, 507
afocal, 261, 261-263, 262 lens clock to measure curvature of
alternate equations for, meridian on, 331-333
263-264 Toric wavefront, 312
for ametropes, 287-288 Torsion, dioptric power and off-axis VDT. See Video display terminal
astronomical, 261 meridians and, 329, 330 Vectographic polaroid, 516
converting to microscope, 264-265 Torus, 306, 329 Vector(s), matrices and, 556
Galilean, 262-263, 288, 412-413, Total field of view, 404 Vector addition, 370-371
413 Total internal reflection, prisms and, dioptric power and off-axis
aperture stop for, 415 356, 357-360, 358-360 meridians and, 345-347, 346,
with reading cap, 265 Total percent gain, 283 347
Keplerian, 261-262, 288, 408, Townes, Charles, 526 Velocity, group versus phase, 506
408-409, 410, 411, 413 Trace, 340 Verdet, 464
aperture stop for, 415 Transfer function Vergence, 16-17, 17
exit pupil for, 412 coherent, 548 conversion factors and, 17, 18t,
low vision, 413 modulation. See Modulation 18-19
meridional, astigmatism and, transfer function derivation of equation for,
322-325, 323, 324 optical, 543 110-112, 111
reading cap on, 264-265 Transition rate, 524 downstream, 19-21, 19-21, 95-96,
spectacle magnification and, 279 Translation matrix, 229, 229-230 96
spectacle mounted, 264 dioptric power and off-axis in Gauss system, 206, 206-208, 208
Terrestrial telescope, 262-263 meridians and, 333, 333-334 generalized or reduced, 21
Theory, in problem solving, 1 Translucent materials, 8 reflection and, 176-177
Thermal radiation, 519 Transmission axis, of polaroid, 508 reversibility and finding the object
Thickness Transmission holograms, 496 and, 77
central, 279-280 Transmittance, 519 thin lenses and, 57, 58
iso-thickness curves and, 392, Transposition, equivalent combinations upstream and downstream changes
392-394, 393 and, 304-306 and, 19-21, 19-21
optical, 472 Transverse wave, 452 Vergence effectivity, 88, 96
Thin film interference, 468, 468 Trauma, endothelial, 502-503 spherocylindrical lenses and,
Thin lens eye models, 81-104 Triad, near, 420 320-321
accommodation and, 88-89, 89 Trial and error, as problem solving Vergence matrix, dioptric power and
through contact lenses, 98-99, technique, 1 off-axis meridians and, 351-352
99 Triangles, similar, 297, 298 Vergence vertex relations, 237, 243-244
Index 577

Vertex, front, matrix to secondary focal eye and, 4, 4 spherical, 29, 30


plane, 238, 238-240, 239 perceptual aspects of, 3-4 toric, 312
Vertex distance, 86 refracting states of eye and, 26, Wavefront aberration, monochromatic,
Vertex neutralization, 150 26-27, 27 431-432, 432
Vertex power, front and back, 148, Visual acuity. See Acuity Wavefunction, of photon, 529, 529
155-156 Visual pathways, laser visual acuity Wavelength, 5
plane refracting interfaces and, testing of, 464 changes in, frequency invariance
147-149, 147-149 Vitreous humor, 4 and, 455t, 455-456, 456
Vertex vergence relations, 237, 243-244 color and, 5t, 5-6
Vertical magnification, 322, 325 Wavicles, 521
Vertical prism imbalance, 397 Water-plastic interface, object space/ Wien's displacement law, 520
Video disk ribs, as diffraction gratings, image space and, 128-129, 129 With motion, 387
492 Wave(s) With rotation, 386
Video display terminal (VDT), basic properties of, 451-453, 452, With-the-rule (WTR) astigmatism, 315,
polarization and, 515 453 317, 317
Vignetting, 406 harmonic, 452-454, 453-454 Wollaston prism, 512
Virtual focal points, 204, 204 high energy, 451-453, 452, 453 Wollaston's meniscus lens, 447
Virtual images, 32-34, 33, 35 intensity of, 454-455 WTR. See With-the-rule astigmatism
Virtual line images, 312-313, 313 light, 451-453, 452, 453
Virtual objects, 34, 34-35 longitudinal, 452, 452
reflection and, 173, 174, 174-176, low energy, 451-453, 452, 453
176 periodic, 451-453, 452, 453 Xanthophyll, light absorption by, laser
Visibility, 535 plane, 15, 16 surgery and, 528-529
fringe, partial coherence and, 459, collimated, 527, 527
462-464, 463, 464 standing, 459, 459
Visible light, 4, 455 transverse, 452
Young, Thomas, 461
Vision, 2 - 3 , 3 Wavefront(s), 13-14, 14, 454
binocular, 355 converging, 14, 15
polarized light and, 516 plane interfaces and, 136,
clear. See Range of clear vision 135-137, 137 Zeiss, Carl, 548
color, 4 spherical, 15-16, 16 Zeiss planar system, 447
psychophysiological cylindrical, 295 Zeiss Tessar lens, 448
dependence of, 6 diverging, 15, 15, 17, 17 Zigzag paraxial ray trace, 195-197,
depth perception and, 3 plane interfaces and, 136, 196-199
double, 396 135-137, 137 Zone plates, 493, 493-494, 494

You might also like