You are on page 1of 178

Volume 9

APPLIED PHYSICS
AND ENGINEERING
An International Series

Geometrical Optics
Hans-Georg Zimmer

Geometrical Optics

With 46 Figures

Springer-Verlag New York· Heidelberg· Berlin 1970


Dr. HANS-GEORG ZIMMER

7082 Oberkochen
Thiersteinweg 4

ISBN-13: 978-3-642-86833-7 e-ISBN-13: 978-3-642-86831-3


DOl: 10.1007/978-3-642-86831-3
The use of general descriptive names, trade names, trade marks, etc. in this publica tiOD,
even if the former are not especially identified, is not to be taken as a sign that such names,
as understood by the Trade Marks and Merchandise Marks Act, may accordingly be used
freely by anyone
This work is subject to copyright. All rights are reserved, whether the whole or part oj
the material is concerned, specifical1ythose of translatioD,reprinting, re-use of illustrations,
broadcasting, reproduction by photocopying machine or similar means, and storage in
data banks
Under § 54 of the German Copyright Law where copies are made for other than private
use, a fee is payable to the publisher, the amount of the fee to be determined by agreement
with the publisher
© by Springer-Verlag Berlin· Heidelberg 1970. Library of Congress Catalog Card Number
72-94095. Title No. 3899
Softcover reprint of the hardcover 1st edition 1970
Foreword
Geometrical optics is no longer fashionable. Research workers do
not expect significant new discoveries to be made in this field of classical
physics. Teachers avoid the subject because its use for many generations
in arid mathematical exercises has robbed it of all freshness and stimulus,
with the result that it no longer seems relevant to a modern physics
course. There remains - and perhaps this has grown in recent years -
the technical significance of geometrical optics. It provides the basis for
the design of optical instruments for use in everyday life as well as for
scientific and industrial purposes.
This small book is intended to treat two aspects of the subject: the
laws of geometrical optics and their application to the design of optical
instruments. The theory is not based on Snell's law of refraction but on
a conservation law for the radiated energy. The subject can then be
treated in a manner appropriate to contemporary physics: auxiliary
geometrical parameters become unnecessary and the singularities
resulting from their use vanish. The laws of geometrical optics can be
formulated much more simply and their physical significance is revealed
more clearly.
I have tried to present the material in a form satisfactory both to
teachers and to workers in the technical applications of optics. The
content has thus been deliberately kept within the limits of a pocket book.
The reader will find neither a complete discussion of the theory nor a
manual for designing optical instruments; but I hope he will be encouraged
to delve deeper into the former and discover the interest of the latter.
I am indebted to Prof. FALK (Karlsruhe) and to the publishers,
Springer-Verlag, for the suggestion that the book should be written;
to the board of management of the firm of Carl Zeiss for their encourage-
ment and, above all, to Mr. HABERMANN and his colleagues in the optical
design department for microscope optics for their help with the theoretical
and experimental investigations.

Oberkochen, May 1967 H.-G. ZIMMER


Translator's note
There are perhaps two extreme attitudes which the translator of a
scientific book may adopt. On the one hand, he may legitimately take
the view that the translation represents a new work and effectively
paraphrase the original; in doing so, he may impart nuances to the
content which are to some extent influenced by his own ideas. The other
extreme implies a purely "passive" attitude: here the translator sticks as
closely as possible to the original text. I favour the second approach and
have not hesitated (in close consultation with Dr. ZIMMER) in certain
cases to make some sacrifice in style to ensure as exact a translation of
the scientific substance as possible. In this connexion, it is a pleasure to
record my debt to Dr. ZIMMER for his constructive criticism of the English
text.
To ensure freedom from error, the numbered equations and formulae
have been taken over unaltered from the German edition. As a conse-
quence, certain abbreviations (notably LLW = linearer Leitwert = linear
conductivity, and R.E. = Rayleigh-Einheit = Rayleigh unit) appear
throughout the English edition in their German form.

Heidenheim, May 1969 R. N. WILSON


Contents
Part 1: Perfect Imagery
§ Light, Radiation and Geometrical Optics .
§ 2Light Tubes . . . 2
§ 3Linear Conductivity 5
§ 4 Difiraction . . . . 7
§ 5 The Rayleigh Unit. 8
§ 6 Definition of Perfect Imagery . 9
§ 7 The Fundamental Processes of Perfect Imagery . 11
§ 8 Relationships between the Fundamental Processes 15
§ 9 Summary of all the Processes of Perfect Imagery 17
§ 10 The Eye . . . . • . . . . . . . 21
§11 Spectacles and Magnifying Glasses 25
§ 12 The Telescope and the Microscope. 27
§ 13 The Photographic Emulsion 32
§ 14 The Photographlc Objective . 34
§ 15 The Calculation of Light Tubes 39
§ 16 The Gaussian Brackets 45
§l7 Graphical Representation of Perfect Imagery 52
§l8 Condnctivity Diagrams of Optical Instruments 62
§ 19 Imagery in Depth . 66
§ 20 Summary of Part 1 . . . . . . . . . . . . 72

Part 2: Theory of Aberrations


§ 21 Aberrations and Geometrical Optics 75
§ 22 Optical Glasses . . . . . . . . . 77
§ 23 The Chromatic Aberrations of Perfect Imagery 85
§ 24 Analysis of the Primary Chromatic Aberrations 92
§ 25 Secondary Spectrum . . . . . . . . . . . 98
§ 26 Plane Rays . . . . . . . . . . . . . . . . 104
§ 27 The Primary Monochromatic Aberrations of Plane Rays 113
§ 28 Analysis of the Aberrations of Plane Rays 123
§ 29 The Secondary Aberrations of Plane Rays 129
§ 30 Skew Rays. . . . . . 137
§ 31 Astigmatism . . . . . 143
§ 32 Analysis of Astigmatism 152
§ 33 Coma . . . . . . . . 156
§ 34 The Influence of Object and Pupil Positions on Aberrations. 162
Conclusion and References to Literature 166
Index . . . . . . . . . . . . . . . . . . . . . . • . 169
PART ONE

Perfect Imagery
We must first consider what is meant by optical imagery and under
what conditions it can be considered perfect. The laws governing perfect
imagery will be derived and their significance discussed in detail. Every-
thing that is necessary for the understanding of the function of optical
instruments follows from these laws.

§ 1. Light, Radiation and Geometrical Optics


Definition of Terms
Light is a sensation which is produced through nerve cells in the eye
and the brain. These physiological and psychological processes are related
to the physical term radiation in the sense that radiation is usually the
cause of the sensation of light. But there are also other causes such as
door posts or alcohol.
Radiation is a fundamental physical term defining the transfer of
energy from an "emitter" to a "receiver". Its nature can vary widely.
In the case of sound it consists of oscillations in matter; radio trans-
mission is effected by means of electromagnetic waves; corpuscular radia-
tion consists of moving particles. For each of these cases, a different
mathematical formulation is necessary because there are fundamental
differences in the physical processes occurring in the matter involved.
The significance of this "interaction between radiation and matter"
becomes particularly clear if we consider the manner in which various
emitters and receivers work or the transmission factor of the matter in
question for the different types of radiation. Because of such effects,
radiation is a much used tool in that branch of experimental physics
which is concerned with the structure of matter. Examples range from
testing of materials with ultrasonic waves to experiments in nuclear
physics.
Geometrical optics is a branch of physics in which radiation is investi-
gated. We are not limited to one particular sort of radiation - that
giving rise to the sensation of light - but are concerned with radiation
in a quite general sense. However, a fundamental assumption is made,
namely, that the interaction of matter and radiation has no effect on the
1 Zimmer, Geometrical Optics
2 Perfect Imagery

transfer of the energy. The investigation is therefore confined to the


undisturbed (with respect to the energy) propagation of radiation or the
"interaction of radiation and space" .
In fact, this premise of geometrical optics is never fulfilled in natural
processes. But in many cases of practical importance the influence of the
matter on the quantity of energy it propagates is so small that it can only
be measured with great difficulty. In such cases geometrical optics gives a
good approximation to physical reality. Even when the matter has an
appreciable effect on the energy transferred, the laws of geometrical
optics can still give a very useful approximation because they are much
easier to manipulate and give a more readily understood picture of the
physical process than the strict theory applicable to the special case.
However, it is important to establish the range of validity of geometrical
optics if we are to avoid drawing false conclusions from such approxi-
mations. For this reason, the following sections are mainly concerned
with the assumptions involved in geometrical optics. In particular, I shall
deal with the radiation which gives rise to the sensation of light. This
seems to me to be justified because the theory of optical instruments is
the principal field of application of geometrical optics and because there
is no change in principle involved with other forms of radiation.

§ 2. Light Tubes
Geometrical Description
of the Light Cone used in Energy Measurements
Suppose the energy transferred by radiation is to be measured. Let the
emitter 0 be considered as two-dimensional, for example a glowing piece
of tin. It radiates light and heat in all directions, some of which reaches
the receiver P, also supposed to be two-dimensional. It could be a piece
of tin blackened with soot and placed in front of a sensitive thermometer
or thermo-element. How the emitter and receiver work is not the concern
of geometrical optics: we simply assume that they function perfectly.
But it is important that they be two-dimensional. Three-dimensional
emitters such as gas discharge lamps obey, strictly speaking, different
laws.
The volume of space connecting all points of 0 with all points of P
by straight lines will now be termed a light tube. To be fully consistent
we should have to call it a radiation tube since, in physical measurements,
we are only concerned with radiation, not with the sensation of light.
Furthermore, the measuring procedure is not confined to the radiation
we receive as light. Nevertheless, the term light tube is convenient in
practice.
Light Tubes 3

The light tube can have a very complex form which can only be
described with the help of differential calculus. However, all the funda-
mental problems of geometrical optics can be treated with the very
simplest model. This model is applicable to centred systems: emitter and
receiver are circular discs perpendicular to a straight line through their
centres. This straight line is termed the axis of the system or instrument.
This .simple case has great practical significance since the vast
majority of optical instruments are, in fact, centred systems or can be
interpreted as such. In the case of the photographic camera, the radiation
receiver (the emulsion on the film) is rectangular in form. It can never-
theless be treated as a centred system if we consider the circle through
its coruers to be the edge of the image. The theory developed in this book
presupposes centred systems even when this assumption is not explicitly

Fig. 1

mentioned. The special features of decentred imagery will only be briefly


discussed in the theory of aberrations.
Fig. 1 shows the light tube of a centred system in a plane containing
the axis. It is a straight frustrum of a cone which is limited on the left and
right by the circular discs 0 and P. The terms left and right may con-
veniently be used since it is a generally accepted convention for diagrams
and calculations in optics that the direction of the energy transfer be
from left to right. The rule breaks down for reflection. In such cases the
departure from the convention must be clearly shown by arrows. It is
convenient to term the direction of energy transfer more simply the
direction of the light.
Three coordinates are sufficient to describe light tubes mathematically
in this special case of centred systems, for example the radii Ro and Rp
of the circular discs 0 and P and the sepflration D between 0 and P.
This coordinate systems is very simple and practical for geometrical
investigations: but the physically important characteristics of light tubes
can be better described in the coordinate system shown in Fig. 2.
We join the centre of the emitter to the edge of the receiver by a
straight line lying in the plane of the paper which we shall call the
1*
4 Perfect Imagery

marginal ray. Similarly, we define a principal ray as the line joining the
edge of 0 to the centre of P. Now we choose a point E lying on the axis
but not necessarily within the light tube. We draw normals A and B
from the point E to the marginal and principal ray respectively. The
length of these normals is taken as positive when the normal lies above
the axis, negative under it. In this way, we have defined two coordinates
for the light tube. Two further coordinates can be obtained from the
angles u and w subtended by the marginal and principal rays with the
axis. The coordinates we choose are U = sin u and W = sin w. The signs
of the angles are not chosen in agreement with the convention of analytical
geometry but exactly the opposite. In Fig. 2, u is therefore negative,
w positive. The signs are assigned in this way to give a positive refracting
power for a converging lens. It is not advisable to change this convention
since considerable confusion would result.
o
p
Marginal ray

Principol/l!Y
o
Fig. 2

The four coordinates A, B, U and W determine the form of the light


tube; for the distance from the point E to the emitter 0 is given by A/U,
that from E to the receiver P by BjW. This is obvious from Fig. 2 since
the distance of E from the receiver P is the hypotenuse of a right-angled
triangle. The formula BjW - AjU = D therefore gives the distance
between emitter and receiver. The radii can be calculated from the
angles wand u.
Although they completely define the light tube, the four coordinates
are not themselves uniquely defined by its form: for if another reference
point is chosen for the construction of the normals, the lengths A and B
will be different. However, the expression D = B/W - AjU is indepen-
dent of the position of E since it defines the distance between emitter and
receiver. Furthermore, the angles and consequently the coordinates U
and Ware not dependent on the choice of E. Thus
-DUW=AW-BU
is an invariant quantity of the light tube.
This is not a coincidence: the coordinates A, B, U, W were, in fact,
chosen specifically with this invariant relationship in view.
Linear Conductivity 5

§ 3. Linear Conductivity
A Measure of the Influence of the Light Tube
on the Energy Transfer
Suppose now the emitter and receiver are set in operation. The energy
which is transferred through the light tube in a given time is the total
quantity of energy, a physically meaningful definition whose technical
unit is the Watt-Second. But we are normally interested in the energy
transferred per unit of time. This is measured in Watts and is termed the
energy flow or the radiating power.
To begin with, we operate emitter and receiver under "normal con-
ditions". This means, for both sound and light radiation, that the energy
transfer takes place in air. For radiation which is strongly absorbed in
air, the normal medium is a vacuum. We now change the sizes of emitter
and receiver without changing their other characteristics, and we also
change their separation. The measured radiating power is thereby changed.
In a second series of experiments we maintain the geometrical form
of the light tube. However, we replace the normal medium between
emitter and receiver by another material which does not absorb the
energy to be transferred but changes the three-dimensional energy density
in the light tube. For sound radiation this would imply a different gas.
A change of temperature or gas pressure has the same effect. For light
radiation one would choose a perfect optical glass.
From these experiments, the basis of which was given by J. H. LAM-
BERT (1728-1777), we can deduce that the radiating power (/J is pro-
portional to the square of the refractive index n of the medium used in
the light tube and to the square of the quantity A W - BU characteristic
of the light tube. Thus
(/J = c n 2 (A W - BU)2, c being the constant of proportionality.
It follows that if two light tubes of different form give the same
numerical value for A W - BU, they will transfer energy in the same
medium with the same radiating power. In different media, the radiating
powers are proportional to the squares of the corresponding refractive
indices. If the first light tube has a value A W - BU twice that of the
second light tube, it will transfer energy with four times the radiating
power.
The constant of proportionality c is a measure of the power of the
emitter and should strictly be given a specific name. But not all light
tubes are centred and it was not our intention to base our measuring
technique on this particular property. Instead, we shall use the expression
c/n2 and call it the radiation density. The denominator n 2 merely signifies
a change from circular to square emitters and receivers.
6 Perfect Imagery

The example of a glowing piece of metal demonstrates clearly the


significance of the term radiation density. A white hot metal source
radiates more energy from the same surface in the form of light and heat
than a source at dull red heat or lower temperature. There is no corre-
sponding theoretical measure for the power of the receiver, for it is a basic
requirement that it absorb the entire radiated power - otherwise our
measurements would be false.
The experimentally determined formula f/J = c n 2 (A W - B U)2 giving
the relationship between the radiating power, the refractive index of the
medium used and the coordinates of the light tube shows that the
expression n (A W - B U) is characteristic of the transfer properties of
the light tube. It is therefore useful to give this quantity a name. We shall
call it the linear conductivity. It is defined by the equation
LLW = AnW - BnU
the abbreviation LLW being derived from the German term "linearer
Leitwert".
The adjective "linear" implies that the square of this quantity appears
in the energy relation and avoids confusion with the "light conductivity"
']1;2 (A W - BU)2 of centred systems.
Before we establish the linear conductivity as the fundamental quan-
tity of geometrical optics, we must be sure that the experiments we have
indicated provide sufficient proof of its validity. In fact, from the point
of view of strict logic they prove nothing at all, for neither a perfect
emitter nor a perfect receiver is a practical possibility in a real instrument.
Furthermore, no measurement is free from error. It is possible that we
could prove with the same experiments that another expression, differing
only very slightly from the LLW, defines the transfer characteristics of
the light tube.
Consider an example: instead of taking the length of the perpendicular
from the reference point E to the marginal ray as the coordinate of the
light tube we could have taken the length of the perpendicular A to the
axis from the marginal ray through the point E. Similarly, B would be
the length of the perpendicular to the axis drawn from the principal ray
through the point E (Fig. 3). As further coordinates we would choose
U =tanu and W = tan w. The equationB/W -A/U =Dis perfectly valid
in this coordinate system for the distance between emitter and receiver;
furthermore, AW- BU is an invariant irrespective of the choice of the
point E. The relationship between the two sets of coordinates is given by
A = A cos u, B = If cos w, U = fJ cos u and W = tv cos w, that is

AW - BU = cos U'cos w (AW - BU).


Diffraction 7

If we wish to decide by experiment which coordinate system is better,


we must choose the angles u and w to be large enough for the product
cos U'cos w to depart significantly from 1. This leads us to the technical
problem of constructing a receiver which will measure accurately over a
o

D )1
Fig. 3

large surface and a large range of angles. The next section shows how
this difficulty can be avoided.

§ 4. Diffraction
A Limit for the Transfer of Energy through Light Tubes
Suppose we choose the point E for the coordinate system A, B, U
and W of the light tube to be in the centre of the emitter. Then A = 0 and
the linear conductivity assumes the special form LLW = - BnU.
If we reduce the diameter of the emitter, Band LLW become correspond-
ingly smaller. If we wish to leave LLW unchanged, we must increase U to
compensate for the reduction in B. But there are limits to this process;
for the value of U = sin u cannot become greater than 1. We can increase
the refractive index up to 2.5 (diamond), but after this there is nothing
we can do that will prevent the linear conductivity being made smaller
by further reduction in the size of the emitter.
The coordinate system A, if etc. mentioned in the previous paragraph
gives fundamentally different results: it does not matter how small we
make the emitter. We can always find an angle u which gives the desired
value for - BnU, since U = tan u tends to infinity when u approaches
a right angle.
If the point E is placed in the middle of the receiver, completely
analogous conclusions follow for the expressions AnW and AnW. We
thus arrive at the following conclusion. If the first coordinate system is
the right one for the light tube, there must exist a fundamental difference
between the transfer of energy from large and small emitters. The same
applies to receivers. If, on the other hand, the system with the barred
letters is correct, no such difference can occur.
8 Perfect Imagery

This very significant difference must enable us to decide experimen-


tally which supposition is false. In fact, the transfer characteristics of a
light tube do change fundamentally when one reduces emitters or
receivers from large to very small sizes. The existence of diffraction
phenomena proves this. We know today that it is quite impossible to
observe an atom with even the most powerful light microscope in the
way we can observe, say, our hand. Similarly, we know that it is impossible
to receive a sharp image of a landscape through a tiny pinhole. This
experiment can be tried by anyone by piercing a piece of card with a
needle and holding it close to the eye. Such experiments prove that the
coordinate system based on the barred letters is useless for the analysis
of the physical processes with which we are concerned. Nevertheless, we
have still not proved that A, B, U and Ware the correct coordinates.

§ 5. The Rayleigh Unit


A Measure of the Minimum Conductivity of a Light Tube
The simple rules for the energy transfer in a light tube break down if
emitter or receiver become very small and diffraction effects become
significant. It is not possible to define clearly the limits of their validity.
Special features of the emitter and the type of observation determine
whether diffraction effects are barely detectable or large enough to be
significant for the accuracy of the measurements. But we know that they
depend basically on the wavelength of the radiation and on the linear
conductivity of the light tube: that is, not only on the sizes of the emitter
and receiver but also on their separation. If we want to see something
clearly, we bring it as close as we can to our eye. The emitter 0, the object
we wish to observe, does not change its size. The size of the eye pupil,
which serves as the receiver P of the radiation, also remains effectively
unchanged. However, the linear conductivity is increased because of
the reduced separation: this gives the physical explanation of our ability
to see the object more clearly.
The influence of diffraction on the energy transfer becomes measurable
when the linear conductivity LLW becomes about as small as the wave-
length. For this reason, it is often useful to express LLW in wavelengths
as units rather than in millimeters or other linear measures.
If the wavelength is used for the unit of linear conductivity, it is
convenient to call it the Rayleigh unit or R. E., the abbreviated form
of the German "Rayleigh-Einheit". In principle we do not require a name,
but it is nevertheless useful as this term is only used for conductivities,
whereas multiples of wavelengths are also used for measuring other
quantities such as phase differences or the length of the standard meter.
Definition of Perfect Imagery 9

The name is given to this unit in honour of Lord Rayleigh. About


1880 he worked on the problem of the relationship between aberrations
and the performance of optical instruments. His work is of significance
to us because he derived the criterion for the influence of the aberrations
from the physical characteristics of the radiation, not from the geometry
of the rays. This sounded the death knell of "classical" geometrical optics
which ascribed physical significance to the rays and considered them to
be energy carriers.
Just how useless the "classical" theory is for the solution of physical
problems was recognized earlier by ERNST ABBE when CARL ZEISS pre-
sented him with the problem, about 1870, of designing a microscope
objective. If rays are energy carriers, he reflected, then it must be possible
to construct a microscope objective with a small angle u and small
receiver which is just as powerful as one with a large angle; for the loss
of intensity can be compensated by a better light source. When he
realized that this theory must be wrong, ABBE became aware of the
significance of diffraction and recognized the limits of performance of
optical instruments. His "sine condition" is today the reason for the
choice of the coordinates A, B, U and W for the light tube.

§ 6. Definition of Perfect Imagery


The Invariance of the Linear Conductivity as the
Fundamental Principle of Optical Theory
and the Rayleigh Unit as a Limit of its Validity
We can now use the basic ideas of the preceding sections to elaborate
the theory of perfect imagery for centred systems. The basic concept is
the light tube with the coordinates A, B, U and W defined in § 2. The
basic principle is the invariance of the linear conductivity LLW = AnW
- BnU. The quantity n in this formula is a pure number. We call it the
refractive index of the medium at the point E of the light tube, but the
theory is independent of a specific meaning of this number. The assump-
tion is made in the entire development of the theory that all media are
homogeneous and isotropic, that is, that the refractive index can only
change at the boundary between two media. This supposition greatly
simplifies the formulation and is fulfilled in all but a few special cases.
In § 2 we showed as an example that LLW does not change if the
point E is moved along the axis, implying that the refractive index does
not change. This shift of the point E is thus a process which fulfills the
fundamental principle of the theory. The purpose of the theory is to find
all the processes which leave LLW unchanged. In this way, we can build
10 Perfect Imagery

up a general picture of all the processes that are possible or even conceiv-
able in connexion with perfect imagery.
In order to avoid conflict with experiment while establishing our
theory of perfect imagery, we must restrict its range of validity to exclude
disturbing diffraction effects. Since there is no natural sharp boundary,
we shall define the limit of validity according to the requirement that the
value of the linear conductivity may not be less than a Rayleigh unit,
i.e. one wavelength.
The limiting performance of optical instruments is often defined by
the resolving power, a quarter of the Rayleigh unit. There is no con-
tradiction here with the arguments above since the resolving power is not
concerned with a correct description of the energy transfer; it merely
indicates when all details of the light tube (for example, form and size
of the emitter) vanish under the diffraction pattern. If LLW lies between
one Rayleigh unit and one quarter of a unit, we can still use the theory
of perfect imagery as a rough approximation. Below the resolving power
the conclusions of the theory become definitely false.
Experimental facts formed the basis for the introduction of the three
fundamental concepts: the light tube, the linear conductivity and the
Rayleigh unit. However, we still have no formal proof that they are
correct. The theory of perfect imagery makes use of these fundamental
concepts and arrives by means of logical developments at conclusions
which far exceed the experimental knowledge implied in those concepts
but which are amply confirmed by the behaviour of optical instruments.
In this way, the basic assumptions find their justification in the final
conclusions. Aberration theory shows that they are not exactly correct
but nevertheless of great practical value. The more exact formulation of
aberration theory is from the start highly complex; but the significance
of these complicated formulae only becomes clear if we maintain the
concept of perfect imagery as a guide line. The principles of aberration
theory also represent an approximation and give only a rough criterion
for the image quality.
The theory takes account of only one property of the medium, the
refractive index. It can therefore give no information at all about polari-
zation or the special optical effects occuring in crystals.
The following three sections give the abstract theory of perfect
imagery, in other words, the formal consequences of the above definition.
Many readers will find the formulae somewhat arid because their signi-
ficance cannot readily be grasped. The abstract form. has deliberately
been chosen in order to separate the pure theory from the applications
which follow.
The Fundamental Processes of Perfect Imagery 11

§ 7. The Fundamental Processes of Perfect Imagery


Shifts of Object and Pupil, Transfers and Refractions
In this section we shall attempt to answer the following questions.
What processes leave the linear conductivity of the light tube unchanged?
How can they be formally described and what physical processes do they
correspond to? We start by choosing two light tubes with a common
axis and reference points E and E*. The reference points may be the
same or different but the linear conductivities LLWand LLW* are taken
to be equal:
AnW - BnU = A*n*W* - B*n*U*. (7.1)
Introducing A*nW - B*nU into both sides and collecting terms:
0= (A*-A)nW + (n*W*-nW)A*-
- (B* - B)nU - (n*U* - nU)B* . (7.2)
If we suppose A, B, nU and nW to be known values, then (7.2) is an
equation with four unknowns A*, B*, n*U* and n*W*. The number of
solutions is infinite but one can identify a number of them quite easily
by trial.
We put A* = A and n*U* = nU. The formula (7.2) reduces to the
relation
(n*W* - nW)A = (B* - B)nU . (7.3)
If neither A nor nU is equal to zero, we can write
n*W*-nW B*-B
nU =-A-=n.
With the help of the proportionality factor n we can express the set
of solutions as
A* =A B*=B+nA
{
II n*U* = nU n*W* = nW + nnU . (7.4)
We also arrive at this form of solution for equation (7.3) if A or nU
is equal to zero. It is not possible for both to become zero at the same
time since LLW would. then also be zero, a value numerically less than
the Rayleigh unit. The case lies outside the range of validity of the theory
and has no physical meaning.
To give some understanding of the meaning of these equations, we
shall limit ourselves to the case for which n = n* and the reference points
E and E* coincide. (This is a severe limitation since the general case has
both theoretical and practical significance. However, it is not easy to
grasp this overall picture and very difficult to describe it in a correct
and readily understandable way). Furthermore, we shall place the refer-
ence points in the centre of the emitter, so that only the special case
12 Perfect Imagery

A = 0 is treated. This is not a serious limitation because the choice of a


different reference point does not affect the conductivity or the geometry
of the light tube. With these assumptions, all coordinates with and
without stars are the same except for W* = W + :rcU. Fig. 4 shows the
parts of the two light tubes lying above the axis.
The most obvious difference is the shift of the receiver and the
associated change in its size. Since the pupil of our eye is an excellent
example of a radiation receiver, the process described by a solution of the
set of equations II will be called a pupil shift. With the special assumptions

Fig. 4

made in the construction of Fig. 4 we can also ascribe a physical signi-


ficance to the parameter :rc, since the distances between emitter and
receiver are given by
B B*
-=D and -=D*.
W W'
But B B* so that
(1 1)
=

:rc= nOW' - nW =nB


- ---
nU nU D* D

is a measure of the change in the reciprocal of the separation between


emitter and receiver. In Fig. 4, D = 2D* and llD* - liD = liD. If an
optically equivalent pupil shift is made in the other direction, then the
equation llD* - liD = - liD must hold. This can only be true if
D* = 00, i.e. the receiver must be at 00, not twice as far away. It is
clear from this example that it is confusing to work with reciprocal dis-
tances. But it is not always possible to avoid this in geometrical optics.
The Fundamental Processes of Perfect Imagery 13

A special feature of the pupil shift must be emphasized: the chauge


in size of the emitter. The perpendiculars B aud B* drawn from the
centre of the emitter to the principal rays have the same lengths. For
this reason, the emitter appears to have the same size when viewed from
the centre of the receiver although the actual sizes of 0 aud 0* are
different if W aud W* are not equal. For comparison, the size of 0 is
shown by the dotted line in the lower half of Fig 4. The nearer we move
the receiver to the emitter, the larger must the emitter become if the
linear conductivity is held constant. This example illustrates very clearly
the difference between "classical" and modem conceptions of geometrical
optics. According to the former, the emitter was to be considered as a
real physical object possessing a definite size. Today we consider the
emitter merely as a concept which, together with the receiver, is useful
in determining the energy trausfer. The geometrically obvious quantities
of the classical theory have been sacrificed to the abstract concept of
energy.
In the theory of perfect imagery, the energy released from the emitter
o is independent of the size of the angle w aud is defined by the projection
B of 0 on to the centre of the receiver. Emitters with this special prop-
erty are called Lambert radiators. A piece of glowing tin approaches this
ideal case quite closely, but other emitters - for example, three-dimen-
sional radiators - have fundamentally different radiation characteristics
and the assumption that they are Lambert radiators is an approximation
only valid for small augles w.
Another set of solutions of equation (7.2) emerges if we set B* = B
and n*W* = nW. By a similar reduction the condition
(A*-A)nW= (n*U*-nU)B
gives the set of solutions

{ A* = A + wB B*=B
Q n*U* = nU + wnW n*W* = nW, (7.5)
to which we can give the physical significance of an "object shift". The
justification is analogous to the first case since we have simply reversed
the roles of emitter and receiver. It should be remembered that in the
theory of perfect imagery the assumption of Lambert radiation applies
equally to the receiver.
For our third set of solutions we shall make the assumption n*U*
= nU aud n*W* = nW giving

LI{ A*=A+bnU B*=B+bnW


n*U* = nU n*W* =nW. (7.6)
14 Perfect Imagery

The geometrical significances will be illustrated for a single light tube


for which we set n = n*. In this special case, the equations (7.6) define
the shift of the reference point E to E*. Fig. 5 shows the construction.
The dotted line makes it clear that the relation A * = A - dU is valid.
The minus sign results from the sign convention for the perpendiculars
and for the angles and from the basic premise that distances in the
direction of the light are considered positive. The corresponding relation
B* = B - dW applies for the other coordinates. It follows that the para-
meter Cl and the separation d are related by the expression d = - nCl.
This process of shift of the reference point has considerable practical
application in geometrical optics. For this reason, it is useful to give
it a shorter and more convenient name. We shall call it simply transfer.
We mean by this the transfer from one reference point to the next, in
practice from a given lens or refracting surface to the next one.
o

Fig. 5

The three sets of solutions object shift, pupil shift and transfer do
not introduce anything fundamentally new, for the first two processes
merely provided us with experimental evidence of the invariance of the
conductivity, while the third is geometrically obvious from the choice of
the coordinate system. But there are further solutions contained in the
equation (7.2). If we put A * = A and B* = B, we find the set of solutions
A* =A B*=B
{
(J> n*U* = nU + cpA n*W* = nW + cpB. (7.7)
This set describes those processes which we call refraction, since the
solution contains the classical lens formula. For if we divide the first
equation of the second row by A * = A we arrive at the relation
n*U* nU
A*'=A+CP' (7.8)

If we take A and A * to be the perpendiculars to the marginal rays from


the centre of a thin lens in air, then n = n* = 1, A*/U* = b, the image
distance, and A/U = - a, the object distance measured negatively from
the lens. Writing cp = l/f we deduce from (7.8) the well-known formula
Relationships between the Fundamental Processes 15

Ija + Ijb = Ijf. Fig. 6 shows the corresponding geometrical relationships.


Considered physically, this figure is incomplete because the corresponding
construction for the principal ray is not included and the form of the
light tube before and after refraction therefore remains undetermined.
Lens

~I~r~,
a ..I b )1
Fig. 6

§ 8. Relationships between the Fundamental Processes


Representation of Object and Pupil Shifts
by Transfer and Refraction
We are now familiar with four single parameter sets of solutions of
the equation LLW = LLW*. If we carry out two such processes, they
yield together a further process which leaves the conductivity unchanged
but changes the coordinates of the light tube. Such combinations of the
fundamental processes give a variety of processes, all of which fulfil the
condition LLW = LLW*. But the question remains open whether every
solution of this equation can be produced by combining the fundamental
processes. It will be shown in § 9 that this is indeed the case, but we shall
deal first - since the proof is then more straightforward - with another
problem, i.e. whether the four fundamental processes are independent.
In fact, they are not. Two theorems apply here, the first being as follows:-
Theorem 1. Every object shift Q and every pupil shift II can be
produced by combining transfers L1 with a refraction rJJ.
At first sight, this theorem is surprising from an algebraic point of
view; for, when combined with the second theorem given later, it implies
that the equation LLW = LLW* has a maximum of two independent
solutions. In general, one expects three independent solutions in an
equation with four unknowns. From a physical point of view the theorem
is trivial, for a lens in the same plane as the receiver does not affect the
receiver and allows any desired object shift. Similarly, a so-called field
lens placed at the emitter can produce any desired pupil shift and has no
effect on the position and size of the emitter. If the emitter lies at infinity
a pupil shift is equivalent to a transfer to another reference point. A cor-
responding significance can be attached to an object shift in the case of
a receiver at infinity. These relationships can be proved as follows.
16 Perfect Imagery

We combine a transfer Ll j, a refraction (]J and a second transfer Ll 2 •


This results in the formulae: -
1. Transfer Llj:
A'=A+<'ljnU
n'U' =nU
B'=B+<'ljnW
n'W'=nW.

2. Transfer Llj combined with refraction (]J:


A"=A' =A +<'lj nU
n"U" = n'U' + IPA' = nU + IP(A + <'lj nUl
B"=B' =B+<'ljnW
n"W" = n'W' + IPB' = nW + IP(B + <'lj nW) .
3. Transfer Llj and refraction (]J combined with transfer Ll 2 :
A'" = A" + <'l2n"U" = A + <'ljnU + <'l2nU + <'l21P (A + <'ljnU)
n"'U'" = n"U" = nU + 17' (A + <'ljnU)
B'" = B" + <'l2n"W" = B + <'ljnW + <'l2nW + <'l21P (B + <'ljnW)
n"'W" 1 = n"W" = nW + 17' (B + <'ljnW) . (8.1)

If these relations describe an object shift Q, the following equations


must hold:-
A'" = A* = A + wB
n"'U'" = n*U* = nU + wnW
B'" = B* = B
n'"W''' = n*W* = nl¥ . (8.2)

Setting these equations equal to the corresponding ones of (8.1), we


can determine the parameters <'lj' (j2 and 17'. If nW =1= 0, the last equation
nW = nW + 17' (B + (jj nW) is satisfied for (jj = -~.
nW
+ (jj nW +
:w.
The last equation but one gives the relation B = B (j2 nW.
This is satisfied by (j2 = - (jj or (j2 = With the given value for <'lj
the second equation
nU+wnw=nU+IP(A- ~~)
yields the solution for 17' as 17' = .
w i~~2 Substitution in the first equa-
Summary of all the Processes of Perfect Imagery 17

tion shows that it is also satisfied. This procedure has enabled us to


determine for nW =1= 0 the transfers and the refraction whose combination
represents the given object shift.
If nW = 0, then 61 = ()) n~ , 62 = 0 and rp = 0 represent a solution
of the set of equations (8.2). This can be checked quite simply by sub-
stitution in (8.1). The denominator nU cannot be zero, otherwise LLW
would be zero.
If the equations (8.1) define a pupil shift, then the following equations
must also hold:-
A'" = A
n"'U'" = nU
B'" = B +:n:A
n"'W'" = nW +:n: nU.
Following the example above we find the solutions
A A (nUl2
61 = - -;u , (l2 = - 61 = nU ,rp = -:n: LL W ' if nU = 0
and
if nU = O.

We have now proved everything stated in Theorem 1. It is also


possible to show that any transfer and any refraction can be produced
by a combination of object and pupil shifts. But this theorem has no
physical significance.

§ 9. Summary of all the Processes of Perfect Imagery


Representation by Transfers and Refractions
Theorem 2. If LLW =1= 0, then, with the aid of the four coefficients

A*nW - B*nU B*A - A*B


a= LLW b= LLW '

n*U*nW - n*W*nU n*W*A - n'U*B


C = -----;;L~L:-::W=--- d= LLW (9.1)

every solution of the equation LLW = LLW* can be represented in the


form:
A*= aA + bnU B* = aB+ bnW
n*U* = cA + dnU n*W* = cB + dnW. (9.2)
2 Zimmer, Geometrical Optics
18 Perfect Imagery

The relation
ad-bc=l (9.3)
exists between the coefficients.
In order to prove this, we will suppose that the quantities A, B, nU,
n Wand A *, B*, n*U*, n*W* are solutions of the equation LLW = LLW*
and that LLW = AnW - BnU ¥= O. We can then calculate the coeffi-
cients a, b, c, d from equations (9.1). The simple reduction
A*'nWA - B*nUA + B*AnU - A*BnU
aA +bnU= LLW

= A*AnW - BnU = A*
LLW
shows that the first equation of (9.2) is satisfied. It is just as easy to
verify the other three equations.
The reduction necessary for the proof of the relation (9.3) is somewhat
longer:
1
ad - be = LLW2 [(A*nW - B*nU)·(n*W*A - n*U*B)-
- (B*A - A*B)·(n*U*nW - n*W*nU)]
1
= LLW2 [A*nWn*W*A - B*n*W*AnU - A*n*U*BnW +
+ B*n*U*BnU - B*n*U*AnW + A*n*U*BnW +
+ B*n*W*AnU - A*n*W*BnU].
The second and seventh terms and the third and sixth terms in the
square brackets cancel out and the remaining terms can be rearranged
to give
(A*n*W* - B*n*U*) (AnW - BnU) = LLW*·LLW = LLW2,
which proves relation (9.3).
With the help of theorems 1 and 2 we can now prove the
Principal theorem of the theory of perfect imagery. Every solution of
the equation LLW = LLW* can be obtained from the combination of
refractions (J) and transfers Ll. Of course, the assumption is made here that
the linear conductivity is not equal to zero, since this is one of the funda-
mental principles of the theory.
The proof can be derived from the set of equations (8.1) which can
be expressed in the following form with A, B, nU and nW as factors:-
Alii = (1 + rp (l2)A + ((11 + (12 + (11 rp (l2)nU
nlllU'" = rp A + (1 + (11 rp)nU
Bill = (1 + rp (l2)B + ((11 + (12 + (11 rp (l2)nW (9.4)
n'"WIII = rp B + (1 + (11 rp)nW .
Summary of all the Processes of Perfect Imagery 19

These equations have the same form as (9.2) and the coefficients
satisfy the relation (9.3):

(1 + rp 152 ) (1 + 151 rp) - rp (151 +152 + 151 rp 152)


= 1 + rp 152 + 151 rp + 151 r 152 - rp 151 - rp 152 - 151 r 15 = 1 .
2

The last relation is self evident without algebraic proof, since transfers
and refractions are solutions of the equation LLW = LLW*.
It now only remains to be proved that 151, 152 and rp can be determined
in such a way that the coefficients assume definite values a, b, c, d as
given by (9.1). These equations are

b = 151 + 152 + 151 rp 152


c = rp d = 1 + 151 rp.
d-l a-I
If c =? 0, then 151 = -c-' 152 = -c- and rp = e are solutions of the
first, third and fourth equations. The second is also satisfied. It can be
very simply checked and follows formally from the condition (9.3). The
principal theorem is thus proved for the special case c =? o.
A physical interpretation of the above result is given by the well
known Theorem oj Gauss: - If the refracting power rp of an optical system
is not zero, there exist two clearly defined reference points Hand H*
in the axis of the system. These are called the principal points of the
system and can be determined from any given reference points by trans-
d-l I-a
fers with the parameters 15 = -- and 15* = -- respectively. If the
rp rp
coordinates of the light tube in front of the system are referred to point H,
those behind the system to point H*, the transfer parameters 151 and 15 2
given above become zero and the effect of the system can be expressed
as a refraction with the parameter rp. Such systems can thus be fully
defined by the two principal points and one other quantity, the usual
choice being the focallengthj = IJrp. The principal points are defined by
the condition a = d = 1 for the coefficients a and d. Since ad - be = 1
and e = rp =? 0 it follows that b = 0 in all these cases. Equations (9.1) then
give the relation A * = A and B* = B as a geometrical characteristic of
the principal points.
The case e = 0 occurs in telescopic systems, of which binoculars and
opera glasses are the best known examples. Their function can be describ-
ed by two refractions - in the objective and in the eyepiece - and one
transfer, which corresponds to the distance between them. To prove this
formally, we transform from the given values A, B, nU, nW by means of
2*
20 Perfect Imagery

a refraction determined by a parameter rp, which can have any value


except zero, to the coordinates
A=A B=B
nU = 1~U +rpA nW=nW +rpB.
We re-write these equations in the form
A=A B= B
nU=nU -rpX nW=nW -rp B
and substitute in (9.2). This gives the set of equations
A* = (a - brp)A + bnU B* = (a - brp)B + bnW
n*U* = (c - drp)X + anU n*W* = (c - d rp)B + dnW .
The coefficient c= c - d rp in these equations cannot be equal to zero
since the premise was made that c = 0, rp =1= 0; furthermore d =1= 0 because
ad - be = 1. According to the first part of the proof, therefore, transfers
and a refraction must exist with the parameters
d-l a-bq;-l
61 = _ dq>' 62 = _ dq; , fj! =- d rp .

The combination of the transfers and the refraction expressed by (9.4)


transforms from the system with bars to the system with stars. This
means that the combination of the refraction with the parameter rp, the
transfer with 15 1 , the refraction with fj! and the transfer with 152 transform
the starting system A, B, nU, nW into the final system A*, B*, n*U*,
n*W*. This provides a complete proof of the principal theorem.
The physical interpretation of the principal theorem is that any
transformation of a light tube for which the linear conductivity remains
unchanged can be achieved by means of refractions and transfers, in
other words, by means of a perfect lens system. It is the task of the optical
designer to calculate the parameters of such systems and to furnish the
necessary information for their technical realisation. The first process
is simple since a complete theory exists but the second process is an art
because the problem can be rigorously solved only in special cases. It may
be impossible to achieve the required refraction by means of a lens with
a given diameter; furthermore, all lenses possess the inevitable defect
that the refracting powers at the edge and the centre are not the same.
The second part of this book indicates some aspects of the designer's art
by means of which these defects are reduced to sufficiently small values.
A simple consequence of Theorem 2 which is useful in practice is
Theorem 3. The law of commutation is valid for perfect imagery and
any object shift Q or any pupil shift II.
The Eye 21

This theorem has the following significance: -


If the image of a light tube in the object space is formed by a perfect
lens system as a light tube in the image space, and an object shift Q is
then performed in this space, this combination of (first) perfect imagery
and (second) object shift gives the same result as (first) the object shift
in the object space followed by the formation of an image of the shifted
object space in the image space.
The proof is very simple, for from (9.2) we have
A* + wB* = aA + bnU + w(aB + bnW)
= a·(A + wB) + b·(nU + wnW) ,
n*U* + wn*W* = cA + dnU + w(cB + dnW)
= c·(A + wB) + d·(nU + wnW).
The proof for the pupil shift is completely analogous.

§ 10. The Eye


The Basic Data of the Human Eye necessary
in Geometrical Optics
Physiological optics is concerned with the structure and function of
the eye. We shall confine ourselves here to a few aspects of experimental
and theoretical work in this field. For complete accounts the reader is
referred to H. v. HELMHOLTZ: Handbuch der Physiologischen Optik
(1866), W. TRENDELENBURG: Der Gesichtssinn (1943) and H. SCHOBER:
Das Sehen (2 volumes, 1957 and 1958), J. P. SOUTHALL: An Introduction
to Physiological Optics (1937).
The eye gives us a sensory impression of radiation in the form of
electromagnetic waves with wavelengths between 400 nm and 700 nm
which we term light. A nanometer (nm) is a millionth of a millimeter.
These wavelengths are therefore extremely small. The perception of light
is normally linked with the perception of colour. A rough indication of
the colours which correspond to various wavelength regions is given in
the following table:
Table 1
Wavelength Colour
innrn

400 to 440 violet


440 to 495 blue
495 to 570 green
570 to 630 yellow and orange
630 to 700 red
22 Perfect Imagery

This table should be read from left to right, not the reverse. The
colour sensation yellow can also occur as a "complementary colour" if the
region around 480 nm is absent from the visible range of radiation. We
shall leave the details of the theory of colour vision to the physiologists,
painters and Goethe and confine ourselves to purely physical aspects.
The eye is markedly different from a physical radiation receiver. It
does not register the quantity of radiation entering it from outside or
the average value of the energy flow. It compares brightness, by which
we mean radiation density, values of which are measured in units which
are not independent of the wavelength. This is necessary because the
sensitivity of the eye is a function of wavelength. But there are consider-
able variations between the eyes of different observers. An international

to V(.t}

O+,~~~~~~~~~~l?
400 500 600 700 nm
violet· blue· green ·yel/ow· red
Fig. 7

standard brightness sensitivity curve has been established from average


values (Fig. 7).
The spectral radiation density Be(J.) must be multiplied by the factor
V(J.) to obtain the corresponding brightness. Integration over the wave-
length range from 380 to 720 nm then gives the required brightness when
multiplied by a constant Km:
720

B = Km SBe (J.). V(J.) bJ. .


380

B = brightness measured in stilb (sb),


B. = spectral radiation density measured in watt per cm2 and per
unit solid angle (sr),
sb
Km = 680 W'cm-2 sr'
The Eye 23

The brightness of a black body at 2042 oK, the solidifying point of


platinum, is 60 sb and is taken as normal brightness. Other examples
are given in the following table: -
Table 2
Light source Approximate
value of
brightness in sb

Fluorescent lamp 0.5


Incaudescent lamp 500 to 1200
Carbon arc 16000
Sun 150000
High pressure mercury
lamp 170000

A further unit in use is the apostilb (asb) defined by 31416 asb = 1 sb.
It is practical for defining the brightness of diffusely reflecting sources of
finite area or for twilight or night vision; for the human eye can adapt
itself to the large range of brightness from 20 sb down to 10-10 sb, i.e.
down to about 3.10-6 asb. Everyone of us has noticed that it is possible
to look at a fluorescent lamp but not at the filament of an incandescent
lamp, and certainly not at the sun. We shall content ourselves here with
these brief comments on photometry and light measurements as distinct
from physical radiation measurements.
In the physical description of vision, the observed object is considered
to be the emitter. The receiver is the hole in the iris which we call the
"pupil". Its apparent radius seen from outside lies between 1 mm and
4 mm. A typical mean value for the pupil radius in daylight vision is
1.5 mm. What size must the emitter have in order that the linear conduc-
tivity of the light tube be just one Rayleigh unit? The wavelength is taken
to be 554 nm = 0.000554 mm since the eye is particularly sensitive to
radiation of this wavelength. The emitter is taken to be at infinity so
that its size is defined by the angle w. The smallest value for geometrical
imagery is then described by the relation
LLW= 1 R.E.
In the plane of the pupil B = 0, in air n = 1, so that
AW = 0.000554 mm
W = 0.000554 mm = 0.00037 .
1.5mm
This corresponds to an angle of about 1.3', a value agreeing well with
experimental results which show that we can recognize details with an
angular diameter of 2.6' when suitably illuminated. This can readily be
24 Perfect Imagery

verified. A slide rule is observed at a distance of about one metre. For


an observer with normal vision the millimeter divisions can be read off
but the logarithmic subdivisions become blurred at those points where
the separation reduces to about 0.5 mm. At a distance of one metre,
an object diameter of 0.74 mm corresponds to an angular diameter of 2.6',
while at a normal observing distance of 250 mm this angle represents an
object diameter of about 0.2 mm.
The resolving power of the eye is, according to the above calculation,
about a quarter of 1.3' or about 20". This corresponds to a length of
0.09 mm perpendicular to the line of sight at a distance of 1 metre or to
a length of 0.02 mm at a distance of 250 mm. A hair is about 0.04 mm
thick. It can readily be seen against a bright background at a distance
of more than 1 m. In this case, the eye can resolve more than the calcu-
lation predicts. However, the form of the object has considerable influ-
ence, in this case its length. Two adjacent points can usually only be seen
separately if the angular distance between them is about 1.5', several
times the resolving power predicted. This inadequate point resolution of
the eye is important in daily life. For example, a square grid containing
6 points per millimetre gives the impression at 250 mm of being a uniform
grey surface. This is made use of in art printing. Even the much coarser
gratings of newspaper pictures or the line pattern on a television tube are
adequate to give the impression of a picture.
These examples show that, under normal conditions, there is no serious
error in applying the rules of geometrical optics to the eye. The anatomical
and physiological aspects only become significant under extreme demands.
In such cases, they can become more important than the physical effects.
For example, a physical interpretation would suggest that half millimetre
divisions must be recognizable at a distance of one meter if the pupil
radius is 3 mm. But if we wait until it is dark and the pupil has opened up
to this size, we find that we can no longer distinguish one millimetre
divisions. This is because in the process of dark adaptation the eye
changes its function.
Strictly speaking, we cannot ascribe a linear conductivity to the eye
since it is not a centred system. In the process of vision the eye rotates
continually about the eye rotation point which is approximately in
the middle of the eye ball. The position of the pupil thus changes with
the direction of view. However, many centred optical systems are used in
combination with the eye so that it is a useful simplification to consider
the eye as centred. This assumption implies that we can cover our field of
view with a fixed eye position. How large is, in fact, the field of view?
Visual impressions such as the perception of brightness or movement are
possible over a very large angular field exceeding 180 0 in the horizontal
direction. But the range over which sharp vision with both eyes is possible
Spectacles and Magnifying Glasses 25

is much smaller, the limit being an angular field of about 30° on each side
of the direction of view.
If the field of view in an optical instrument is restricted to less than
20° on either side of the axis, the observer has the feeling that he is
looking through a tube. But it is worthwhile to construct eyepieces with
a field of more than 30° to the axis because we are used to having a
region of sharp vision surrounded by a larger field in which structures
and movements can be perceived.
For the schematic eye defined in this way with a fixed head position
we have nW = l'sin 30° = 0.5. If we place the reference point E in the
middle of the pupil, then B = 0 and A = 1.5. It follows that the linear
conductivity is given by
LLW = AnW - BnU = 1.5 x 0.5 mm = 0.75 mm .
This value is important when the question is raised whether the light
tube of an instrument can be adapted to the eye.

§ 11. Spectacles and Magnifying Glasses


The Simplest Aids to Vision
We can only detect differences in brightness if the emitter is placed
at a sufficient distance from the eye. An adult with normal vision cannot
see the end of his nose clearly; a myopic adult has a blurred picture of
distant objects. The so-called range of accommodation over which one
can see clearly varies from person to person and also depends on age.
The normal range is taken to lie between 250 mm and infinity. This is an
arbitrary definition which applies to a person with normal vision at the
age of about 45.
If we wish to see objects clearly which are not within the range of
accomodation, we must bring them within this range. Furthermore,
we must change their size according to the change in distance if the eye
is to see them under the correct angle w. This means that we must
operate an object shift as explained in § 7. This is rarely possible and
fortunately not necessary because any object shift can be achieved by
appropriate refracting power in the plane of the pupil. This follows from
Theorem I for the special case B = O.
Spectacles and magnifying glasses are technical devices for doing
this by means of lenses or lens systems. They differ from one another in
design and range of application but not in the principle of their function.
For anatomical reasons the lens cannot be placed in the plane of the pupil
itself: it is placed in front of it. In this position, the effect is not a pure
object shift. But the additional effects only become significant when the
26 Perfect Imagery

distance of the lens from the eye becomes comparable with the distance
of the object. This particular point will not be discussed further here.
The object shift caused by a spectacle lens or a magnifying glass is
determined by the refracting power rp. It has the dimensions of a reci-
procallength and can therefore be expressed as mm-i or m-i . The latter
unit is called a dioptre (dpt). Another convention is to use the reciprocal
of the refracting power which is called the focal length f and is expressed in
mm or m. A lens or lens group with the refracting power rp = 0.01 mm-i
1
= 10 m-i = 10 dpt has a focallengthf= 0.01 mm = 100 mm = 0.1 m.
For some purposes the first system is more practical, for others the second.
It is easier to picture a focal length but calculations can be carried out
better with refracting powers.
What is the magnifying power of a spectacle lens or a magnifying
glass? Strictly speaking, there is no magnification since the lens should
really be placed in the plane of the pupil. If this were possible, the eye
would see the object under exactly the same angle with and without the
lens. This is made clear by Fig. 4 if the roles of emitter and receiver are
reversed. 0 and 0* are to be considered as the eye pupil. An object at p*
cannot be seen clearly since it is too close to the eye. An object shift
brings it to P in the range of accommodation of the eye.
The definition of the magnifying power N for a magnifying glass is
based on a convention. It is
250mm
N = - - , - = 250 mm·rp

in which f and rp are the focal length and refracting power respectively
of the magnifying glass.
The basis of this definition is the relation proved in § 8 between an
object shift Q and a refraction if> in the plane of the receiver:
(nW)2
rp= w LLW ·
In the special case of the eye pupil we have n = 1 and B = 0 so that
W
rp=w-X.
For the normal range of accommodation we have an object shift of
u* - u A (1 1) A
Wo = -W-- = IV 250 mm - -;;;; = W. 250 mm·
Th·· hi d b h
IS IS ac eve yt e
refracting power
W 1
rpo = Wo -X = 250 mm = 4 dpt
and is produced in the eye by a change in shape of the eye lens situated
immediately behind the pupil.
The Telescope and the Microscope 27

A magnifying glass with a magnification of n-times thus produces an


object shift of n times the aperture angle corresponding to the normal
range of accommodation. With a magnifying glass of unit power, a person
with normal vision can see clearly in the range from 125 mm to 250 mm
in front of the eye; with a magnifying power of 10 x the range of sharp
vision is 22.73 mm to 25 mm. If eye defects are present or if the range of
accommodation is abnormal, the position and length of the range of
accommodation in which one can see clearly with a spectacle lens or
magnifying glass are changed. A short-sighted person can see distant
objects with a - 2 x (dispersive) magnifying glass, whereas such a lens
is useless to a person of normal vision. He could only use it to observe
virtual objects in the range - 250 mm to - 125 rom which appear to lie
behind his pupil.

§ 12. The Telescope and the Microscope


Instruments which Increase Resolving Power
Spectacles and magnifying glasses are aids to clear vision outside the
range of accommodation. They do not change the position or size of the
pupil and consequently do not influence the angular size W of the smallest
perceptible object. Appreciably finer detail can only be made visible by a

0'
"I ...... ...... ,
p
I ' ,
...... ,0·

L2
'--v--/
LT Eyepiece
Objective
Fig. 8

corresponding increase in pupil size. For the naked eye, an increase in the
diameter of the pupil beyond 3 mm gives no improvement in the resolution
of fine detail. The reason for this lies in the construction and function of
the eye. This difficulty is overcome by the use of an optical instrument,
a telescope, which produces an enlarged image of the eye pupil. This
enlarged image serves as receiver for the radiation. The principle of such
an instrument is shown in Fig. 8.
28 Perfect Imagery

The lens Ll is the receiver P of the radiation emanating from the


object O. It can have a very much larger diameter than the eye pupil.
Large reflecting telescopes possess mirrors as pupils which are several
meters in diameter. The "objective" Ll functions as a magnifying glass
and shifts the object from 0 to 0'. Of course, it is immaterial whether
this shift is achieved by means of a lens or a mirror. In the plane of the
intermediate image 0' there is a "field lens" L2. This shifts the pupil
from P to P*. The refracting power of the field lens must be such that
the diameter of p* is roughly equal to the diameter of the eye pupil.
At P*, the plane of the eye pupil, there is a second magnifying glass
L3 whose purpose is to bring the intermediate image 0' into the range
of accommodation by an object shift to 0*. The field lens L2 and the
magnifier L3 are constructed as a unit which is called the eyepiece. In
practice, it is important to leave a gap between the last lens of the
eyepiece and the eye pupil so that the movement of the eyelids is not im-
peded. Wearers of spectacles need an even larger gap since eyepieces do not,
in general, have an adjustment to compensate for eyesight defects. This
leads to technical difficulties which have been deliberately ignored in
Fig. 8 in order to make the principle of the telescope easier to understand.
The telescope magnification r is defined as the ratio of the perpendi-
culars A and A * in the plane of the "entrance pupil" P and the "exit
pupil" P*. By virtue of B = 0 and B* = 0 for the pupils and since the
refractions and transfers leave the linear conductivity unchanged, we have
AnW = A *n*W*. It follows that the telescope magnification is given by
the relation
A nOW"
r= A* = nW . (12.1)
If the length of the perpendicular in the exit pupil is a fifth of that in
the entrance pupil, the angle of view in air presented to the eye is five
times greater than the field angle W in front of the instrument. The
object therefore appears linearly magnified five times although it is
the same distance away and fine detail can be more readily observed than
with the naked eye. The fact that the object appears to be nearer to the
eye is an illusion based on our experience of apparent size in relation to
distance.
Hand telescopes constructed according to the principle of Fig. 8 can
be specified by their magnification and the diameter of the objective
in mm. An 8 x 30 binocular has a telescope magnification of 8 x and a
clear diameter of the objective of 30 mm. If the instrument is correctly
designed, this is also the diameter of the entrance pupil. Since the object
is usually so far away that the angle u is only a few degrees, the length of
the perpendicular A is effectively the same as the radius of the pupil, in
this example 15 mm. The perpendicular A * in the exit pupil is then
The Telescope and the Microscope 29

15 rnm/8 = 1.875 mm. This is rather more than the radius 1.5 mm of
the eye pupil in daylight. This excess is deliberate and takes account of
vision in twilight. The angle of view of the eye W* = 0.5 = sin 30°
corresponds to a field angle of
W = 0.5/8 = 0.0625 = sin 3.5°
for the instrument. We can thus only observe an angular field of 7° with
this telescope but can recognize fine detail of angular diameter one eighth
of 2.6', i.e. about 20".
The simple opera glass has a design which is basically different from
that shown in Fig. 8. Its objective diameter is not the determining factor
in its resolution. We shall deal with this instrument in § 18.
The basic principle in the design of the telescope, as shown in Fig. 8,
is also used in the microscope, but the objective is specially designed for
the observation of very near objects. Suppose a magnifying glass of
strong refracting power is placed in front of the telescope. Compared with
a magnifying glass directly in front of the eye, this system has the
advantage that the dimensions of the lenses are technically feasible from
a manufacturing point of view. For a given desired object shift w we would
need a magnifying glass with refracting power rp* = w : : if it were
placed directly in front of the eye, but placed in front of the telescope
with a magnification of k times we only need the refracting power rp = w ~

= w ~; ~k = -Iz rp* . The telescope magnification of a microscope depends


on the choice of the eyepiece and normally varies between 5 times and
20 times.
The basic data for optical imagery in the microscope are given separa-
tely for the objective and the eyepiece because it is convenient to be able
to change objective and eyepiece independently as desired. The eyepiece
is specified by the magnifying power given in § 11, whereas the objective
is specified by the magnification ratio f3 and the numerical aperture nU at
the object. The magnification ratio is defined as the ratio of the perpendi-
culars B' and B in the planes of the intermediate image 0' and the
object O. It is a function of the position of the intermediate image 0' in
the instrument. This plane is often determined by a given tube length
which indicates the separation between the objective and the eyepiece.
Correct use of the instrument requires that this separation be kept con-
stant. It follows from the invariance of the linear conductivity that
nUB = n'U'B'. The magnification of the objective is therefore given by

(12.2)
30 Perfect Imagery

The total magnification of the microscope is given by the product {J , N


of the magnification of the objective (for the correct tube length) and the
magnifying power of the eyepiece.
It is possible to calculate the size A * of the exit pupil from the data
of the microscope if we suppose that the eyepiece shifts the intermediate
image to infinity, for then n*U* = O. The design of eyepieces is based
on this assumption. Equation (9.1) gives the relation
n*U*n'W' - n*W*n'U' n'U'
If = LLW = - -A-*-

for the refracting power of the eyepiece. With the definition 250 mm'lf = N
for its magnifying power it follows that
A* = _ n'U'·250mm
N
and from (12.2)
A* = _ nU·250mm
(12.3)
p·N .
If, for example, we take an objective 40/0.65 and a 10 x eyepiece,
then {J = - 40, nU = 0.65 and N = 10 so that
A* = 0.65·250mm = 041
40.10 . mm.

The exit pupil of this microscope is much smaller than the eye pupil
in daylight vision. The resolving power is therefore only limited by the
design of the instrument and not by the eye. Microscopists usually choose
total magnifications of the instrument between 500 and 1000 times the
numerical aperture at the object. The corresponding values for the pupil
radius A * are from 0.5 mm to 0.25 mm, Le. from one third to one sixth of
the radius of the daylight eye pupil. Smaller values than this are not
advisable since the natural inhomogeneities of the eye become noticeable.
Larger pupils P* are only possible if the total magnification is reduced;
but this reduces the angular size of the image since the linear conductivity
of modem microscopes seldom exceeds - n'U'B' = 0.03 x 15 mm =
0.45 mm. The effect of increasing the exit pupil of the instrument to the
size of the daylight eye pupil is to reduce the useful visual field angle W*
below the value 0.3 = sin 17.5° with the result that the observer has the
feeling that he is looking into a tube.
If, on the other hand, the exit pupil of the instrument is smaller than
the eye pupil, extended objects which can also be seen with the naked
eye appear darker in the microscope. There are two quite different
reasons why this must be so. Firstly, radiated energy is absorbed in the
glass and reflected at the lens surfaces. These effects do not really belong
to the field of geometrical optics and can be substantially reduced bv the
The Telescope and the Microscope 31

choice of suitable glasses and anti-reflexion films. But even if they could
be entirely removed there still remains a loss in brightness. This second
effect arises from the fact that the radiated power per uuit of angle of
vision registered by the eye at constant radiation density is proportional
to the square of the pupil aperture.
The conditions governing the brightness of very small objects which
cannot be resolved in the instrument are fundamentally different. The
rules of geometrical optics break down here. Instead of an image, we see
a diffraction pattern whose size depends only upon the effective pupil
diameter. If the exit pupil of the instrument is not smaller than the eye
pupil, the diffraction pattern seen by means of the instrument has the
same size as that seen by the eye alone. Given a telescope magnification
r = k, the area of the entrance pupil of the telescope is k2 times that of
the eye pupil. But the radiated power received is proportional to the area
of the entrance pupil so that the diffraction pattern in the telescope
appears k 2 times brighter than it does to the naked eye.
If brightness losses through absorption and reflexion are neglected,
the sky appears to have the same brightness through an 8 x 30 binocular
as it does to the naked eye with a pupil of not more than 1.875 mm radius.
But the stars are 82 = 64 times as bright and can thus be much more
readily seen than with the naked eye.
If the radius of the eye pupil is a-times larger than that of the exit
pupil of the instrument, then the diffraction pattern seen in the instru-
ment is a-times larger than that seen directly by the eye. The radiated
power received by the entrance pupil of the instrument is distributed
over a larger area. Theoretically, the same loss of brightness occurs as
for extended objects. In fact, this can only be observed when the radius
of the exit pupil has been reduced to about a quarter or a fifth of that of
the eye pupil. It is not possible to give an exact value as the effect depends
on the physiological nature of the eye. It is used to increase the contrast
between diffraction objects and background so that details of low inten-
sity can be more readily seen. This is the reason why it is usual in astro-
nomical and microscopical observations to work with "excess magnifica-
tion", that is, with an exit pupil smaller than the eye pupil.
The resolving power of the microscope at the plane of the object is
given by BnU = 0.25 R.E. With a fixed wavelength, as nU is increased,
smaller and smaller objects can be resolved. U = sin u cannot exceed 1;
in practice it rarely exceeds 0.9. The resolution is improved if the object
is not observed in air (n = 1) but immersed in a medium of higher refrac-
tive index. The commonest immersion liquids are water (n = 1.33) or
special oils (n = 1.52). Monobromnaphthalene (n = 1.68) and methylene
iodide (n = 1.74) are only used in special cases.
32 Perfect Imagery

With a numerical aperture nU = 1.3, the smallest resolvable radius B


of the object for a wavelength of 554 nm is
B = 0.25'0.000554 mm = 0.000107 mm •
1.3
Objects with a radius four times this value, that is a diameter of
0.00085 mm are imaged according to the rules of geometrical optics. Such
objects appear to the eye to subtend an angular diameter of 2.6' if the
exit pupil of the microscope has a radius A * = 1.5 mm, or a correspond-
ingly larger angle the smaller A * becomes.
Since the resolving power is proportional to the wavelength, finer
detail can theoretically be seen if radiation of a shorter wavelength is
used. In practice, this solution is not as easy as it seems at first sight on
account of the very low sensitivity of the eye at wavelengths below 400nm.
However, the application of suitable receivers or image converters
enables one to use the improved resolution in the ultra-violet region down
to 200 nm. However, the lenses can no longer be made of glass since it
strongly absorbs radiation with wavelengths less than 360 nm.
A much better resolution is provided by the electron microscope in
which the radiated energy is transferred by electrons. If they are accele-
rated to energies corresponding to 50000 volts, they possess a wavelength
of about 0.0055 nm = 0.0000000055 mm. This is a hundred thousandth
of the wavelength 550 nm for which the eye has maximum sensitivity.
Nevertheless, the resolution in the electron microscope is not, in practice,
one hundred thousand times better. For, although it is possible to produce
refracting power for electron radiation with electric or magnetic fields,
it is not possible to achieve the perfection of design that can be attained
with lens systems. For this reason, the numerical aperture is usually
restricted to nU = 0.005 or less. With this value, we have from the
formula BnU = 0.25 R.E. a theoretical resolving power of B = 0.0000003
mm = 0.3 nm. This order of magnitude has indeed been reached. It is
about 350 times the resolution of the normal microscope.

§ 13. The Photographic Emulsion


Exposure to radiated energy produces chemical changes in certain
substances. This property is made use of in photography. Silver bromide
decomposes on exposure to light into silver and bromine. We shall not
discuss here the physical, chemical and technical problems of film
manufacture, nor the processes of development, fixing and printing which
render the photographic image visible. The reader who is interested in
photography should refer for details to the many popular accounts or
technical works on the subject.
The Photographic Emulsion 33

The diameter of a grain of silver bromide lies between 0.0003 mm


and 0.003 mm for normal emulsions, depending on the purpose for which
the film is to be used. For simplicity, we shall suppose a radius A of
0.0005 mm. The sensitivity of the film, measured in terms of the quantity
of silver produced per watt-second per square millimeter, depends on
many factors including the grain size and the wavelength of the radiation.
The maximum sensitivity of silver bromide lies under 400 nm, that is
outside the visible region. By "sensitizing" the emulsion with dyes, it is
possible to produce films which are sensitive to wavelengths greater than
600 nm. Without this chemical trick, photographs could not give us a
natural impression of a scene since all red objects would appear too dark
and all blue or violet objects too bright.
The radiated energy received by the film depends on the radiation
density of the emitter, on the conductivity of the light tube and the
duration of the irradiation. This duration is generally called the exposure
time. It is regulated in a camera by means of the shutter.
Within the framework of geometrical optics, a single grain of silver
bromide cannot be considered as a radiation receiver; for the expression
LLW = AnW represents the linear conductivity in the plane of the grain
and, with A = 0.0005, n W must have a value almost equal to one if LLW
is to be one Rayleigh unit. In practice, this is impossible. Furthermore
the silver bromide grains do not lie in a mathematically defined plane
but are embedded with a statistical distribution in a gelatine layer 0.01 mm
thick. Within this layer the radiation spreads out because of scattering
and reflexion at the lower surface. Because of these effects, practical
results have shown that it is realistic to assume a value A = 0.01 mm as
the smallest receiver radius in the emulsion. The value nW = 0.05 is then
sufficient for imagery according to geometrical optics. The corresponding
aperture in a cameraisj/lO, lying between the engraved values of 8 and 11.
The above figures concerning the "grain" of the emulsion do not mean
that one can see black grains of 0.02 mm diameter in the nega-
tive after exposure, development and fixing. The blackened areas are
much smaller and correspond to the positions of the silver bromide grains
before exposure. But the region of blackening has a mean diameter of at
least 0.02 mm. It is therefore not the grains of silver which cause the user
of 35 mm format so much trouble. These individual grains only become
visible when they are magnified 100 times so that the 24 mm x 36 mm
format fills a screen 2.4 m x 3.6 m. What is recognized as "grain" at
normal magnifications results from statistical clumps of blackened silver
grains.

3 Zimmer, Geometrical Optics


34 Perfect Imagery

§ 14. The Photographic Objective


The principle of the photographic objective is usually described in the
following way: - A lens produces an image of a distant object roughly in
the region of the focus where it is recorded by the film. This description
is particularly easy to confirm, for we need only introduce a piece of
"ground glass or a piece of parchment into the film plane in order to see
the image.
What do we see if we dispense with these devices and simply bring
the eye pupil into the region of the film plane? Nothing at all, or at least
no sign of the objects which are imaged in the focal plane. The reader
can try this out when loading his camera. He will only see the "hole"
through the objective defined by the edge of the stop. The hole is coloured
and uniformly light or dark according to the colour and brightness of
the object towards which the camera is pointing. In normal cameras the
edge of the stop is not sharply defined because it is too near the eye.
It can be focused sharply with a magnifying glass without changing the
other effects.
From a physical point of view, the function of the photographic
objective can be described as follows. The emitter is the object which is
either emitting or reflecting radiation, the receiver of the radiation is the
entrance pupil of the objective. For simplicity, we shall assume that the
objective consists of a thin lens and that there is no stop placed in front
of or behind it. The lens mount defines the entrance and exit pupils. The
lens thus serves as radiation receiver. But since the lens cannot absorb
energy, it must be radiated further: the lens becomes an emitter and
radiates with the same radiation density as the object because the radiated
power and the conductivity of the light tube are not changed by the
(ideal!) lens. Then why do we need a lens, for a hole in an opaque screen
can serve equally well as receiver and emitter of the radiation? This
interpretation is quite correct. But the light tubes emanating from two
separate objects cannot be separated behind the hole if their separation
is less than the diameter of the pupil. This is shown in Fig. 9. It is, of
course, possible to measure the radiating power of the two sources behind
the hole but we cannot distinguish between the two. To do this, we should
have to push the receiver through the hole near to the plane of the object.
This difficulty can be overcome by using a lens which shifts the object
behind the hole (Fig. 10).
In Figs. 9 and 10 the light tubes are decentred and are therefore
represented by the outer rays, not by the principal and marginal rays.
If we have a complex objective instead of a thin lens, the following
generalisation of the above result is valid. The emitter is represented by
the exit pupil of the objective. This is the optical image of the entrance
The Photographic Objective 35

pupil and is illuminated by the object. It radiates (apart from losses


from refiexion and absorption) with the radiation density of the object.
The radiated power registered by the film is therefore proportional to the
area of the exit pupil. The size of the exit pupil is controlled by an iris
diaphragm situated within the objective. The objective images the
emulsion layer of the film into the object space - not the reverse, for
the emulsion and the pupil are physically determined, not the object. On
account of its grainy structure, the film not only registers the radiated
power of the entire light tube but also the radiated power per element of

0,

0;

Fig. 9

:~;" Lens
Fig. 10

area of the emulsion, that is the intensity of irradiation. Since the


geometrical data of the light tube and its conductivity hardly change
between neighbouring surface elements, the intensities of irradiation are
proportional to the radiation densities. The radiation densities are
invaliants in optical imagery if absorption and refiexion are ignored and
therefore give us a visual impression of the brightness distribution in the
region of the object space corresponding to the image of the emulsion
through the objective.
It is fortunate that photography was discovered a long time ago,
for the above explanation would certainly have frightened away any
possible inventor. This way of looking at the process has been given here
because it is instructive for the reader who is familiar with practical
photography to realize how complicated even a familiar daily process
3*
36 Perfect Imagery

becomes when it is analysed strictly in accordance with its basic physical


principles.
The local blackening of the film is not only determined by the
exposure time but also by the intensity of irradiation at that point. It
is proportional to the radiation density and can be increased by better
illumination. If the radiation density is fixed - as in landscape photo-
graphy - the intensity of irradiation is proportional to the square of the
linear conductivity of the light tube between the exit pupil and the
irradiated element of the surface divided by the irradiated area F.
For small areas, the difference between the radius and the perpendicular
B' is negligible so that the intensity of irradiation is proportional to
(LLW) 2 (B'n'U')2
- - F - "" nB'2 ,that is proportional to (n'U')2. Here u' is the angle

Lens as
pup/7 Image in the
fi7m plane
Fig. 11

formed by the marginal ray from the edge of the exit pupil to the centre
of the elementary area in the film plane. A typical case is shown in Fig. 11
in which an infinitely distant object (U = 0) is imaged by a thin lens
in the plane of the pupil. It follows that it is not the absolute size of the
stop which is of fundamental importance in a camera, but the angle u'
it subtends at the film plane. The space between objective and film
contains air so that n' = 1.
The speed of a photographic objective is defined by its largest "rela-
tive aperture" 2t. The quantity 2h defines more exactly the diameter
of its entrance pupil. For u = 0, we have h = A as the perpendicular to
the marginal ray and rp = If/the refracting power of the objective. Hence,
from equation (9.1)
n'U'nW - n'W'nU n'U'
rp = LLW =--x- fornU = 0.

Consequently, 2 t = 2 A rp = 2 n'U' is a linear measure of the intensity


of irradiation in the film plane for the imagery of objects at infinity. The
stop numbers are the reciprocals of 2 n'U'. The following table gives the
The Photographic Objective 37

normal stop numbers and the corresponding values of 2 n'U'. The factor 2
has no special significance in photography. The field angle is usually
given as 2W rather than W.
Apart from the bracketed intermediate value 3.5, two successive stop
numbers differ by the factor V-Z = 1.41. This is convenient, since the
intensity of irradiation in the image of extended objects is proportional
to the square of n'U' and is therefore halved by a change from any stop
number to the next higher number, doubled by a change to the next
lower number. If the radiation density is to remain unaltered, we require
an exposure time twice as long and half as long respectively.

Table 3

Stop number 0.5 0.7 1.0 1.4 2 2.8 (3.5) 4.0

2n'U' 2.0 1.4 1.0 0.7 0.5 0.35 (0.28) 0.25

Stop number 5.6 8 11 16 22 32 45 64


2n'U' 0.18 0.125 0.088 0.062 0.044 0.031 0.022 0.016

How is the intensity of irradiation changed if we focus on near objects?


The following equation follows quite generally from (9.1): -
'u' LLW n'W' U
n = f'nW + nW n .

In the entrance pupil, B = 0 and LLW = AnW. Here, A is the perpendi-


cular from the centre of the entrance pupil on to the marginal ray and
can differ slightly from the radius of the entrance pupil. Therefore

2 n 'u' 2A
=T+ n'W' 2 U
nW n . (14.1)

This equation shows that there is no generally applicable answer to the


'W'
above question. Essentially, the quotient _n- determines the results, i.e.
nW
the pupil magnification. For many objectives n'W' "" nW. In this case
the intensity of irradiation becomes less for near objects since U and U'
have opposite signs. This normal case can be readily understood if one
remembers that the area of the exit pupil is unchanged while the distance
to the film plane is increased by focusing on near objects. In the case of
telephoto and wide angle objectives, Wand W' are not roughly equal.
Here the user must rely on trial and error or exposure times supplied by
the manufacturer, as information on the position and magnification of the
pupil is not usually available. Equation (14.1) shows that the irradiation
38 Perfect Imagery

intensity is not a function of the object plane if W' = 0, that is, if the
exit pupil is placed at infinity. This is technically possible but is of no
significance in practical photography.
Modern miniature camera objectives have a focal length f = SO mm,
i.e. a refracting power rp = 0.02 mm-I = 20 dpt and a relative aperture
2 h/f = 1/2.8. The smallest stop number is 2.8 giving a diameter of the
entrance pupil 2 A = SO mm/2.8 = 18 mm. For an angular field 2 w = 46°
we have in air nW = -0.39. The minus sign corresponds to Fig. 11 but is,
of course, arbitrary. In the plane of the entrance pupil, B = 0 so that
LLW=AnW= -9 x 0.39mm=- 3.Smm.
A comparison with the linear conductivity of the eye (LLW = 0.7S
mm) shows that it is not possible with the tools of geometrical optics to
adapt the light tube of the camera objective to the eye. Consequently,
no viewfinder can exist which enables the eye to see the image as it
appears in the film plane. We can only view the whole image field if we
reduce the aperture in the viewfinder to one fifth. This implies photo-
graphing with a stop number of 14. Alternatively, we can use the whole
aperture but only see a fifth of the image field. In practical viewfinders
the full aperture is abandoned in favour of a view of the whole image
field. For this reason, the image quality cannot be checked in the view-
finder and incorrect focusing can only be detected by a trick. This consists
of introducing into the plane of the emulsion a ground glass screen which
scatters the radiation more or less uniformly. It follows that the intensity
distribution over the ground glass screen can be seen from more or less
any position and it is no longer necessary to locate the eye in the pupil
plane. This destruction of the light tube by scattering has the disadvan-
tage that the rules of geometrical optics are no longer valid. The radiation
density becomes smaller, the image on the ground glass screen appears
markedly darker than the object. This is the reason why photographers
used to cover their heads with a black cloth and why it is still necessary
today to use the largest stop aperture when focusing with a ground glass
screen viewfinder.
Finally, we shall refer briefly to a curiosity of photography, the
pinhole camera. As Fig. 9 showed, a geometrical separation of the light
tubes behind the pinhole is not possible. Nevertheless, the pinhole camera
can yield good pictures if the exposure times are long enough and the
size of the pinhole is suitably chosen. In Fig. 12, suppose the object
distance a is large compared with the radius r of the pinhole. If this is
the case, the geometrical spread of the light tube compared with the
central projection through the centre of the hole is of the order of
magnitude of r. The greater the separation b of the pinhole from the
film plane, the less significant this spread becomes in relation to the
central projection. On the other hand, the angle u' subtended by the
The Calculation of Light Tubes 39

pinhole at the film plane becomes correspondingly smaller, the image


becomes darker and the geometrical shadow boundary more and more
obliterated by diffraction images of the pinhole. For this reason, the

Olyect FIlm {one


Pinhole
....~~~~~
, ..........
...... -r
""'-
...................... ~

Axis
Central projection
of the object
~, -;-(' -;::, -:.: --- -r
a ;,1 b
"I
Fig. 12

distance b is best chosen to be such that the geometrical spread r is of


the same order of magnitude as the diffraction pattern:
A
r = sinu' ,
r sin u' = A,

There is no difference between the sine and tangent of this small angle
so that we can write sin u' = rib giving:
r2
b=A, r = VbA.

For b = 300 mm and A = 0.00055 mm we have r = 0.41 mm; for


b = 50 mm and the same wavelength r = 0,17 mm. Lord Rayleigh gave
r = 0.95 Vb1 as the best value. This difference is negligible in practice
because of the finite spectral range.

§ 15. The Calculation of Light Tubes


Iterative Application of the Formulae
for Transfer and Refraction
The essentially qualitative treatment in the above sections is not
adequate for an exact analysis of optical instruments and the special
technical features of their design. If the parameters of an optical system,
that is the refracting powers and separations, and the form of the light
tube in front of the system are known, then the change of the light tube
caused by its passage through the system can be calculated by applying
formulae (7.6) and (7.7) for a transfer Ll and a refraction ([J respectively.
40 Perfect Imagery

This process is repeated for every lens or refracting surface of the system
until we arrive at the form of the light tube in the image space behind
the last surface.
Numerical calculations are not popular because they require consider-
able self discipline. Optical designers facilitate their daily work in no
small measure by using a fixed scheme for the calculation of the necessary
quantities, i.e. the values corresponding to fixed symbols are calculated
strictly in a given order. The choice of such a scheme depends on what
machines are available for calculation, the type of design and personal
factors: in principle, it has no effect on the results. Experience has shown,

Im.age plane

Stop
Exif PPpI7.', ,.,En/ranee pIJpl7
,,
Axis.
H'H

,,
T T
I ,

Fig. 13

however, that strict adherence to a given scheme is the surest way of


avoiding mistakes.
In preparing the calculation, the lenses or refracting surfaces are
numbered in order in the direction of the light. These numbers serve as
indices for the parameters of the system and the coordinates of the light
tube. The refracting power of the first refracting surface is thus CPI' the
radius of curvature of the third r3 , and so on.
The quantities immediately preceding the refraction at lens no. i
receive the index i; the corresponding values following the refraction
also have the index i but with a "dash" attached. A z is the perpendicular
from the pole of the second refracting surface (or from the centre of the
second lens) on to the marginal ray before the refraction, A~ the perpendi-
cular after refraction. It follows from this rule that all separations,
refractive indices and angles have two symbols, for example n~ = n 3 ,
Uj = UHl etc. This causes no confusion.
The Calculation of Light Tubes 41

The first example we shall take is a "triplet", a photographic objective


consisting of three lenses. The numbered parameters will refer here to
lenses which will be considered as "thin", an ideal case. Fig. 13 shows a
section through the system and the light tube. The system parameters
are taken to be:
CfJl = 0.016 mm-1 62 = - IOmm
CfJ2 = - 0.029 mm-1 63 =-l1mm.
CfJ3 = 0.019 mm-1
The dimensions mm and mm-1 are usually left out: it is assumed that
dimensions are in mm unless otherwise stated. The system thus consists
of a concave lens placed between two convex lenses with separations of
10 mm and 11 mm, for in air n = 1. We will suppose that the object lies
at infinity and has an angular diameter of 2 w = 2 x 23.6° = 47.2°. Then
n 1 U1 = 0 and n 1WI = - 0.4. We place the first reference point at the
first lens and start from the values Al = 14 mm, Bl = - 5 mm.
The calculation through the system consists of the following steps: -
1 a) Transfer from the first reference point to the first lens:
Al = Ao + 61 n 1 U 1 Bl = Bo + 61 n 1 WI .
If the first reference point Eo with perpendiculars Ao and Bo already
lies in the first lens, then 61 = 0 and this step may be omitted.
1b) Refraction at the first lens:
A~=Al B~ = Bl
n;U~ = n 1 U 1 + CfJIAI n~W~ = n 1W 1 + CfJIBI .
2 a) Transfer to the second lens:
n2U2 = n~U~ n2W 2 = n~W;
A2 = A~ + 62n2U2 B2 = B~ + 62n2W 2 .
2b) Refraction at the second lens:
A~ = A2 B~ = B2
nfU~ = n2U2 + CfJ2A2 n~W~ = n2W2 + CfJ2B2'
It is clear from these steps how the process is continued for further
elements. Since the equations are used repeatedly in the same form, one
soon becomes used to them. The trivial relationships A; = Ai and
niH UiH = n; U; have only been included for the sake of completeness
since they have no significance for the numerical calculation. The values
are noted down in a special tabular arrangement consisting of two blocks
for each lens, one for the data of the marginal ray, one for the principal
ray. Fig. 14 shows this formal layout and the values for the triplet given
above.
42 Perfect Imagery

As a check, the linear conductivity is calculated in front of and behind


the system: LLW = 14 -( - 0.4) - (- 5) -0 = - 5.6,
LLW' = 13.04744-( - 0.378892) - 5.0162-0.130861 = - 5.599998.
The refracting power of the objective is
= n~U~n,W, - n~W~n, U, = n~U~ = 0.130861 = 0009347 -I
rp LLW A, 14mm' mm.
The reciprocal of this is the focal length f = 1/rp = 106.98 mm. The relative
aperture is 2n~U~ = 0.261722,.. 1/3.8. The image plane is determined by
the point of intersection of the marginal ray with the axis and therefore
lies A3/U~ = 99.7 mm behind the third lens.
The position of the entrance pupil is determined by the point of
intersection of the principal ray in the object space with the axis. Itlies

Tabular arrangement Example


of calculation
n, U, n,W, o - 0.4

1. A, B, 1.
14 -5
n~U~ n~W~ 0.224 - 0.48

2. A2 B2 2.
11.76 - 0.2
n~U~ n~W~ - 0.11704 - 0.4742

A3 B3 13.04744 5.0162
3. 3.
n~U~ n~W~ 0.130861 - 0.378892

Fig. 14

B1/W1 = 12.5 mm behind the first lens. The exit pupil lies B3/W~ =
- 13.24 mm behind the third lens; or, following the sign rule for separa-
tions, 13.24 mm in front of the third lens. How long is the perpendicular
AAP in the exit pupil? We can calculate this by means of a transfer
with 15 = 13.24 (in the opposite direction to the incident light) from the
relation AAP = A~ + t5n~U~. But the following method is more elegant.
At the exit pupil, B = 0 by definition. It follows from the invariance of
the linear conductivity that AAP n~W~ = LLW, so that
AAP = - 5.6/ - 0.378892 = 14.78 mm .
The magnification between the entrance pupil and the exit pupil
is AAP/A EP = 14.78/14 = n l WI/n~W~ = - 0.4/ - 0.378892 = 1.0557. A
real stop must be introduced in the plane where the principal ray (not
its extension) cuts the axis. This plane lies Bz/Wf = 0.42 mm behind the
second lens. At this point, the perpendicular on to the marginal ray has
a length A B = LLW/Wf = 11.81 mm. The marginal ray has an inclination
to the axis sin u = 0.117, giving a radius of the stop R = AB/cos U =
11.81 mm/0.9931 = 11.89 mm.
The Calculation of Light Tubes 43

To determine the principal points we evaluate the coefficients a and d


from equation (9.1):
= A 3n,W, - B 3 n,U, = ~ = 13.04744 = 0932
a LLW A, 14 .
d = n~W~A, - n~U~B, = (- 0.378892)·14 - 0.130861·( - 5) = 0830
LLW - 5.6 . .
According to the equations derived in §9, 6 = (d - 1) ·f gives the transfer
from the first reference point to the "front" principal point. The distance
ofHfrom the firstlensis therefore -nj (d - 1) f = 18.1 mm. 6* = (1 -a)f
gives the transfer from the last reference point to the "back" principal
point H*, the distance being - n~(1- a)f = - 7.3 mm. The "back" prin-
cipal point thus lies 7.3 mm in front of the third lens, i.e. in front of the
"front" principal point. In practice, the calculation of the positions of
the principal points is of no great importance.
In the above example, the refracting powers rp and transfer para-
meters 6 were given quantities. Often the only lens data available are
the refractive indices and the geometrical parameters. For example, a
lens may have the refractive index n = 1.62287, a central thickness
d = 15 mm, one spherical surface of radius r = 200 mm and the other
surface plane. How can we calculate the effect of this lens?
Since the lens is centred, the central thickness gives the separation
along the axis of the two refracting surfaces. The corresponding transfe r
parameter is 6 = - din (d. Fig. 5 in § 7), in this case - 15/1.62287 =
- 9.2429 mm. For the calculation of the refracting power of a surface
from its radius r and the refractive indices nand n' in front of and behind
the surface, we use the equation
n' - n
rp=-y-= (n'-n)'(!, (15.1)

in which (! = l/r is the curvature of the surface. Calculation with radii is


clearer but curvatures are formally more convenient. We cannot prove
equation (15.1) to be correct in the general case. In fact, the aberrations
in optical instruments show that it cannot generally be valid. Lenses
provide only an approximation to the refracting powers needed for perfect
imagery and their function depends in a highly complex way on the
coordinates of light tubes. Aberration theory will show that the relation
(15.1) is indeed only an approximation, although a remarkably good one;
for if the design of the optical system is such that the aberrations are no
longer significant, it gives the correct results. On the other hand, if (15.1)
gives quite wrong values, the lens system is useless on account of its
aberrations and cannot yield an approximation to perfect imagery. If the
aberrations are ignored, (15.1) is the correct formula. With this assump-
tion we shall derive it formally in § 26.
44 Perfect Imagery

Radii are to be considered positive when the surface is convex to the


incident light, that is when the centre of curvature lies to the right of the
pole of the surface. No special rule is necessary for a plane surface since
it is considered to be a sphere with an infinite radius. A plane surface
alway has a refracting power ffJ = O.
In the above example, the lens is surrounded by air with its convex
surface directed towards the object. Then
ffJt = (n; - ntl/r = 0.62287/200 = 0.00311435 mm-t ,
ft = l/ffJt = 321.09 mm .
For the plane surface ffJ2 = O.
We shall now work through the following example whose application
will be discussed later. The object has a diameter of 30 mm and lies 6 m
in front of the lens. A stop of 30 mm diameter at the pole of the lens
surface will serve as entrance pupil. Because of the large object distance
it is not necessary to distinguish between sine and tangent of the angles.
The radii of object and pupil can therefore be taken as the lengths of the
perpendiculars to the rays. With this simplification we have U t = -15/
6000 = - 0.0025 = Wt. The signs result from the assumption that the
marginal ray is drawn to the upper edge of the pupil and the principal
ray from the lower edge of the object. The first reference point is placed
at the pole of the first surface. For this point At = 15 and B t = O. The
calculation gives the following values for the individual surfaces:

Table 4
- 0.0025 - 0.0025

15.0 0
1. 0.044215 - 0.0025

2. 14.591324 0.023107
0.044215 - 0.0025

The image is formed A2/U~ = 330.0 mm behind the second surface;


the magnification is given by n t Ut/n~U~ = - 0.05654. The object of
30 mm diameter thus gives an image of only 1.7 mm diameter. Since the
magnification is negative, the image is reversed both vertically and hori-
zontally. Where is the image after refraction at the first surface? In
considering this, one must remember that the whole space behind the
first surface is filled with glass. The formula for the image distance is
At/U;. But in the second line of the first column of Table 4 we have n;U;
n;
with = 1.62287. Thus the correct value is not 15/0.044215 but 1.62287
x 15/0.044215. The image is thus formed in glass 550.6 mm behind the
first surface. It is clear from this result that the plane surface affects the
The Gaussian Brackets 45

position of the image although rp2 = o. However, it does not alter its
size, for n 1 Udn~U~ has the same value as n 1 Udn~U~.

§ 16. The Gaussian Brackets


An Explicit Technique of Computation
The method of computation given in the preceding section enables
us to calculate the effect of an optical system on the form of a light tube.
The optical designer is often confronted with the obverse problem. He is
presented with a prescribed form of the light tube in the object and
image spaces and is required to determine the optical system capable of
effecting this transformation. In this case, the unknown quantities are
not the coordinates of the light tube but the parameters of the system.
An explicit calculation technique is advantageous for the determination
of these parameters; the recursive method given in the preceding section
is less suitable. However, it should be mentioned that the practical
significance of the Gaussian brackets is not very great since any experi-
enced optical designer can determine the refracting powers and separations
necessary for a given problem by "rule of thumb" methods followed by
an exact determination of the values by trials with the recursive formulae.
The explicit technique is based on Theorem 2. This states that any
type of optical imagery is defined by four coefficients a, b, c, d. Equations
(9.1) show how these coefficients depend on the coordinates of the light
tube. The Gaussian brackets express the coefficients in terms of the
parameters of the optical system. This gives the link between the light
tube coordinates and the system parameters.
In deriving the formulae we make the assumption that the system
parameters are given in the order bl , rpl' b2 , ••• , b", rp", b"+1. This does
not imply any restriction of generality since it is quite permissible for
certain parameters to be zero. If, for example, the refracting powers rpi
and rpHl follow each other without separation, we can either describe
their effect by the refracting power rpt + rpHl or by the parameters rpl,
bl +1, rpHl in which bi +1 = o.
If only one transfer L1 takes place without a refraction, equations (7.6)
are applicable to the transformation of the light tube:

A*=A+bnU B*= B+bnW


n*U* = nU n*W*= nW.
Comparison with equations (9.2) gives the coefficients
a = 1, c=O and d = 1.
46 Perfect Imagery

We have already derived the transformation equations (9.4) for a


refraction (/), and two transfers LI, and LIz in the proof of the principal
theorem and determined the coefficients
a= 1 + CPt b2 = [cpl> b2 ] ,
b= b, + (1 + b,cp,)b2 = [b l , CPt, b2 ] ,

c= CPt = [cpt] ,
d= 1 + b,cp, = [b" CPI] .

The square brackets should at this stage simply be considered as abbre-


viated forms of the expressions on the left hand side. If the expression
a = [cpt, b2 ] means nothing more than that a is a function of CPI and b2 , then
it is of little practical value. Nevertheless, all the formal properties of
these brackets imply that they can be used as the basis of a formal scheme
of calculation. For, if we can write [cpt] instead of CPI and [b" CPt] instead
of 1 + b, CPt' then we can also write [b,] + [b" cpd' [b 2 ] instead of
bl + (1 + b, CPI)b2 . If this formulation is logically consistent, then
[b" CPI' b2 ] must be equal to [b,] + [b" CPI] • [b2 J. This example shows the
reduction rule for the representation of a brackets expression by the sums
and products of brackets with few elements. Such a rule is generally valid.
We shall prove this by considering the change of the four coefficients
caused by a further refraction and subsequent transfer: -
Refraction with the refracting power CPk:

A*=A* = aA + bnU,
n*U* = n*U* + CPkA* = (c + acpk)A + (d + b CPk)nU .
The corresponding equations for the refraction of the principal ray need
not be given, as they have the same coefficients.
Transfer with the parameter bkH:

A* =:4* + bkH n*[J* = {a + (c + acpk)bkH}A +


+ {b + (d + b CPk)bk+l}nU ,
n*U* = n*U* = (c + a !pk)A + (d + b cpk)nU .
The new coefficients are therefore
c= c + a !pk [!PI' b2 , ••• , CPk-l] + [!PI' b2 , ••• , CPk-l, bk] '!pk
=
[!PI' b2 , ••• , bk, !pk] ,
=

d= d + b CPk = [b" CPt, .. " !Pk-l] + [b" !p" ... , CPk-l, bk] '!pk
= [b" CPt' ... , bk, CPk] ,
Ii = a + CbkH = [!PI' b2 , ••• , bk] + [!PI' bz, ... , bk, !Pk] ·bkH
= [cpt, b2 , ••• , CPk, bkH] ,
The Gaussian BFackets 47

b = b + d15k+1 = [15 1, rpl' ... , 15k] + [151, rpl' •.. , 15k, rpk] ,15k+1
= [15 1 , rpl' ... , rpk, 15k+1] .
This inductive process has led us to the desired result, expressed by
equation (16.1), which represents the transformation of the light tube in
terms of the system parameters:
Ak+1 = [rpl' 152 , ••• , rpk, 15k+l]Ao + [15 1 , rpl' 152 , ••• , rpk, 15k+1]noU o '
nk+1 Uk+1 = [rpl' 15 2 , ••• , 15k, rpk]A o + [15 1, rpl> 152 , ••• , 15k, rpk]noUo '
Bk+1 = [rpl' 152 , ••• , rpk, 15kH]Bo + [151, rpl' 152 , ••• , rpk, 15k+1]noW o ,
nkHWkH = [rpl,15 2, · · · , 15k, rpk]B o + [15 1, rpl' 15 2 , ••• , 15k, rpk]noWo '
(16.1)
The perpendiculars and apertures with the suffix refer to reference
point no. 0, the origin of the first transfer 151 ; the quantities with the
°
suffix k + 1 to reference point number k + 1, the terminal point of the
last transfer 15k+1.
The expressions given in the square brackets are called "Gaussian
brackets" because C. F. GAUSS drew attention to this formulation in the
course of his researches into the theory of numbers. But I do not know
of any reference to show that he used the technique in his numerical
researches into optical imagery. The following rules of calculation apply
to the Gaussian brackets: -
[] = 1; a bracket which does not contain any parameter has
the value one. This is a formal definition which is of practical
value for the reduction rules. (16.2)
[x] = x. (16.3)
[XI' x 2, ... , xn] = [XI' x 2' ... , X"-2] + [XI' x 2' ..• , Xn-l] 'X n . (16.4)
(16.4) is the reduction rule, proved above, expressed here for a general
case of parameters XI to X". The following "splitting" rule is a generali-
sation of the reduction rule: -
[XI' x 2' .•. , x,,] = [XI' x 2, ... , xi-d' [XI+2' ... , x,,] +
+ [XI' x 2' ... , Xi]' [Xl+l' ..• , X,,]. (16.5)
By using induction according to the number of the "split" parameters
Xl+l to Xn , it is possible to prove equation (16.5). This proof will not be
given because it does not yield any new physical insight.
The differentiation of the Gaussian brackets represents an application
of the splitting rule which is of interest in optics. Suppose that
K = [XI' x 2' •.. , Xi-I, Xi + LlXi' Xl+l, •.. , Xn] .
Then splitting gives
K = [XI' ... , Xi-I]' [XI+2' •.. , X,,] + [XI' •.. , Xi + Llxd' [Xl+l' ..• , X,,] •
48 Perfect Imagery

The third bracket can be transformed using the reduction rule, giving
[Xl' ... ,Xt+Llxt]= [Xl' •.. ,XH]+ [Xl' .•• ,XH]·(Xt+LlXt)
= [XI' ..• , Xi] + eXt, .•. , Xi-I] ·LlXt ;
for we can combine two additive terms, after expansion to remove the
round brackets, by reversal of the reduction rule. If this result is sub-
stituted in K and terms are collected after applying a reversal of the
splitting rule, we obtain
K = [Xl' X2' ... , Xn] + [XI' ... , Xt-I]· [Xi+!, .•. , Xn] ·Llx! .
Comparison with the first expression for K yields the simple rule for
differentiation:
a
ox, [Xl' X2' ... , Xn] = [XI' ... , Xi-I]· [XHI •.. , Xn] . (16.6)

When a Gaussian bracket is differentiated with respect to a parameter,


it splits up into a product of two brackets which either end before this
parameter or begin after it. The parameter itself no longer appears in the
differential: the Gaussian bracket is therefore a linear function of the
paramater.
To illustrate the use of Gaussian brackets, we will deduce the para-
meters of a doublet photographic objective with focallengthf = 100 mm
and relative aperture 2 h/f = 1/5.6. Suppose it is a telephoto objective,
so that the length of the system from the first lens to the film plane is
shorter than the focal length, say 70 mm instead of 100 mm. The first
reference point is placed in the first lens, the last in the film plane.
Then <5 1 = 0, n l U I = 0, AI = 100/11.2 = 8.95, n3 U 3 = 1/11.2 = 0.0895.
The requirement for the length of the system is <52 + <53 = - 70. The film
plane is to be the image plane, which means that the marginal ray must
cut the axis in this plane:
A3 = 0 = [IP1' <52 , IP2' <53 ]A I + [<5 1, IPI' .•. , <53 ]n1U1 •
The prescribed relative aperture gives
n3U 3 = 0.0895 = [IP1' <52 , IP2]A I + [<5 1, IPI' <52 , IP2]n1U1 .
If the given light tube coordinates AI and n 1 U 1 are inserted in the last
three relations we arrive at three equations
-70 = <52 + <53 ,

o = [IP1' <52 , IP2' <53 ] (16.7)


0.01 = [IPI' <52 , IP2]
for the four unknown parameters <52 , <53 , IPI and IP2' A unique solution
requires a further condition, for example specified pupil imagery. Alter-
natively, one could prescribe that the refracting powers be in some
The Gaussian Brackets 49

given ratio which simpliftes the correction of certain aberrations. To begin


with, we shall assume that the separations are free parameters and shall
eliminate the refracting powers from (16.7). Expanding the Gaussian
brackets according to the reduction rule gives
om = [911' "2' 912] = 911 + [911' "2] '912 and
0= [911' "2' 912' "3] = [911' "2] + [911' "2' 912] '''3 . (16.8)
The first bracket of the second expansion is again removed by the
reduction rule while the second is set equal to 0.01. This gives
o = 1 + 911 "2 + om "3 ,
911"2 = - (1 + om "3) = om 152 - 0.3 since 152 + 153 = - 70. (16.9)
The value for 1 + 911 "2 = [911' "2] is inserted in the first line of (16.8)
giving. after multiplication with 152
0.01 "2 = 911 "2 - 0.01 153 "2 912 ,
and further from (16.9)
912 (0.01 "2'''3) = - (1 + om 152 + om "3) .
Since "2 + "3 = - 70. this yields finally
912"2"3=-30. (16.10)
In practice, if no further conditions are to be fulfilled, one would plot
the refracting powers 911 and f!!2 against 152 or else tabulate the function.
and then decide on a suitable solution. Here. only one case will be eva-
luated, i.e. that for 152 = - 30. Then 153 = - 40 and the corresponding
separations are d2 = 30 and d3 = 40. Hence
911 = - 0.6/ - 30 = 1/50 = 0.02,
912 = - 30/1200 = - 1/40 = - 0.025 .
The first lens is convergent with a focal length of 50 mm, the second
divergent with a focal length of - 40 mm.
A second example of formal calculation with Gaussian brackets will
illustrate the differentiation of the light tube coordinates with respect
to the system parameters f!! and b. Here we suppose that the coordinates
Ao. noUo etc. in the object space are fixed in advance. How do the
coordinates in the image space change if a parameter is changed, for
example the refracting power 911? Applying equation (16.6) to the first
equation of (16.1) we have
8Ak+l
-~-
urp,
= [911' 152 , ••• , "t]' [bHl •... , ".l:+l]A o +
+ [151,911' ... , bt]' [15tH • ...• bkH]noUo .
4 Zimmer. Geometrical Optics
50 Perfect Imagery

After removing the factor [dHl, ••. , dk+1] the remaining terms can
be collected together to give At. Furthermore, equation (16.1) shows that
[dt+1' .•. , dk+1] is the coefficient b of the light tube transformation from
the reference point E, in the plane of the refracting power cP' to the
reference point Ek+1 behind the system. By virtue of equation (9.1)
B k+lAI - A>+lB,
[dHl' ... , dk+1] = LLW '

from which
OA>+l _ A Bk+lA, - A>+lB,
~- t LLW .

In a similar way, we find


OA.+1
~= [CPI' d2 , ••• , CPi-l]' [cpt, ... , dk+1]Ao +
+ [dl , CPI' ... , !Pt-l] • [!pt, ... , dk+1]noUo
A k +l n,W, - Bk+l n,U,
= njU, [cpj, ... , dk+1] = ntU, LLW .

The equations become particularly simple if the final reference point


is placed in the image plane because A k +1 = 0 for this plane. It follows
that

- (n,U,)2 B k+l (n,U,)2


LLW = nk+lU'+l .

These equations give the size of the geometrical discs of confusion in the
original image plane when one parameter is changed by one unit. We
suppose, of course, that the image plane lies at a finite distance. If not,
the changes in the aperture nk+1Uk+1 must be investigated.
The complete set of equations for the differential coefficients of the
light tube coordinates in the image space with respect to the system
parameters!pt and d, for given coordinates in the object space is:
OAk+l A B k +lA,-Ak+lB,
~=! LLW '
OB.+1_ B B>+lA, - A'+lB,
~- t LLW '
on'+l U>+l _ A n>+lWk+l A, - nk+lUk+l B,
o 0lP' -! LLW '
(16.11)
on>+lW>+l B n>+lWk+l A, - nk+lUk+l B,
o 0lP' = j LLW '
OA>+l _ n U 4_'j-l n,WI - B'+l nlU,
00 - i I LLW '
The Gaussian Brackets 51

(16.11)

an.+1wk+l _ W nk+l U k+l n.W. - nk+1Wk+1 n.U.


iJt5. - nj i LLW •

These formulae are very useful when one wishes to change an existing
system or a rough design to fulfil specified conditions. A calculation
which follows the procedure described in § 15 with the correct starting
coordinates A o, B o, noUo and noWo gives all the intermediate values A j , B j
etc. If these are inserted in the formulae given above the necessary para-
meter changes can be calculated.
Since the Gaussian brackets are linear functions of each parameter,
the differential coefficient has the same value as the finite difference
quotient. If only one parameter is free to be changed, this change is
given by equation (16.11) not merely as an approximation but exactly,
even for large changes. Thus one can arrive at the correct result without
iteration if, for example in a rough starting design, a refracting power has
been given a wrong sign or a completely incorrect value. The situation
is different when several parameters are changed simultaneously by finite
amounts. In this case, the changes in the coordinates of the light tube are
only approximately described by (16.11) because second order cross terms
do not disappear. In practice, this non-linearity does not worry us much
since a second iteration leads to a sufficiently accurate solution. Never-
theless, one cannot blindly rely on the convergence of this method: there
are cases where it diverges.
We have assumed up to now that the starting data A o, Bo etc. are
fixed values independent of the system parameters. This is almost always
the case for Ao and noUo because the data for the object are given. On
the other hand, a physical stop inside the system is often present rather
than a fixed entrance pupil. If the system parameters in front of the stop
are changed, the position of the entrance pupil is also changed. In applying
the differentiation formulae one must take this into account.
The stop is placed in lens or refracting surface number j. This does
not imply a restriction of generality because a lens of refracting power
zero can be introduced at any reference point. Then from (16.1):
Bk+1 = [!pjH, ... , !5k+1]-Bj + [!5j+1' ... , !5k+1]njWj,
nk+1Wk+1 = [!pjH, .•• , !Pk] -Bj + [!5j+1' ... , !Pk]njWj .

Since Bj = 0, LLW = AjnjWj and consequently njWj = LLWjAj .


It follows that

52 Perfect Imagery

nlc+l
W k+l = LLW ["1+1, ... , <po]
AI '
.
In W
hi h A
C ,oft

smce
B
,= 0 .
If the differentials with respect to system parameters in front of the
stop are formed, we have
8B>+1 LLW ["1+1, ... , ">+1] • aAl
8tp, = - A1 8<p,
i ;;;;, j.

The last differential is derived from the set of equations (16.11).


Hence, with B, = 0 we have
LLW ["1+1, ... , ">+1] A - B
8B>+1 _
a;p;- - - A: i LLW
AlB, _
-
B>+lA,
! LLW .

A formally identical analysis gives


an'+lW>+l _ B . nk+'W>+l • A,
iJ<pl - i LLW '
8n>+lW>+1 _ W - no+,W>+l n,U,
8.5, - ni t LLW .

These formulae enable the changes in the exit pupil corresponding


to a fixed stop position to be calculated. It is clear that they represent
formally a special case of equations (16.11). If parameters behind the
stop are changed, there is no change in the entrance pupil: in this case
equations (16.11) are applicable. The position of the stop has no influence
on the coordinates of the marginal ray so that no special formulae are
necessary for their differential coefficients.

§ 17. Graphical Representation of Perfect Imagery


Section through the Light Tube
and the Conductivity Diagram
Perfect imagery can be represented graphically by a longitudinal
section containing the axis through the light tube. This form of represen-
tation has been repeatedly used in this book, for example in Figs. 2, 8,
11 and 13. It describes the light tube by the paths of the marginal and
principal rays. A significant advantage of this method is that it provides
a simple geometrical model which enables all angles, perpendiculars,
diameters and separations to be specified. A transfer to another reference
point can be followed very easily since the form of the light tube remains
unchanged. This is no longer the case if refraction occurs at a surface or
at a thin lens. As our geometrical model cannot provide a direct measure
of a refracting power, we are obliged to determine its effect by calculation
Graphical Representation of Perfect Imagery 53

or by a special geometrical construction. The computation procedure was


given in § 15. It is much quicker and more precise than the corresponding
geometrical construction which is given here for the sake of completeness.
We place the reference point E on the axis in the plane of the thin
lens or at the pole of the refracting surface. Then A' = A and B' = B;
the marginal and principal rays before and after refraction are therefore
tangents to the circles with radii A and B respectively centred on the
reference point E. For the sake of clarity, this procedure is shown in
Fig. 15 although it does not, on its own, completely describe the process
of refraction.
The lens or refracting surface is not shown. The marginal ray, the
perpendiculars A and A' and the arc of the circle about E with radius A

E
Fig. 15

are represented with continuous lines; the principal ray and the con-
struction corresponding to it are shown with dashed lines.
Even if the marginal and principal rays are defined by known values
before the refraction, the above construction is not adequate to define
the refracted rays because the tangent points of these rays to the given
circle are not known. They are determined by means of the aperture
relation of the law of refraction
n'U' = nU + tpA , n'W' = nW + tpB .
If the first equation is multiplied by B, the second by A, we have
(n'U' - nU)B = (n'W' - nW)A = tpAB. The following construction is
based on this relation.
We draw the perpendicular 5 to the axis through E. The four angles
u, u', wand w' are defined with respect to this line. The other lines
defining these angles are the perpendiculars A, A', B and B'. Fig. 16
54 Perfect Imagery

shows these angles, which correspond to values given in Fig. 15. In


Fig. 16 we can represent the sine of an angle as a length perpendicular
to S, i.e. parallel to the axis. If we wish to construct nUB, we measure
off the length nB along the perpendicular A to the point Q starting from
the point E and drop the perpendicular to S from Q. The point on the
line S is denoted by Q*. The length of QQ* is nUB.
In Fig. 16, nUB is shown by a dotted line for n = 1 while n'U'B' is
the corresponding dotted line Q'Q*' for n' = 1.5. According to the sign
convention for angles, the distances are to be considered positive when
they are measured from the point on the line S in the same direction as
the light. Thus n'U' B is positive. (n'U' - nU)B is therefore the distance
from Q to Q' measured in the direction of the light, i.e. in the direction of
the axis. For a refraction this must be equal to (n'W' - nW)A, that is

A :4'
Q'

E
Fig. 16

equal to the distance between the points P and P', again measured in
the direction of the light. P P' is found by an equivalent construction
in which the distances nA and n'A' are drawn on to the perpendiculars
Band B'.
The construction becomes particularly simple for a thin lens in air
for which n = n' = 1, and it is only in this case that it possesses any
practical significance. For here Q and Q' are the points of intersection
of the perpendiculars A and A' (or their extensions) with the circle of
radius B centred on the reference point E; and, correspondly, P and P'
are the points of intersection of the perpendiculars Band B' (or their
extensions) with the circle of radius A centred on E. Fig. 17 shows'the
complete construction for this special case. It is supposed that the axis
with reference point E and the principal and marginal rays before
refraction are given. The perpendiculars A and B can then be constructed
and circles drawn about E as centre. To continue the construction, we
Graphical Representation of Perfect Imagery 55

will suppose that the refracted marginal ray is also given, for example
by the angle u' necessary for a given radiation intensity in the image or
for a prescribed magnification. Alternatively, this ray may be defined
by a given intersection point with the axis which, in turn, is determined
by a particular image position. In any event, we will avoid at this stage
any explicit assumptions about the refracting power. We can now con-
struct the perpendicular A' and determine Q and Q' as the points of
intersection of the perpendiculars A and A' with the circle of radius B;
and P as the intersection of the extension of the perpendicular B with
the circle of radius A. The desired point P' lies on the same circle and is
the same distance from P in the direction of the axis as the distance

--..!¥

E
Fig. 17

from Q to Q'. The line joining P' to E gives the perpendicular B' which
thus defines the refracted principal ray.
If the refracted principal ray is given, the construction can be carried
out in an analogous manner with the roles of A and B, P and Q reversed.
If the refracting power rp of the lens or surface is given, the distance n'/rp
is measured off along the axis from the point E giving the back focal point
F'. A tangent is drawn from this point to one of the two circles, for ex-
ample the circle with the radius B. The perpendicular from the reference
point E to this tangent intersects the circle with radius A in R. The pro-
jection of ER on to the axis is the distance (n'W' - nW) A =
(n'U' - nUl B = rpAB which is necessary for the construction shown
above. This can readily be seen if one remembers that the tangent used
in the construction may be considered as a principal ray which was
parallel to the axis before refraction and for which the relation n'W' =
rpB must therefore hold.
56 Perfect Imagery

Fig. 17 demonstrates quite clearly that a single thin lens cannot


produce perfect imagery. For if one maintains the planes of the object,
pupil and image fixed while repeating the construction for smaller and
smaller angles u and w, the "comer points" where the marginal and
principal rays change direction move nearer and nearer to the reference
point E, their locus being two different curves which meet only at the
reference point. The most favourable form of the lens is thus not identical
for both the object and pupil imagery: it is not possible to produce an
acceptable image of both with a single lens.
If A or B is equal to zero, we must indroduce an auxiliary circle with
a non-zero radius.

Fig. 18

The construction given above for light tubes is useful in that it can
be readily understood from the rules of elementary geometry; but it
rapidly becomes unwieldy when applied to optical systems consisting of
several lenses with small separations. For such cases, the "conductivity
diagram" introduced by E . DELANO in an article in Vol. 2 of "Applied
Optics" (1963) is more practical. This is much simpler to construct and,
although its significance is less clear than that of the section through
the light tube, it nevertheless permits a geometrical interpretation of
the process.
DELANO considers the marginal ray and the principal ray as two
projections of a line, skew to the axis and joining the edge of the emitter
to the edge of the receiver, on to two planes which are perpendicular to
each other and contain the axis. To illustrate this, we shall suppose one
Graphical Representation of Perfect Imagery 57

plane to be vertical, the other horizontal. We join the "upper" object


edge 0 in the vertical plane to the "front" pupil edge P in the horizontal
plane. The line OP is skew to the axis. Its projection on to the vertical
plane is identical with the path of the principal ray from the object
edge to the centre of the pupil; the projection on to the horizontal plane
is identical with the marginal ray from the object centre to the pupil edge.
The three-dimensional effect is shown in Fig. 18.
H the skew line OP is projected on to a plane perpendicular to the
optical axis, the "perpendicular line" of the conductivity diagram is
obtained. This shown by the dot-dashed line in Fig. 18. But this definition
of the perpendicular line is not really correct; for, strictly speaking, it is
the lengths of the perpendiculars A and B which are shown in the con-
ductivity diagram, whereas the dot-dashed line in Fig. 18 gives distances
perpendicular to the axis, that is to say the radii of the object and the
pupil. Apart from this limitation, which is of no significance in practice,
the perpendicular line demonstrates the behaviour of the skew ray shown
by the continuous line as it passes through the optical system. The
observer is considered to be looking in the direction of the optical axis.
At first sight, compounding the principal ray and the marginal ray into
a single line in this way appears to have the disadvantage that the
separations, refracting powers and angles can no longer be read off from
the diagram. But this is not the case.
In the conductivity diagram the perpendiculars and apertures are
constructed to two coordinate axes which are perpendicular to each
other. Usually A and nU are horizontal (corresponding to the horizontal
plane), Band nW are vertical. The scale for A and B must be the same,
as well as that for nU and nW, but it is almost always convenient to
choose different scales for perpendiculars and apertures.
For any transfer from one reference point to another there is a
corresponding line in the "perpendicular diagram" with the slope
B' - B nW
(17.1)
A' -A nU

and a corresponding point in the "aperture diagram", since n'U' = nU


and n'W' = n W. Conversely, for every refraction there is a corresponding
point in the perpendicular diagram with the coordinates A' = A and
B' = B and in the aperture diagram a line with the slope
n'W'-nW B
n'U' -nU =A".
(17.2)

In Fig. 20 the conductivity diagram is shown for imagery by a thin


lens L. The heavy continuous line OLO' is the perpendicular diagram,
the heavy dot-dashed line IX {3 the aperture diagram. For comparison,
58 Perfect Imagery

Fig. 19 shows the longitudinal section through the light tube for the
same system.
The apertures and perpendiculars have deliberately been chosen to
be very large to make the significance of the diagrams clear. In practice,
imagery as shown here would not be possible with a single lens.


<I\~
o ."cW
?II'.

IReal exit pllpl7


9, , , ,ff, , , ,19
Scale
Thin lens
I Virtllal entrance pllpil
Fig. 19

The following relations exist between the conductivity diagram with


M as origin of coordinates and the system parameters: -
If C and D are the end points of a line in the perpendicular
diagram, the area of the parallelogram formed by MC and
MD is <5. LLW, in which <5 = - din is the reduced distance
between the reference points on the axis corresponding to C
and D. (17.3)
If IX and (J are the end points of a line in the aperture diagram,
the area of the parallelogram formed by MIX and M{3 is
cp. LLW, in which cp is the refracting power of the lens or re-
fracting surface. (17.4)
If the corner point L in the perpendicular diagram indicates
the reference point of a lens, the line ML is parallel to that
line in the aperture diagram which describes the aperture
change arising from the refraction by the lens. (17.5)
If the corner point IX in the aperture diagram indicates the
apertures corresponding to a transfer the line MIX is parallel
to that line in the perpendicular diagram which represents the
transfer. (17.6)
Graphical Representation of Perfect Imagery 59

[nw. B

~

IT?
"\
~

10 1,0 \
[\
1.'1 \
1\
111
i n~ \ '-.
I~ ,; '\

"
,;

h V r'\

" -.- ~\ . b
~ -12 ML/ 2 6 p A
:0 IS ,; .~

"- GIS 1\ ~p ,; nU

'- '- V'.\


"-"
,;
1-:>" ,;
,; f',
,; V

"
-~
'1-. 1-.
1-. ./
j=ll5 ~.
1-. ~
-8 1/
'l/
-.'I
V
./
_Il 1.A'1'
0

Fig. 20

Fig. 20 shows examples. If we wish to calculate the linear conductivity,


the reference point is placed in the centre of the object. Then A = 0,
B = - 10 and nU = - 0.5 giving LL W = - 5. In the perpendicular
diagram the points 0, Land 0' indicate the planes of the object, lens
and image; in the aperture diagram IX gives the apertures before the
60 Perfect Imagery

refraction and (3 the apertures after the refraction. It follows from (17.6)
that MIX is parallel to OL and M(3 parallel to LO'; from (17.5) that ML
is parallel to 1X(3. The area of the parallelogram formed by MIX and M(3
is calculated as the product of the base and height of the triangle MIX(3,
giving <>IXM(3 = 0.4167. From (17.4), rp = 0.4167/5 = 0.0833 = 1/12. The
sign can only be obtained from this construction if the area contained
by the parallelogram is given a sign according to the direction of move-
ment round M. The parallelogram formed by MO and ML has the area
<>OML = 100. As we are supposing that the lens is in air, the distance
from object to lens is, according to (17.3), 20 units. Since <> LMO' = 150
it follows that the distance from the lens to the image is 30 units. The
pupil planes are given by B = 0, that is by the intersection points of the
lines on the perpendicular diagram, or of these lines produced, with the

Fig. 21

A-axis. The plane of the entrance pupil is indicated by the point P.


Since <> LM P = 25, it follows that this point lies 5 units behind the lens.
The exit pupil is indicated by P' and lies 3.5 units behind the lens, a
result given by <>LMP' = 17.6.
The magnification for the object imagery is MO' /MO = - 1.5; that
for the pupil imagery MP'/MP = 0.7. The apertures on the object side
nU = - 0.5 and nW = - 0.4 correspond to apertures on the image side
n'U' = 0.33 = - 0.5/ - 1.5 and n'W' = - 0.57 = - 0.4/0.7.
The proof of (17.5) and (17.6) follows directly from equations (17.1)
and (17.2) which were used in the construction of the conductivity dia-
gram. The theorem (17.3) is given by the elimination of b from the transfer
relations
A' -A = bnU, ,
B' _ B = bnW B A - A B = b(AnW - BnW) .
A and B are the right angled coordinates of C, while A' and B' are those
of D. Fig. 21 shows clearly that B'A - A' B gives the area of the parallelo-
Graphical Representation of Perfect Imagery 61

gram formed by MC and MD. For if one constructs the triangles MB'D
and MCA, which form part of the rectangle of area B'A and lie outside
the parallelogram, then the area of the parallelogram is obtained if the
area of the upper rectangle A' B is subtracted from that of the rectangle
B'A. The proof of (17.4) is completely analogous and follows from the
elimination of rp from the relation (7.7) for the refraction@.
There remain a few points concerning the conductivity diagram.
There can be no line or extension thereof on the perpendicular or
aperture diagram which goes through the origin of coordinates M, since
LLW ¥= O. Any polygon, whose sides (or extensions thereof) do not pass
through M, may be considered as a perpendicular diagram. Furthermore,
we may use (17.5) and (17.6) to construct the corresponding aperture
diagram from any specified starting point, and the reverse process applies

equally well. For every such conductivity diagram there is a corresponding


case of perfect imagery whose parameters 6 and rp are uniquely determined
by (17.3) and (17.4). The conductivity diagram is therefore a complete
geometrical equivalent of the principle of the invariance of the linear
conductivity.
An object shift [J is defined by the equations (7.5)
A* = A +wB B*=B
n*U* = nU + wnW n*W* = nW.
Since, according to theorem 3, the parameter w is unchanged by perfect
imagery, the object shift [J can be represented in the conductivity diagram
by a simple coordinate transformation, i.e. a "shearing" of the B-axis.
As shown in Fig. 22, the B-axis is rotated about M and its scale thereby
increased so thatfor lines parallel to theA-axis, B* = Band n*W* = nW
The new coordinates A* and n*U* are read off parallel to the B*-axis
and therefore have the values A* = A + wB and n*U* = nU + wnW.
62 Perfect Imagery

If a line parallel to the new B*-axis is drawn through the point with
the coordinates A = 0 and B = 1, then this line cuts the A-axis at a
distance w from the origin M. In Fig. 22, w is negative. If a conductivity
diagram has been constructed and we wish to investigate the effect of
an object shift, we can leave the figure as it is and simply "shear" the
B-axis. The changed perpendiculars and apertures as well as the new
positions and sizes of the object and image can immediately be read off
in the new coordinate system.
As an example of an object shift represented by a "shearing" of the
B-axis, we shall show that the points F and F' in the perpendicular
diagram of Fig. 20 indicate the front and rear foci respectively of the
lens L. If the B-axis is "sheared" until it is parallel with OL and therefore
passes through the point IX, then nl*UI * = O. The object lies at infinity
and has an apparent angular diameter of nrWr = - 0.4. The point F'
has the new coordinate A * = 0 and is thus the image point of the
infinitely distant object, i.e. the rear image point. The image size is given
by B* = 4.8 and the aperture on the image side is n*U* = 1.03. Such
apertures in air are, of course, meaningless and serve to remind us that
Fig. 20 is concerned with a theoretical demonstration of the conductivity
diagram and not with a practical case of image formation. The demon-
stration that F is the front focal point of the lens is analogous to that
for F' above. The distances of the focal points from the lens can be
calculated according to (17.3) from LLW and the area of the parallelo-
gram formed by ML and MF or by ML and MF'.
Any pupil shift II can be represented in the conductivity diagram by
a "shearing" of the A-axis. The proof is completely analogous to that
for an object shift above.

§ 18. Conductivity Diagrams of Optical Instruments


Examples
Fig. 23 shows the conductivity diagram of a 5 x magnifying glass
whose position is indicated by the letter L. The linear conductivity is
determined by the eye, so that LLW = 0.75 mm. The focal length
follows from the definition of the maguifying glass magnification as one
fifth of 250 mm, i.e. f = 50 mm. The refracting power is cp = 1/50 mm =
0.02 mm-I. This value can also be derived from Fig. 23. The parallelogram
constructed from the end points of the dot-dashed aperture diagram
has the area 0.5 x 0.03 = 0.015 which, after division by 0.75 mm,
gives cp = 0.02 mm-I.
The eye pupil is indicated by the point of intersection P' of the
perpendicular diagram with the A-axis. The position and size of the
Conductivity Diagrams of Optical Instruments 63

entrance pupil are given by the point of intersection of the perpen-


dicular line OL, produced, with the A-axis. In Fig. 23 the eye pupil
is 9.6 mm behind L. For a pupil shift of a further 10 mm away from the
magnifying glass the A-axis must be "sheared" until it passes through
the point P* with the coordinates A = 1.5 and B = - 5. This pupil
position requires a lens diameter almost twice as large. By contrast,
the object shift to a distance of 250 mm brought about by accommodation
of the eye hardly produces any effect. The corresponding effect in the
diagram is a small tilt of the B-axis of about 3/4 0 in a clockwise direction.

nW 8
o
l-
I 20

0,3

0.1 L

-5 5 A
-0.1 p' 0.1 nU

+P'
Fig. 23

Fig. 24 shows the conductivity diagram of an 8 x 30 binocular. The


linear conductivity has the value LLW = - 0.975 mm, taking account
of twilight vision. The 8 x telescope magnification can be deduced either
from the ratio 15 to - 1.875 of the perpendiculars A and A' in the plane
of the entrance pupil P and the exit pupil P' respectively, or from the
ratio 0.52 to - 0.065 of the angles of view W' and W behind and in front
of the instrument.
The line joining P and 0' (plane of the real intermediate image)
passes above the actual perpendicular line of the binocular and therefore
forms a larger area with the axis than that formed by the perpendicular
line. This implies that the geometrical distance from P to 0' is shorter
than the focal length of the objective: the objective has a weak "telephoto
effect", as this reduction in length relative to the focal length is called.
The effect is important in the design of photographic objectives.
64 Perfect Imagery

The conductivity diagram is particularly striking for the eyepiece.


A field lens behind the intermediate image 0' and an eye lens would be
sufficient for perfect imagery. Their effect on the perpendicular diagram
is shown by the dotted line in Fig. 24. The "detour" which takes place
in the actual instrument is only necessary because the aberrations have
to be corrected. It wastes the reduction in length achieved by the tele-

nW B
a2 4

-2 P' (P A
-al nU
Eyepiece I
.. I
I
I
10 I
P'
.... \
~.,..-.-.-
\i
a5 I
I
I
\i I
\i ·-a5 -10
I
\i

,
I
\i
\i
II
-a5 -10

Fig. 24 Fig. 25

photo effect of the objective and would be quite pointless if one could
achieve perfect imagery with two lenses.
Fig. 25 shows the conductivity diagram for a simple opera glass with
2.5 x magnification. The scale is the same as that of Fig. 24. The linear
conductivity LLW = - 0.39 and the angle of view of the eye is only
utilized up to W' = 0.264. This telescope does not have a real intermediate
image, i.e. there is no intersection of the perpendicular diagram with the
B-axis. Of the two pupils, only the exit pupil P' is real and is determined
in normal use by the position of the eye pupil. The entrance pupil (P)
Condnctivity Diagrams of Optical Instruments 65

is virtual and appears in the case shown 128 mm behind the objective,
or 87 mm behind the eyepiece. Since the position of the exit pupil depends
upon how the user holds the instrument, the values found.in practice
may differ widely from those calculated. The variations in the pupil
position can be represented in the diagram by "shearing" the A-axis
The useful field is very sensitive to such variations. A "shearing" in the
clockwise direction corresponds to an increase in the distance of the
eyepiece from the exit pupil. In this case, the objective must be given
a larger diameter if vignetting of the light tube by its rim is to be avoided,
although the diameter of the entrance pupil remains unchanged at 7.5 mm.

nW B

0.3 6

MagnifYing glass
-Qj -6

-5 P A
><
....
'-
~:

-0.5 '-'-'-'~' nU -.:~~


Objective
Fig. 26 Fig. 27

Fig. 26 shows the conductivity diagram of a microscope consisting


of an objective with a magnification fJ of 40 x and a 10 x eyepiece. The
numerical aperture at the object is nU = - 0.65, the object size Bo =
- 0.2 mm, the linear conductivity LLW = - 0.13 mm. The eyepiece
consists of 2 lenses, of which one is situated just in front of the real
intermediate image 0' and functions as a field lens while the other
functions as a magnifying glass for the observation of the second inter-
mediate image 0" produced from 0' by the field lens. In microscopes,
the eyepieces are simpler than those used in binoculars because the exit
pupil and the angular field are both considerably smaller. A comparison
of Figs. 24 and 26 makes this clear.
The striking zigzag paths in the aperture diagram are of no conse-
quence for perfect imagery. They follow from the requirement of the lens
system that the aberrations caused by the refractions shall be sufficiently
small that they do not disturb the imagery.
5 Zimmer, Geometrica Optics
66 Perfect Imagery

Fig. 27 shows the conductivity diagram of a Tessar 2.8/50. This


objective, for use in miniature photography, has a focallengthJ = 50 mm
and a relative aperture 2h/J = 1/2.8. It consists of a converging lens and
a diverging lens in front of an iris diaphragm, which regulates the intensity
of irradiation in the film plane, and a converging cemented component
consisting of a diverging lens and a converging lens behind the iris
diaphragm. The linear conductivity LLW = - 3.63 mm.
The perpendicular diagram cuts the B-axis at B = 20.3 mm. If the
initial and final lines are produced (the dotted lines) they meet at the
principal point H. This marks the position of an ideal thin lens which
would give the same perfect imagery as the whole objective. There is
exactly one such point: a distinction between front and back principal
points is only made when their geometrical positions are given relative
to the front or back surfaces of the objective. The points P and pi
indicate the position and size of the entrance and exit pupils. The position
and size of the iris diaphragm are given by the point of intersection of
the perpendicular diagram with the A-axis. In the case of the Tessar,
as in the other examples, the departure from the dotted diagram enables
the aberrations to be removed.

§ 19. Imagery in Depth


Depth ofImagery, Depth of Focus,
Depth of Accommodation
All the above theory has been based on the assumption that the light
tube is defined by a two-dimensional emitter and a two-dimensional
receiver. In practice, objects are rarely confined to a plane: the normal
case is three-dimensional. In the theory of perfect imagery, this can be
taken into account by investigating the imagery of objects with different
distances from the receiver. Once again, the assumption is made that all
these objects are two-dimensional. Thus imagery in depth is not concerned
with three-dimensional radiators but with plane radiators disposed in
depth. The theory requires that they be considered as transparent which
means, in practice, that they may be treated as objects lying adjacent
to each other in parallel planes.
The transfer from an object 0 to another object 0* centred on the
same axis with the same conductivity of the light tube is defined by an
object shift Q. If the parameter w is removed from the formulae
A*=A +wB
n*U* = nU + wnW
Imagery in Depth 67

by multiplying the first equation by - nU and the second by A and the


forming the sum, we have
w(AnW - BnU) = An*U* - A*nU. (19.1)
For every case of perfect imagery, both the expression AnW - BnU
and (from theorem 3) the shift parameter w, are invariant. It follows that
the expression on the right hand side of equation (19.1) is independent
of the choice of reference point and has the same value in the object
space and image space. Furthermore, since it is independent of the sizes
of the object and image, it is well suited to the analysis of object depth
and image depth. However, it is convenient to relate a positive object
depth to the accommodation of the eye, so we reverse the sign of the
expression and define depth by
T = A*nU - An*U*. (19.2)
Depth has the dimension length and is expressed in mm or Rayleigh units.
o
o· p

Fig. 28

The depth of imagery T = ± 0.25 R.E. has considerable importance


in optical imagery. It has the same significance for the depth of the
emitter as the resolving power has for the height of the emitter perpendi-
cular to the axis and is closely related to it. In Fig. 28 an object 0 is
converted into 0* by an object shift. The figure shows T as positive.
There is no clearly defined region in the plane 0 corresponding to the
radiation from 0*. The effect is rather that for every point in 0* there
is a corresponding circle of confusion in o. All these circles of confusion
have about the same size. The optical effect of the one corresponding to the
centre of 0* is measured by the length of the perpendicular A * from the
centre of 0 to the marginal ray of the object 0* which, together with
the given pupil P, thus gives the linear conductivity - A *nU; for at
this point A = 0 and A * assumes the role of B for the circle of confusion.
If the reference point is placed in the centre of 0*, then A * = o. The
circle of confusion for the centre of 0 in the plane of 0* is measured by the
perpendicular A on to the marginal ray of the object 0 and gives the
linear conductivity - An*U*. If we change from this specific reference
S*
68 Perfect Imagery

point to some other one, then the conductivity naturally retains its
numerical value and assumes the form given in (19.2). Accordingly, we
can interpret the expression for the depth T as the linear conductivity
with which the circles of confusion from 0 are imaged in the plane of 0*;
or (apart from the sign) as the conductivity with which the circles of
confusion from 0* are imaged in the plane of O. Thus if the absolute
value of T is less than one quarter of a Rayleigh unit, then the unsharp-
ness of the first object is less than the Rayleigh limit when the other is
sharply imaged. This means that the unsharpness is less than the lower
limit for the rules of perfect imagery. It follows that the geometrical
difference in the position of the objects will not be rendered visible by
optical imagery.
Care is necessary in interpreting depth of imagery in practice. In
special cases, objects whose separation is less than the depth of imagery
can nevertheless be separated optically by means of interference techni-
ques. But the relative amount of energy originating from the two emitters
can no longer be measured. This implies that the two objects are effectively
seen as one. If two objects are separated by more than ± 0.25 RE. from
a mean focus position - that is, more than half a Rayleigh unit from
each other - then they are, in many cases, visibly separated. This is
important in the theory of aberrations. Suppose a lens system yields a
red image separated from a blue image; then this colour error can be
detected when the separation between the two images is more than
0.5 RE. The effect only becomes serious with larger values. Our eye is
fortunately less critical here than a physical measuring instrument. The
special characteristics of the eye make it impossible to give a precise
limit for the depth at which two objects can be recognized as separate.
An example of such special visual effects is the different resolution of
points and lines to which we have already referred. The data given by
E. ABBE and M. BEREK for the calculation of depths of focus in the
microscope correspond to a depth of about 2 RE. This is probably a
reasonable value for microscopic objects with low contrast.
As an example, we shall calculate the depth of imagery of the human
eye when observing an infinitely distant object. We thus have the con-
dition nU = O. The reference point is placed in the centre of the pupil.
In this plane for daylight vision we have A = A * = 1.5 mm. From
T = - An*U* = ± 0.25 RE. = ± 0.00014 mm
and n* = 1 it follows that U* = ± 0.0000923. The limit of the depth of
imagery is at A*JU* = ± 16250 mm = ± 16.25 m. The positive limit
gives us no practical advantage since it corresponds to an object lying
behind our head. The above result shows that we can see objects clearly
without accommodation down to a distance of just over 16 m, when our
Imagery in Depth 69

eyes are focused on very distant objects. If we focus on an object 16.25 m


away, we can make use of the depth of imagery in both plus and minus
directions and see objects between infinity and 8.12 m sharply without
accommodation. If we focus on an object at a distance of 250 mm, then
nU = - 1.5/250 = 0.006 and n*U* = - 0.006 ± 0.0000923. The depth of
imagery then indicates that we can observe objects in sharp focus which
lie within the range 246 mm to 254 mm. It is not easy to confirm these
values experimentally because we cannot normally prevent an involun-
tary accommodation. Nevertheless, simple observations demonstrate that
our eyes have a finite depth of imagery: for example, in taking a bearing
by lining up two marks at different distances from the eye, or in observing
through the holes of a net curtain. In the latter case, the eye is little
disturbed if the curtain is very close to the eye, but the observation of
distant objects can become impossible if the curtain is at an appreciable
distance.
The values given above for the depth of imagery are apparently
contradicted by our ability to estimate whether an object is 10 m, 100 m
or 1000 m away. But this is a consequence of binocular vision combined
with experience in comparing the sizes of known objects and allowing
for the effects of atmospheric haze. These are physiological and psycho-
logical effects which do not belong to the physics of optical imagery.
If the pupil is reduced in size, the depth of imagery extends over a
larger region: it increases as the reciprocal square of the pupil perpendi-
cular. With a third of the normal daylight pupil (i.e. A = 0.5 mm) the
depth of imagery extends for a mean accommodation of 1.80 m from
infinity to 0.90 m; with a pupil perpendicular A = 0.25 mm and the eye
focused for 0.44 m, the range is from infinity to 0.22 m. It is easy to
demonstrate this effect, at least qualitatively. A pinhole is made in an
opaque card and is pressed as close as possible to the eye. It is best to
observe through the window of a room which is not too brightly illumi-
nated, since the eye is then presented with sufficiently bright objects
while maintaining a pupil large enough for the experiment.
If we look through an 8 x 30 binocular in normal daylight, the
effective pupil perpendicular A in front of the instrument is 8 x larger
than that at the eye. The depth of imagery is the same in front of and
behind the binocular, but the corresponding geometrical distance in front
of it is only 1/64 of that behind it. If the mean focus of the binocular
is 1040 m, the region from infinity to 520 m appears sharp; but for a
focus point 10 m away the range is only from 9.9 m to 10.1 m.
For the calculation of the depth of imagery in the microscope we
place the reference point in the object plane for which A = O. The
entrance pupil is so far away from the object that we can assume
nU = n*U*. It follows from the expression A*nU = ± 0.25 R.E. that
70 Perfect Imagery

the depth of imagery is given by the perpendicular to the edge of the disc
of confusion A * = - 0.00014 mm/nU, and the geometrical extent of the
depth of imagery is thus A */U* = ± 0.00014 mm/nU2. In Table 5 values
for this depth are given in the fourth column as a function of the refractive
index and the numerical aperture nU. The last four columns refer to the
depth of accommodation in the microscope and will be explained later.
The size of the eye pupil has no importance for the depth of imagery in
the microscope provided it is larger than the exit pupil of the instrument.
This is usually the case. The values given are the smallest values according
to the theory. In practice, values between twice and four times as great
may be assumed, depending on the object structure.

Table 5
Refrac- N urn. Depth of Total Depth of Total Depth of
tive aperture imagery magnif- accomo- magnif- accommo-
index ±A*,U* ication dation ication dation
n nU U [mm] (3N= A*,U* (3N = A*,U*[mm]
500nU [mm] 1000nU

1 0.16 0.16 0.00541 80 0.03906 160 0.00976


1 0.32 0.32 0.00135 160 0.00976 320 0.00244
1 0.64 0.64 0.00034 320 0.00244 640 0.00061
1 0.90 0.90 0.00017 450 0.00123 900 0.00031
1.5 0.96 0.64 0.00023 480 0.00163 960 0.00041
1.5 1.35 0.90 0.00011 675 0.00082 1350 0.00021

In photography, the depth of imagery can formally be calculated exactly


as in microscopy. But since the perpendicular A * = ± 0.00014 mm/nU
only becomes as large as the film grain with very small apertures,
the depth of imagery is of no practical significance. It is replaced by the
"depth of focus". This is defined by allowing a disc of confusion up to
A'* = ± z in the plane of the image (A' = 0). The value of z depends on
the film material used, the type of effect the photographer wishes to
produce and, of course, the sUbjective nature of the term "unsharp".
Nevertheless, the value of z is always at least as great as the radius of a
film grain. The depth of focus in the image space ST = - A'*n'U' =
± zn'U', in which n'U' represents the aperture on the image side.
Because of the invariance of this expression, the depth of focus on the
object side must naturally have the same value. The corresponding
ranges must be calculated using the expression ST = A*nU - An*U*
from the given values of A and U. The photographer makes use of
depth of focus tables or depth of focus scales on his camera. Consider a
camera focused at infinity as a simple example of such a calculation.
Then nU = 0, n'U' = A//, in which f = 1/l{! is the focal length of the
Imagery in Depth 71

objective. Then ST = - An*U* = ± zA/J. For the limits of the range


of sharpness we have A*/U* = ± A*f/z. In the plane of the entrance
pupil A * = A and 2A/f = l/b is the relative aperture of the objective,
b being the corresponding stop number. Then ~: =± 2~b. For a mini-
ature camera objective with f = 50 mm, stop number 5.6 and z = 0.01
mm, the resulting value is A*/U* = ± 22.3 m. The positive value is
useless because it lies "beyond infinity". Focused in this way the camera
yields "sharp" images in the range from infinity to 22.3 m. If your
camera indicates instead a value 7 m, about a third of our value, then the
manufacturer has assumed a disc of confusion three times as large,
i.e. 0.03 mm.
When an instrument is used visually, the depth of accommodation of
the eye must be taken into account as well as the depth of imagery.
This is the depth which the normal eye can accommodate and corresponds
to an object shift from infinity to a point 250 mm in front of the eye.
If the reference point is placed in the centre of the eye pupil, then
A * = A, nU = 0 and n*U* = - A */250 mm. The depth of accommoda-
tion thus has the value
AT = A2/250mm.
It is proportional to the square of the pupil aperture and has the value
0.009 mm = 16 R.E. for daylight vision with A = 1.5 mm. For daylight
vision its value is thus 32 times larger than the depth of imagery; for
A = 0.5 mm it is 3.5 times larger and for A = 0.25 mm it is only 0.9 times
the depth of imagery. With an 8 x 30 binocular it is possible to see
sharply over the range from infinity down to 16 m on account of the depth
of accommodation, and this range is independent of the size of the eye
pupil. This can be demonstrated as follows. Let A' be the perpendicular
in the eye pupil and let the reference point for the calculation be placed
r
in the entrance pupil; then in this pupil A * = A = A' in which is r
the telescope magnification. With nU = 0 we have
AT = - An*U* = A'2/250 mm from which
A*/U* = - AA*·250 mm/A'2 = - P·250mm = - 16m for r= 8.

In the microscope, the depth of accommodation at the object covers


the range A */U* = n x 250 mm/{{J·N)2. n is the refractive index in the
plane of the object, {J·N is the total magnification of the microscope.
Numerical examples are given in Table 5 on page 70. To derive the formula
we place the reference point in the object, for which A = 0, nU = n*U*
so that
AT = A*nU = A'2/250 mm.
72 Perfect Imagery

The expression (12.3)


A' = _ nU·250mm
p·N
gives the relation between the perpendicular A I in the exit pupil, the aper-
ture nU and the total magnification. Thus A * = nU x 250 mm/({J' N)2
from which, after division by U* = U, we arrive at the formula given
above.
Depth of imagery, depth of focus and depth of accommodation have
considerable importance in the application of optical instruments to
observation and particularly to measurement. Nevertheless, it is q:uite
wrong to follow slavishly the given values for these quantities since they
are not only affected by large variations in individual visual capacity
but also depend on the term "sharpness", which is based on subjective
perception.

§ 20. Summary of Part One


In the preceding sections, I have given the basic theory of perfect
imagery and a number of examples which seemed necessary to its under-
standing. The versatility of the theory and its wide range of application
can only be appreciated when it is illustrated by many practical examples.
The theory is indeed of little value unless it can be widely and usefully
applied to practical instruments. However, I have not attempted to
write a compendium for specialists in optical instrumentation: my aim
was rather to describe a theory of optical imagery which is adapted to
practical problems.
Although the development of this theory is not fundamentally new,
it nevertheless departs radically from the normal presentation. For this
reason, it seems useful to summarize here the basic points.
Radiation has the property of transferring energy through space.
During this process interaction occurs, in the general case, between the
radiation and the matter filling the space. For example, part of the radia-
tion energy may be absorbed and appear as heat. If the matter has no
influence on the amount of energy transmitted by the radiation and only
changes the geometrical shape of the part of the space it traverses, then
very simple laws emerge for the propagation of the:'radiation. These are
the laws of geometrical optics. '
In general, the purpose of geometrical optics can be expressed as the
determination of the form of the volumes of space in which the propaga-
tion of the radiation' takes place, assuming conservation of energy. The
problem can be solved with the aid of elementary rules if the emitter and
receiver are plane 'and the light tube formed between them is centred.
In these conditions, the invariance of the linear conductivity LLW =
Summary of Part One 73

AnW - BnU, which is defined by the refractive index n of the material


and the coordinates A, B, U, W of the light tube, is a criterion for the
conservation of energy during the propagation of the radiation.
The invariance of LLW can be demonstrated experimentally. The
most satisfactory proof is provided by the existence of optical instruments
such as telescopes and microscopes of such excellence that they yield
all the information which can theoretically be expected. A formal proof
is not possible: for the invariance of the linear conductivity is analogous
to the constancy of mass in mechanics and only represents a good
approximation for the range over which we normally conduct our experi-
ments. The invariance breaks down when the value of the conductivity
becomes very small. Below one Rayleigh unit the departure from geo-
metrically perfect imagery becomes evident in the form of diffraction
effects; at one quarter of a Rayleigh unit, corresponding to the resolving
power of optical instruments, the form of the actual image is markedly
different from the geometrical image.
The analysis in Part 2 of this book is concerned with aberration theory
and shows that the invariance of the linear conductivity also breaks down
when the conductivity becomes large. This occurs when the emitter and
receiver are both large and near to each other.
The technical application of optical imagery is based on the assump-
tion that LLW = const. The relevant theory is elementary and is given
in sections 7 to 9. The two most important points are that the refraction
formula (7.7) follows as a consequence of the invariance of LLW and
that any transformation of the light tube involving no loss of energy can
be achieved by a combination of refractions and transfers. The abstract
theory thus indicates how technical realization of a design project may
be possible.
In sections 10 to 14 several optical instruments are analysed to
illustrate the theory of perfect imagery. The method given in elementary
text books for the determination of the position and size of an image by
means of the foci and principal planes has deliberately been avoided
because it leaves the role of the pupil obscure. I have emphasized the
energy conditions and tried to show how they determine the basic design
of an optical system. It becomes clear that the position and size of the
pupils are just as important as the corresponding values for the object
and the image. Only when image and pupil are considered together - that
is, when the changes in the form of the light tube are followed through
the system - is it possible to define the energy conditions correctly and
avoid the danger of formulating requirements which contradict the laws
of nature just as radically as perpetual motion machines. The ultimate
purpose of any optical system is to adapt a given light tube in the object
space to the characteristics of the receiver. For this reason, we included
74 Perfect Imagery

the receiver in our discussion of optical instruments and chose in 'our


examples those in most frequent use - the human eye and the photo-
graphic emulsion.
Sections 15 to 18 are concerned with methods of design of ideal optical
systems. The aberrations which occur in any practical instrument using
refraction by lenses are not considered. The most important technique
for practical work is the iterative computing scheme given in section 15,
or some equivalent depending on the computer aids available. The explicit
calculation technique using Gaussian brackets gives the powers and
separations for systems consisting of two or three lenses; but even with
three lenses it becomes unwieldy. Consequently, I have thought it better
to omit this treatment. The derivation of the formulae is not difficult.
The differential formulae (16.11) are useful if electronic computers are
available for design work.
Graphical methods have been discussed in detail to an extent which
their practical importance hardly justifies. But this is the only opportu-
nity we have in "geometrical" optics to indulge in "geometry", and
I make no attempt to hide my weakness for this approach. Several
relationships in aberration theory only became clear to me through study-
ing the conductivity diagram of Delano. The fact that any example of
perfect imagery can be completely defined by a two-dimensional diagram
demonstrates the basically simple structure of this theory. This simplicity
is expressed algebraically by the "principal theorem" proved in section 9,
which states that any form of perfect imagery can be represented by two
fundamental processes.
One assumption of perfect imagery, that the object is plane, is rarely
true in practice. It is thus important to establish the rules applicable
to imagery in depth. The concept of object shift leads to the definition
of an optical measure of depth. It is an invariant of the imaging process
and therefore has the same value in the object and image spaces. The
geometrical extent of this depth can readily be calculated. In applying the
formula, it is important to distinguish between the depth of imagery
of ± 0.25 Rayleigh units determined by the nature of the radiation and
that governed by the properties of the receiver. Examples of the latter
which are of special significance are the depth of focus in image recording
on photographic emulsions and the depth of accommodation in visual
instruments. The depth between two images of the same object has
considerable significance in the theory of aberrations: for, together with
the departure of the conductivity from strict invariance, it provides the
link between the properties of imagery by a real lens system and those
of perfect imagery.
PART TWO

Theory of Aberrations
Aberrations describe the difference between the effects of lenses and
the predictions of perfect imagery. They are thus a measure of the
imperfection of practical imagery with lens systems; their relationship
with the system parameters is the basis of optical design. The following
sections give an inductive introduction to this theory.

§ 21. Aberrations and Geometrical Optics


Significance and Limitations of the Theory
The term "theory of image aberrations" is basically misleading since
geometrical optics is not normally concerned with the problem of the
formation of a physical image. An analysis of image formation in the
physical sense requires information concerning the nature of the radia-
tion: in the case of electromagnetic radiation we must know not only
the form of the waves but also their amplitude and polarisation. The
change in these quantities must be investigated for every refraction in
the system. Such calculations are indeed performed in certain special
cases, but the word "aberration" has normally a very much simpler
meaning: the departure from perfect imagery which arises if a specific
assumption is made with regard to the kind of image formation. For this
reason, there are many different types of aberration and the theory of
chromatic aberrations is quite different from that of monochromatic
aberrations. Furthermore, the theory of monochromatic aberrations of
light microscopes is quite different from that of electron microscopes.
In the modern sense of the word, an aberration is not a property of
the physical image but a theoretical aid to its analysis. A complete
analysis requires knowledge of the interactions of all conceivable aber-
rations. For this reason, it is difficult to say what a single aberration
"looks like" or to define its physical significance. Its appearance in a
physical experiment or in an optical instrument varies enormously
according to the special characteristics of the object, the illumination
and technique of observation. In general, it cannot be analysed by
geometrical optics because this theory does not possess the necessary
physical substance.
76 Theory of Aberrations

In practice, calculations are limited to a small number of aberrations.


The "art" of good optical designers is the ability to produce the desired
performance by manipulating a minimum of aberrations. The reason for
this is not - as is often supposed by the layman - the effort involved
in the calculation of the aberrations, since this hardly of any consequence
if modem computers are used. The real reason is that it is only possible
to analyse the results of experiments and discover unknown relationships
if the theory is clear and easy to manipulate - and this is not compatible
with a large number of aberrations. This is the technical significance of
aberration theory in geometrical optics. The systematic development of
new types of high quality optical systems has only been possible through
the application of this theory in spite of its inability to describe the
physical processes of image formation.
The optical designer approaches his problem rather as a mechanical
engineer designs a new machine by considering the statics and dynamics
of solid bodies rather than the latest discoveries concerning the physical
structure of the materials. The engineer then estimates whether the parts
used are capable of withstanding the forces involved. The optical designer
starts from the basic principles of perfect imagery. Then he tries to
achieve a sufficiently good approximation to this with lenses whose form,
position, refractive index and dispersion are determined according to the
rules of aberration theory. Because of the necessary accuracy, the work
is not carried out at the drawing board but by calculation; but it is just
as futile to try to replace the optical designer by a computer as to try to
replace a mechanical engineer by a programmed drawing machine. The
aberrations are calculated for the lens system the designer has set up
and are assessed with reference to the quality of imagery required.
In practice, the development of an optimum system does not depend on
the amount of calculation performed but essentially on the skill of the
experienced designer in controlling the aberrations and his knowledge of
the relationship between aberrations and image quality.
This knowledge is gained from precision measurements and extensive
calculations based on physical theory for certain special cases. During
the last hundred years, a considerable amount of such material has been
accumulated so that there is today little risk of a false assessment of the
aberrations. Nevertheless, when faced with new or unusual problems the
designer must investigate the relationships afresh, either theoretically
or experimentally. (It may be a relief to the reader if we disclose on the
sly that new systems are always good: the aberrations only become
noticeable or objectionable when a competitor brings out a better
instrument!)
As in every technical discipline, the results obtained from mathe-
matical formulation (aberration theory) do not agree exactly with the
Optical Glasses 77

observed image. This discrepancy is bridged over by empirical methods.


Of course, numerous attempts have been made to provide a theoretical
solution. Criteria of image quality such as resolving power, the Strehl
intensity ratio, wave aberrations and the modulation transfer function
have been developed. Each of these criteria is satisfactory in certain
cases but all of them have the common weakness that they are not
universally applicable. In this book, I shall confine myself to a treatment
of elementary aberration theory.

§ 22. Optical Glasses


Specification by Refractive Index,
Dispersion Number and a-Values
Glass is produced by heating a mixture of silicic acid (in the form of
pure quartz sand) with alkali and metal oxides to a temperature of
about 1400°C. Silicates are then formed which combine with silicic acid
to give a complex structure which is still not fully understood. When
the mass is carefully cooled, the hard amorphous material we call glass
results. For use in prisms and lenses the glass must be particularly pure
and homogeneous. This is why optical glasses are manufactured in special
plants, the most important producer in West Germany being the Jenaer
Glaswerk Schott & Gen. in Mainz.
Before considering the essential optical properties of the glasses, we
shall briefly mention other points that concern the manufacturer and the
optical designer. Firstly, bubbles in the block of glass. These cannot be
entirely avoided because the escape of gases from the melt is an essential
part of the production process. The manufacturer attempts to keep their
size and number to a minimum. Often they have only an aesthetic
importance in optical instruments, for generally "lenses are there to be
looked through, not at" (FRAuNHoFER). More serious for the imagery
are striae, inhomogeneities which may have a threadlike form or may
cover a certain area. The change in colour of the light due to absorption
of certain spectral regions can become very significant with thick lenses
or prisms. For some types of glass a yellow colour due to absorption of
the shorter wave region is inevitable because of the chemical composition.
Absorption in the UV is a more general form of the same effect. All glasses
absorb UV light, but a few still let more than 50% of the radiated power
through at 300 nm with a glass thickness of 5 mm. At the other extreme,
some glasses let through less than 1% at 360 nm. A very important
property is the resistance of the polished surface to chemical change, in
particular weathering caused by dampness in the air and solution by
acids. Sensitive glasses are even attacked by the acid contained in the
78 Theory of Aberrations

sweat from human hands. For this reason, the polished surface of a lens
should never be touched. There are even a few glasses which contain
thorium and are therefore radioactive. In Germany the protection laws
are so strict that it is virtually impossible to use these. Finally, a charac-
teristic of great importance is the price. We shall quote two examples,
even though these prices may no longer be correct when this book appears.
Since the density varies between 2.28 g/cm3 and 6.18 g/cm3 , a kilogram is
contained in cubes with sides of lengths between 76 mm and 55 mm.
If one purchases from the Schott glass works, the lowest price for a
kilogram lies between $ 3.65 and $ 5.40, according to the terms of sale.
For the most expensive glass, including supplementary costs for fine
tolerances in the refractive indices, the price is $ 455 per kilogram.
The optical properties of glasses are defined by the refractive indices
n(..1.) for chosen wavelengths..1.. The refractive index is defined as the ratio

n(..1.) = s~n i
SIn 1.'

of the sines of the angles of incidence i and i', which are formed by the
wave normals and the surface normals before and after refraction at a
plane polished surface separating air (at 20°C and 760 Torr) from the
glass. The measurement of the refractive index is thus reduced to an
angle measurement carried out in air under normal conditions. This is
convenient because these are the conditions in which lenses are normally
used. In special cases, the above definition must be borne in mind; it
may be necessary to adjust the refractive indices for another reference
medium. For . 1. = 546 nm the refractive index of air at a pressure of
760 Torr relative to a vacuum is 1.00029 at oDe and 1.00027 at 20°C.
In the visible region, it is less sensitive to changes of wavelength than to
normal temperature variations.
Precision measurements of refractive indices can only be made if
monochromatic sources with small band width and high radiation density
are available. Formerly, sodium vapour lamps or even salt in a Bunsen
burner flame were used. The strong sodium yellow doublet (..1. = 589 mm)
was the standard for many years and was termed D. The corresponding
refractive indices were given the symbol nD. Because of its width this
doublet is not capable of giving the accuracy required of modern measure-
ments. Up to a decade ago, the commonest lines for measurements were
the helium d (.J. = 588 nm), the red hydrogen line C (..1. = 656 nm) and the
blue hydrogen line F (..1. = 486 nm).
Since then metal vapour lamps have tended to replace Geissler tubes
because of their greater technical convenience. The most convenient lines
for measurement with these lamps are given in the following table: -
Optical Glasses 79

Table 6
Wavelength Symbol for Chemical Colour
(nm) the line element

365 i Hg (ultra-violet)
405 h Hg violet
436 g Hg blue-violet
480 F' Cd blue
546 e Hg green
644 e' Hg red
1014 t Hg (infra-red)

The choice of these lines has the advantage that they can all be produced
with a single cadmium-mercury lamp.
In general, the refractive indices are measured at the glassworks and
this information is passed on to the purchaser. Modern techniques can
reduce errors of measurement to as little as 10-6 , but this accuracy can
rarely be exploited in practice because the optical homogeneity is con-
siderably worse in spite of the utmost care in manufacture. The refractive
indices vary from melt to melt, or even for different pieces of the same
melt, up to ± 10--4. Consequently, the values are normally measured
with accuracy of ± 10-6 • Unless one is prepared to pay the manufacturer
a supplementary charge for the selection of pieces which lie particularly
close to the catalogue value, variations up to ± 1 x 1()-3 between diffe-
rent deliveries must be accepted. These examples show what can be
achieved by modern techniques, but also provide a salutary reminder
that it is useless to push a design to the theoretical limit without consider-
ing the effect of normal refractive index variations and lens manufacturing
errors on the quality of the system.
Fig. 29 shows the refractive indices of Schott glasses F2, SK16 and
K50 as a function of wavelength. It is evident that these "dispersion
curves" show marked differences in position and slope. Many formulae
have been suggested for these curves, of which the best known is due to
DRUDE who assumed resonant frequencies for electrons in the ultra-violet
and for ions in the infra-red. He derived the formula

ABe
+ ,,.. _ ).2 + r-=1:2 '
2_
n - 18 UY

in which the coefficients A, B, C, AIR and Auv must be determined in such


a way that best agreement is obtained with measured values of the
refractive indices. This is not easy in practice and it is only worth while
if very precise measurements are available. Furthermore, this formula
cannot be used for extrapolation beyond the range of the measured values
80 Theory of Aberrations

since the model on which it is based represents a considerable simpli-


fication of the real dispersive properties of glasses.
A particularly simple and convenient formula is that due to HART-
MANN:

n(J.) = no + J. ~ J.o.
This is an empirical formula which has no direct physical basis. Never-
theless, experiment has proved that it is capable of giving the dispersion
curves of optical glasses in the visible region from I. = 405 nm to I. = 644
nm with a maximum error of ± 1 x 10-4 • The best agreement is found
for crown glasses, with a maximum error of ± 2 x 10-5 • The error

F2
~65

SKI.
~60 F2

~5 400 600 800 l000nm


Fig. 29

increases to ± 5 x 10-5 for flint glasses, the worst agreement being


for the heavy flints. However, extrapolation beyond the range given
above leads rapidly to large errors. The coefficients no, c and 1.0 can be
readily calculated from the refractive indices for three wavelengths, a
convenience which accounts for the wide use of the formula in practice.
The coefficients for the dispersion curves shown in Fig. 29 are given in
Table 7:-
Table 7
Glass no c [nm] J. o [nm]

K50 1.503755 7.872 169.4


SK 16 1.5981 II 9.328 169.4
F2 1.587059 12.543 207.7

The theory of chromatic aberrations is not based on dispersion


formulae. Values of the aberrations for a few fixed wavelengths are quite
Optical Glasses 81

adequate. Intermediate values can readily be determined to a sufficient


accuracy by interpolation.
Refractive indices for a number of wavelengths completely define the
dptical properties of a glass, but their significance does not emerge
clearly from the indices themselves. Auxiliary parameters are therefore
constructed which give a clearer picture of the optical properties. The
first is the mean refractive index, often referred to for the sake of brevity
as the refractive index. Nowadays, n. has replaced nd, which had, in turn,
replaced nD. The choice is naturally somewhat arbitrary but can be
justified on the grounds that this line lies in the region of the maximum
sensitivity of the eye and can be readily produced with high intensity by
a mercury vapour lamp.
The mean refractive index defines one point on the dispersion curve.
The slope of the curve is defined by the difference of refractive index for
two chosen wavelengths. Earlier the lines C and F were used, nowadays
C' andF'. The mean dispersion nF' - ne' is defined as the second auxiliary
parameter. This quantity occurs in the formulae which will be derived later
for the chromatic aberrations and is therefore useful for their calculation.
However, the real task of the optical designer is not to determine the magni-
tude of the chromatic aberrations but to decide which glasses are capable
of removing them or, at least, reducing them to acceptable values. He is
obliged to use commercially available glass and cannot simply select a
convenient value of the mean dispersion for a given desired value of the
mean refractive index. When the mean refractive index is changed, the
corresponding parameters rp and t3 which had been determined to give
the optical system the necessary characteristics for the "principal wave-
length" (A = 546 nm) are also changed. In practice, one needs therefore
a definition of the dispersion property which is independent of the choice
of the mean wavelength. In the general case, this is not possible. But the
requirement can be met for a thin lens in air, or for a refracting surface
between glass and air, by the simple expression
np' - net
n. - 1

The proof will be derived later from formulae for the chromatic aber-
rations. Since the influence of the lens thicknesses on these aberrations
can be neglected in very many cases, practising designers find the above
formulation the most useful.
It is somewhat inconvenient that the numerical values of

(nF' - nd/(n. - 1)
are so small. Values are given in the following table: -
6 Zimmer, Geometrical Optics
82 Theory of Aberrations

Table 8
Glass Mean Principal np' - ne' Abbennmber Dispersion
refractive dispersion n. - 1 n. - 1 number
n. nF' - no' np' - no' np'- ne'
1831~

K50 1.52465 0.00875 0.01668 60.13 30.53


SK 16 1.62287 0.01037 0.01665 60.06 30.48
F2 1.62410 0.01729 0.02770 36.10 50.73

For this reason the reciprocal


n. - 1
Ve =
nF' - no'

known as the Abbe number, was introduced. The Abbe number of optical
glasses lies between 80 and 20 and is used, because of this convenient
range of values, as the second coordinate for the graphical representation
of the glasses. It should be noted that a small Abbe number means a high
dispersion and vice versa. The convenient range of values is, unfortuna-
tely, the only practical aspect of the Abbe number. It is very incon-
venient not to be able to compare the effects of two glasses on the
chromatic aberrations directly from the difference of the Abbe number:
to do this, we must calculate the difference between the reciprocals.
But since estimates of such effects are constantly required, it is better
to use the expression (nF' - nO')/(ne - 1) and bring the values into a
more convenient range by multiplying by an appropriate factor. I suggest
the number 1831 = 1/0.000546, since this is the factor needed to express
the chromatic aberrations in Rayleigh units. This conversion must any-
way be undertaken whenever we wish to check what effect the chromatic
aberrations will have on the image. Thus, we kill two birds with one stone
if we decide to introduce this factor from the start into our definition of
dispersion. According to this scheme, the dispersion number of an optical
glass has the value
me = 1831 np' - ~c' . (22.1)
n. -
Thus defined, the dispersion numbers lie - similarly to the Abbe
numbers - in the range between 20 and 90; but glasses with a low
dispersion have a low number, those with a high dispersion a high number.
Optical glasses with dispersion numbers smaller than 36.6 are termed
"crown glasses", those with a higher dispersion "flint glasses". Fig. 30
shows the mean refractive indices ne and dispersion numbers me of the
optical glasses available from Schott. In addition, the position of fluorite
is shown at the extreme bottom left hand comer. The arrows attached
to many of the glasses will be explained later.
Optical Glasses 83

The dispersion number can be readily calculated from the refractive


indices of a glass. On the other hand, knowledge of the mean refractive
index and the dispersion number does not define the indices nF' and no'
but only their difference. The indices themselves are defined if a further
quantity is known, for example the difference nF' - n •. How is it that
we can normally manage without this third quantity in our glass classi-
fication? Suppose we construct a dispersion number
mZll = 1831 nm - nv
n, - I
analogous to me for two lines x andy, for example for F' and e, then we
can determine coefficients IXZll and fJZll in such a way that the equation
(22.2)
is roughly true for all glasses.
Equation (22.2) is not a physical law. There are transparent materials
- an important example is fluorite but there are also some optical
glasses - which show marked deviations. But the equation is satisfied
within the accuracy of measurement for the majority of optical glasses.
Such glasses are called "normal glasses". The following table gives the
coefficients of the normal glasses for several pairs of lines: -

Table 9
Line x Liney (X • • {l ••

h F' 1.1567 - 7.052


g F' 0.5698 - 2.691
F' 0.5417 - 0.989
C' 0.4583 0.989
C' - 0.5749 - 8.966

It must be emphasized that these coefficients have been fixed more


or less arbitrarily. They are the result of the practical requirement that
as many commonly used glasses as possible should be "normal" rather
than of an abstract mathematical least squares process.
If the formula (22.2) were rigorous, it would be possible to calculate
the refractive indices of glasses for any wavelength from their values ne
and me, and Fig. 30 would give a complete optical specification. But in
fact they satisfy the equation
(22.3)
in which (]ZlI is not a universal constant but a quantity which varies from
glass to glass. Since (]ZlI is negligible for many glasses over the visible
spectral region and is a monotonic function of the wavelength outside
6*
Zinc crown. . . . . . . .. ZK 0
Barium light brown . BaLK + JS 'IIJ 9S 50 5S GO 6S 70
Short flint ......... KzF 0 /6
Short flint spec. glasses KzFS D.. US 11 1 l1..9S :?:
All other glasses ..........•
m. ~ 1831 nF'-nO' ~ 1831 ,,"
n,-l v, (.901 1 I l1..9Q
Secondary spectrum: -- /1
Deviations <T, and <T. TIe
~ from the normal values 1.8S ,~ 1.8S
St=8..97
~ S,~8.97 for H- G' and
«> S.~7.05 for h-7-P' are
>='
shown as follows by t80I-f--+-.c+-.c'---+------'f----+-----+--L.....~....L-.,L.+_----
arrows;- 21-
~o ..L'1"i .Q LaF
e. 1 1 1o;.'3,,_7.0S /"
1.7S ~ I~H~7_~____+_---~~=r=-.--~---

_Sf
'KJ
t 1.701 .~---+
ISU ss
___~f--~~--~.QL-4---~H~~~~--___~~~~~_+-~~_+---~
!
g,
i
~
Po
~
'-<
g f
~
~

w ~

w _

! !! !!! I! I 't80
80 8S .90 .
~ JS 'IIJ % 50 me
me
The Chromatic Aberrations of Perfect Imagery 85

the range from C' to pI, two suitably chosen a-values are sufficient to
give a clear and accurate picture of the properties of a glass. In Fig. 30,
the symbols all and at indicate the values allP' and ate' and are shown as
the horizontal and vertical components of an arrow. The length and
direction of the arrow give the magnitude and direction of the departure
of the glass from the n. - m. plane of the normal glasses. The signi-
ficance of the a-values for the chromatic aberrations will be explained
in § 25.

§ 23. The Chromatic Aberrations of Perfect Imagery


Primary Longitudinal Aberration and Difference
of Magnification
The definition of the primary chromatic aberrations is based on the
approximation that the optical imagery of the system can be accomplished
by means of the parameters If! = (n' - n)lr and t5 = - dIn, as explained
in § 15. The fact that the complete effect of a refracting lens surface
cannot be defined by a parameter If! which is independent of the coordi-
nates of the light tube will be ignored for our present purposes. It follows
that the results of this theory will only agree with experiment if other
aberrations are not present. If this is not the case, there is no point in
reducing the chromatic aberrations to very small values; indeed, this
can in certain circumstances make the image seem worse.
The parameters If! = (n' - n)lr and t5 = - dIn always represent perfect
imagery, but since they are functions of the refractive indices of the
glasses, the images for lines x and y do not generally coincide. Suppose
we start from the same fixed light tube in the object space for both
wavelengths and compute the light tubes in the image space using first
the refractive indices nx and then n y. We arrive at the coordinates
A~, B~J n~U~, n~W~
and

for the two wavelengths if the same reference point is taken. With these
coordinates we construct the chromatic aberrations
CHLOxy = 1831 (A~n~U~ - A;n~U~) R.E. ,
CHVO
xy
= 1000 (A:n!W!
LLW
- B~n;U; _ 1) 0100'
CHLPxy = 1831 (B~n;W; - B;n~W~) R.E., (23.1)
CHVP
xy
= 1000 (A~n;W;
LLW
- B;n~U~ _ 1) 0100·
86 Theory of Aberrations

These are called the longitudinal chromatic aberration and the chromatic
difference of magnification of the (perfect) object or pupil imagery. (In
modern literature, these aberrations are often simply termed the longi-
tudinal and transverse colour aberrations). The subscripts xy can be
omitted if it is clear from the context which pair of lines has been used.
Equations (23.1) follow immediately from the rules of perfect imagery.
In the object space the object with the "colour" x coincides with that of
"colour" y, but this is not necessarily the case in the image space. CHLO",y
gives the depth explained in § 19 between the image of colour x and that
of colour y. The value is in Rayleigh units for the wavelength A = 546 nm
because of the factor 1831 = 1/0.000546. For any given line, the linear
conductivity of the light tube is an invariant of the image forming
process. The chromatic differences of magnification give the changes in
the "coloured" conductivities per mill of LLW in terms of the object
coordinates of one colour and the pupil coordinates of the other colour.
To make this clear, we shall place the reference point in the image plane
for the colour corresponding to the line y. In this plane, A; = 0 so that

CHVO",y = 1000 (!i - 1) %0' Thus the chromatic difference of magni-


fication of the object imagery indicates by how much per mill, in the
plane of the image of colour y, the perpendicular B~ drawn to the prin-
cipal ray for colour x is larger or smaller than the perpendicular B;. If
we neglect the difference (generally small) between the length of the
perpendicular B' at the image plane and the image size perpendicular
to the axis, we can take the CHVO to be the difference of the image sizes
per mill. This explains the name given to this chromatic aberration. The
interpretation of CHLP and CHVP as the chromatic aberrations of depth
(longitudinal) and size (transverse) of the pupil imagery is entirely ana-
logous.
In any medium without dispersion (i.e. n", = ny), the chromatic
aberrations are independent of the choice of the reference point for the
light tube coordinates. This independence was implicit in our interpreta-
tion of the chromatic difference of magnification. It can readily be
checked by applying the transfer formulae A = A' - dU', etc. The proof
will not be given here because it follows later as a special case of equa-
tions (24.2).
Does the expression (23.1) give all the chromatic aberrations of perfect
imagery arising from two wavelengths? The answer, in a certain sense,
is that it does, although there are many possible ways of constructing
other chromatic aberrations from the light tube coordinates. But such
a procedure is valueless since it provides no further information.
This must be the case because the four aberrations CHLO, CHVO,
CHLP and CHVP, together with the coordinates of a light tube for
The Chromatic Aberrations of Perfect Imagery 87

the line y, for example, completely define the coordinates of the other
line:
A , A' CHVP A' CHLO B'
'" - 1/ = 1000 • 1/ - 1831 LLW' r'

B' B' CHLP A' CHVO B'


",- Y=1831LLW' 1/+ 1000 . 1/'

'U' - 'U' - CHVP . CHLO . 'w'


'u' _ 1831 (23.2)
n",,, nil 1/ - 1000 nil 1/ LLW nv 1/'

'W'
n" ",-n1/
'w'1/= 1831LLW'
CHLP n 1/'u'1/+ CHVO
1000
'w'11'
'nll

These expressions are obtained by solving equations (23.1) for A~ and B~.
If we suppose that the object is free from chromatic aberration and
that the conductivities for both wavelengths are the same, it follows from
the invariance of the linear conductivity for any perfect imagery that the
relation
(23.3)
must hold. This implies that the four chromatic aberrations of equations
(23.1) cannot be independent of each other. If now the values of (23.2)
are substituted for A~, B~ etc. in (23.3), the following relation results
for the chromatic aberrations: -

( CHVO ) (CHVP ) CHLO CHLP


(23.4)
1000 + 1 . 1000 + 1 - 1831 LLW • 1831 LLW = 1 .
Multiplying out the brackets, we have
CHVO CHVP CHVO· CHVP CHLO'CHLP
1000 + 1000 + (1000)2 (1831 LLW)2 .
(23.5)

In systems intended for practical use the chromatic aberrations must be


small. The product terms of two aberrations may then be neglected
compared with the aberrations themselves giving
CHVO"", - CHVP , (23.6)
an approximation which is useful in practical design work.
For a system intended for visual use, CHLO should be less than 1 R.E.
and CHVO less than 10%0 (one per cent) between the lines C' (red) and
F' (blue). If this is the case, it is not necessary to perform separate
computations of CHVO and CHVP. In practice, CHLO is often zero or
very small; for this is normally the first requirement of chromatic correc-
tion. If this has been achieved, the tracing of principal rays to determine
B' and n'W' may be omitted and the chromatic difference of magnifica-
tion calculated from the coordinates of the marginal ray. This is possible
88 Theory of Aberrations

when CHLO = 0, since the first equation of (23.2), together with (23.6),
gives the relation
A' - A'
1000· ~ = CHVP"", - CHVO for A~ # 0 and CHLO = 0 . (23.7)
v

These two examples illustrate how an optical designer makes use of


theory to save computation.
At first sight, it may seem illogical that equations (23.1), which are
valid for any wavelength pair, should contain in all cases the factor 1831
derived from the wavelength J. = 546 nm. However, it is convenient to
have the same basic reference system for the region of visible radiation:
minor adjustments taking account of the sensitivity curve of the receiver
or the special requirements of the system may be undertaken at a later
stage. The spectral region round J. = 546 nm is of particular importance
in the visual testing of optical instruments because of the maximum
sensitivity of the eye.
As examples, we shall compute the chromatic aberrations for the
imagery of a single lens as described at the end of § 15. We shall suppose
that one wavelength is defined by the mercury g line (J. = 436 nm), the
other by the mercury e line (J. = 546 nm). The glass is SK 16 which has
the refractive indices ng = 1.63314, ne = 1.62287. The calculation of the
light tube coordinates for the e line was given in § 15. Table 10 reproduces
these results for e and gives the equivalent values for g. The reference
point is in both cases the point of intersection of the plane surface with
the axis.
Table 10
Line A' B' n'U' n'W'

g 14.586819 0.022962 0.044986 - 0.0025


14.591324 0.023107 0.044215 - 0.0025

The chromatic aberrations are thus:


CHLO = - 20.9 R.E., CHVO = - 0.2%0 ,
CHLP= 0.0 R.E., CHVP = 0.2%0 .
Of these aberrations only CHLO, the longitUdinal chromatic aber-
ration of the object imagery, is of consequence. It will be shown later that
all other aberrations have a negligible effect on the image quality; It is
therefore possible to study the appearance of the aberration CHLO
experimentally without confusion with other aberrations.
Fig. 31 shows schematically how the imagery described in § 15 may
be realized in practice.
The Chromatic Aberrations of Perfect Imagery 89

A slide of 30 mm diameter serves as object O. It consists of opaque


strips separated by transparent (or, at least, translucent) strips. The
opaque and clear strips are 5 mm wide. This pattern is imaged according
to the rules of geometrical optics with the linear conductivity LLW =
2.5 x 0.0025 mm = 0.00625 mm "'" 11 R.E. The pupil P is a diaphragm
of 30 mm diameter placed directly in front of the lens L, whose data and
function were described in § 15. Imagery is only possible if the object 0
is illuminated. The mercury vapour lamp p* and the condensor K provide
the illumination. It is not sufficient simply to place the lamp in front of
the slide because its radiating area is smaller than the area of the slide
and cannot illuminate it evenly without the condensor. Apart from this,
the lamp radiates approximately equal amounts of energy in all direc-
tions, so that the pupil P at a distance of 6 m would receive very little
energy without the condensor K. It consists of one or two lenses whose
diameters must be larger:than 30 mm and images the radiating surface p*
p

p'
IL
~ iI
Fig. 31
~
of the lamp in the pupil P. Correct illumination thus requires that p* be
the object, P the image and the slide 0 the pupil; or, more exactly, the
exit pupil. In the example given, the focal length of the condensor may
conveniently be taken to be 200 mm. It then forms an image in P of the
radiating surface p* with about 30 x magnification. Since there is a stop
with diameter about 30 mm in P, a radiating spot of about 1 mm diameter
is necessary at P*. A mercury lamp has roughly the required dimensions.
The slide 0 is evenly illuminated because it serves as pupil for the source
and condensor.
To adjust the illumination, a screen is placed in the plane P, 0 is set
up at the correct distance (6 m) from P, K is placed directly to the left
of 0 and p* is moved until the best image of the source appears on the
screen. It is often better to remove the slide in order to increase the
brightness of the image: in fact, if the non-opaque parts of the slide are
mat, it is essential to remove it. The lens mount of the condensor then
serves as pupil, but the small pupil shift is of no consequence for the
source imagery provided the condensor does not have a significantly
larger diameter than the slide.
After the above adjustment, we remove the screen from P, replace the
slide in the correct position and can then consider 0 as the emitter and P
90 Theory of Aberrations

as the receiver of the radiation. The source itself may now be ignored. The
illumination has been considered here in some detail because it is usually
treated very cursorily. Often, a light source is set up somewhere behind
the object with the idea that all will be well if the source is powerful
enough. But the decisive factor is that the light source must be properly
adapted to the light tube of the object imagery. If this is not correct, no
amount of complication with lenses can redeem the situation because any
wrong energy distribution in the light tube will be faithfully transmitted
by the optical system.
An image of the slide 0, reduced about 18 times, is formed in 0'
about 330 mm behind the lens L. But we cannot assess its quality without
further apparatus because of the position and size of the exit pupil. This
is the image of P and lies in the lens L, in front of 0', and is much larger
than the eye pupil. We therefore need an observing microscope, shown
schematically in Fig. 31 behind 0'. It must possess a total magnification
of at least 25 x and the numerical aperture of the objective must be at
least 0.05. In practice, on account of the brightness of the mercury lamp,
it is more agreable for the eye to observe with a considerably greater
magnification than this minimum, for example with a 10/0.22 objective
and a 10 x eyepiece, giving a total magnification of 100. This also has the
advantage that the accommodation depth of the eye becomes negligible.
The observing microscope yields a strongly reduced image P' of the
exit pupil P in a plane behind the eyepiece which is accessible to the eye.
Furthermore, it forms an enlarged image of 0' at a considerable distance
in front of the eye, thus enabling 0' to be viewed without difficulty. The
observing microscope has no influence on the image quality, provided
that a reputable make is available and that the eyepiece is suitable for
use with the objective; for the design, manufacture and quality control
of good microscopes guarantee that their aberrations will be completely
negligible for the purposes of our experiment.
Before we can begin, the whole apparatus must be carefully centred
and adjusted. This is quite essential, otherwise the results will be value-
less. An old knotted thread can transmit mechanical force just as well as a
new one, twisted wires can conduct electricity perfectly well; but a lens
held between two fingers instead of being properly mounted is useless. If
its plane surface is tilted so that the normal subtends 3 0 with the axis,
the results of the experiment will be quite invalid. But it is not possible
to centre the lens with an accuracy of 1 0 by eye. Sensitive systems
require adjustments to a few thousandths of a millimeter and angular
centring errors of a few seconds of arc. Such refinement of adjustment
demands time and considerable skill, but without this there is no hope
of experimental verification of the effects predicted by the theory.
Fortunately, in our experiment we are investigating a very large effect:
The Chromatic Aberrations of Perfect Imagery 91

a relatively course adjustment of the order of magnitude of a degree and


a millimeter is sufficient.
The front edge of the observing microscope is set up at a distance of
about 350 mm from the plane surface of the lens L. Through the eyepiece
we can then see a bright field of view but without sharp boundary or
structure. Now we shift the observing microscope along the axis towards
the lens. At some point, the grid structure of the slide becomes visible.
The background is bright and yellowish in colour, the opaque strips have
a bluish edge and a dark centre. If the microscope is pushed 1.1 mm
further towards the lens, the edges of the strips become temporarily
blurred, then sharp again, but the edges are now coloured purple. The
centre remains dark. A further shift produces first of all blurred edges
and then the entire structure disappears in the bright background. But
after 5.6 mm shift the structure again becomes visible, this time with
sharply defined strips coloured uniformly lemon yellow without a dark
centre. The contrast against the slightly bluish white background is so
poor that the effect can easily be missed if the microscope is moved too
rapidly. Nevertheless, the difference in colour is sufficient to permit sharp
focusing. If the microscope is moved still further, the pattern disappears
and does not reappear. The focusing error caused by the depth of imagery
is ± 0.07mm.
What is the relationship between these observations and the calculated
chromatic aberrations? The shift of 5.6 mm from the purple to the yellow
image is equal to the calculated separation between the images for the
green e line and the blue-violet g line. But why does the last image
observed appear yellow? If we observe it through a blue filter placed
behind the eyepiece, the edges of the strips appear black and sharply
defined against a deep blue background. In the part of the field of view
occupied by the strips, no light of the g line is transmitted but the other
colours do pass through and enter the eye giving the visual effect of the
complementary colour to g, i.e. yellow. The image with the purple edges
can be explained by an analogous effect with the e line; that with the
pale blue edges is similarly related to the yellow mercury line with
A = 578 nm. But in these cases only the edges of the strips appear in the
complementary colour because the two images are so close to each other
that light from the e line cannot appear in the centre of a strip when the
focus is correct for the yellow line.
The longitudinal chromatic aberration between the two wavelengths
A = 546 nm and A = 578 nm is 4 R.E. for a pupil diameter of 30 mm, and
0.5 R.E. for a pupil diameter of 10.6 mm. With an adjustable iris dia-
phragm, it is easy to determine what amount of longitudinal chromatic
aberration an observer can just detect.
92 Theory ot Aberrations

Suppose now that the longitudinal chromatic aberration between two


wavelengths is zero, but that a chromatic difference of magnification is
present. The two images appear sharp at the same focus but are of diffe-
rent sizes. This causes coloured edges, the width of the colour fringes
being proportional to the image size. They are thus widest at the edge
of the field and vanish in the centre. This characteristic enables one to
distinguish between chromatic difference of magnification and longitudi-
nal chromatic aberration. Normally, when colour error is observed, a
mixture of both aberrations is present.

§ 24. Analysis of the Primary Chromatic Aberrations


Surface and Thickness Contributions,
Approximate Formulae
Equations (23.1) enable us to calculate the chromatic aberrations of
perfect imagery, but they give no information about the origin of the
aberrations or how they may be corrected. This section will deal with this
problem.
The coordinates of the light tubes in the image space are calculated
from the coordinates in the object space by repeated application of
equations (7.6) and (7.7) for transfers LI and refractions CP. It is therefore
desirable to investigate the effect of a single process on the chromatic
aberrations. We shall begin with the effect of a transfer with the para-
meter ij = - d/n on CHLO. For the sake of clarity, the coordinates of
the new reference point are denoted by symbols with a star, those of the
old reference point by plain symbols. Then

A;n;U; - A;n;U; = (A", - :. n",u",) nyUy - (Ay - :. nyUy) n",Uz


= (AznyUy - AynzUz) + d UzUy(nz - ny) .
Multiplying by 1831 we have
CHLO:II = CHLOzy + 1831 dUzUy(nz - nil) . (24.1)
The effect of the transfer on the chromatic aberrations, that is
CHLO* - CHLO, will be denoted here by the inverted delta sign V.
Equation (24.1) and similar equations derived for the other chromatic
aberrations give
V CHLO = 1831·dUz Uy (nz - ny) R.E. ,
V CHVO = - lOOO·dWzUy (nz - ny)/LLW %0' (24.2)
V CHLP = 1831·dWz W y (nz - ny) R.E. ,
V CHVP = lOOO·dUzWy (n", - ny)/LLW%o.
Analysis of the Primary Chromatic Aberrations 93

The indices x and y refer to the spectral lines used and d gives the
geometrical distances of the reference points. It is taken to be positive
if the transfer occurs in the same direction as the light.
To calculate the effect of a refraction with the parameterrp = (n' - n)e
on the chromatic aberrations, we place the reference point at the pole of
the surface and refer to light tube coordinates before the refraction by
plain symbols, those after it by dashed symbols. Thus
A~n~U~ - A~n~U~
= Aa;· [nyUy + (n~ - ny)e·Ay] -Ay·[na;Ua; + (n~ - nx)e·Aa;]
= AxnyUy - Ayna;Ua; - Aa;Aye{(n~ - n;) - (na; - ny)} .
From this, the longitudinal chromatic aberration follows as
CHLO' = CHLO - 1831·Aa;Aye{(n~ - n;) - (na; - ny)}. (24.3)
The effect of the refraction on the chromatic aberrations is represented
by the upright 6. symbol, for example 6. CHLO = CHLO' - CHLO. From
(24.3) and similar equations for the other chromatic aberrations, we can
derive the equations:
6. CHLO = - 1831· Aa;Aye{ (n~ - n~) - (na; - ny)} R.E. ,
6. CHVO = lOOO·Ba;Aye{(n~ - n~) - (na; - ny)}/LLW %0'
6. CHLP = - 1831· Ba;Bye{(n~ - n~) - (na; - ny)} R.E. , (24.4)
6. CHVP = - lOOO·Aa;Bye{(n~ - n~) - (na; - ny)}/LLW %0.

Here 12 = 11r denotes the curvature of the refracting surface, x and y refer
to the spectral lines used.
By virtue of the linearity of equations (24.1) and (24.3), the longi-
tudinal chromatic aberration of an optical system is the sum of the
contributions V CHLO and 6. CHLO for all transfers and all refractions:
CHLO = 2: V CHLO + 2: 6. CHLO . (24.5)
The other chromatic aberrations can also be expressed in terms of
sums of individual contributions. It is for this reason that the quantities
given by (24.2) are termed thickness contributions to the chromatic aber-
rations, while (24.4) gives the surface contributions. They are functions of
the separations ("thiclmesses") d or the curvatures 12, the dispersions
(na; - ny) and the coordinates of the light tube. The surface and thickness
contributions are therefore changed if the imagery is changed. In other
words, the same lens system has different surface and thickness contri-
butions if the manner in which it is used is altered.
The changes in the contributions resulting from shifts of the object
or the pupil are readily calculated. The object shift Q as given by equa-
tion (7.5) is substituted in the first equation of (24.2) giving
94 Theory of Aberrations

V CHLO* = 1831 d(U;c + ooW;c) (Uy + ooWy) (n;c - ny)


= 1831 dU;cUy(n;c - ny) +
+ 1831 doo(U",Wy + W",Uy) (n", - ny) +
+ 1831 dw 2 W",Wy(n", - ny)
= V CHLO + 1.831 oo(V CHVP - V CHVO) ·LLW +
+oo2VCHLP.
By a similar reduction from the first equation of (24.4) we arrive at
~ CHLO* = ~ CHLO + 1.831 00 (~ CHVP - ~ CHVO) ·LLW +
+oo2~CHLP .
If all the equations for the thickness and surface contributions to a
given aberration are added, the resulting equations for the effect of the
object shift Q on the chromatic aberrations of the entire system are
CHLO* = CHLO + 1.831 00 (CHVP - CHVO)· LLW + 00 2 CHLP ,
CHVO* = CHVO - 0.54600 CHLP/LLW,
CHLP* = CHLP, (24.6)
CHVP* = CHVP + 0.54600 CHLP/LLW,
the last three equations being derived by reductions analogous to that
for CHLO. The corresponding equations for the effect of a pupil shift II
on the chromatic aberrations are
CHLO* = CHLO ,
CHVO* = CHVO - 0.546n CHLO/LLW, (24.7)
CHLP* = CHLP + 1.831 n (CHVP - CHVO)LLW + n 2 CHLO,
CHVP* = CHVP + 0.546n CHLO/LLW.
Equations (24.6) and (24.7) are also applicable to the individual
thickness and surface contributions of the chromatic aberrations. They
show that the chromatic aberrations of an optical system are not con-
stants of the system but change with shifts of object and pupil. There is
only one case for which the chromatic aberrations of a system are inde-
pendent of the position of object and pupil: a light tube for which all
chromatic aberrations are zero. In the following sections we shall show
that, in practice, freedom from chromatic aberrations for a finite spectral
range cannot be achieved. The special case of zero aberration thus has no
practical significance. The system is corrected for chromatic aberrations
for the desired imagery and may not be used for other conditions of
imagery which deviate markedly from this. There is, however, consider-
able practical significance in the fact that the difference of magnification
between two wavelengths is independent of the position of the pupil (or
that of the object) if the longitudinal aberration of the object (or that of
Analysis of the Primary Chromatic Aberrations 95

the pupil) is zero. This follows immediately from equations (24.6) and
(24.7).
The rigorous formulae (24.2) and (24.4) for the thickness and surface
contributions explain the theoretical origin of the chromatic aberrations
but are not very suitable for numerical calculation. Instead, approximate
formulae, derived from them, are used in the following way. The blue
cadmium line F' (A = 480 nm) is taken as x, the red cadmiun line C'
(A = 644 nm) as y and the chromatic aberrations corrected for this line
pair. "Corrected" does not necessarily mean reduced to zero, but brought
to a given desired value. The light tube coordinates are not calculated
for these lines but for the intermediate "principal wavelength" e (the
green mercury line with A = 546 nm) and these values are inserted as
approximate values for A"" Ay etc. in equations (24.2) and (24.4). For a
reason which will be explained later, the expressions on the right hand
side are multiplied with (ne - 1) so that the term 1831· (nF' - ndj(ne-I)
can be replaced by the dispersion number me. Equations (24.2) for the
thickness contributions then yield the approximate formulae:
n, - 1
V CHLO "" - . - ·d· (nU)2me RE. "" 0.23·d· (nU)2me R.E. ,
ne
ne - 1
VCHVO"" - 0.546· --.- ·dnUnWmejLLWOfoo""
n.
"" - 0.I3dnUnWmejLLWOfoo, (24.8)
n, - 1
V CHLP "" - . - ·d· (nW)2' me RE. "" 0.23·d· (nW)2'me RE. ,
n.
V CHVP "" - V CHVO .
The factor 0.23 in these expressions represents a rough approximation
to (ne - I)jn;. The following table gives the range of values which occur
in practice.
Table 11
n (n - 1)/n2 n (n - 1)/n2

1.4 0.204 1.7 0.242


1.5 0.222 1.8 0.247
1.6 0.234 1.9 0.250

In the analogous approximate formulae for the surface contributions,


it is useful to consider the special case for which n = 1 or n' = 1, i.e. a
surface bounded by air. Since /2(n' - 1) = 'P or /2(1 - n) = 'P from (24.4),
we have
~ CHLO "" - 'P A2me R.E. ,
~ CHVO "" 0.546'P A BmejLLW Ofoo , (24.9)
~ CHLP "" - 'P B2 me R.E. ,
~ CHVP "" - ~ CHVO .
96 Theory of Aberrations

These equations are not applicable to a refracting surface between


two glasses, a so-called "cemented surface". (In practice, the two glass
surfaces are held together by a thin layer of a tough cement). However, the
formulae may still be applied if the two surfaces are supposed separated
by an air film of zero thickness. These two equivalent surfaces then yield
the right result, since the air film has no effect on the chromatic aber-
rations.
Equations (24.9) are valid not only for a surface bounded by air but
also for a thin lens in air. The term "thin" means that the coordinates
A and B do not change their values between the two surfaces within the
accuracy required for the calculation. Theoretically, this is only possible
if the axial thickness of the lens is zero. In this case, At = A 2, B t = B2
and the total refracting power of the lens Cf! = Cf!t + Cf!2' the sum of the
refracting powers of the surfaces. The chromatic aberrations of a thin
lens are thus independent of the form of such a lens, that is independent
of the distribution of refracting power between the two surfaces. The
simplified formulae (24.9) for thin lenses in air are a very useful aid to the
optical designer in the correction of the chromatic aberrations of an
optical system. The reason for the definition of the dispersion number me
in § 22 is its practical convenience in these equations: for the thickness
contributions, the principal dispersion nF' - nc' would be more sui-
table.
In practice, the procedure is as follows. The refracting powers Cf! of
the lenses and their separations are given the necessary values to produce
the desired transformation of the light tube for the principal wavelength e.
A computation following the scheme described in § 15 gives for this
wavelength the coordinates A, B, nU, nW for all the lenses or refracting
surfaces. Equations (24.9) may then be used to estimate what dispersion
number me a lens must have in order to produce a desired effect on the
chromatic aberrations. In general, the in:fl.uence of the thickness is ignored
at this stage; for we must first of all get a rough idea whether the glasses
necessary for the chromatic correction exist on the glass diagram (Fig. 30).
If not, the basic design of the system must be changed. At this rough
design stage, the approximate formulae (24.9) can be used without any
reserve, since we are only concerned to establish a region in which
practical solutions may be found. If now the stage has been reached where
suitable glasses have been determined, the exact chromatic aberrations
of the system are calculated with the refractive indices nF' and nc' and
finite thicknesses of the lenses from equations (23.1). The result will
certainly show a discrepancy from the sum of the approximate values;
but this difference can readily be corrected because the formulae (24.9)
give quite reliable information about the effect on the chromatic aber-
rations of a given change in the dispersion number.
Analysis of the Primary Chromatic Aberrations 97

An experienced optical designer will, in practice, short cut the above


procedure. Since the primary chromatic aberrations are the simplest to
correct, he will normally set up the basic design of a complex system with
other design problems in view, the necessary conditions for chromatic
correction being relatively easy to achieve and maintain. Nevertheless, the
procedure described is quite typical of the techniques used in correction.
It has been given here because it is particularly easy to follow and can
be performed as an independent operation.
We shall now consider the example of a chromatically corrected
objective consisting of two cemented lenses, a so-called achromat. The
focal length I = l/cp is taken to be 100 mm, and the object is at infinity.
The pupil is defined by the lens mount and has a diameter of 25 mm,
giving a relative aperture of 1/4. The field angle is n 1WI = - 0.1. If the
lenses are supposed thin, then we have A = 12.5 and B = O. Since B = 0,
all chromatic aberrations are zero apart from CHLO. If we stipulate that
this aberration must also be zero between the lines F' and C, then (24.9)
gives the condition for the lenses 1 and 2:
0= - CPl A~ m 1 - CP2 A~ m2 = - (12.5)2. (CPl m 1 + CP2 m2) •
In addition, to give the right focal length, the condition CPl + CP2 = 0.01
must be fulfilled. Combining these equations gives the condition
0= 0.01 m1 + CP2 (m2 - m 1) •

It can be satisfied in many different ways, for example by deciding


on a glass pair for which m 1 #< m2 • We shall take the glass of the first lens

SKIt; F2

Axis

1i 12 fj
Fig. 32

to be SK 16 with me = 30.5, that of the second lens to be F 2 with


me = 50.7. Then CP2 = - 0.305/20.2 = - 0.0151, CPl = 0.0251. The first
lens is positive with a focal length II = 39.8 mm, the second negative
with 12 = - 66.2 mm. Each lens produces a longitudinal chromatic
aberration of almost 120 Rayleigh units. The contribution of the positive
lens is negative, that of the negative lens positive, the sum thus being
zero. It should be remembered that this compensation must be achieved
7 Zimmer, Geometrical Optics
98 Theory of Aberrations

not only on paper but also in practice within about half a Rayleigh unit.
This gives an idea of the care necessary in the manufacture of the glass
and the lens.
The following data are for a design using these glasses:-
r1 = 90.00 mm,
33 13 d2 = 8.5 mm, glass SK 16, f = 39.9 mm ,
r2 = - . mm, d3 -- 4 •0 mm, g1ass,F2 f = 64.9 mm.
r3 = - 190.00 mm,
Fig. 32 shows a section through the objective. The numbers refer, of
course, to surfaces, not to lenses. The lens thicknesses have been delibera-
tely made larger than necessary for manufacture in order to demonstrate
their effect more clearly. The following table gives the thickness and
surface contributions to the chromatic aberrations, calculated from
equations (24.2) and (24.4). The cemented surface is divided into two
surfaces bounded by air to show the individual contributions of the two
lenses.
Table 12
Contributions for CHLO [R.E.] CHVO [0/00] CHLP [R.E.J CHVP [%oJ

1st surface - 32.97 0 0 0


2nd thickness 0.46 - 0.23 0.61 0.23
2nd surface; SK 16
into air - 83.18 - 1.58 - 0.16 l.58
2nd surface; air into F2 138.69 2.63 0.26 2.64
3rd thickness 0.35 - 0.18 0.48 0.18
3rd surface - 23.34 - 0.66 - 0.10 0.67
Sum for both lenses 0.01 - 0.02 1.10 0.02

Sum for 1st lens - 115.69 - l.81 0.45 l.81


Sum for 2nd lens 115.70 l.79 0.55 - l.79

For CHLO, the influence of the thicknesses has been compensated by


the refracting powers; but this is not possible for CHLP because of the
small value of the perpendicular B which results in the thicknesses having
a predominant effect.
It must be emphasized that only the chromatic aberrations are cor-
rected in this system which is useless in practice because of other aber-
rations.

§ 25. Secondary Spectrum


Secondary Longitudinal Chromatic Aberration
We now calculate the longitudinal chromatic aberration of the achro-
mat given in the last section for a number of spectral lines using equations
Secondary Spectrum 99

(23.1). The reference line y is the green mercury line e. The other lines,
their wavelengths and the corresponding calculated values for CHLO are
given in Table 13: -
Table 13
Line h g P' C' A'

Wavelength [nm] 405 436 480 644 768 1014


CHLO [R.E.] 11.9 5.6 1.4 1.4 5.4 14.9

This table shows that the longitudinal chromatic aberration does not
necessarily disappear for another line pair when it is zero for F' and C'.
This is equally true for the other chromatic aberrations of perfect imagery.
It will suffice if we confine the treatment of secondary spectrum to CHLO,
since the effect is formally identical for the other chromatic aberrations.
The aberration was given the name "secondary spectrum" for the
following reason. Suppose the primary chromatic aberrations for a given
case of imagery are uncorrected, then the image of a suitable object shows
red and blue fringes on each side of the yellowish green region which is
perceived as "the image" because of its brightness. If now the chromatic
aberrations between F' and C' are corrected so that the "primary colours"
red and blue are eliminated, small colour errors remain which cause green
and violet fringes, the "secondary colours". These secondary effects are
much smaller than the primary fringes. If the achromat described above is
replaced by a single thin lens of SK 16, the longitudinal chromatic aber-
ration between F' and C' is - 48 Rayleigh units. By comparison, the
achromat has only - 1.4 R.E. between e and F' and 4.2 R.E. between g
and F'. Nevertheless, the secondary aberrations significantly worsen the
image quality.
The starting point for the analytical treatment of the secondary
spectrum of longitudinal chromatic aberration is the equation
n - 1
CHLOF'a' ,." - L A 2 t:p me + L ----;;> . d· (nU)2 me . (25.1)

This follows immediately from (24.5), (24.8) and (24.9). The symbols
and conditions of Validity of these equations were explained in section 24.
Equation (25.1) can be applied with exactly the same assumptions to any
line pair x and y giving
n - 1
CHLO",y"" - LA2t:pm",y+ L----;;>d.(nU)2m",y. (25.2)

Here, the symbol m",y = 1831· (n", - ny)/(ne - 1) represents a genera-


lization of the dispersion number me. Its relationship to me is given by
m",y = rx",y me + {l",y + a"'l1 ,

100 Theory of Aberrations

in which the coefficients !Xzy and /3zy are functions only of the spectral
lines x and y and are the same for all glasses. The quantity azy is depen-
dent on the glass but is zero for many glasses - the so-called normal
glasses.
If (22.3) is substituted in (25.2) and combined with (25.1), we have
n - 1
CHLOzy "" !Xzy·CHLOF'C' + /3ZY {- L: A2 t:p + L: --.;;2d· (nU)2} +
n- 1
+ {- L: A2 t:p azy + L: --.;;2 ·d· (nU)2·aZY}. (25.3)

If /3ZY = 0 and azy = 0, then mzy = !Xzy me and the secondary spectrum
is zero; for it follows that if CHLOF'C' = 0, then CHLOzy = 0 for any line
pair apart from small inaccuracies due to the replacement of A",Ay by A~
and n",UznyUy by (n eUe)2. The second term in brackets on the right hand
side of equation (25.3) is determined by the constructional parameters of
the system and can hardly be influenced by the choice of glasses because
the value of (ne - 1)/n~ varies little over the range of useful refractive
indices. This was shown in Table 11 (page 95). The third term only gives a
contribution to CHLOzy if glasses are used for which azyoF 0, i.e. glasses
which are not "normal". Such glasses exist. They are indicated in the
glass diagram (Fig. 30) by arrows with components at = ate' andal! = al!F'.
In general, they have the disadvantage of being expensive, difficult to
work, coloured or liable to striae. It is also particularly unfortunate that
they do not exist at the very points of the glass diagram where they are
most needed. It is therefore clear why optical systems for which the
secondary spectrum is considerably reduced or completely removed are
much more expensive than simple achromats.
Following equation (25.3), the secondary spectrum of the longitudinal
chromatic aberration between the lines x and y is defined by the expression
SSP(CHLO)zy = CHLO",y - !XzyCHLOF'C' . (25.4)
Instead of SSP(CHLO)zy, one usually writes simply SSP",y since the
secondary spectra of more than one chromatic aberration are rarely
needed at a given stage of the design. The above definition is very useful
in practice. In the special case of CHLOF,c' = 0 it gives CHLOzy as the
secondary chromatic aberration. However, there is no restriction that
the primary aberration shall be zero specifically between the lines F'
and C'. Such a definition, based on the special nature of visual obser-
vation, would be very unsatisfactory because of the varied requirements
of photographic emulsions and other receivers. Equation (25.3) shows
that within the accuracy specified for this relation, SSP is independent of
the normal glasses contained in the system and how they are used to
correct the primary chromatic aberration. The value of SSP nevertheless
Secondary Spectrum 101

takes account of all effects achieved by crystals or special glasses. In


practice, the values SSPeF' and SSPUF' are preferred because they
correspond to the maximum sensitivities of the eye and the photographic
emulsion. When only normal glasses are used, the last term in the brackets
{} disappears, giving the relation SSPuF,/SSPeF' = {lUF'/ {leF'. If the values
from Table 9 (page 83) are inserted, we have the relation
SSPuF' = - 2.72 SSPeF,
for normal glasses.
There is an extremely simple and clear procedure for the graphical
representation of the primary and secondary chromatic aberrations as a
function of wavelength. The aberrations are based on the principal wave-
length e and plotted in a rectangular coordinate system. The horizontal
axis represents the chromatic aberrations in Rayleigh units (or, alter-
natively, %0). The vertical axis represents the spectral lines as wave-
lengths in units defined by the coefficient exxe of the lines x and e. The
scale is immaterial for the representation of the primary aberrations
alone, but if we want to read off the secondary spectrum correctly we
must have the correct scale. For if we draw a straight line p through a
point (y; CHLOye) of the aberration curve and parallel to the line joining
the points (F'; CHLOF'e) and (C'; CHLOc'e) , then the horizontal dis-
tances between the aberration curve and the straight line p give the
secondary spectrum referred to the spectral line y. This follows at once
from the definition of the scale of the ordinate because exF'C' = 1.
In Fig. 33 the longitudinal chromatic aberration of achromats with
f = 100 mm is represented in this way. The continuous curve shows the
CHLO value of the system given in § 24, the reference line p for the
secondary spectrum being identical with the line joining the points
(F'; CHLOl"'e) and (C'; CHLOc'e). The dashed curve is for another
achromat with f = 100 mm, performing identical imagery, but with the
longitudinal chromatic aberration corrected between F' and e instead of
between F' and C'. The design data are
11 = 95.025 mm
12 = - 34 . 100 mm d2 = 8.5 mm glass SK 16, f = 41.3 mm,
13 = _ 170.300 mm d3 = 4.0 mm glass F 2, f = 69.1 mm .
The corresponding reference line p is also shown dashed. The hori-
zontal distances between this line and the aberration curve are effectively
identical with those between the continuous curve and the basic axis.
The differences are so small that it is a waste of time to recalculate the
form of the aberration curve resulting from a change in the primary
aberration. Suppose now an achromat with f = 100 mm is constructed
from the glasses K 50 and F 2 for the same imagery as that described at
102 Theory of Aberrations

the end of § 24 and such that CHLOF'a' = O. The aberration curve is


practically identical with the continuous curve of Fig. 33, the only
noticeable difference being one Rayleigh unit for the t line (J!. = 1014 nm).
This discrepancy is due to the abnormal dispersion in the infra-red region
of some crown glasses with low refractive indices. The effect becomes
marked for some of the boron crown glasses in the region of K 50 (see glass
diagram, Fig. 30). It follows from this stability of the secondary spectrum
that it is only necessary in special cases to compute the chromatic
aberrations for more than two lines.
The additional influence of "abnormal" glasses or crystals on the
secondary spectrum of longitudinal chromatic aberration is defined, using
the quantity axy, by the expressions
n - 1
- A2 rp axy and --;;2 ·d(nU)2 axy·

These follow from (25.3). The first gives the additional effect due to
a refracting glass-air surface or a thin lens in air, the second the additional

p ~LO CHLO
./ h
./
./
./
/ 9

F'

CHLO CHLO
-5 15 [R.£] -5 e 12 [R.£l

C'

--
Fig. 33 Fig. 34

effect of a glass plate of thickness d. In general, it is not possible to say


whether positive or negative values of a will have a favourable effect;
it will depend on the values of the secondary spectrum which have to
be compensated. For an achromat with positive refracting power rp
consisting of two cemented thin lenses, the secondary spectrum of the
lines hand t relative to F' is always positive. From (25.3) we have
SSP"F' = 7.05 X A2 rp RE., SSPtF' = 8.97 x A2 rp R.E. ,
Secondary Spectrum 103

giving for the example with A = 12.5 mm and!p = 0.001 mm-1


SSP"F' = 11.0 RE., SSPtF, = 14.0 RE.
For the achromat with J= 100 mm using thick lenses given in § 24,
Fig. 33 gives the values:
SSPliF' = 10.5 RE., SSPtF, = 13.5 R.E.
If abnormal glasses are used in the thin lenses, the additional effect
on the secondary spectrum is
- A2!p! a! - A2!p2 a 2 .

Compensation of the positive secondary spectrum of the complete achro-


mat can be achieved if a glass with positive a value is used for the positive
lens and one with negative a value for the negative lens. Two materials
which are particularly favourable in this respect are fluorite (CaF2) for
the positive lens and the Schott glass KzFS 2 for the negative lens.
Refractive indices for fluorite have been given by T. RADHAKRISHNAN
and G. N. RAMACHANDRAN in Revue d'Optique tMorique et instrumen-
tale, volume 32 (1953). From these values we calculate aliF' = 2.4 and
atF' = 3.7. Similarly, from the refractive indices given by Schott for
KzFS 2 we have aliF' = - 0.5 and atF' = 2.4. If these two materials are
used to correct the longitudinal chromatic aberration between F' and C',
a refracting power !p! = 0.0232 mm-! is necessary for the positive lens,
!P2 = - 0.0132 mm-! for the negative lens. With A = 12.5 mm the result-
ing additional effect of the positive lens is - 8.7 RE. for SSPliF' and
- 13.4 RE. for SSP tF ,; that of the negative lens is -1.0 RE.
for SSPliF , and - 4.9 RE. for SSP tF ,. The secondary spectrum of the
achromat is thus slightly undercorrected between hand F' and over-
corrected between t and F'. If the primary longitudinal aberration is
changed so that CHLOF,c' becomes negative, the aberration curve shown
in Fig. 33 "rotates" in an anti-clockwise sense. By this means, the longi-
tudinal chromatic aberration can be made small for the whole visible
spectral region. The following data are for an achromat designed to have
such a chromatic correction with J = 100 mm and a relative aperture
ofJ/4: -
r! = 56.4 mm
r2 = - 26.8mm
d2 = 8.0 mm Fluorite J= 43.0 mm
d3 = 3.0 mm glass KzFS 2 J= - 73.9 mm .
r3 = -79.0mm
In Fig. 34 the longitudinal chromatic aberration is shown for the
region 405 nm to 1014 nm.
104 Theory of Aberrations

§ 26. Plane Rays


Notation, Law of Refraction and Ray Tracing Formulae
In the theory of chromatic aberrations we supposed that every surface
of every lens in the optical system was capable of perfect imagery. From
now on, our analysis is based on the law of refraction for rays.
A ray is a mathematical abstraction, a straight line which determines
a direction. The ray can be defined by a three-dimensional, right-angled
coordinate system following the convention of analytical geometry. It is
convenient to define the x-y-plane as the plane of the paper. Departure
from this convention in exceptional cases must be clearly stated. The
positive x-direction is from left to right, the positive y-direction upwards.
The z-axis is perpendicular to the plane of the paper, positive above it.

y y

x x
(z)
Perspective
Plane section diagram
z

Fig. 35

When we represent the three-dimensional coordinate system in perspec-


tive, we draw the z-axis leftwards and downwards to show that the posi-
tive direction lies above the plane of the paper.
The relationship between the coordinate system and the optical
imagery is defined as follows: -
1. Since we are limiting our analysis to centred systems, there is
always a system axis. This axis is taken to coincide with the x-axis and
is positive in the direction of the light.
2. By virtue of the rotational symmetry of the imagery about the
system axis, the y-axis and z-axis are completely equivalent. If, in spite
of this, we choose to limit the generality by further assumptions, it is only
because the limitation leads to a better understanding of a special case.
Thus we lay down the general rule that a "principal ray" from the edge
of an object to the centre of the pupil shall lie in the x-y-plane; and that
a "marginal ray" from the centre of the object to the edge of the pupil
shall lie in the x-z-plane. This rule corresponds exactly with the scheme
shown in Fig. 18 in connexion with the conductivity diagram of Delano.
Plane Rays 105

For reasons which will be explained later in the treatment of astigma-


tism, the x-y-plane of the coordinate system defined in this way is called
the tangential section and the y-direction the tangential direction; the
x-z-plane is the sagittal section of the system and the z-direction the
sagittal direction.
Rays which lie in the tangential or sagittal plane will be called plane
rays, for short. They may be defined by two coordinates. The reference
point E is chosen to lie on the axis. One coordinate is the length of the
perpendicular from E on to the ray, the other is the sine of the angle
formed by the ray with the axis. There is thus no formal difference
between these coordinates and those defined in § 2 for the light tube.
The sign convention is also identical. To distinguish the coordinates of a
ray from those of perfect imagery, the ray coordinates in the sagittal
section are given the symbols a and sin u, those in the tangential section
b and sin w. The line with the coordinates A and U in the light tube of
perfect imagery will from now on be termed the ideal marginal ray, that
with coordinates Band W the ideal principal ray. In dealing with these
ideal rays we shall always use the rules of perfect imagery, not the ray
tracing formulae now to be derived for plane rays.
The law of refraction defines the change in direction of propagation
of waves caused by their passage through the surface separating two
media with refractive indices nand n'. If the rays 5 and 5' are wave
normals before and after the refraction and N is the normal to the surface,
the law of refraction can be formulated geometrically as follows: -
The three rays 5,5' and N pass through a point. (26.1)
The three rays lie in a plane, the so-called plane of refraction. (26.2)
If 5 forms the angle i with Nand 5' the angle i' with N, then
the relation n'sin i = n'· sin i' is valid. (26.3)
No special rule is necessary for the sign of the angles of inci-
dence i and i'. It is sufficiently defined by the fact that both
angles are less than a right angle and that both are measured
in the same sense between the positive directions of the rays
and the positive direction of the normal.
In the theory due to Maxwell the radiation is assumed to consist of
electromagnetic waves. This theory, which will not be treated in this
book, confirms the above geometrical interpretation of the law of refrac-
tion at the surface separating two insulating materials. However, it also
shows that the law is quite inadequate as a complete definition of the
physical processes involved in refraction. The quantities defining electro-
magnetic waves are phase, amplitude and polarisation. Of these, the law
of refraction only defines the phase relationships at the surface separating
the two media: it gives no information concerning the amplitude and
106 Theory of Aberrations

polarisation. This information is given by the Fresnel formulae which do


not belong to the field traditionally classed as geometrical optics because
they are only valid for transverse waves, not for radiation in general.
It follows that the law of refraction is not the equivalent of the
theorem of conservation of radiation energy. Thus calculations in geo-
metrical optics, which are based only on the law of refraction, can - in
principle - give no physical information regarding the quality of the
image. They can only satisfy one prerequisite of good imagery, namely,
that the necessary phase conditions are fulfilled. However, the following
experiment shows that correct phase relationships alone do not necessarily
give good imagery. A good quality microscope with a high aperture
objective and plane polarised illumination is used. An analyser behind

Q6

Q2

20 40
Fig. 36

the microscope with polarisation direction perpendicular to that of the


polariser does not give the extinction in the field of view of the eyepiece
to be expected from the geometrical theory because the refractions in the
objective introduce amplitudes in directions which were not present in
the illumination.
In practice, the restriction of our calculations to phase relations is
not so serious as one might suppose at first sight. This is because the
amplitude of the refracted waves varies little over the range of angles of
incidence i that occur in lens systems which are technically feasible.
Fig. 36 demonstrates this. The ordinate is the transmission factor {}, the
ratio of the transmitted to the incident amplitude. This is plotted against
incidence angle for refraction at the perfectly polished surface of a glass
plate with refractive index 1.6. The symbols II and ~ indicate that the
curve applies to the case for which the waves are polarised parallel with
or perpendicular to the plane of refraction. On account of the aberrations
introduced by each refraction, the angles of incidence at glass-air surfaces
in optical systems are normally less than 45°. Nevertheless, the example
of the polarisation microscope cited above shows that it is possible in
Plane Rays 107

special cases to detect not only the reduction in amplitude but even the
small amplitude differences between different polarisation directions.
Nowadays, lens surfaces bounded by air are usually "coated" with
anti-reflexion films. These films are about 100 nm thick and are produced
by evaporating magnesium fluoride (n = 1.38). The transmission factors
of a coated surface are larger than those of a pure glass surface but the
differences are so small that they would hardly be discernible in the curves
of Fig. 36. However, the significance of anti-reflexion coatings is much
greater than this would indicate because the measured intensities are
proportional to the squares of the amplitudes: an amplitude difference
of 2% gives an intensity difference of 4%. For a single surface it may be
unimportant whether 95% or 99% of the incident intensity is transmitted

R%
40
I
30 I
I
20 I
/
10 /
/
,/

20 40 60 80 j.
Fig. 37

and therefore 5% or 1% reflected. But for ten such surfaces it is very


important whether 59.9% or 90.4% of the intensity is transmitted.
Even if this difference can be compensated by better illumination, the
stray light caused by double reflexions will give reduced contrast in the
image plane. For this reason, the invention of a practical process for the
anti-reflexion coating of polished glass surfaces (A. SMAKULA, 1935) was
of great significance for the industrial production of optical systems.
Values of the reflected intensities R, calculated from the Fresnel formulae
for unpolarized light of wavelength 550 nm, are given in Fig. 37 as a
percentage of the incident light and plotted against the angle of incidence i.
The continous curve applies to the pure surface of a glass with n = 1.6,
the dashed curve to the same surface with a 100 nm thick anti-reflexion
layer of magnesium fluoride.
The equations for the transfer from one reference point E to another
E* have the same form for the coordinates of plane rays as for the coordi-
nates of the light tube:
a* = a + (J n sin u, b* = b + (J n sin w ,
n*sin u* = n sin u, n*sin w* = n sin w . (26.4)
108 Theory of Aberration

Here, <5 = - din in which d is the distance from E to E*, and n = n*


is the refractive index of the medium. In practical ray tracing, it is more
convenient to leave out the factor n from n·sin u and n'sin w and calcu-
late simply sin u and sin w. The formulae corresponding to (26.4) are then
a* = a - d sin u , b* = b - d sin w ,
sin u* = sin u , sin w* = sin w . (26.5)
The proof requires only elementary geometry as for the light tube
coordinates and follows directly from Fig. 5.
The derivation of the formulae for the refraction of a plane ray at a
centred spherical surface is illustrated in Fig. 38 for a ray in the sagittal
section (x-z-plane). To make the figure clearer, the refracted ray is not
shown. The centre M of the sphere lies on the axis. The refracting spherical
surface with radius r cuts the axis in the point E ---,- which is taken as the
reference point for the refraction equations - and the ray in the point P.
The straight line PM is the normal to the refracting surface. It forms the
angle i with the ray and the angle cp with the axis. Perpendiculars E PI = a
and MP 2 = r·sin i are dropped on to the ray. If MP2 is produced beyond
M to the point of intersection P 3 with the line EP3 parallel to the ray,
then MP3 = r·sin u. In the rectangle EP1 P 2 P 3 , since the opposite sides
are equal,
a=rsinu+rsini.
Multiplying by e = l/r we have
a e = sin u + sin i , (26.6)
a relation which is also valid for the limiting case r = 00.
The signs of a, sin u and e were laid down in § 2 and § 15. That of
sin i is defined by equation (26.6). Since the normal to the surface PM
is the same for the incident and refracted rays, a corresponding equation
must also apply after refraction:
a' e= sin ~t' + sin i' . (26.7)
In addition, Fig. 38 shows that the relation
cp=u+i (26.8)
exists between the angles. The sign of cp is defined by (26.8). For the
refracted ray
cp' = u' + i' , (26.9)
and, of course,
cp=cp'.
The procedure for numerical computation is as follows. From a, sin u
and e = l/r we calculate sin i from equation (26.6), then sin i' from the
Plane Rays 109

law of refraction n sin i = n' sin i' (26.3). If a calculating machine is


available which calculates square roots automatically, we compute
cos i = V1 - sin2 i and cos i' = V1 - sin2 i', the positive roots being
taken since the angles of incidence are less than 90°. From equations
(26.8) and (26.9) the following angle functions are calculated: -
sin cp = sin (u + i) = sin U'cos i + cos u'sin i,
cos cp = cos (u + i) = cos U'cos i-sin u'sin i,
sin u' = sin (cp - i') = sin Cp'cos i' - cos Cp'sin i' ,
cos u' = cos (cp - i') = cos Cp'cos i' + sin cp'sin i' .
The shorter equation cos cp = V1 - sin2 cp is not valid in the general
case since the angle cp can become larger than 90°. This is actually the

case with microscope objectives. If calculating machines or computers


are not available, the values of u, i and i' are found from sin u, sin i and
sin i' in trigonometrical tables. Then cp = u + i, and u' = rp - i' are
calculated and sin u' found from the tables. Finally, a' is computed from
a' = r (sin u' + sin i') .
This formula is only applicable for finite radii and gives significant
rounding errors for radii greater than 1000 mm. The radius can be
eliminated by forming the quotient of (26.6) and (26.7):
, sin u' + sin if
.
a =a' Slnu "
+Sln't
.

However, this equation also has a singularity; for both nominator and
. sin u' + sin if
denommator of the factor . . . become zero when r .... 00. Apply-
Slnu + sin 1.
110 Theory of Aberrations

ing the sums and products formulae and the relation u' + i' = u + i, the
singularity of the factor can be removed by the relation
sin 1,1,' + sin i' COSu' + cos i'
sinu +sini cosu + cosi .

Instead of a formal computation, we shall give here a geometrical


derivation using Fig. 38. Let P 4 be the second point of intersection of the
sphere with radius r and centre M with the axis; and P s the point at the
foot of the perpendicular from P 4 on to the ray. Then PP2 = ,·cos i and
P 2 P s = r·cos u. The second of these equations becomes evident when a
line is drawn through P 4 parallel to the ray and the relation M P 4 = ris
borne in mind. The dashed lines EP and PPs are perpendicular to each
other (THALEs' theorem). It follows that the corresponding sides of the
triangles EPP j and PP4 P s are perpendicular to each other and the
triangles are therefore similar. From the ray theorem we then have
a l"cosu+r'cosi . . . a
EP = PP s for the mCldent ray and correspondingly EP =
""cosu' + y·cosi' - --
--~~-- for the refracted ray, since the lengths EP and PP s
PP s
have the same significance for both rays. Division gives finally
, cos 1,1,' + cos i'
a =a' .. (26.10)
cosu + COSt

This relation is also valid for plane surfaces because cos u = cos i i= 0
whene = O.
Apart from the coordinates of the rays, one often needs those of
the point of incidence P. Referred to the point E, these are given by
x = r·(1 - cos rp) and z = r·sinrp. Once again, the radius can be eliminated
from these formulae giving relations valid for plane surfaces. In Fig. 38,
we drop the perpendicular from P on to the axis, the point of intersection
being P 6 • Similar triangles EPP6 and PP4 P 6 give the ratios
EP z
----=--
pP. Y + Y'cos rp
Y'sin rp
which, combined with
EP a
PP. -Y'cosu+Y'cosi'
gives
sin rp l+cosrp
x=acosu+cosi
and z=a cosu+cosi
(26.11)

For the sake of clarity, it is useful at this stage to re-state the complete
set of formulae for tracing a ray in the sagittal section through a refracting
spherical surface with curvature e = l/r: -
Plane Rays 111

sin i = a e - sin u, cos i = V1 - sin2 i ,


sin cp = sin U'cos i + cos u'sin i,
cos cp = cos wcos i-sin u'sin i, (26.12)
sin i' n ..
=;;;·sm~, cos i' = V1 - sin2 i' ,
sin u' = sin cp·cos i' - cos cp·sin i' ,
cos u' = cos cp·cos i' + sin cp·sin i' ,
I cos U' + cos i' _ . sin u' + sin i'
a =a' cosu . ( fore# 0),
+ cos. -a sinu +sini
sinq:> 1 + cos q:>
X= a· cosu+cos.
., Z = a' cos u + cos i .

The reference point E is the pole of the refracting surface. The same
formulae are also valid for a ray in the tangential section; but the
quantities represented by a, U, i, cp and z are given the symbols b, w, j,
cp andy.
At first sight, there seems to be no similarity between these formulae
and the refraction formulae (7.7) for perfect imagery. However, a
similarity is made evident by a simple rearrangement. Equation (26.6) is
multiplied by nand (26.7) by n'. Then, applying the law of refraction,
nt'a'e - n'·sin u' = n'·sin it = n sin i = n·a e - n·sin U J

giving
n"sin u' = n'sin U + e (n'a' - n a) .
The condition a' = a is only fulfilled for rays in special cases. But
if we write a' = a + IX in which IX = a' - a is the difference in the lengths
of the perpendicular before and after refraction, then we arrive at the
following formulae for the refraction of plane rays: -
a' = a + IX,
n'·sin u' = n·sin u + e (n' - n)·a + en' IX ,
b' = b + p, (26.13)
n'·sin w' = n·sin w + e(n' - n)·b + en'P.
These are analogous to those derived for perfect imagery.
These formulae are useless for practical computation because they
give no indication how IX and p are to be determined. Nevertheless, they
are very significant in aberration theory. For refraction at a spherical
surface, they show that the difference between light tube coordinates
and plane rays is given by the two quantities IX and p. A plane ray has
only one degree of freedom relative to the corresponding light tube
coordinates, not two, because the "angle difference" elX or eP is the
product of the surface curvature e and the "perpendicular difference"
IX or fl. When setting up a new system, it is important for the optical
112 Theory of Aberrations

designer to have a clear idea of the range of values of IX and {J which


occur or might be attainable in the desired imagery. The following
product representation of IX and {J makes this possible.
We transform the expression
, sin u' + sin if
lX=a -a=a· sinu +sini -a
= a. sin u' + sin i' - (sin u + sin i)
sinu+sini

using the sums and products formulae for trigonometrical functions into
u' + if u' - if u +i u - i
2 sin - 2 - . cos - 2 - - 2 sin -2- . cos -2-
IX = a·
u+i u-i
2 sin -2- cos -2-

Since t~ +i = rp = u' + i', this reduces to


u' - if U - i
cos - 2 - - cos -2-
IX = a· ----u----:-i--
cos-2-

u' - i' u' + it i' rp


Now - 2 - is replaced by - 2 - - 2· "2 = "2 - i' and, correspondingly,

u ; i by t- i. The difference in the numerator then gives


. i_if
'P
cos ( 2- .,) ( 'P
t -cos 2- t
•) 2· 'P - i' - i ·SIn-
=- SIn 2 2- ,

so that
2a . u' - i . if - i
IX= - - - . · S I n - - ' S I n - -
u-< 2 2 '
cos-
2-
2b . w'-i . i'-i (26.14)
{J = - -w--1. ·SIn-- ·SIn--.
2 2
cos-
2-

The significance of this formulation will be explained later in the


tratment of aberrations.
Furthermore, the set of equations (26.13) now provides us with a
justification for equation (15.1) in which the ideal refracting power of a
spherical surface was defined by rp = (n' - n)/r = (n' - n)e. For this is
precisely the result given by (26.13) as the refracting power for rays which
are refracted without "perpendicular difference", i.e. without aberration.
The Primary Monochromatic Aberrations of Plane Rays 113

§27. The Primary Monochromatic Aberrations of Plane Rays


Spherical Aberration and Distortion
for Object and Pupil Imagery
If we assume the light tube coordinates of a ray in the sagittal section
to be A and nU in the object space, then we can consider this ray as the
normal of a spherical wave propagated from the centre of the object.
Because the ray passes through the edge of the pupil, it is the ray in this
section with the greatest inclination to the axis (or, for objects at infinity,
the ray with the largest height above the axis). For this reason, it is
termed the marginal ray of the spherical wave. According to this defini-
tion, there is no difference in the object space between the path of this
marginal ray and the ideal marginal ray of the light tube: The ray in the
tangential section with the coordinates b = Band n' sin w = n W is called
the principal ray. In the object space, it is identical with the ideal
principal ray of the light tube. It can be interpreted as the central
normal ("principal axis") of a spherical wave which is propagated from
the edge of the object and passes through the centre of the pupil. This
interpretation explains the name given to the ray. However, we can
equally well consider the principal ray as the marginal ray of a spherical
wave which is propagated from the middle of the pupil and passes through
the object surface (or, in reverse direction, is propagated from it). Simi-
larly, the marginal ray defined above can be considered as the central
normal of a spherical wave from the edge of the pupil which passes through
the object surface. This demonstrates that a ray in optical imagery does
not possess a special function: the same ray simply plays a different role
in the object and pupil imagery.
If we compute the light tube coordinates from equations (7.6) and
the refraction formulae (7.7), and trace the plane rays through the system
using the sets of equations (26.5) and (26.12), then we arrive at the coordi-
nates A', B', n'U', n'W' and a', b', n' sin u', n' sin w'
in the image space for a common reference point. In general, the two
sets of coordinates will not be the same. The following formulae, analogous
to those for the chromatic aberrations, define the monochromatic aber-
rations: - SPHO = 1831 (a'n'U' - A'n' sin u') R.E. ,
SVO = 1000 (A' n' sin w' - b' n' U' _ 1) 0/
LLW 00'
SPHP = 1831 (b'n'W' - B'n' sin w') R.E. , (27.1)
SVP = 1000 (a' n' W' - B' n' sin u' _ 1) 0/
LLW 00'
The aberration SPHO is called the spherical aberration of the object
imagery or of the marginal ray. The name has historical significance. This
8 Zimmer, Geometrical Optics
114 Theory of Aberrations

aberration was the first to be discovered by trigonometric tracing of rays


through spherical surfaces. At that time, it was thought that it was the
most important aberration introduced by lenses. It can be removed by
making the lens surfaces aspheric, a possibility which has attracted many
theoreticians and amateurs since it was pointed out by R. DESCARTES
(159&-1650). Aspheric surfaces only have practical significance in special
cases, in particular for astronomical instruments. For example, the high
cost of manufacture of Schmidt plates can be justified by the enormous
advantage they bring to instruments for scientific research. Aspheric
surfaces are also used in condensors for illumination purposes. Here, high
quality imagery is not required but a compact, low-cost design, implying
a minimum number of lenses, is important. The precision of these aspherics
is much lower than that readily attainable with spherical surfaces. Aspheric
surfaces are used in spectacle lenses, but the purpose here is the correction
of astigmatism in the eye, not spherical aberration. The surfaces of these
"toric" spectacle lenses have different radii of curvature in the tangential
and sagittal sections: they are not surfaces of revolution with an axis
of symmetry and represent a type of optical instrument for which the
theory in this book is not applicable. A special theory and a special
computation scheme are necessary. Industrial mass production (as op-
posed to expensive one-off manufacture) is possible for toric spectacle
lenses because only a small area of the lens - the area of the eye pupil-
is used at any given time so that small departures from the calculated
form are not noticeable.
Why can spherical surfaces be produced more easily than aspherics?
Because glass is a homogenous and isotropic material, a spherical surface
results automatically from the grinding process if this is randomly oriented.
One can observe a similar effect with pebbles on the sea-shore, but the
tidal movement here usually produces ellipsoids of revolution because its
direction is not random relative to the coastline. Evenness of wear is
achieved by the grinding and polishing of lenses because the lens and the
grinding or polishing surface are in free contact during their relative
movement. This free contact guarantees a spherical form as the statistical
mean surface but the radius of curvature is not determined. The form of
the working tool must be such that it gives the desired radius. It is possible
to produce surfaces which depart from the required form by only a small
fraction of a wavelength. A further advantage of the sphere from the
point of view of manufacture is that any straight line through the centre
is an axis of symmetry. This means that the second spherical surface of
a lens can be produced without worrying about its alignment to the first
because the line joining the centres of the two spheres is the axis of the
whole lens. This is only the case for spherical surfaces.
Apart from the technical difficulties of production, there are other
The Primary Monochromatic Aberrations of Plane Rays 115

reasons why aspheric surfaces are not in general use. Their significance is
also limited in the design stage because aberrations resulting from the
spherical form can be corrected with spherical surfaces if additional lenses
are introduced; furthermore, it may not be necessary to remove them
entirely as the image quality is often impaired by other aberrations not
dependent on the form of the surfaces. Chromatic aberrations, parti-
cularly secondary spectrum, are examples of such aberrations. Other
examples are the change in spherical aberration with wavelength of the
radiation (Gauss error) and field curvature, an aberration which will be
dealt with later in connexion with astigmatism.
SPHO can be interpreted, exactly like CHLO, as the depth explained
in § 19 between the image given by the marginal ray and that arising from
perfect imagery. What is the effect of spherical aberration on the image
quality? Are there two separate images which can be observed separately
as in the case of chromatic aberration? If the image suffers from spherical
aberration then A' = 0 and a' ¥= 0 apply in the plane of the ideal image.
Is the image blurred because an ideal image point is replaced by a circle of
confusion with radius roughly equal to a'? Where is the image formed:
in the plane of the perfect image, in the plane where the marginal ray
cuts the axis or somewhere between? The theory cannot answer these
questions; we must again be guided by experiment.
We shall use the experimental set-up described in § 15 and § 23 and
shown in Fig. 31. Since the chromatic aberrations would mask the effects
we wish to investigate, we use a filter which only transmits the green
mercury line e. The filter is placed in front of the object 0 or behind the
eyepiece of the observing microscope. The coordinates of the light tube
for perfect imagery were calculated in § 15. We now compute the coordi-
nates of the marginal and principal rays from equations (26.5) and (26.12).
In the object space they are identical with the light tube coordinates;
in the image space the computation gives
a' = 14.592826 mm, n'sin u' = 0.0443065 ,
b' = 0.023107 mm, n' sin w' = - 0.002500 .
The reference point is again the point of intersection of the plane
surface of the lens L with the axis. Equations (27.1) give for the aber-
rations:
SPHO = - 2.3 R.E., SVO = 0.0%0 ,
SPHP = 0.0 R.E., SVP = 0.2%0 .
The only value of consequence is the spherical aberration of the object
imagery. We shall now investigate this aberration more closely.
At first, we use our apparatus exactly as described in § 23, but with
a green filter. There is no difficulty in finding the black strips and focusing
S'
116 Theory of Aberrations

them sharply against the green background. The depth of focus is about
± 0.1 mm and is in good agreement with the theoretical value of the
depth of imagery. The marginal ray cuts the axis 0.6 mm in front of the
plane of the perfect image, but it is quite impossible to detect images in
different planes along the axis.
We now replace the stop P by an iris diaphragm or by a series of stops
with different diameters. Spherical aberration varies as the fourth power
of the pupil diameter, but this result cannot be seen directly from
equations (27.1). It could, of course, be confirmed by a numerical com-
putation, but this is not necessary as a simple proof is given in the next
section. The significance of this fourth power law for our experiment is
that the pupil diameters of 68 mm, 34 mm and 17 mm give - 64 RE.,
- 4 RE. and - 0.25 RE. of spherical aberration respectively. If we
slowly reduce the pupil diameter from an initial value of 30 mm to 15 mm,
we can be quite certain that the spherical aberration will vanish under
the depth of imagery. What happens to the image during this process?
It becomes darker because the exit pupil behind the microscope is smaller
than the eye pupil and becomes even smaller when the aperture is reduced.
We have already commented on this effect in connexion with perfect
imagery: it has nothing to do with the spherical aberration. Apart from
the brightness, there is no change in the image. In particular, the sharp-
ness of the strips is unaltered and refocusing does not bring any impro-
vement for small apertures.
Suppose now the diaphragm is slowly opened. The image becomes
brighter but a special effect appears when the diameter exceeds 34 mm:
light starts to flood in like a bright fog from the edges into the dark parts
of the strips. The effect occurs so suddenly that the radius of the critical
diaphragm can be quite precisely determined, within a millimetre. With
this diameter, the spherical aberration has the value 4 RE. Further
increase in diameter produces a strong fogging effect with the result that
the brightness difference between the strips and the background is
rapidly reduced, i.e. the contrast of the image is worsened. With 200 RE.
of spherical aberration, careful observation is needed to find the pattern
at all! In spite of this, the boundary between the background and the
strips remains clearly defined at the same focus as that found for small
apertures, although the marginal ray cuts the axis at a distance from this
focus equal to about a hundred times the depth of imagery.
This observation contradicts the view generally held that spherical
aberration is an aberration causing unsharp images and that the best
image lies between the point of intersection of the marginal ray and the
ideal image plane. The fact that rays from an object point do not combine
in a point when spherical aberration is present does not prove that
The Primary Monochromatic Aberrations of Plane Rays 117

unsharp images will result; for rays are merely a mathematical abstrac-
tion defining the direction of propagation of the radiation.
It is very instructive to carry out the experiment described above
using a "point" object. This is easily done by replacing the test pattern
by a metal screen with a pinhole of about 0.5 mm diameter. Together
with a diaphragm of 30 mm diameter, this gives a linear conductivity of
1 R.E., i.e. just at the limit of geometrical imagery. In the plane where
the test pattern was sharply in focus we now see a bright spot whose
diameter corresponds to that of the pinhole. If the diameter of the stop
is reduced to 15 mm, the linear conductivity is 0.5 R.E. and diffraction
effects become clearly visible. The spot has become somewhat larger and
forms the centre of a system of bright and dark concentric rings. The
brightness of the outer rings falls off rapidly. The first can always be
observed quite readily and a dark-adapted eye can see between three and
five rings. By closing the diaphragm further to a diameter of 7.5 mm we
reduce LLW to 0.25 R.E. The central spot is now twice as large as the
geometrical image of the pinhole; the rings are also stronger and wider.
The appearance of the diffraction image is completely altered if the focus
of the microscope is now changed. The outer ring becomes bright and wide
while a number of narrower rings surround the central spot. This spot may
be either bright or dark, depending on the plane of focus of the microscope.
If, starting at the image plane 0', the object plane of the microscope is
moved away from the lens L, then the rings appear to originate at the
centre and spread out from it. If the focus direction is reversed, the rings
vanish into the centre. The movement of the rings is the same whether
the diffraction pattern is observed in front of or behind the plane 0'.
To investigate the effect of the spherical aberration on the diffraction
pattern, we replace the object 0 by a smaller pinhole with only about
0.1 mm diameter. Even with a large aperture, this object is not imaged
geometrically. The necessary pinhole can be produced quite simply by
punching a few holes in a metal foil with a fine needle. The exact diameter
is not critical. The observing microscope is first focused using a stop
diameter of 30 mm with the strip test object which is then replaced by a
metal foil with the fine pinholes. The diffraction pattern can now be
observed with the large aperture. Stopping down produces no noticeable
change, but when the aperture is increased beyond 34 mm a bright,
blurred ring structure is superposed on the diffraction pattern. As with
the strip test object, this effect becomes visible when the spherical aber-
ration reaches about - 4 R.E., but the precise aperture at which it occurs
is not so easy to determine.
A special effect appears when the observing microscope is defocused.
The appearance of the image is no longer the same for intra- and extra-
focal settings at the same distance from the image plane. In the extra-focal
118 Theory of Aberrations

direction (further from L than 0') the image rapidly becomes so blurred
that the ring system can hardly be seen. In the intra-focal direction, on the
other hand, there is a setting where the image appears sharp. It lies
halfway between the ideal image plane and the intersection point of the
marginal ray, or about 0.5 mm in front of the ideal image plane when the
diameter of the diaphragm is 40 mm. The image here is at its smallest and
the ring system very sharply defined: it is perceived as "the best image".
The central spot has become smaller and darker, the first bright ring
broader and as bright as the central spot. Further defocusing in the intra-
focal direction causes a broadening of the whole pattern into a slightly
blurred ring system.
This asymmetry relative to the focus position can even be observed
when no unsharpness is detectable in the diffraction image. Since the real
structure of the object is seldom known, there is a definite danger that the
focus will be wrongly chosen in favour of the "best image". For example,
if the focus for the "best image" is found with the pinhole of diameter
0.5 mm, it will be found to have a clearly defined bright ring round the
centre. But at the focus chosen for the strip test object, the ring is
definitely weaker. This is the reason why the spherical aberration
of high quality microscope objectives is kept below 4 R.E. although
this quality brings no improvement in the images of coarser struc-
tures.
The situation regarding the effect of spherical aberration on the quality
of photographic images is quite different and not entirely understood.
I ts effect is a reduction in contrast and a shift of the plane of the sharpest
image from the ideal image plane towards the point of intersection of the
marginal ray. The effects which were observed separately for geometrical
imagery and diffraction are now combined. The interruption of the light
tube by the photographic emulsion provides a plausible explanation
which can be tested experimentally by replacing the microscope in the
set-up shown in Fig. 31 by a fine ground glass screen in the plane of the
image 0'. The screen is observed with a lO-times magnifying glass. The
strip test slide is used as object. Of course, the pupil imagery has now
been destroyed, but the effect corresponds to the conditions of obser-
vation of a photograph. We adjust the pupil diameter to 25 mm and
focus the image sharply on the ground glass screen. Then the pupil is
opened up to 50 mm diameter. The background of the image is now
brighter and the edges of the strips are blurred. A shift of the screen of
about half a millimetre towards the lens restores the sharpness but the
contrast between the bright and dark parts of the image remains signi-
ficantly reduced. Focusing is more difficult because of the grain of the
screen than it was when the image was observed directly with the micro-
scope.
The Primary Monochromatic Aberrations of Plane Rays 119

The second aberration SVO in equation (27.1) is the distortion or,


more exactly, the sine distortion of the object imagery*. This is readily
interpreted in the same way as the chromatic difference of magnification.
If the ideal image is at a finite distance, it may be taken as the reference
point. Then at this point A' = 0 and SVO = 1000 (~ - 1) %0 gives the
difference in the lengths of the perpendiculars on to the real and ideal
principal rays in per mill of the length of the ideal perpendicular. If the
ideal image is at infinity, then U' = 0 and SVO = 1000 (Si~~' - 1) %0
gives the difference in the sines of the real and ideal principal rays in per
mill of the sine of the ideal ray. These expressions show clearly the signi-
ficance of SVO in two special cases of practical importance. But, as with
the expressions for spherical aberration, they do not tell us what an image
with or without distortion looks like. In fact, the subjective impression
depends very much on the special characteristics of the imagery, on the
object structure and the choice of pupil.
We will consider a few examples. It is quite possible with a suitable
technique to photograph a field of view of more than 360 0 in a continuous
image of perhaps 100 x 600 mm. If this image is observed at a distance
of somewhat less than a metre, the impression of the scene is quite
unnatural: distant objects seem too small, near ones too large. If the
photograph was taken from the centre of a space with a square boundary,
the boundary lines appear curved in the image, intersecting with angles
greater than a right angle.
These effects are not due to the technique of photography but to an
incorrect position of the eye pupil relative to the image. If the eye is
moved nearer to the image, the impression becomes more natural; but
this soon presents difficulties because the eye can no longer see the image
sharply. The use of a 6-times magnifying glass to extend the accommoda-
tion of the eye enables the observer to move so close that his nose almost
touches the image. He will be surprised to observe that the impression
of the scene has become quite natural, that distant and near objects are
viewed in the correct perspective and that the boundary lines of the
square are straight and intersect at right angles. The magnifying glass
has achieved this "miracle" indirectly simply by bringing the pupil nearer
to the image. Its direct influence on the distortion is small if it is held
near to the eye pupil.
This example illustrates quite clearly that the distance of the pupil
from the image is quite fundamental to the impression it gives. When
viewing photographs in their original size, the eye pupil should be at the
* The abbreviation SVO is derived from the German expression "Sinusver-
zeichnung des <2bjekts". - -
120 Theory of Aberrations

same distance as the rear principal point of the objective with which the
picture was taken. This distance is the focal length of the objective, or
somewhat more if near objects were photographed. Enlargements should
be viewed at a correspondingly greater distance. A picture taken by a
miniature camera with! = 50 mm and a format 24 mm x 36 mm cannot
be viewed properly with the naked eye because a normal adult cannot
accommodate for an object only 50 mm away. To view the picture
correctly at the normal distance of 250 mm, an enlargement of five times
is necessary, i.e. a format of 120 mm x 180 mm.
The word "distortion" implies failure of the image forming system
to reproduce the exact form of the object. This interpretation was,
indeed, the origin of the name given to the aberration: imagery was only
accepted as correct if straight lines in the object appeared as straight lines
in the image. But this definition was based on the assumption that
Euclidean geometry and the associated perception of space represented
the only possible framework of reference. In 1827, C. F. GAUSS and
N. LOBATCHEVSKY showed that this assumption was false and that other
geometries were perfectly feasible. In 1863, the physicist CLAUSIUS re-
cognized in the course of his theoretical investigations into heat radiation
in space that imagery produced by radiation cannot produce similarity
of form in the sense implied by classical Euclidean geometry. If two
figures F and F' in the same plane are observed under the angles wand
w', then according to the rules of Euclidean geometry F' is twice as large
as F if tan w' = 2·tan w. The constancy of the ratio tan w'/tan w for all
corresponding parts of F' and F is the criterion for the geometrical
similarity of these figures. The departure from constancy is expressed by
the "tangent distortion"
TVO = 1000 (A' n' sin w' - b' n' U' . cos w _ 1) 0/ •
LLW cos w' 00

The sine distortion is based on the constancy of the ratio sin w' /sin w
and is only identical with the tangent distortion when w = w', that is
when both figures appear to subtend the same angle at the eye. Apart
from this special case, one cannot expect optical imagery to produce
geometrical similarity between object and image.
The condition w = w' is roughly fulfilled for imagery in our eye, that
is for the refraction by the cornea and the crystalline lens. Since more or
less sharp vision only extends over an angular field of about 5° diameter
for a fixed direction of view, it is quite impossible to decide with the
maked eye whether or not Euclidean geometry is applicable to our per-
ception of space.
If it is important to have a geometrically correct reproduction of the
object, the system is designed so that the condition w = w' is roughly
The Primary Monochromatic Aberrations of Plane Rays 121

fulfilled. There is then no significant difference between the tangent and


sine conditions. However, the requirement that wand w' be equal implies
that the entrance and exit pupils of the optical instrument must have the
same size. In telescopes or microscopes - instruments intended to
increase resolving power - the entrance pupil must be significantly
larger than the exit pupil. It follows that these instruments are funda-
metally quite incapable of yielding images which are geometrically similar
to the object.
One apparent way out of this difficulty is to produce geometrical simi-
larity by deliberately not fulfilling the sine condition. Such lens systems are
possible but have severe limitations in practice because other aberrations
are linked with the distortion. In 1873, working independently, E. ABBE
and H. v. HELMHOLTZ discovered that if the distortion of the pupil
imagery does not obey the sine condition, sharp imagery of the object is
impossible. This is unquestionably the most important theorem in aber-
ration theory. It will be discussed in detail in § 30 and § 33.
ABBE and HELMHOLTZ had both announced the theorem for the special
case of microscopes. Its significance here can be illustrated by a very
instructive experiment. A "dry" objective (Le. not an immersion objec-
tive) of high aperture is used, say a 63/0.90 objective. It gives sharp
images which means that SVP, the sine distortion of the pupil, must be
small. If the eyepiece is removed, the pupil imagery can be observed by
viewing the upper end of the objective. The exit pupil lies in this plane
and is the image of an entrance pupil at infinity - in practice, 30 or 40 em
is sufficiently far away. If the microscope is now pointed towards distant
objects, the image seen in the exit pupil is free from sine distortion. Fig. 39
shows the image of a chess board formed in this way. The correct observing
distance is 1.5 m. We see this figure as seriously distorted. Why is this?
The figure is Euclidean and cannot itself be responsible for the distortion.
Nevertheless, the straight lines appear curved: in fact, they are portions
of ellipses, the straight line at infinity appearing as a circle. This is shown
by dashes in Fig. 39 and corresponds to the limiting aperture with
sinu=1.
Within this circle, for any "straight line" through a given point there
is a "parallel line" , i.e. an elliptical curve which does not intersect the
given "straight line". Similarly there is a "perpendicular". With this
figure, it is possible to demonstrate all the laws of Euclidean geometry
which we learnt at school and have since forgotten. If we look through
the microscope in this way, we can still measure and locate objects
although everything in the field appears tilted and distorted.
I believe that the impression of distortion is due to a large extent to
the fact that the angular measurements established by visual experience
of such figures do not agree with the angles of rotation of the eyeball. This
122 Theory of Aberrations

explanation is confinned by the considerable reduction in the impression


of distortion which occurs when the observing distance is made small by
means of a magnifying glass, or when Fig. 39 is imagined to represent
a ball.
The reverse effect occurs in binoculars. Here, the real angular separa-
tion of objects is considerably smaller than the eye movement necessary
to observe them through the instrument. Straight lines therefore appear
to be curved outwards. If a departure from the sine condition is delibera-
tely introduced to make the straight lines appear straight, then "local"
distortion appears in the image. When the binocular is moved, one has

Fig. 39

the impression that the image is curved into a spherical form. In practice,
a compromise between these two extremes is adopted. A certain amount
of distortion is introduced so that the straight lines appear somewhat
curved but the curvature of the image is not immediately obvious.
The effects discussed in this section do not contribute in any way to
the development of the theory. Nevertheless, their significance is so great
for an understanding of optical instruments that I have thought it
justified to deal with them in some detail.
As with the equivalent formulae for the chromatic aberrations, equa-
tions (27.1) express all the aberrations ofthe two rays; for the coordinates
of the rays are completely determined by the coordinates of the light tube
for perfect imagery and the four aberrations. This is proved by the
following equations which can be readily derived from (27.1): -
Analysis of the AbelTations of Plane Rays 123

, A' SVP A' SPHO B'


a - = 1000 • - 1831'LLW' ,
b' B' SPHP A' B'
- = 1831'LLW' + SVO
1000' , (27.2)

,. ,
n sm u - n
'U' SVP n 'U'
= 1000' SPHO
- 1831.LLW·n
'w' ,
,. ,
n sm w - n
'W' SPHP
= 1831'LLW' n
'u' + SVO
1000'n
'w' .
However, a relation between these aberrations analogous to (23.5)
does not exist because the expression a'n' sin w' - b'n' sin u' is not an
invariant for the coordinates of the rays.

§ 28. Analysis of the Aberrations of Plane Rays


Surface Contributions and the Approximate Formulae
of Seidel
We shall now concern ourselves with a question of great importance
for the design of optical instruments: the origin of the aberrations of plane
rays and possible methods of correcting them. Vlfe can conclude at once
from equations (27.1) that the aberrations they define are independent
of the choice of reference point. This is a consequence of the fact that the
transfer equations (7.6) for light tube coordinates and (26.4) for rays are
formally identical. As an example, we shall consider the effect of a transfer
on the spherical aberration of the object imagery. By virtue of the rela-
tionsn*U* =nU,n* sin u* = nsinu,A * =A +bnU and a* = a +bnsinu,
we have
SPHO* = 1831 (a*n*U* - A *n* sin ~e*)
= 1831 {(a + b n sin u)nU - (A + b nUl n sin tt}
= 1831 (a nU - A n sin u) + 1831 b(n sin unU - nU n sin u)
= SPHO,

since the last bracket is always zero. The proof is completely analogous
for the other aberrations. The monochromatic aberrations therefore differ
from the chromatic aberrations in that the thickness contributions of the
monochromatic aberrations of plane rays are all zero. As this result can
also be extended to the aberrations of skew rays, it follows that only
surface contributions need be considered in the analysis of monochromatic
aberrations.
The effect of a spherical refracting surface on the spherical aberration
can be calculated quite easily. Combining the refraction equations (7.7)
with f{! = (} (n' - n) and (26.13) we have
124 Theory of Aberrations

SPHO' = 1831 (a'n'U' - A'n' sin 1/)


= 1831{(a + IX) [nU + e (n' - n) A] -
- A [n sin u + (n' - n) a + en'IX]}
e
= 1831 (a nU - A n sin u) - 1831 IX (n A e- nUl .
For the marginal ray, ae - sin u = sin i is the sine of the angle of
incidence. We now use the light tube coordinates to construct the formally
analogous expression Ae - U = I and call it the "incidence" of the ideal
marginal ray. The law of refraction n' l' = nI is also formally valid
because (n' - n)e·A = n'U' - nU. Using this symbol we may write
SPHO' = SPHO - 1831 IX nI, (28.1)
and arrive at the expression
L1 SPHO = SPHO' - SPHO = - 1831 IX nI . (28.2)
for the surface contribution to the spherical aberration.
Expressions for the other aberrations can be derived in a similar way.
The following set of equations defines completely the contributions of a
centred refracting surface with the curvature e = l/r separating two
media with refractive indices nand n': -
L1 SPHO = - 1831 IX nI R.E. ,
L1 SVO 1000 fJ nI/LLW %0'
L1 SPHP = - 1831 fJ nj R.E. , (28.3)
L1 SVP = - 1000 IX nj/LLW %0'
In these equations, I and j are defined by
nI = nAe - nU = n'l' = n'Ae - n'U',
nj = nBe - nW = n'l' = n' Be - n'W' , (28.4)
while IX and fJ were defined by equations (26.14).
Under what conditions does a surface yield a zero contribution to an
aberration? Equations (28.3) show that this is the case if at least one of
the quantities I, j, IX and fJ is zero. We shall now consider these four
possibilities in turn.
1=0 reduces SPHO andSVO to zero. This means thatA/U = A/U' = r,
i.e. the centres of the ideal object and image lie at the centre of
curvature of the spherical refracting surface. The position of the object
is not changed by this imagery and the magnification nU/n'U' = n/n' is
always positive.
j = 0 has a completely analogous significance for the pupil imagery
and reduces the pupil aberrations to zero. The centre of curvature of the
refracting surface now coincides with the centre of the pupil. A surface
can never fulfil both the conditions I = 0 and j = 0 because object and
pupil can never coincide.
Analysis of the Aberrations of Plane Rays 125

IX = 0 implies that at least one factor must disappear in the expres-


sion (26.14)
2a . u' - i . if - i
IX = ---u=i 'sm - 2 - ·sm -2- .
cos-
2-

Since u and i always have values less than 90°, it follows that cos U ; i

can never be zero. There remain, therefore, three possibilities: -


1. a = O. The marginal ray traverses the refracting surface at its pole,
the point we took as reference point for the refraction equations. If the
marginal ray aberration is not large, it follows from a = 0 that A "" 0
also holds. This leads us to the following approximate rule. A lens placed
in the plane of the object or the image introduces neither spherical aber-
ration of the object imagery nor distortion of the pupil imagery.
2. sin i' ; i = 0, that is, i' = i and sin i' = sin i. But since the law of
refraction n' . sin i' = n'sin i must also be fulfilled, this case subdivides
into two further possibilities: -
2. a) n' = n. The refractive indices are identical on both sides of the
refracting surface: the surface thus has no effect. The dictum "He who
does nothing does no harm" applies just as much in aberration theory as
it does in ordinary life!
2. b) n' # n. This case is only possible when sin i' = sin i = 0 and has
the same significance as ae - sin u = 0 = a'e - sin u'. The marginal ray
traverses the centre of curvature before and after refraction and is not
deviated.
In all the above cases, it is geometrically obvious from the course of
the marginal ray that no aberrations can be caused: there is nothing
surprising about the results. Geometrical evidence is, however, wanting
in the last remaining case of aplanatic imagery which is defined by the
condition: -
3. sin u'; i = 0, that is, u' = i. Since u +i = u' + i', the condition
also implies that u = i'. Combining this with the law of refraction gives
. .. "
ae - n'.
sm u = sm ~ = -; sm ~ = -; sm u.
n'.

The following relations thus apply before and after refraction respec-
tively: -
ae - (1 + ~) sin u = 0,
a' e - (1 + ;?) sin u' = 0 . (28.5)
The case e= oresults in sin ~t=sin u' = Oandis therefore without inter-
est. But if e # 0, (28.5) indicates that the marginal ray cuts the axis in
126 Theory of Aberrations

the distances a/sin u = r' (1 + n'ln) and a'/sin u' = r' (1 + nln') respec-
tively from the refracting surface. These two points are called the
aplanatic points of the refracting surface. If the centre of the object lies
in one of these points, it will be imaged in the other without SPHO
and SVP. This imagery is by no means trivial from a geometrical point
of view, for the two aplanatic points have a separation

r(nln' - n'ln) .
The magnification is

n sin uln' sin u' = n sin i'/n' sin i = (nln')2 .


It may be greater or less than one, but is always positive. It follows that
it is impossible to produce a real image of a real object by surfaces which
refract according to aplanatic imagery.
f3 = 0 gives conditions for zero values of Ll SVO and Ll SPHP which
are completely analogous to the conditions for zero values of Ll SVP and
LI SPHO following from IX = O.
The special cases discussed above are rarely ideal conditions for the
design of good optical instruments in practice. But they are very impor-
tant as indications of the effect of a given surface on the aberrations.
Fig. 40 shows the surface contributions to the spherical aberration as a
function of ae and sin u for n = 1, n' = 1.6 and an object free from
aberration. The values for sin i = ae - sin u are constant along the
straight line at 45°. The diagram shows the lines representing constant
values of eLI SPHO. The value can be read off directly in Rayleigh units
if r = 1 mm; otherwise it must be multiplied by the radius. The three
straight lines

1. ae=O, 2. b) ae = sin u and 3. ae = 2.6 sin u


correspond to the special cases discussed above which give zero contri-
bution to the spherical aberration.
Equations (28.3) enable us to calculate the exact contribution of each
surface to the aberrations. However, they can only be applied if the
coefficients IX and f3 have been detennined for the corresponding rays by
tracing their paths using equations (26.12). This may involve too much
work to make it worth while. In practical designing, an approximation
which is much simpler to calculate may be sufficient.
If the trigonometric relation
. x -y sin x - siny
srn-2- = x+y
2'cos-2-
Analysis of the Aberrations of Plane Rays 127

:g. ~ c!;. a. ~ C)

~.
",OJ
"
.~

C)•

...
",'

:g.

'"
C).

~.

a.
'"""
<;- <;-

C) <>

9~ ~'

~ '"'i'
~ 9"
?' 9'
9' '"9'
<0
..:.:.
§1 <0

. .
9' 0 9'
og
.!
~
'
::l
9' .S 9'
·sCIJ, l5' CIJ

~' 0- ~ ~,
t::J" II ~'
10'0
~I~~
0

9' " 0' " :;t


" ::l a: 7
.... 0 ·sCIJ ll.i ~.~ ~
~ :::. ~,
"'.
0\
~. ...
"'. :g• "'oo. c!;. a. ~ <;- C)

Fig. 40
128 Theory of Aberrations

is inserted into the product formula (26.14) for IX, the expression
a· (sin u' - sin i) . (sin i' - sin i)
IX=
u - i u' + i i' + i
2 cos -2- ·cos-
2 - ·cos-
2-
results.
This formula can be simplified if we suppose that all the angles
occurring in it are small. Then the cosines in the denominator are
roughly 1 and the equation reduces to
IX "" a (sin u' - sin i) (sin i' - sin i)f2 .
This approximation is of little value in practice because the coordi-
nates of the marginal ray are still present. The next approximation is
based on the assumption that the difference between the coordinates
of the actual and ideal marginal rays is not large. This assumption
has no theoretical justification, for the exact theory is equally valid for
large differences. But large differences mean very large aberrations which
are not acceptable in practical optical systems. Thus, the fact that aber-
rations must be small in cases of practical interest justifies the assumption
that only small differences exist between the coordinates of the rays and
those of the light tube. This approximation gives

IX "" 4.(U' - I) .(1' - I) ,

a result which only contains the coordinates of perfect imagery. It is


no longer necessary to trace the actual rays. A further simplification
follows if we take n' outside the first brackets, replace I by ~ l' and take
n
n'1' = nI outside the second brackets, giving

IX "" -An'I' (U'n- ' - -n'I) . (1 - -n')


2 n
=

= :t...nI(~
2 n'
- ~
n
-~+~),
n' n

and, since Ifn' = I'fn and U' + l' - I = U, finally

A2 (U'n' U)
IXI'>I-nI - - - .
n
(28.6)
Similarly: B (WI -:nW) .
(J "" 2 n] 7 -

If these approximations are inserted into equations (28.3) for the


surface contributions, we obtain:
The Secondary Aberrations of Plane Rays 129

A (U U)
Ll SPHO"" - 1831 "2 (nI)2· ;;;-- -;; R.E.,
I

1000B (WI W)
Ll SVO "" 2.LLW nI nJ· 11- -;; %0'
B
Ll SPHP "" - 1831 2 (nJ)2· (WI W) R.E.,
11- -;; (28.7)

Ll SVP "" _ 1000 A nI nJ.


2.LLW
(!!. _~) 0/
n' n 00
.
These equations are equivalent to the well-known formulae given by
L. SEIDEL in 1856.
The aberrations expressed by (28.7) can readily be calculated. The
sum of the contributions from all its refracting surfaces gives an approxi-
mation to the aberration of the complete system.
In the general case, it is not possible to say how good this approxi-
mation is. In very high aperture microscope objectives, the Seidel values
for Ll SPHO can amount to as little as one third of the exact values, an
error of 500 R.E. The sum of the contributions may have not only a
wildly incorrect value but even a wrong sign. On the other hand, the
Seidel approximation for Ll SVO is so accurate in microscope systems
that ray tracing is unnecessary. The situation is the reverse in systems
with large field angles and small apertures. For most systems, however,
the Seidel formulae give at least a rough approximation to the effect of
a single surface and require relatively little calculation. Numerical
examples are given in the following sections.

§ 29. The Secondary Aberrations of Plane Rays


Zonal Aberrations, Graphical Representation and Examples
Because of their simple form, the Seidel formulae (28.7) are also
important theoretically. Their significance can be more readily under-
stood than that of the exact formulae (28.3). As an example, we shall
now consider how the aberrations vary with aperture and field. For the
sake of simplicity, we shall suppose that the object and the pupil are both
at finite distances. If this is not the case, nj U j must be replaced by Aj
when njUj becomes zero, or njWj by B j when njWj becomes zero. The
results are then quite generally correct.
If we take the aperture on the object side to be njU j * = k'njU j
instead of nj Uj and retain the same object plane, then At = k·A j also
applies. Since the imaging formulae (16.1) are linear, all perpendiculars A
and apertures nU within and behind the optical system are also multiplied
9 Zimmer, Geometrical Optics
130 Theory of Aberrations

by the factor k. Furthermore, because of the relation 1= Ae - U, all


incidence angles and therefore LLWare also multiplied by k. Thus

L1 SPHO* "'" - 1831 A2* (nI*)2 (~* - ~*) =

= k4 • { - 18311; (nI)2 (~' -~)}.


The Seidel values for L1 SPHO are therefore proportional to the fourth
power of the aperture and become sixteen times larger if the aperture is
doubled. This is one reason why the maximum aperture is considered to
be a measure of the power of all optical systems and why optical designers
never exceed the aperture necessary for the solution of a given problem.
Suppose now the pupil plane is kept fixed and the field angle nj W j on
the object side is multiplied by the factor 1. Then all perpendiculars B,
field angles nW, incidence angles] as well as LLW are multiplied by 1.
This has no effect on L1 SPHO. In L1 SVP the factor appears in both the
numerator and the denominator and cancels out. If similar calculations
are carried out for the other aberrations, we arrive at the following result.
For fixed object and pupil planes, the Seidel approximations for the
surface contributions (and their sums as an approximation to the aber-
rations of the complete system) vary with aperture nj U j and field nj W j
in the following way: -
SPHO is proportiqnal to (n j U j )4 ,
SVO is proportional to (n j Wj )2 , (29.1)
SPHP is proportional to (nj W j )4 ,
SVP is proportional to (n j U j )2 •
In fact, the aberrations grow faster than equations (29.1) indicate.
This difference is caused by the second order aberrations, usually called
the zonal aberrations. The way they are defined is - as with secondary
spectrum - essentially a matter of practical convenience. One could, for
example, calculate the actual surface contributions and their Seidel
approximations and take the zonal aberration to be the difference. This
shows quite clearly what the values of the zonal aberration are but
requires the calculation of the Seidel contributions as well as the ray
trace.
Since the marginal ray with the largest angle always has the largest
aberration, it is essential in practice to trace this ray through the system.
The same applies to the principal ray with the largest angle. The ray
trace gives the correct values of the surface contributions and the total
aberration for the edge of the pupil and the edge of the image. One assumes
that they would decrease with the angle of the ray according to the rules
of (29.1) for the Seidel approximation: that is as the square or the fourth
The Secondary Aberrations of Plane Rays 131

power. If, for example, the principal ray has the inclination sin w and the
object distortion SVO while a zonal principal ray has the inclination
l·sin w and the distortion SVO*, then the zonal aberration of the distor-
tion is defined by the expression
(SVO) = SVO* - l2 ·SVO .
Similarly, if the marginal ray has the inclination sin u and spherical
aberration SPHO and a zonal ray through the centre of the object the
inclination sin u* = k sin u and the spherical aberration SPHO* referred
to the light tube coordinates A * = kA and U* = kU, then the zonal
aberration must logically be expressed by
SPHO* - k4 ·SPHO.
Nevertheless, it is usual to divide this by k 2 giving
(SPHO) = k-2 ·SPHO* - k2·SPHO.
There are both historical and practical reasons for this which we shall
discuss in connexion with the graphical representation of the aberrations.
The complete set of formulae for the zonal aberrations is
(SPHO) = k- 2·SPHO* - k2·SPHO,
(SVO) = SVO* - l2 ·SVO , (29.2)
(SPHP) = l-2·SPHP* -l2·SPHP,
(SVP) = SVp* - k 2 ·SVp ,
in which k = n sin u*jn sin u = A*jA and l = n sin w*jn sin w = B*jB
are the ratios of the apertures or perpendiculars for the zonal rays and the
marginal ray or principal ray. The calculation of the aberrations is per-
formed with the values of the light tube coordinates for the corresponding
rays in the object space.
Equations (29.2) have the advantage of economy of calculation - the
evaluation of the Seidel approximations becomes unnecessary. Since
they are linear functions for fixed values of k and l, the total zonal
aberration is equal to the sum of the zonal aberrations from each sur-
face. Thus the optical designer can calculate the amount of zonal aber-
ration originating at the surfaces of the system and take the necessary
steps to eliminate it.
From a mathematical point of view, there are objections to selecting
the rays with the largest aberration as the reference rays for zonal
aberration because they depart most markedly from the Seidel approxi-
mation. In practice, however, this objection is of little consequence. For
if the primary aberrations cannot be adequately corrected, there is no
point in calculating the zonal aberrations. If, on the other hand, they
have been adequately corrected, they represent better reference quan-
tities than the calculated values of the Seidel approximation because

132 Theory of Aberrations

they have more physical significance. From the point of view of practical
optical designing, there is no reason to depart from the definition of (29.2).
Graphical representations are convenient in that they give a clear
picture of the aberrations of an optical system. All optical designers are
agreed on this point although they hardly ever agree on the best way of
representing them. This disagreement arises because individual designers
want to read off values relevant to their specific problems. Aerial survey
photographs must be free from distortion: differences b' - B' in the per-
pendiculars of the images are only acceptable to the extent that they
remain smaller than the grain size of the film material. This is the same
over the entire film plane and independent of B'. It is therefore con-
venient to plot the quantity 1000· (b' - B') = B' . SVO rather than SVO
against sin WI. This curve is roughly a third order parabola. Other design-
ers favour a direct plot of SVO because this quantity emerges directly
from the theory and can be obtained by simple addition of the surface
contributions. Of course, it is equally valid to plot tangent distortion
rather than sine distortion.
There are even more possible ways of representing spherical aber-
ration. If the value calculated from equations (27.1) is plotted as a func-
tion of sin u, (or a, for an infinitely distant object), the result is a fourth
power curve or higher. This curve shows at once the effect of pupil size
on the spherical aberration or, in other words, the effect of stopping down.
A disadvantage of this method of plotting spherical aberration is that
the curve virtually vanishes into the ordinate when the aperture becomes
small, so that information about the contrast and resolving power of the
system can hardly be obtained from the plot.
Equation (27.1) gives the spherical aberration of a ray relative to the
diffraction image of perfect imagery with the same aperture. But it is
quite permissible to give a value relative to the smallest diffraction
pattern, that is the pattern corresponding to full aperture. This curve
is obtained by plotting SPHO/k as a function of n, sin u" where
k = nsin u*/n sin u is the ratio of the aperture of the ray in question to the
aperture of the marginal ray. The curve is a third order parabola and is
used for photographic objectives, because here the grain size rather than
the diffraction pattern is usually the quantity determining the acceptable
aberration.
For optical systems which must yield the theoretical resolving power
it is convenient to plot SPHO/k2 as a function of n, sin u, (or a, for an
infinitely distant object). This curve is a parabola for the Seidel approxi-
mation and has the advantage that the zonal aberration emerges more
clearly than in the other methods of plotting given above. This is the
reason for the definition of (SPHO) in (29.2). Apart from this, SPHO/k2
represents - to an accuracy sufficient for all practical purposes - the
The Secondary Aberrations of Plane Rays 133

geometrical distance between the perfect image and the point of inter-
section of the ray with the axis. Thus the curve has a geometrically
evident significance. This was the reason for the choice of this form of
representation quite early in the history of geometrical optics.
Apart from Rayleigh units or pro mill (%0), aberrations have been
expressed in mm, p" inch, diopters and other units. Similarly, the inde-
pendent variable is not always taken to be the sine of the angle: it may,
for example, be the tangent or the square of the sine. Inevitably, with so
many different methods of representation, the plotted curve of an aber-
ration has little value unless the axes are defined. It is dangerous to
compare curves from different sources.
We shall now illustrate the theory given in sections 27 to 29 by three
objectives, each consisting of two cemented lenses. These are very simple
examples which are not intended to demonstrate what is technically
possible today: they are chosen to illustrate the fundamental problems
encountered in lens designing. We shall therefore assume that the positive
lens consists in each case of the crown glass K 5, and the negative lens of
the dense flint glass SF 1, although these glasses do not necessarily give
the best chromatic correction in all cases. The objective mount will be
assumed to be the stop so that, for the calculation, the centre of the
entrance pupil may be taken to be the pole of the first refracting surface.
This simplifies the computation and the small difference has no significant
effect on the results. In each case, the field angle is n 1 WI = - 0.1, the
perpendicular from the centre of the entrance pupil on to the marginal
ray Al = 12.5 and the focal length of the systemj = 100 mm. The green
mercury line e is used for the calculation of the monochromatic aber-
rations.
The first objective is intended to image an object at infinity. The
design data are: -
r1 = 61.5 mm
d2 = 6 mm glass K 5 j= 52.0mm
r2 = - 47.4 mm
d3 = 3 mm glass SF 1 j = - 107.0 mm .
r3 = - 125.6 mm
Table 14 gives the surface contributions to the aberrations and their
sums. The line below shows the equivalent Seidel approximation SN.
The results show that there is no difference between the exact values
and the Seidel approximation for SVO and SPHP; furthermore, the
values of the surface contributions are so small that they have no effect on
the image quality. They have not been corrected during the design pro-
cess: they are simply the values emerging from the final form. The fact
that they are negligible could be predicted from the small values of the
perpendiculars b on to the principal ray. On the other hand, the small
134 Theory of Aberrations

Table 14
Aberration 1st Surface 2nd Surface 3rd Surface Sum

SPHO [R.E.] - 22.2 SO.6 - 5S.3 0.1


SN: - 21.7 73.2 - 56.6 - 5.0
SVO [°/00] 0 0.1 0.3 - 0.2
SN: 0 0.1 0.3 - 0.2
SPHP [R.E.] 0 0.0 0.3 0.3
SN: 0 - 0.0 0.3 0.3
SVP ["/00] 4.S 6.2 - 10.9 0.1
SN: 4.7 5.6 - 10.5 - 0.2

values for the sums of SPHO and SVP are achieved by suitable compen-
sation of the surface contributions. Comparison of these contributions
in the table will show that the correction of SVP is by no means auto-
matic if SPHO is reduced to zero.
The correction of both these aberrations simultaneously requires the
use of both the degrees of freedom of the achromat, i.e. the form or
"bending" of the achromat and the distribution of the total refracting
power between them. The distribution of refracting power necessary for
the correction of the monochromatic aberrations is only compatible with
the correction of the chromatic aberrations for a few special pairs of
glasses. K 5 and SF 1 give an adequate correction of the chromatic
aberrations.
The exact surface contributions and the Seidel approximations are
not identical for SVP, but the small differences are of no consequence in
practice. This is not the case for the spherical aberration. The Seidel
approximation gives a residuum of five Rayleigh units, an aberration
which would be clearly visible. The actual sum remaining is negligible.
This is an example of zonal aberration bringing a significant departure
from the Seidel approximation. Fig. 41 shows the aberration curves.
The reduced spherical aberration (i.e. SPHO/k2 as explained above) is
plotted against the incident perpendicular at - since sin u t = 0 - for
various wavelengths. The most important curves for our present purposes
are the e-curve and the corresponding Seidel curve, both of which start at
the origin. Since the spherical aberration at the edge of the pupil is
effectively zero, the "belly" of the curve is a direct measure of the zonal
aberration for various values of the incident perpendicular. The value
reaches a maximum of about - 1.4 RE. for at RO 9 mm.
This example demonstrates clearly how the Seidel value of an aber-
ration can be reduced by higher order aberration. Inevitably, zonal aber-
ration remains. This is quite small in the above example, but it can
become very serious in the case of a microscope objective of high aperture
The Secondary Aberrations of Plane Rays 135

and must be further compensated by aberration of a higher order. The


corresponding aberration curves cut the axis representing apertures or ray
perpendiculars twice or even three times, apart from the origin. However,
a complex curve of this sort is not possible for a system consisting of two
cemented lenses.
In Fig. 41 the spherical aberration curves are also shown for other
spectral lines. In general, they do not pass through the origin because
CHLO =F 0 between the spectral line concerned and the line e. The
horizontal distance from the origin gives CHLO in R.E. If Fig. 33 is
rotated through 90° so that its plane is perpendicular to that of Fig. 41,
the intersection points with the abscissa in Fig. 41 are the projections of
the corresponding points of Fig. 33 on to its abscissa. Fig. 41 shows that

Seidel a rmmI
,~
,
approximation

'\
C' F' ft
9

'\
\
~

SPHOlk 2
-5 o 5 [REI
Fig. 41

the spherical aberration of a two lens achromat varies with wavelength.


The difference between e and F' is not significant but between C' and g
it is sufficient to be visually objectionable. This difference in the value of
SPHOjk2 for two different spectral lines (F' and C' are usually meant
unless otherwise stated) is called the Gauss error of the spherical aber-
ration. Although we shall not concern ourselves further with this aber-
ration, it demonstrates clearly that it is not sufficient in high quality
systems to correct the monochromatic aberrations for only one spectral
line.
The objective considered above was designed for an infinitely distant
object. Is it capable of forming a satisfactory image of an object at a
finite distance? This is possible for distances greater than about 2 meters.
If the object distance is reduced to 198 mm from the first surface of the
system, the ideal image lies 196 mm behind the last surface and the
magnification {J is - 1. But the image is of poor quality with SPHO =
= - 36 R.E. and SVP = 10%0. The other aberrations have not changed
significantly. The imagery is even worse if the object distance is reduced
136 Theory of Aberrations

to 98.4 mm giving an image at infinity. We can achieve exactly the same


effect by turning the objective round so that the dense flint lens faces
an object at infinity. This arrangement gives SPHO = - 121 R.E. and
SVP = - 19%0' This example illustrates quite clearly that it is not
sufficient simply to select an objective with the correct focal length:
the precise conditions for which it has been designed must also be borne
in mind.
The following data are for an objective with focallengthf = 100 mm
designed for f3 = - 1: -
r1 = 123.0mm
r2 = - 33.05 mm
d =6mm
2
glass K 5 f = 50.3 mm
d3 =3mm glass SF 1 f = - 103.9 mm.
r3 = - 61.27 mm
For this objective SPHO = 0.1 R.E. and the Seidel approximation
is larger than before, - 6.8 R.E. It follows that the zonal aberration is
also larger with the value - 1.9 R.E. The value of SVP is 0.1%0 and the
Seidel approximation - 0.4%0' The other monochromatic aberrations
remain virtually unchanged.
Design data for an objective with f = 100 mm, a front lens of SF 1
and an object at infinity are as follows: -
r 1 = 41.055 mm
r 2 = 25.280 mm
d2 =3mm glass SF 1 f= - 98.9mm
d3 =6mm glass K 5 f= 48.2mm.

Table 15 gives the aberrations, their surface contributions and the


Seidel approximation.
Table 15
Aberration 1st Surface 2nd Surface 3rd Surface Sum

SPHO [R.E.l - 82.4 94.5 - 12.0 0.1


SN: - 78.7 82.9 - 11.8 - 7.6
SVO [0/001 0 0.0 - 0.2 - 0.1
SN: 0 0.0 - 0.2 - 0.1
SPHP [R.E.l 0 0.0 0.3 0.3
SN: 0 0.0 0.3 0.3
SVP [0/001 11.8 - 7.6 - 4.3 - 0.1
SN: 11.3 - 6.7 - 4.2 0.4

It is interesting to compare these results with those of Table 14 for


the first objective. The plane surface of the third objective introduces
roughly 10 R.E. less spherical aberration than the corresponding (first)
Skew Rays 137

surface of the first objective; but the other two surfaces must therefore
yield higher contributions, in fact 14 R.E. and 24 R.E. more respectively.
This results in a larger discrepancy between the Seidel approximation
and the true value giving a zonal spherical aberration of - 2.1 R.E.
compared with - 1.4 R.E. for the first objective. The third objective is
also inferior from a manufacturing point of view. Although the plane
surface can be produced rather more cheaply than spherical surfaces,
the other two surfaces are steeper than in the first objective so that fewer
lenses can be ground simultaneously on a "block". A further disadvantage
is that the manufacturing tolerances are smaller because of the larger
surface contributions to the aberrations.

§ 30. Skew Rays


Ray Tracing Formulae, the Sine Condition and their
Relationship to the Invariance of the Linear Conductivity
The refracted path of skew rays through an optical system does not
lie in a plane containing the axis. They are defined from a given reference
point by four coordinates. These are based on a right three-dimensional
coordinate system as described in § 26. The reference point E lies on the
axis. There are two possible ways of defining a ray. The first method is to
project it on to the tangential plane (the x-y-plane) and the sagittal plane
(the x-z-plane). The projections are plane rays and can be defined in the
usual way by the coordinates b and sin w or a and sin u. This coordinate
system is important in abelTation theory because it demonstrates the
relationship between plane and skew rays and between the law of refrac-
tion and the invariance of the linear conductivity. But available formu-
lations are too unwieldy for numerical calculations using this coordinate
system. The following coordinates are preferred for ray tracing. A plane
is drawn through the reference point perpendicular to the axis (not to
the ray!). The ray cuts this plane in the point (0; Y; Z), Y and Z being
two coordinates of the ray. Suppose the ray forms the angle v with the
positive x-direction. The quantity cos v will be denoted by~. The symbols
'YJ and, have a similar significance. Thus three further coordinates have
been defined but they are not independent of each other because of the
relation

Simple geometrical relationships exist between the two coordinate


systems. They will be given here without proof. For a and b:
a = Z CoSU, b=Ycosw. (30.1)
138 Theory of Aberrations

For the trigonometrical functions:


. -C ,;
SInU = Vf" + C2 ' cos U = Vf" + c" '
. -'Y} ,;
SIn w =V,;2 +;]2' cosW=Yf"+'Y}2 ' (30.2)

; = cos wcos W - sin w'cos u c= -sinu'cosw


Yl-sin2 usin2 w' 'Yj = Vl - sin2 usin2 w' Vl - sin2 usin2 w'

If d is the separation between the reference points E and E*, measured


positively in the direction of the light, the transfer equations for skew rays
in a homogeneous medium are identical with those for plane rays:
a* = a - d sin u, b* = b - d sin w ,
sin u* = sin u, sin w* = sin w . (30.3)
In the other coordinate system:

y*= Y+d:l
,; , Z* =Z+ dl..
,; , (30.4)
;* =;, 'Yj* = 'Yj, C* = C·
All the above relationships follow from the definition of the coordina-
tes and the principles of elementary geometry. The difference in sign is a
result of the fact that our sign rule for sin u and sin w is the opposite of
that normally used in mathematics.
In the refraction formulae we only use the coordinates Y, Z, ;, 'Yj, C.
The refracting surfaces are normally planes or spherical caps. Later we
shall limit our equations to such surfaces, but at this stage we shall
merely postulate that the surface separating the media with refractive
indices nand n' is rotationally symmetrical to the axis and that it posses-
ses a normal at every point. The latter supposition is true for any "con-
tinous" surface. By virtue of the symmetry of rotation, all normals either
cut the axis or are parallel to it: they cannot be skew to it. To simplify
the treatment, we shall not discuss the case of normals parallel to the
axis and will merely note here that quite simple considerations show that
this special case leads to the same results as the general case. The direc-
tion cosines of the normals are denoted by lX, fl, y.
If the ray cuts the refracting surface in the point (xp; yP; zp) then
Xp = l;, yp= Y + l'Yj, zp =Z + lC,
in which l is the distance from the reference point (0; Y; Z) to this point
on the surface. Suppose the normal at this point cuts the axis in the point
defined by the coordinates (r*; 0; 0) and that this point is at a distance r
from the surface point. Then the coordinates of the surface point can be
Skew Rays 139

represented by the relations XF = r* - rlX, YF = - r{J, ZF = - ry. Com-


bined with those given above, these lead to
19 =r* - rlX,
1'YJ =
Y - r{J,
-
tt;,=-Z -ry.
We now eliminate 1 from these equations by multiplying the second
one by l; and the third by 'YJ and forming the difference. Thus
0=- Yl; + Z'YJ - r ({Jl; - Y'YJ)
and - with e= 1Jr-
e(Yl; - Z'YJ) = (Jl; - 'Y'YJ '
e(zg + r*l;) = 1Xl; - yg , (30.5)
e(yg + r*'YJ) = IX'YJ - {Jg ,
the second and third equations resulting from analogous reductions.
These relationships are purely geometrical and are therefore not con-
cerned with physical processes. It should be noted that the right hand
sides only contain direction cosines, not point coordinates. We shall make
use of this property. But first we shall give corresponding equations for a
ray with the coordinates V', Z', g', 'YJ', l;' which is incident on the refrac-
ting surface at the same point (XF; YF; ZF):
e(Y'l;' - Z''YJ') = (Jl;' - 'Y'YJ' ,
e(Z' g' + r*l;') = 1Xl;' - yg' , (30.6)
e(Y' g' + r*'YJ') = IX'YJ' - (Jg' •
These equations fulfil the condition (26.1) of the law of refraction.
The other two conditions can be expressed by the equations
n'({Jl;-y'YJ) = n"({Jl;'-y'YJ').
-n'(IXl; - 1'$) = - 1t' • (1Xl;' - yg'j , (30.7)
n' (IX'YJ - (Jg) = n' •(IX'YJ' - (J$') .
This follows since a ray with the direction cosines ({Jl; - 'Y'YJ), (yg - 1Xl;)
and (IX'YJ - (J$) is perpendicular to the ray 5 and the normal N, a conse-
quence of
({Jl; - 'Y'YJ)$ + (yg - 1Xl;)'YJ + (IX'YJ - {Jg)l; = 0 and
((3l; - y'YJ)1X + (yg - 1Xl;){J + (IX'YJ - (3g)y = 0,
relations which can readily be derived. Equations (30.7) express the fact
that 5, 5' and N lie in a plane. Here, 1t and 1t' represent any two real
numerical values which are not zero. It is specifically with regard to
the refractive indices that the third condition (26.3) of the law of
140 Theory of Aberrations

refraction is fulfilled by equations (30.7); for it follows from the rules


of analytical geometry (the vector product of two unit vectors) that
({Jt; - Y'Yj)2 + (lXt; - y~)2 + (lX'Yj - n)2 = sin 2 i .
Combination of the three sets of equations (30.5), (30.6) and (30.7)
yields the following relations for the corordinates of the rays 5 and 5' as
a consequence of the law of refraction: -
ne(Yt; -Z'Yj) = n'e(Y't;' - Z''Yj') ,
ne(Z~ + r*t;) = n' e(Z'~' + r*t;') ,
ne(Y~ + r*'Yj) = n'e(y'r + r*'Yj').
If e is cancelled out in the first equation and er* = k introduced into
the other two, the fundamental equations of geometrical optics for the wave
normals in centred systems emerge as:
Ynt; - Zn'Yj = Y'n't;' - Z'n''Yj' ,
n(Ze~ + kt;) = n'(Z'e~' + kt;') , (30.S)
n(Ye~ + kYJ) = n'(Y'e~' + k'Yj') .

The first equation shows that the expression Ynt; - Zn'Yj is invariant
for any refraction in a centred system. With the help of the transfer
equation (30.4) it can be readily demonstrated that it also remains
invariant if the reference point is changed. It follows that the expression
Ynt; - Zn'Yj has the same formal property for any skew ray in any
centred system as LLW = AnW - BnU for perfect imagery, i.e. the
property of invariance. Thus we can take over the entire formulation of
perfect imagery, although the quantities do not have the same signifi-
cance as the light tube coordinates.
It may well be asked at this stage why different coordinates were
chosen for the light tube if the new coordinates Y, Z etc. are clearly
better adapted to the law of refraction and, therefore, to physical reality.
The difference between LLW and the expression Ynt; - Zn'Yj is made
clearer if we introduce the conversion formula (30.2) into the latter and,
using (30.1), obtain
Y nt; _ Zn = a n sin w - b n sin u (30.9)
'Yj V 1 - sin2 u sin2 w
Apart from the purely formal difference between the ray coordinates
and light tube coordinates, the numerator on the right-hand side of (30.9)
is identical with the linear conductivity. The denominator confirms that
the tangential and sagittal components of a ray are not independent of
one another during the process of refraction. The function relating them
is solely determined by the direction of the ray: it is immaterial whether
the direction is a result of refraction at spherical or aspheric surfaces.
Skew Rays 141

The ray tracing formulae for plane rays and the calculation of the light
tube coordinates are simpler precisely because the coordinates in one
section (for example, a and sin u) and those in the other (for example,
b and sin w) are independent of each other. Equation (30.9) shows that
this procedure must be correct for the two principal sections since the
denominator then has the value unity. For rays lying close to the prin-
cipal sections, the rules of perfect imagery yield the correct limiting
values because the departure of the denominator from 1 is proportional
to sin2 u ·sin2 w. This is the real basis for the link between ray coordinates
and light tube coordinates in the formulae (27.1) defining SVO and SVP.
One can imagine the light tube coordinates to be infinitesimally small
components perpendicular to the plane of the ray. The disappearance of
the aberrations then has the significance that skew rays in a small region
perpendicular to the plane of the traced ray behave in a stable manner
similar to that of the plane ray. This is very important in practice, for it
enables the optical designer to proceed with his work on the basis of
relatively few rays. In many cases the tracing of skew rays, which
requires more computation than that of plane rays, is unnecessary.
The invariant
ansinw-bnsinu
(30.10)
V 1 - sin2 u sin2 w
is the most general form known at present of the Abbe sine condition.
It is valid for coordinates of the sagittal and tangential components of a
skew ray in a centred optical system, as explained at the beginning of this
section. In practice, it is mostly applied in the special cases for which the
components in the sagittal or tangential direction are infinitely small.
Even with this limitation, it saves much calculation and gives valuable
information regarding the relationship between aberrations. Although
HELMHOLTZ - and, some ten years earlier, CLAUSIUS - noticed this
invariant at about the same time as Abbe, it is justifiable to name the sine
condition after Abbe because it was he who first recognized its practical
significance for optical instruments and thereby laid the foundation of
design techniques for high quality systems which reduced the amount
of computation necessary to acceptable proportions.
The sine condition is not equivalent to the law of conservation of
energy for transverse waves, although this is often assumed to be the
case. It is a consequence of the law of refraction and thus only defines
the continuity of the phase relationship in the region of the ray.
For numerical computation it is necessary to express equations (30.8)
in a different form. The case Yn~ - Zn1J = 0 may be ignored because the
skew ray lies in a plane containing the axis and may therefore be treated
as a plane ray. If Yn~ - Zn1J oF 0, then according to Theorem 2 in § 9, four
142 Theory of Aberrations

coefficients a, b, e, d with ad - be = 1 exist which lead to the following


relations analogous to (9.2) between the coordinates before and after
refraction: -
Y' = aY + bn1] , Z' = aZ + bnC,
n'1]' = eY + dn1] , n' C= eZ + dnC . (30.11)
Formally, the same relationship between the coordinates also exists in
front of and behind the entire optical system. The procedure for the
determination of the coefficients is indicated by the principal theorem
in the theory of perfect imagery. We divide the process of imagery into
three parts. The first is a transfer from the plane containing the pole of
the refracting surface to a new reference plane containing the point on
the surface common to the three rays S, S' and N. This process does not
change the direction cosines, but the point coordinates are transformed
according to the rule YF = Y + L1], ZF = Z + LC in which L is the
distance between the two reference points measured along the ray. We
now calculate the change in direction of the ray caused by refraction for
the new coordinates. Since the point coordinates remain unchanged,
equations of the form
n'1]' = n1] + CPYF, n'C' = nC + CPZF
must emerge, in which cP is the "oblique refracting power" of the surface.
In the third and final part of the process we must go back the distance L'
measured along the refracted ray S' to the original reference plane.
The coefficients a, b, e, d in equations (30.11) for the refraction of a
skew ray at a surface therefore assume the general form
L'
a=I- 7 <P,

e = <p, (30.12)
In the special case of spherical surfaces with curvature e = l/r we have
yz +Z2
L = r'(~ - YI21J - ZeC - cos i) = e' (~_ YQt} -zQC) + cosi '
cP = - e (n' 'cos i' - n'cos i) , (30.13)
L'=Lt,
in which the values
sin2 i = e2(YC - Z1])2 + (Y e~ + 1])2 + (Ze~ + C)2 ,
. '/ n ..
SillZ =-'SillZ (30.14)
n'
cos i = + V1 - sin2 i , cos i' = + V1 - sin2 i' ,
n'~' = n~ + <P' (L~ - r)
must be calculated beforehand for use in (30.13).
Astigmatism 143

For plane surfaces, a = d = 1, c = b = O. These relations can be proved


from elementary geometrical principles and do not give further physical
insight into the problem. For this reason the proofs will not be given here.
The amount of computation may be reduced by changing the form
of the equations. We insert the coefficients (30.12) into equations (30.11)
and then multiply the equations for Y' and Z' by n' t;'. If, after multi-
plying out the brackets, the last equations from (30.13) and (30.14) are
used, we obtain
Y'n't;' = Ynt; + (n' ·cos i' - n'cos i) '(Y + LTJ),
Z'n't;' = Znt; + (n' ·cos i' - n'cos i) ·(Z + LC),
n't;' = nt; - e(n' ·cos i' - n'cos i)·( - r + Lt;), (30.15)
n'r/ = nTJ - e(n' 'cos i' - n'cos i) ·(Y + LTJ),
nT = nC - e(n' ·cos i' - n'cos i)· (Z + LC) .
The number of computations can be still further reduced if the
products of the right-hand sides are multiplied by nt; and all terms
expressed in the reduced coordinates Ynt; and Znt;. This formulation is
given by M. HERZBERGER in "Modern Geometrical Optics", Interscience
Publishers, New York 1958. HERZBERGER also gives the proofs which
have been omitted above.

§ 31. Astigmatism
Coordinates of Skew Pencils,
Lambert's Radiation Law,
Ray Tracing Formulae for Astigmatism,
Astigmatism and Field Curvature
If the four monochromatic aberrations SPHO, SVO, SPHP and SVP
have been corrected or reduced to sufficiently small values by suitable
lens shapes, it follows that for the given wavelength the centres of the
image and the exit pupil are at the correct points on the axis and that
their edges are imaged with the correct size relative to their centres. But
this does not necessarily imply that the edges are sharp and in the correct
position. This is only true if further aberrations are corrected, namely
astigmatism, field curvature and coma.
The formulae for these aberrations are more complicated than for
those treated previously. To avoid the danger of confusion through too
much detail, we shall confine the analysis to the object imagery only.
The derivation of the corresponding formulae for the pupil imagery is
without exception analogous if the symbols for pupil imagery are used.
144 Theory of Aberrations

Since the size of the image is completely determined by the distortion,


we are concerned solely with the image sharpness at its edge. This depends
on the behaviour of a cone of light whose apex in the object space is at
the edge of the object in the tangential section and whose base is the
surface of the entrance pupiL The principal ray forms the axis of the cone.
The sections perpendicular to the principal ray are no longer circles, as
was the case for imagery of the object centre, but ellipses if the entrance
pupil is perfect. In real optical systems the cone is also deformed by lens
edges and parts of the mounts. As a result of this vignetting the entrance
pupil is formed by two intersecting circles - rather like the eye pupil of
a cat. Since the type and extent of the vignetting is considerably influ-
enced by mechanical requirements arising from lens manufacturing tech-
niques, it is better to ignore vignetting during the computation of aber-
rations and leave its consideration to the final design stage when the
image quality is to be assessed. The loss of brightness and resolution must
also be borne in mind.
We introduce coordinates into the oblique light cone analogous to
those of the marginal and principal rays in the axial case. They are
referred to the principal ray which lies in the tangential section and
forms the angle W with the axis. For the rays lying in the tangential
section of the cone we use the coordinates
at = ± a·cos W , n sin ~It = ± n sin u·cos2 w. (31.1)
Fig. 42 demonstrates the significance of these coordinates. The mar-
ginal ray is also shown in the tangential section.

Fig. 42

This definition has served two purposes: -


1. Measured in the direction of the axis, the centre and edge of the
ob]ect h avet h esame d·Istancesinu=cosw·sinu-cos2w
· a a·cos W from t h e reer-
f
ence point.
Astigmatism 145

2. The cones of the central pencil and the oblique pencil intersect
each other roughly at the edge of the entrance pupil. This is only strictly
true if U or w have small absolute values, i.e. for a distant object or pupil.
But it would be splitting hairs to worry about this discrepancy since the
effect of vignetting is usually very much greater in practice.
In the sagittal direction we shall use the coordinates

a8=± a, n sin Us = ± n sin u"cos w , (31.2)

which also satisfy the above conditions. Of course, it is essential that the
coordinates in the tangential direction change continously into those of
the sagittal direction. Clearly this is possible if the relations

af} = ± a V1 - sin2w cos2{} ,


n sin Uf} = ± n sin u"cos w VI - sin2w"cos2{}, (31.3)

apply, in which {} is the angle between the direction under consideration


and the tangential plane, a and sin u are the coordinates of the marginal
ray of the central pencil and w is the angle of the principal ray with the
axis. The four different combinations of sign correspond to the four
quadrants of the light cone. {} = 0° and {} = 180° yield equations (31.1)
for the tangential direction, while {} = ± 90° yields equations (31.2) for
the sagittal direction"
The energy relationships are significant for imagery with oblique
pencils. If we take an object element Bs in the sagittal direction - that
is a small length 2 B. perpendicular to the plane of Fig. 42 - and join
its edge to the centre of the pupil, then this new "principal ray" has the
coordinates b. and n sin y."cos w relative to the axis of the light cone, in
which the factor cos w is given by the geometry of the system, that is by
the increased distance to the pupil. By virtue of (31.1) the linear conduc-
tivity for the imagery of this sagittal object by the tangential pencil is
given by

at"n sin Y."cos w - b.·n sin Ut = (a"n sin Y. - b·n sin u) "COS2 w"

It is reduced by the factor cos2w compared with the corresponding value


for the centre of the image.
Conversely, if we image an object element B t in the tangential direc-
tion by the sagittal pencil, its "principal ray" has the coordinates
be"cos w and n sin yt"cos 2 w relative to the axis of the cone. Combined
with the sagittal "marginal ray", it yields using (31.2) the linear con-
ductivity

a."n sin yt"cos2 w - bt "cos won sin u. = (a"n sin Yt - bon sin u) "COS2 w .
10 Zimmer, Geometrical Optics
146 Theory of Aberrations

If the corresponding analysis is performed for any two mutually per-


pendicular directions, the resulting factor with which the linear conduc-
tivity of the image centre must be multiplied is
cos w· V1 - sin2 W +i sin 4 W·sin2 2{} "'" COS2 W + l- sin4 W sin2 2{} ,
in which {} is the angle defined above for equation (31.3). The additional
term only becomes significant compared with cos w if w is very large.
This term has no special physical basis: it arises because the coordinate
system consisting of the perpendiculars a, b and apertures n sin u, n sin w
is less well adapted to oblique rays than the coordinate system x, y, z,
g,1), Cgiven in § 30.
So far as practical designing of optical systems is concerned, one can
generally ignore these finer points of the theory and simply assume that
the linear conductivity in the oblique pencils falls off from the centre
to the edge of the image with the factor cos2 w. Since the radiating power
is proportional to the square of the linear conductivity if the radiation
density is constant, the power falls off towards the edge of the image
according to cos4 w. This fall off in brightness is a consequence of Lam-
bert's radiation law and is occasionally termed "natural vignetting" in
connexion with optical instruments. It is significant even for quite
"normal" field angles. In miniature photography the normal field angle
is w = 23° for which cos w = 0.92 and cos 4 w = 0.72. Thus in the corners
of the image there is less than three quarters of the radiation power
present in the central pencil. In practice, this will be still further reduced
by vignetting losses due to the edges of lens mounts. However, this loss
of brightness at the edge of the image is only serious in special cases
because normal emulsions can compensate significant differences in
exposure.
All rays of the oblique pencil lying outside the tangential section are
skew rays and ray tracing requires much more computation. We shall
deal with this general case in the next section. For the moment, we shall
assume that the aperture of the pencil is sufficiently small that the path
of the rays can be obtained by differentiation of the refraction formulae
for the principal ray. The differentials of the coordinates will be given
the symbols At, nUt, As and nUs. This terminology assures a formal
connexion with the coordinates of the oblique pencil. In the object space
we give the differentials the same numerical values as the coordinates of
the corresponding rays of the oblique pencil.
As a first step in the derivation of the ray tracing formulae for the
differentials we shall take the refraction normal N as the reference axis
y = 0, Z = o. An alternative way of expressing the same reference system
is to suppose that the principal ray cuts the surface at exactly the same
point as the axis of the system; for the axis is also a surface normal in
Astigmatism 147

centred systems. As before, we assume the principal ray to be in the


x-y-plane and give the incidence angles the symbols j and j'. Since we do
not wish to confine the differentiation to the plane of refraction, we must
apply equations (30.15) for skew rays to the principal ray. In these
equations L = 0, Y = Y' = 0, Z = Z' = 0 so that

n'~' = n~ + (n' cosj' - n cosj) . (31.4)

Differentiation of the last two equations of (30.15) yields

n'dr/ = n dn - e(n' cosj' - n cosj) dY,


n'dC' = n dC - e(n' cosj' - n cosj) dZ, (31.5)

since dL = o. This is geometrically evident because the coordinates dY


and dZ define a point in the tangential section of the surface. By differ-
entiation of the first equation of (30.13) it follows formally that dL = o.
Differentiation of the first two equations of (30.15) gives, when combined
with (31.4), the geometrically evident and trivial results dY' = dY,
dZ' = dZ.
The differential formulae (31.5) are formally the same for the tangen-
tial direction (y-direction) and sagittal direction (z-direction). Indeed, the
result is the same for any direction; for we chose the x-y-plane for the
principal ray only because it is convenient for our three-dimensional
treatment but did not suppose that C must necessarily be zero in the
derivation of the formulae.
The unsymmetry which is characteristic of astigmatism for different
directions perpendicular to the principal ray is a consequence of a
coordinate transformation. It becomes evident if we take the principal
ray instead of the normal to the surface to be the axis. The transformation
formulae are
dY = Atfcosj, dZ=A s ,
dn = - Ut"cosj, dC= - Us" (31.6)

Here it has been assumed that the principal ray lies in the x-y-plane and
consequently C= o. They can be derived formally if the last two equa-
tions of (30.2) are differentiated for u = 0 and w = - j and these differ-
ential coefficients are inserted in the definitions (31.1) and (31.2) of the
coordinates for oblique pencils. Fig. 43 gives a geometrical interpretation
for the tangential section. The reader able to interpret three-dimensional
effects will have little difficulty in picturing the corresponding relations
in the sagittal section.
10·
148 Theory of Aberrations

Inserting (31.6) in (31.5), we arrive at the so-called astigmatic rejrac-


tionjormulae for rays in the immediate neighbourhood of a principal ray:
A:Jcos j' = At/cos j ,
n'U; ·cosj' = nUt·cosj + e(n' cosj' - n cos j)At/cos j ,
A; = As , (31.7)
n'U; = nUs + e(n' cosj' - n cosj) As.
The corresponding transfer equations emerge directly from the defini-
tion of the coordinates for oblique pencils:
A~ = At + DnUt , A: = As + DnUs ,
n*U~ = nUt, n*U: = nUs, (31.8)
in which D must now be measured along the principal ray. Its relation
to the projected distance d along the axis is given by
D = - din cos w .
The angle w gives the inclination of the principal ray to the axis.
However, d is not the separation between the poles of two refracting
surfaces as in the case of ray tracing for plane rays but the axial separa-
tion of the incidence points of the principal ray at these two surfaces.

Fig. 43

Its numerical value can be determined from the axial separation and the
"sagittae" x [see Fig. 38 and formulae (26.12)J.
If these formulae are applied to all the separations and refracting
surfaces of a system, the coordinates A;, n'U;, A; and n'U; in the image
space may be calculated from the corresponding coordinates in the object
space. Thus astigmatism is defined by the formula
AST = 1831· (A; n'U; - A; n'U;) R.E. , (31.9)
a definition which is analogous to that for spherical aberration.
It is a measure of the tangential unsharpness A; in the plane A; = 0
of the sagittal image in Rayleigh units of the sagittal pencil aperture;
Astigmatism 149

or what amounts to the same thing, a measure of the sagittal unsharpness


A: in the plane A; = 0 of the tangential image in Rayleigh units of the
tangential pencil aperture. Here, the word "image" only has the formal
significance that A; = 0 or A: = 0; for the presence of astigmatism may
imply that no true image of the object can be formed.
We can use the apparatus shown in Fig. 31 to investigate these effects
experimentally. Chromatic aberrations will be suppressed by green filters,
the spherical aberration by limiting the diameter of the stop P to 25 mm.
At the edge of an object with height 15 mm (30 mm diameter) there can
be no sign of astigmatism since the calculated value lies beneath 0.01 R.E.
In this experiment our interest is confined to the edge of the object. It is
thus permissible to increase the principal ray angle w by setting up the
object 0 excentrically to the axis. To obtain agreement with our conven-
tions regarding sagittal and tangential directions, 0 should be displaced
in the vertical direction, for example downwards. The observing micro-
scope would then have to be moved upwards. In practice, it is simpler to
couple the stop P and the lens L together and to tip both about the
horizontal axis of symmetry of the lens. Normal lens holders on optical
benches only permit a rotation about the vertical axis of symmetry.
Physically this amounts to the same thing but can cause confusion with
the symbolism. For this reason, we shall suppose that the lens is tilted
about its horizontal axis of symmetry so that the tangential plane is
vertical and the sagittal plane horizontal.
We shall use the foil with fine pinholes as object, the same object that
we used for investigating the influence of spherical aberration on diffrac-
tion patterns. The observing microscope is first focused on these diffraction
images with the lens in the vertical position, then the lens together with
the stop is slowly tilted. The diffraction pattern loses its rotational
symmetry when the angle of tilt reaches 1.4 0 , its core becoming a line
in the vertical (tangential) direction. If the microscope is pushed towards
the lens we find a plane for which the appearance is identical except that
the pattern is rotated through 90 0 so that the line lies in the horizontal
(sagittal) direction. The length of the line is the same as the diameter of
the first diffraction ring. Between these two focal planes the core of the
diffraction pattern is round but the diffraction rings are interrupted by
dark areas in the directions at 45 0 to the sagittal and tangential direc-
tions. This pattern thus has four axes of symmetry whereas the figures
with the lines, corresponding roughly to the sagittal and tangential image
points, have only two axes of symmetry. With this angle of tilt of 1.4 0
the astigmatism amounts to - 0.5 R.E., an aberration which is approxi-
mately at the threshold level of visual detection in the diffraction image.
In order to see the effect of astigmatism on the geometrical imagery,
we replace the foil with its pinholes by the slide with the bar chart.
150 Theory of Aberrations

An effect only becomes visible with an angle of tilt of about 2 0 which


gives astigmatism of - 1.0 R.E. The effect is that the sagittal (horizontal)
bars are unsharp in the sagittal focus whereas the tangential (vertical)
bars are sharp. In the tangential focus, which lies 0.4 mm nearer the
lens L, the reverse is the case. These effects may seem very confusing to
the layman. One can picture the effect in the sagittal focus by considering
the symmetry of the ray path with respect to the tangential section.
In the tangential focus, on the other hand, an analogous explanation is
only valid with certain reservations because the corresponding rays in
the upper and lower halves of the pupil do not necessarily arrive in the
plane of the tangential focus with the same phase.
Unlike spherical aberration, which reduces the contrast at the edges
of geometrical images, astigmatism causes unsharpness, an aberration
which is usually more serious because it makes the resolution of fine detail
not merely more difficult but impossible.
In many books the origin of astigmatism is explained as a consequence
of different surface curvatures in the tangential and sagittal sections at
the point where the oblique pencil is refracted. This explanation seems
plausible at first sight but is not correct. Since the astigmatism of the
lens L used above is negative, the curvature in the tangential section
would have to be stronger than in the sagittal section. But the tangential
section represents a great circle of the sphere and a section of larger radius
than this does not exist. A simple experiment demonstrates that the above
explanation must be wrong. A plane parallel plate of thickness about
10 mm is introduced into the system between the lens L and the image 0'
of our previous set-up. The lens is not tilted but the plate is inclined at 45 0
to the system axis. This arrangement gives a strongly astigmatic image
although the curvatures of the refracting surfaces of the plate are identical
in all directions, in fact zero.
The law of refraction expresses the conservation of the phase relations
in the case of normals to a refracting surface. This cannot therefore be
generally true for other rays. This is really all that we can say regarding
a simple geometrical explanation of astigmatism, and it is exactly what
the symmetrical formulae (31.5) and the unsymmetrical formulae (31.7)
express.
The astigmatic refraction and transfer formulae are identical with
equations (7.6) and (7.7) for axial object points because cos j = cos w = 1.
Thus an axial image point of a centred system cannot show astigmatism.
It is for this reason that the presence of astigmatism in the centre of an
image is a characteristic and sensitive measure of centring errors in the
most general sense, namely the tilt of a lens axis relative to the system
axis and the departure of a refracting surface from rotational symmetry.
The destruction of the rotational symmetry of the diffraction image by
Astigmatism 151

astigmatism is the most sensitive centring criterion in setting up a lens


system on an optical bench.
If astigmatism is corrected for extra-axial image points, the sagittal
and tangential image points coincide. But this does not necessarily imply
that they lie in the plane perpendicular to the axis coinciding with the
plane of the ideal image. Usually the image surface is curved. This can
be readily seen in a microscope if a fairly simple objective (an achromat)
is used: the centre and edge of the image cannot be sharply focused
simultaneously. This effect is not particularly disturbing in visual obser-
vations because the point of interest can be focused as required. But it
becomes very serious if the image is to be recorded photographically on a
film or plate, for no refocusing is possible in this case.
The field curvature we are now discussing should not be confused
with the apparent image curvature which was mentioned in connexion
with distortion in binoculars.
Field curvature is not a physical aberration: it has influence neither
on the appearance of the diffraction image nor on the quality of imagery
of coarse structures. It conflicts solely with a geometrical ideal, the
concept that planes should be imaged in planes. It follows that instead
of a physical basis for its definition, we have the geometrical condition
that the image point should lie in the plane perpendicular to the axis
containing the ideal image point of the central pencil. This requirement
can be applied equally well to the sagittal or tangential point, or to some
other point which is related to them both. This is only a matter of prac-
tical convenience depending on the requirements of the system. The
sagittal field curvature is defined by the equation
SAG = 1831· (A; n'U' 'cos w' - A'n'U;) R.E. (31.10)
This formula gives the distance (depth) between the sagittal image
point and the ideal image centre in Rayleigh units. The quantities A'
and n'U' without suffixes are light tube coordinates for perfect imagery.
The factor cos w' transforms the depth to the axial direction.
The apparatus shown in Fig. 31 is not very suitable for the demon-
stration of field curvature because the values of SAG and AST are similar
for the lens L and SAG can only be reliably observed if the object and
observing microscope are moved perpendicularly to the axis without any
tilt of the lens.
152 Theory of Aberrations

§ 32. Analysis of Astigmatism


Surface Contributions, Seidel Approximations,
Petzval Curvature
The values given by equations (31.9) and (31.10) for the astigmatism
and the sagittal field curvature are independent of the choice of reference
point for the calculation. This follows from the form of the equations,
but it can also be checked using the transfer equations (31.8) and (7.6).
The contribution of a refracting spherical surface to the astigmatism is
L1 AST = AST' - AST ,
= 1831 {(A; n'U; - A; n'U;) - (At nUs - A. nUt)}. (32.1)
If we introduce equations (31.7) for astigmatic refraction, we obtain after
reduction
L1 AST = - 1~31 A: (n sin J')2 (U~n' _U,) + (COS ~ -1) AST . (32.2)
COS1'COS]' n cos]'

In contrast with the formulae given previously, the surface contri-


bution contains the value AST before the refraction.
Equation (32.2) is of no practical use because the difference for-
mula (32.1) is much more convenient for computation. Nevertheless,
(32.2) is valuable in aberration theory because it leads at once to the
Seidel approximation formula. We suppose that the field angle wand the
perpendiculars b are so small that the error in setting cos w = 1 and
cos j = 1 is negligible. Then At = A, Us = U and the second term
disappears. If we make the further approximation n sinj = nj, the Seidel
approximation emerges as
(U' U)
L1 AST"" - 1831· A . (nj)2. ;[ - -; R.E. (32.3)

This equation shows that there is a close relationship between


astigmatism and spherical aberration. Apart from the factor Y2, the only
difference between the two approximate formulae is in the factors (nI)2
and (nj)2. Thus the spherical aberration and astigmatism contributions
of a surface always have the same sign. Within the range of validity of the
Seidel approximation, L1 AST becomes zero when the object is situated
either at the pole of the refracting surface (A = 0) or at an aplanatic
point (U'ln' = Uln), or if the stop lies in the centre of curvature of the
sphere (nj = 0).
Equation (32.3) also shows that for fixed object and pupil positions
the astigmatism is, to a first approximation, proportional to the square
of the aperture n l 'sin u l or of the incident perpendicular ai' and that it
grows from the value zero in the centre of the image as a quadratic func-
tion of the field angle if the aperture remains fixed. For large field angles
Analysis of Astigmatism 153

higher powers become significant. Their influence is defined as zonal


aberration, exactly as with plane rays:
(AST) = AST* - l2. AST . (32.4)
Here, as in § 29, AST* is the astigmatism for a smaller field angle W*
and l = nj sin wj*jn j sin Wj = bj*jbj is the ratio of the apertures or per-
pendiculars of this ray and the principal ray corresponding to the edge
of the image.
The effect of a refracting surface on the sagittal field curvature is
given by the expression
,1 SAG = SAG' - SAG (32.5)
= 1831 {(A; n'U' ·cos w' - A'n'U;) - (A.nU . cos W - AnUB)} •
There is no known explicit formula for ,1 SAG analogous to those
given for the surface contributions of aberrations of plane rays and for
the astigmatism. The only formula available is the Seidel approximation
which is thus not derived by the simplification of an exact relation but
by suitable approximations and reduction of (32.5). The result is
A (LLW)2 (1
(U' nU) + 1831-- 1)
,1 SAG"", - 1831"2 (nJ)2 ;; - 2-e ;;- n R.E.
(32.6)
The reader who is interested in the derivation of (32.6) is referred to
a paper by H. A. UNVALA: "Quasi-Invariants in Ray-Tracing", Optik,
Vol. 19, 1962. The first term is one half of the Seidel surface contribution
to the astigmatism and the second is called the surface contribution to
the Petzval curvature, in honour of the Hungarian mathematician and
optical designer J. PETZVAL (1807-1891).
Equation (32.6) yields a very important result for the design of optical
systems. If the astigmatism is corrected so that
2:,1 AST = 0,
then within the Seidel approximation the sagittal field curvature is
given by
2: ,1 SAG"", 915.5 (LLW)2. 2:e (ljn' - ljn) R.E.
Neitherthe coordinates of the lighttube nor those of the rays appear in this
expression which may be evaluated very simply from the linear conduc-
tivity of the system, the surface curvatures and the refractive indices.
The Petzval curvature
PTZ = 915.5 (LLW)2 2:e (~ - ~) (32.7)

is therefore a criterion which is easy to manipulate and gives immediate


information as to whether or not a flat field is possible.
154 Theory of Aberrations

For a thin lens made from a glass of refractive index n and surrounded
by air, the quantity 2/1 (,I, - ~) may be written
2/1 (~
n
-~)
n
=~
1"1
(~n - 1) + ~ (1 - ~)
1"2 n
= (~ -~) (~) = - £. ,
1'1 1"2 n n
in which rp is the refracting power of the lens. This means that the Petzval
curvature is negative for a positive (converging) lens so that the image
is curved towards the incident light. The reverse is the case for a negative
lens. It is evident that a single lens can never produce a flat image of a
flat object.
The lens thicknesses do not appear in the expression for the Petzval
curvature. Nevertheless, it is precisely these thicknesses which form an
invaluable aid for the correction of field curvature in practical systems.
This appears surprising at first sight but is readily explained by the fact
that the thicknesses influence the refracting power of a lens according
to the equation rp = rpl + rpz + fJrplrpZ in which fJ = - din. As an example,
let us consider a "thick meniscus", a lens in air which has identical radii
r 1 and r2 both in value and sign. According to the equation above, its
contribution to the Petzval curvature is zero. The refracting power of
the meniscus is positive and has the value rp = drpUn because rpl = - rp2'
Thus it functions as a positive lens without producing negative Petzval
image curvature, a property which is used extensively in optical design.
Our first numerical example will be imagery by the plane-convex lens
with focallengthj = 321 mm which was described in § 15. However, we
shall assume a stop diameter of 25 mm instead of 30 mm and the field
angle will be increased from - 0.0025 to n l sin WI = - 0.035 (WI = 2°).
This corresponds to an increase of the object diameter from 30 mm to
419 mm. Table 16 gives the numerical values for two cases.
In the first case the curved surface (r = 200 mm) of the lens is turned
towards the object. We used the lens in this way to demonstrate the
chromatic aberrations, the spherical aberration and the astigmatism.
In the second case the lens is turned round so that its plane surface faces
the object. The values given are the surface contributions to the spherical
aberration, astigmatism and sagittal field curvature and their sums.
Table 16
Case Aberration First surface Second surface Sum

SPHO [R.E.] - 0.8 - 0.3 - 1.1


AST [R.E.] - 0.4 - 0.6 -1.1
SAG [R.E.] - 0.6 - 0.3 - 0.9
2 SPHO [R.E.] - 0.0 - 4.3 - 4.3
AST [R.E.] - 0.0 -1.1 -1.1
SAG [R.E.] - 0.0 - 0.9 - 0.9
Analysis of Astigmatism 155

The Seidel approximations agree exactly with these values within


the accuracy of the computation. The zonal aberrations are therefore
negligible. The Petzval curvature is PTZ = - 0.3 R.E. in both cases and
arises from the refraction at the spherical surface. The contribution at
the plane surface is zero since (! = o.
The tabulated values show that it is better to have the curved lens
surface directed towards the distant object. The two arrangements are
effectively identical as far as the effect on the astigmatism at the edge of
the field is concerned, although the individual contributions of the two
surfaces show marked differences.
The second example we shall consider is the achromat with f = 100 mm
given in § 29. The first of the three systems given will be used. Table 17
gives the surface contributions to the astigmatism and to the sagittal field
curvature, their sums and the Seidel approximations SN. The last two
lines give the Petzval curvature and one half of the Seidel approximation
SN(AST) for the surface contributions to the astigmatism. The sum of
these two lines gives the Seidel approximation for the surface contri-
bution to the sagittal field curvature.

Table 17
Aberration First surface Second surface Third surface Sum

AST [R.E] - 10.5 4.6 - 20.5 - 26.4


SN: - 10.6 4.6 - 20.7 - 26.6
SAG [R.E.] - 13.3 4.6 - 15.0 - 23.7
SN: - 13.3 4.6 - 15.2 - 23.9
PTZ [R.E.] - 8.0 2.3 - 4.8 - 10.5
'/2 SN (AST): - 5.3 2.3 - 10.4 - 13.4

Although the aberrations of plane rays are very well corrected, the
sums show that this objective is of little value for the field of n, . sin w,
= - 0.1 which was assumed in § 29 on account of the astigmatism
and field curvature. The field of reasonable correction is limited to about
n, . w,
sin = - 0.02, i.e. about one fifth of that assumed above. This
reduces these aberrations to about 1/25. If very good image quality is
required, the field must be further limited to about n, . sin W, = - 0.014,
an angle w, of less than 1°.
If the surface contributions to the astigmatism are compared with
those to the spherical aberration SPHO (Table 14 in § 29), it is clear that
the effect of the cemented surface on the astigmatism is quite inadequate
to compensate that of the outer surfaces. The marginal ray is refracted
downwards by the first surface and meets the cemented surface with a
large perpendicular and consequently a much larger incidence angle i
156 Theory of Aberrations

than that of the principal ray which is moving upwards and strikes the
cemented surface with a small perpendicular. This difference between
the incidence angles gives a simple geometrical picture of the very
different effect of the cemented surface on astigmatism and spherical
aberration. From this it may be thought that a different stop position
which yields larger perpendiculars for the principal ray within the system
will give more favourable values for the astigmatism. However, this is not
the case. Within the range of validity of the Seidel approximation, a pupil
shift has no effect on the astigmatism of this objective because SPHO
and SVP are corrected for the pupil position chosen initially. We shall
return to this point in § 34.

§ 33. Coma
Aberrations of Oblique Pencils of Finite Aperture and
their Relationship to the Spherical Aberration of the Object
Imagery and the Distortion of the Pupil Imagery
If the infinitesimally narrow pencils about the principal ray are
opened up to give pencils with a finite aperture, a further aberration
appears in addition to astigmatism and field curvature. This is coma.
Its origin is fundamentally similar to that of spherical aberration and
distortion for plane rays. Indeed, it is closely related to these aberrations.
Nevertheless, all the resulting effects are more complicated and can only
be analysed in terms of approximations because the symmetry of the
central pencil is absent in oblique pencils. Our present knowledge of
coma is comparable with what was known of the aberrations of plane rays
before 1850. We have formulae which enable us to trace any desired skew
rays through the system and can therefore determine all ray coordinates
with the necessary accuracy; but there is no exact theory available which
can explain and correlate the many effects and problems that arise. In
any event, we do not posses a theory which is sufficiently simple, general
and flexible to provide a working tool in practical optical design. In
particular, we lack reliable rules for calculating the influence of coma
on image quality. The broad effects are known both empirically and
theoretically but even the experienced designer must rely on guesswork
and experiments to decide whether finer effects, which may be near the
limit of visibility, are acceptable or not.
For these reasons, no attempt is made in this section to give a com-
plete account of present knowledge of coma. This would not only be
outside the scope of this small book but also beyond its author's powers.
Instead, the treatment will be limited to a geometrical description which
Coma 157

is easy to understand and to an attempt to interpret fundamental prop-


erties in terms of rule-of-thumb formulae.
The aberrations which we term coma exist for every ray of the oblique
pencil round the principal ray, but they are best treated by considering
the two directions perpendicular to the principal ray which have already
been defined, i.e. the tangential and sagittal directions.
All the rays which are tangential to the principal ray lie in a plane,
the tangential section. Ray tracing of these rays can therefore be carried
out with the simple formulae for plane rays if the coordinates of these
rays are referred to the axis and not to the principal ray. After tracing
the rays through all the refracting surfaces of the system, the coordinates
are again transformed to the principal ray as reference axis. This is
irritating but not serious. In this way the coordinates a; and n'sin u;
of a tangential coma ray are determined and these give the tangential
coma of the ray, defined according to the same principle as the spherical
aberration, as
TK = 1831 (a; n'U; - A; n' sin Ut) R.E.
The reference ray here is a limitingly close neighbouring ray to the prin-
cipal ray in the tangential section whose coordinates in the object space
agree numerically with those of the coma ray. There is a difference in the
meaning of the large and small symbols in the above equation: a; and
n'sin u; are true coordinates whereas A; and n' U; are differential coordinates
whose values were given by equations (31.7) and (31.8) in connexion with
astigmatism. The reference point for the two pairs of coordinates must
naturally be the same.
The tangential coma can be represented graphically in the same way
as the spherical aberration. If TKjk2 is plotted as a function of n sin Ut
in the object space, k being as before the ratio of the aperture of the
traced coma ray and the maximum aperture defined by the edge of the
stop, the result is the tangential coma curve. The form is similar to the
curve for SPHOjk 2 , but tilted. Fig. 44 gives an idea of this effect.
The "tangential coma" contains, in fact, two aberrations. The first
is a symmetrical part corresponding to spherical aberration which can be
determined numerically as t (TKo + TK u ), the suffixes 0 and U referring
to rays in the upper and lower halves of the pencil respectively. The
second is an unsymmetrical part given by t (TKo - TKu) which defines
the tilt of the curve. There is no corresponding aberration for the central
pencil. However, this division of the tangential coma into symmetrical
and unsymmetrical parts is not generally carried out since in many cases,
particularly for systems with high aperture or field or both, the curves
for tangential coma and spherical aberration show much greater differ-
ences than that indicated by Fig. 44. The division then no longer helps
158 Theory of Aberrations

very much and it is better to consider the effect of the whole curve on the
image qUality.
The geometrical significance of the coma curve in Fig. 44 is simply
that the tangential rays focusing from the space above the principal ray
cut the latter in front of the tangential inlage point (defined by A; = 0)
whereas those from the space below the principal ray cut it behind the
tangential image point. Thus all the rays cut an image plane containing
U Ut

Fig. 44

the tangential image point below the principal ray and form an unsym-
metrical ray pattern as shown in Fig. 45. This suggests at once that the
image will be unsharp on one side. In fact, the visible tail of unsharpness
in the image of point objects is the origin of the name given to this aber-
ration. However, the matter is not so simple as the geometrical ray pattern
would suggest. For the wavefronts represented by the upper and lower
coma rays do not necessarily have the same phase at the point of inter-

Fig. 45

section of the two rays. For this reason, it is not possible to deduce the
distribution of intensity in the image from the ray pattern alone. In
practice, the tilt of the coma curve always has an unfavourable effect on
the image quality and the designer must try to reduce it to small values;
but there is no generally valid rule for the amount of tangential coma
permissible before the resolution of the diffraction image or, in the case
of coarser structures, the sharpness or contrast begin to be affected.
Coma 159

As an example, the tangential coma of the achromat with! = 100 mm


given on page 133 is plotted for the principal ray angles sin WI = - 0.05
and sin WI = - 0.1. A comparison of these curves with the spherical
aberration shown in Fig. 41 reveals relatively little unsyrnmetry. Thus
the tangential coma of the achromat is well corrected, even for field
angles for which the objective is useless because of astigmatism. This is
not a coincidence but a consequence of the relationship between coma
and SVP. We shall return to this point after dealing with sagittal coma.
The rays which must be traced for the determination of the sagittal
coma lie in a plane in the object space containing the principal ray and
perpendicular to the tangential section. They are skew to the axis and

a {mm1 a{mm]

-2

sin w,=-o.05 sin w,=-o.1


Fig. 46

do not generally remain in the initial plane after refractions have taken
place. The tracing of these rays must therefore be carried out with the
formulae for skew rays given in § 30. The amount of computation is
greater and the geometrical effects are correspondingly more complicated.
In spite of this, sagittal coma is easier to understand than tangential
coma because the invariance formula (30.10), known as the Abbe sine
condition, is valid for skew rays:
ansinw-bnsinu
Y1---sin2 u'sin2 w = const .
The condition that rays leaving an object point on the principal ray
of the oblique pencil shall meet in an image point on the principal ray
behind the system can only be satisfied for quite definite coordinates
sin u'. For a' = 0 for the image point if the image is at a finite dis-
tance and b' has a fixed value determined by the principal ray. Thus,
apart from an error which is quite insignificant in practice and which
arises from the quantity sin w' in the square root of the denominator,
sin u' can be calculated from the sine condition. By virtue of the sine
160 Theory of Aherrations

condition, the projection of the skew rayon to the sagittal plane therefore
gives us complete information as to whether the rays in the sagittal sec-
tion of the oblique pencil meet in a point on the principal ray or not.
The degree to which the rays fail to meet in a point is expressed by
two aberrations. The first is the longitudinal sagittal coma which is
defined by the expression
SLK = 1831 (a; n'U; - A; n' sin u;) R.E. ,
a; n' u;
in which and sin are the coordinates of the rays in oblique pencils
defined in § 31 and A;, n'U; the corresponding coordinate differentials of
a neighbouring ray lying limitingly close to the principal ray. The analogy
with the tangential coma is purely formal but there is a close relationship
with the spherical aberration of the object imagery. On account of the
symmetry of the ray paths to the tangential section, the corresponding
rays in the front and rear halves of the pencil must always meet each
other with the same phase. The curve for SLK must therefore be sym-
metrical for the two halves of the pencil and reduces to the curve repre-
senting SPHO when the field angle becomes small.
The second aberration is the transverse sagittal coma. This corre-
sponds to the tilt of the tangential coma and is defined by the expression
SQK = 1000 (a'n' sin w' - b'n' sin u'
a'n' sin w~ - b~n' sin u'
-1) 0/
00 .

The quantities b~ and n' sin w~ are characteristic of the path of the
principal ray in the tangential plane. The other four quantities are the
coordinates of skew rays as explained at the beginning of § 30, not the
coordinates for oblique pencils referred to the principal ray. The value
of S QK is independent of the choice of the reference point, a consequence
of the form of the equation. If the reference point is placed at the inter-
section point of the skew ray with the tangential plane, then a' = 0 and
the expression SQK = 1000· (b'/b~ - 1) %0 gives the "perpendicular"
error between the skew ray and the principal ray in pro mill of the length
of the perpendicular b~ on to the principal ray. This is the justification
for the name given to this aberration.
Since the computation of sagittal coma requires the tracing of
skew rays, it is important for practical design work to possess approxi-
mate formulae which can be applied to small angles. Obviously SPHO
meets this requirement in the case of longitudinal sagittal coma. For
transverse sagittal coma the approximation is provided by - SVP.
This becomes clear if we divide the numerator and denominator in S QK
V
by 1 - sin2 U' 'sin2 w'. The numerator is then an invariant for the
whole system which means that the coordinates of the object space may
be inserted. In the special case of small angles WI' the numerator reduces
Coma 161

to AnW - BnU = LLW. This does not mean that A' is an approxi-
mation for a' or U' for sin u' behind the system: our only information is
that LLW gives an approximate value for the numerator, even behind
the system, and that this is a consequence of the sine condition. In the
denominator we take B' to be an approximation for b;' and n'W' for
tt' sinw;..This is based on the assumption that the principal ray has only
small aberrations behind the system. This is always true if the angle WI
is made sufficiently small because (29.1) shows that SVO and SPHP are
proportional to the square and the fourth power of n l sin Wj respectively.
As an approximation for a' and n' sin u' we use the coordinates of the
marginal ray in the sagittal plane, an assumption which is certainly
valid for small angles WI in centred systems because of the continuity of
the function. In this way we derive for S QK the approximate expression

K = 1000 (,
an
'w' - LL; I.
nSlllU
,- 1) 0/00 ,

which is called the coma number. If this result is compared with the
equation in (27.1) for SVP, we see that the numerator and denominator
are reversed. Thus we have the relation
SVp·K
SVP + K + --woo = 0
or, in practice, K"" - SVP because the third term is negligible if the
coma is kept small, a necessary condition for adequate image quality in
practical optical systems.
We conclude that there exists a relationship between the distortion
of the pupil and the transverse sagittal coma analogous to that established
between the chromatic differences of magnification for the object and
pupil imagery. This is not a coincidence. Just as the ideal object and
pupil imageries are related to each other by the invariance of the linear
conductivity, so are the real imageries linked by the invariance of the
sine condition for skew rays. The skew ray from the edge of the object
to the edge of the entrance pupil, and from the edge of the exit pupil to
the edge of the image, is the sagittal coma ray of the object imagery and
the pupil imagery. This becomes clear if one imagines the conventions
for the sagittal and tangential planes with respect to the marginal and
principal rays, as described in § 30, to be reversed. A small pencil aperture
corresponds to a small field angle in the pupil imagery. If the field angles
are made so small that no aperture aberrations of the spherical aberration
type can appear in the sagittal section of the pupil imagery, it follows
that SVP must be the only aberration present in this imagery. If the
pencil aperture of the object imagery is large, spherical aberration will
be present, given by SLK. In the sense that the sagittal coma rays cut
the tangential plane above or below the principal ray, SQK will also be
11 Zimmer, Geometrical Optics
162 Theory of Aberrations

presentasa type of distortion. It is a consequence ofthe sine condition that


the relationship given above between these two distortions is formally the
same as that existing between the chromatic differences of magnification.
The approximation SQK "'" - SVP is particularly important in prac-
tice. For SVP is not only simple to compute; it has the further advantage
that the surface contributions LI SVP or their Seidel approximations
enable the designer to observe the changes in the transverse sagittal coma
that take place within the optical system and thereby obtain the infor-
mation necessary for the correction of this aberration.
Tangential and sagittal coma are not independent but the relationship
between them can only be determined empirically in the general case.
But a functional relation is available within the range of validity of the
Seidel approximation. The tangential unsymmetry is % (TKo - TKu) =
= 3 ·1.831 . LLW . S QK. This relation, which will not be proved here,
shows why the tangential coma curve represented in Fig. 46 is so sym-
metrical, for SVP and therefore SQK are very small. This is the reason
why the pupil distortion was corrected in the design of the objective.

§ 34. The Influence of Object and Pupil


Positions on Aberrations
The Effect of Object and Pupil Shifts on Aberrations and
the Impossibility of Perfect Imagery
Up till now we have always assumed that the object and the pupil
are in fixed pre-determined positions. It is important in the design and
use of optical instruments to know how the aberrations are changed by
object or pupil shifts. Explicit formulae for these effects only exist for
the Seidel approximation. This information is usually sufficient in
practice because high quality systems, whose correction is largely
determined by higher order aberrations, can normally only be used
under the conditions for which they were designed.
The Seidel approximations for the surface contributions of the five
aberrations of the object imagery are:
LI SPH = - (U' U)
915.5 A (nI)2 n' - -;; R.E.,

LI SV -
_ 500 (W' W)o
LLWB nln] --;:;;- --;; 100'

LI AST = U' - -;;U) R.E.,


- 1831 A (n])2 (n'

LI PTZ = 915.5 (LLW)2 (! (d- - ~) R.E. , (34.1)

LI K -_ LLW
500 (U' U)
A nI n] n' - -;; a100'
The Influence of Object and Pupil Positions on Aberrations 163

In these equations the subscript 0 denoting object imagery has been


omitted; furthermore an equals sign has been used for the sake of
simplicity although the formulae are only approximations.
For a pupil shift II, as defined by equations (7.4), it follows from
B* = B + nA and n*W* = nW + nnU that n*J* = B*n*e - n*W* =
= (Bne - nW) + n(Ane - nUl = n] + nnI. Inserting these relations
in (34.1) we obtain the effect of the pupil shift on the surface contribu-
tions. Clearly ,1 SPH* = ,1 SPH and ,1 PTZ* = ,1 PTZ since these aber-
rations are independent of the coordinates B, n Wand nj. In other words,
the spherical aberration of object imagery and the Petzval curvature are
independent of the position of the stop.
The only pupil coordinate present in the surface contribution ,1 K
for the coma number is nj. Therefore the pupil shift gives
AK* _ LLW
i..l -
500 A I ] (U' U) 0/
n n ;;; - -;- 00 + .n LLW
500 A (U' U) 0/
(n 1)2 ;;; - -;- 00'

= ,1 K - n 0.546 ,1 SPH .
LLW

The surface contributions to the coma number are thus a linear


function of the pupil shift and the changes are proportional to the surface
contribution to the spherical aberration. If all the surface contributions
are added together, it follows that the coma number of a system is
independent of the stop position when the system is free from spherical
aberration, i.e. when L ,1 SPH = O.
A similar treatment gives the astigmatism
,1 AST* = ,1 AST - n 7.324 LLW·L1 K + n 2 2,1 SPH .
Its surface contributions are proportional to the first and second
powers of the pupil shift parameter n. Apart from numerical factors, the
coefficients are the surface contributions to the coma number and spher-
ical aberration respectively. If both these aberrations are zero for a
given system, the stop position has no influence on the astigmatism. This
explains why it is not possible to improve the astigmatism of the achromat
described in § 29 by a shift of the stop.
The surface contributions to the distortion are expressed by a third
degree polynomial in the parameter n:
0.546
,1 SV* = ,1 SV + na 1 + n 2 a2 - n 3 LLW·L1 SPH.
The coefficients a1 and a2 of the linear and quadratic terms in this
equation are given by
(W' W) + Bn1n] (U';;;--;-U) +An1n] (W'
500 { B(n1)2 -:;;---;-
a 1 =LLW W)} ,
-:;;---:;;

a2 = 500 {A n1 n] (U'
LLW U) + B (n1)2 (U';;; - -;-U) + A (n1)2 (W'
;;; - -;- W)} ,
-:;;- - -:;;
11·
164 Theory of Aberrations

and cannot readily be interpreted as aberrations. The first term of a2 is


the Seidel approximation for the surface contribution to the coma number.
The third term can be transformed using the relation

A -(WI W)
n'
--
n
=B -
12.'
--
12.
(U' U)
+LLW -
12.'2
--
n2
, (1 1) (34.2)

an elementary consequence of the invariance of the linear conductivity.


The second term then appears twice and can be reduced to the form of
the first term, which now appears three times, by using the relation
BnI= A n] + LLW (34.3)
which follows from the definition of nI and n]. These transformations
leave the expression

LLW[(nI)2 (n~2 - ~2) + 2 nI (g - ~)]


for the content within the outer brackets in the equation above
for a2. Using n'1' = nI and 1= Ae - U, this expression reduces to
- LLW (U'2 - U2). Hence a2 = 3Ll K - 500 (U'2 - U2).
The first term of at is the Seidel approximation for the surface con-
tribution to the astigmatism of the pupil imagery, not the object imagery.
This can be seen from (34.1) if the roles of object and pupil are reversed.
A transformation using the relations (34.2) and (34.3) yields by a similar
reduction

LLW {23 [
at = - 0.546 - 1831 (W' W)] + 915.5 LLW2e (11-
B (nI)2 -:;;;- - -;; 1 -;;I)} +
+ 500·(U'W'- UW).
This formulation is unsatisfactory because the influence of the astigma-
tism of the pupil imagery on the object imagery is not clear. Within the
Seidel approximation, the relation between the two is very simple. It is

(W'n' W)12.
(U' U)
B(nI)2 - - - =A (n])2 - - - - LLW (U'W'- UW).
12.n
This follows from (34.2) and (34.3) by elementary reduction. If the
contributions of all the surfaces are added, the products UW cancel out
in pairs apart from the first term Ut Wt in front of the system and the last
term U~Wk behind it. There remains
AST P = AST 0 + 1831 LLW (UkW~ - UtW t ). (34.4)
The astigmatism of the object imagery and that of the pupil imagery
always differ by 1831 LLW (UkWk - UtW t ) R.E., an expression which
only depends on the form of the light tube in front of and behind the
system. It is immaterial how many lenses are contained in the system
and what form they have.
The Influence of Object and Pupil Positions on Aberrations 165

It follows from (34.4) that a pupil shift brings

Ll SV* =LlSV -:n;~~~(iLl AST +Ll PTZ) +:n;2 3Ll K- :n;3~~~Ll SPH
- lOOO:n; (U'W' - UW) - 500:n;2(U'2 - U2).
The effect of a pupil shift on the spherical aberration of the pupil is
expressed by a fourth degree polynomial in:n;. This will not be given here.
Instead we shall derive the effect of an object shift A * = A + wB,
n*U* = nU + wnW, n*I* = nI + wn] on the Seidel approximation for
the surface contribution
(U' U)
Ll SPHO = - 915.5 A (nI)2 ;;;- - -;- R.E.

to the spherical aberration of the object imagery. If the object shift


formulae are introduced, we obtain
Ll SPHO* = Ll SPHO + w ci + w 2 c2 + w3 c3 + w 4 Ll SPHP ,
in which the coefficients are:
{ (UI U) (U I U)
ci = - 915.5 2AnIn] ;;;- - -;- + B(nI)2 ;;;- --;- + A (nI)2 -;;;---;- (WI W)}
= - 1.831 LLW·4Ll K + 915.5 LLW (U'2 - U2) I

c2 = - 915.5 {A (n])2 (;;;-


I U) +
U' - -;-U) + 2 BnIn] (U;;;---;-
+ 2 AnIn] (7WI - -;- W) + B (nI)2 (WI
7--;- W)}
= 3Ll AST + 2Ll PTZ + 3·915.5 LLW (U'W' - UW)
= (I Ll AST 0 + Ll PTZ) + (i Ll AST P + Ll PTZ),
{ (UI U) (WI W)
c3 = - 915.5 B(n])2 ;;;- - -;- + 2 BnIn] 7 - -;- + A (n])2 -;;;---;- (WI W)}
= - 1.831 LLW·4 Ll SV - 915.5 LLW (W'2 - W2) .
These expressions are obtained from (34.2) and (34.3) after a similar
reduction process to that given above. The result for c2 also makes use
of (34.4) and renders clear the symmetry existing between the coefficients
for the object and pupil imagery respectively.
The sum of all the surface contributions Ll SPHO* of a system gives
SPHO* = SPHO - 4 w 1.831 LLW· K + 6 w 2 (t AST + t PTZ) -
- 4 (031.831 LLW 'SV + (04.SPHP + (34.5)
+ (0 915.5 LLW {(U1- Ui) + 3w (U~ W~ - U I WI) - (02 (W~ - ~)}.
An interesting law regarding the existence of perfect optical systems
emerges from (34.5). If a system is to form an aberration-free image of a
finite volume of space, then the system must at least be capable of forming
an aberration-free image of a plane section perpendicular to the axis
166 Theory of Aberrations

which cuts that volume of space. Suppose this section be taken as the
object plane, then the coefficients SPHO, K, AST, PTZ and SV in (34.S)
are all zero. If the new image corresponding to any object shift within
a finite interval measured from the basic object plane is to be free at
least of spherical aberration, then SPHP must be zero and also
U'; - U~ = 0, U~W~ - U1W 1 = 0 and W~ - m= 0 .
This implies that U;. = ± U j , WI. = ± Wj' Furthermore, because of the
second of the above three conditions, the same sign must be taken in both
cases: the combination U;. = Uj and WI. = - Wj is not acceptable.
Equations (g.!) then give the refracting power as

in other words, a finite volume of space can only be imaged aberration-


free within the Seidel approximation if the system has zero refracting
power.
Finite volumes of space can be imaged without aberration by simple
light tubes without lenses or by a plane mirror. The above law states that
any perfect system must have an equivalent effect on the light tube to
that of these trivial systems. Systems which can show us more than we
can see with the naked eye can never yield an aberration-free image of a
three-dimensional object.
This law regarding the non-existence of perfect systems is, however,
only depressing for the theoretician; for it also has the consequence that
an appropriate system must be developed for every new application. In a
sense, therefore, the law is the justification for the existence of optical
designers and the optical industry.

Conclusion and References to Literature


The reader who feels tempted, after reading this book, to try his hand
at optical design should be warned: the theory given here is not an
adequate preparation. The necessary formulae have been given but the
information concerning computation techniques is not adequate. Without
a suitable computation scheme; without rules indicating what values are
to be calculated in which formulae and in what order, and without
experience to guide his judgment of those values that are specially
important, the beginner has no chance of success. On the other hand,
a well-tried computing technique enables much to be done without
knowledge of a great deal of the formal theory. Optical design is a field
to which years of computing work must be devoted, otherwise it is a
waste of time to start. It is only when one has reached the stage at which
Conclusions and References to Literature 167

normal computing work becomes a routine rather than a nuisance that


one is in a position to understand the basic behaviour of optical systems.
Otherwise there is a serious danger of drowning in a sea of figures.
No attempt has been made in this book to present a technique for
optical computation, simply because no generally accepted scheme exists.
This is partly due to the fact that optical design does not form part of a
normal university physics course: optical designers learn their trade
through working in the optical industry. A further reason is that no single
computing technique can be suitable for the very different educational
levels and talents to be found among practising optical designers. One
designer may seek analytical solutions using aberration theory; another
may prefer purely numerical methods, while a third may concentrate on
graphical approaches.
Numerical calculations are greatly simplified and speeded up by
electronic computers. If many such calculations have to be made, a
computer is certainly justified economically. However, the enormous
computing potential of modern machines is apt to lead to unnecessary
computation, neglect of aberration theory and trial and error methods.
In view of these dangers, only time will reveal to what extent the use of
modern computers brings economic returns.
Finally, a few references to literature may be helpful for the reader
who wishes to delve deeper into the problems of geometrical optics. It is
accepted as good practice in scientific circles to make the list long and
give no comments on the merits of the work. An admirable book from
this point of view is CZAPSKI-EpPENSTEIN: "Grundziige der Theorie der
optischen Instrumente", third edition, Leipzig 1924. The references to
literature cover 83 pages. It is an invaluable source of information on the
history of the subject which may be read with great profit by the reader
interested in the older sources of original work. In fact, it is rarely
necessary in geometrical optics to consult newer works, unless one wishes
to compare different treatments of the same matter. CZAPSKI-EpPENSTEIN
is very detailed and must be considered today as difficult to read.
Nevertheless, it represents the golden age of the development of geo-
metrical optics and is an indispensable aid to many specialists in the field.
In 1929 the first part of "Applied Optics and Optical Design" by
A. E. CONRADY appeared. This is available as a Dover book (Dover
Publications, New York, 1957). Part two has also been published, post-
humously, by Dover (1960). This work gives a detailed and elementary
introduction to the theory and is illustrated by many practical examples.
CONRADY makes more extensive use of geometrical interpretations and
relations than I have done in this book: his treatment is therefore partic-
ularly easy to follow. In 1930 the well-known book by M. BEREK
"Grundlagen der praktischen Optik" (Verlag W. de Gruyter, Berlin und
168 Theory of Aberrations

Leipzig) appeared. This book was specifically written with the optical
designer in mind. BEREK and CONRADY were both active not only in the
theory of geometrical optics but also as practical optical designers. This
explains why their books are still in wide use after several decades.
Those who work in geometrical optics should not neglect the physics
associated with the subject. In my opinion, the best book in this respect
is R. W. PORL: "Optik und Atomphysik", ninth edition, Springer 1954.
Astigmatism is explained in a debatable manner, but this is the exception
which proves the rule that one can rely on PORL.
A better understanding of the finer points of image quality may be
gained if the wave nature of light is taken into account in considering the
aberrations of geometrical optics. An introduction is given by H. H.
HOPKINS: "Wave Theory of Aberrations", Oxford, 1950. A useful supple-
ment is given in a paper by J. L. RAYCES: "Exact relations between wave
aberration and ray aberration", Optica Acta, Vol. 11, 1964. Wave aber-
rations lend themselves to simple physical interpretation and are there-
fore useful for assessing image quality. But they are more difficult to
compute than geometrical optical aberrations and there is no existing
formulation for the surface contributions to the total aberration. For this
reason, it is the geometrical aberration theory which is used in practice
for the computation and design of optical systems. It allows an exact
analysis of the system and usually gives the designer adequate informa-
tion regarding the image quality. The wave aberrations contain the
entire information concerning image quality but only give a rough indi-
cation what changes in the system are necessary to improve its perform-
ance.
Author- and Subject Index
Abbe 9,68, 121, 141 dispersion 79, 96
Abbe number 82 distortion 113, 119, 156
aberration 9,75,85, 113, 156, 162 Drude 79
accommodation 25
achromat 96
Emitter 1, 6, 23
anti-reflexion films 31, 107
energy flow 5,22
aplanatic 125
entrance pupil 28, 34, 42
aspheric surfaces 114
eye 12,21
AST 148
eyepiece 28, 64
astigmatism 143, 148, 152
exit pupil 28, 34, 42
axis 3
explicit calculation 45

Berek 68
Field curvature 143, 151
binocular 19,63, 69, 122
brightness 22 - lens 15, 28, 64
- of view 24
flint glass 82
Calculation of light tubes 39 fluorite 103
centred system 3, 24 focal length 19,26,42
CHLO (- CHL) 85 fundamental equations 140
CHLP 85
chromatic aberrations 80, 85
CHVO (- CHV) 85 Gauss 47, 120
CHVP 85 - error 115, 135
Clausius 120, 141 - , theorem of 19
coated lenses 107 Gaussian brackets 45,47,49
coma 156 geometrical construction of a refraction
- number 161 53
conductivity diagram 52, 56 - optics 1,75
Conrady 167 glass 77,96
coordinates for light tubes 3,6, 9 grain of silver bromide 33
- - plane rays 105 graphical representation of aberrations
- - skew rays 137 132
crown glass 82 - - - perfect imagery 52
crystal optics 10
Hartmann formula 80
Decentred imagery 3 Helmholtz 21, 121, 141
Delano 56 Herzberger 143
depth of accommodation 66, 70 Hopkins 168
- - focus 66, 68
- - imagery 66, 70 lilumination 89
differential formulae 47, 49 image plane 42
diffraction 7, 9 - quality 76
170 Autbor- and Subject Index

imagery in depth 66 pupil shift (II) 11, 15, 62


immersion 31 - - , effect on aberrations 94, 162
intensity of irradiation 35
invariance of LLW 9,72
iterative computing scheme 39 Radiating power 5, 6
radiation 1, 13, 22, 143
- density 5,22
K= coma number 161 ray 104
- tracing formulae 104, 110, 143
Rayces 168
Lambert 5, 143, 146
Rayleigh 9
Lambert radiators 13
Rayleigh unit (R.E.) 8
law of refraction 104, 105
receiver 1, 6, 12, 23, 33
lens formula 14
recursive calculation 39
light 1
reference point 4, 11, 107
- conductivity 6
refractive index 5, 10, 43, 78, 81
- direction 3
refraction (<1» 11, 14, 17, 108
- tube 2,8, 10
- , geometrical construction 53
linear conductivity 5, 6, 8, 23
-,lawof 104
LLW (defined) 6,8
refracting power 19, 43, 112
resolving power 10, 24, 27, 31, 77
Magnification ratio (f3) 29
magnifying glass 25, 62
- power (N) 26 SAG = sagittal field curvature 151
marginal ray 4,42, 105, 113 sagittal section 105
mean dispersion 81 secondary spectrum 98
meniscus 154 - aberrations 129
microscope 8, 27, 65 Seidel formulae 123, 129, 152
sign convention 4
sine condition 9, 141
Normal glasses 83, 100 skew rays 137
numerical aperture (n U) 29, 32 SLK 160
Smakula 107
Object shift (Q) 11, 13, 15, 61 spectacles 25
- - , effect on aberrations 94, 162 speed of photographic objective 36
opera glass 19, 29, 64 spherical aberration 113, 124, 156
optical designer 45, 76, 166 SPHO (= SPH) 113
SPHP 113
splitting rule 47
Perfect imagery, definition 9 SQK 160
- - , non-existence of 162 SSP 100
Petzval 153 stilb 22
Petzval curvature 152 stop number 36
photographic emulsion 32 SVO (= SV) 113
- objective 34,41,48,66 SVP 113
pinhole camera 38 surface contribution 92, 123, 152
plane rays 104 a-values for glasses 83, 100
Poh1 168
polarization 10, 106
principal points 19,43,66 Tangent distortion 120
- theorem of perfect imagery 18 tangential coma 157
- ray 4, 13, 105, 113 - section 105
PTZ 153 telephoto objective 48
Author- and Subject Index 171

telescope 27 Unvala 153


- magnification (F) 28
telescopic system 19 Viewfinder 38
thickness contribution 92, 93, 123
transfer (Ll) 11, 14, 107 Watt 5
transmission factor 106 wavelength 8, 22, 79
triplet 41
TVa 120 Zonal aberrations 129
Applied Physics and Engineering Series
1. STUPOCHENKOjLoSEV/OSIPOV: Relaxation in Shock
Waves . . . . . . . . . . . . . . . . . . $ 19.80jDM 72,-
2. Re-Entry and Planetary Entry: Physics and Tech-
nology. Part I: Dynamics, Physics, Radiation,
HeatTransferandAblation.EditedbyW.H.T.LoR $ 19.80jDM 79,20
3. Re-Entry and Planetary Entry: Physics and Tech-
nology. Part II: Advanced Concepts, Experiments,
Guidance-Control and Technology. Edited by
W. H. T. LOR. . . . . . . . . . . . . . . $ 12.40/DM 49,60
4. V AKMAN: Sophisticated Signals and the Uncer-
tainty Principle in Radar. Edited by E. JACOBS. $ 14.80/DM 59,20
5. RApOPORT: Dynamics of Elastic Containers Parti-
ally Filled with Liquid. . . . . . . . . . . . $ 14.80jDM 59,20
6. MorSEYEV/RuMYANTSEV: Dynamic Stability of
Bodies Containing Fluid . . . . . . . . . . . $ 16.60jDM 66,40
7. Jet, Rocket, Nuclear, Ion and Electric Propulsion,
Theory and Design. Edited by W. H. T. LOH . $ 29.80/DM 99,20
8. HOYAUX: Arc Physics . . . $ 14.50/DM 58,-
9. ZIMMER: Geometrical Optics $ 9.40jDM 34,-

You might also like