Introduction to Optics 3rd
Introduction to Optics 3rd
“They could but make the best of it and went around with woebegone faces, sadly com-
plaining that on Mondays, Wednesdays, and Fridays, they must look on light as a wave; on
Tuesdays, Thursdays, and Saturdays, as a particle. On Sundays, they simply prayed.”
The Strange Story of the Quantum
Banesh Hoffmann, 1947
INTRODUCTION
The words cited above—taken from a 1947 popular primer on the quantum
world—delighted many readers who were just then coming into contact with
ideas related to the nature of light and quanta. Hoffmann’s amusing and in-
formative account—involving in part the wave-particle twins “tweedledum”
and “tweedledee”—captured nicely the level of frustration felt in those days
about the true nature of light. And today, some 60 years later, the puzzle of
tweedledum and tweedledee lingers. What is light? What is a photon? Indeed,
in October of 2003, The Optical Society of America devoted a special issue
of Optics and Photonic News to the topic “The Nature of Light: What is a
Photon?” In this issue,1 a number of renowned scientists, through five pen-
etrating essays, accepted the challenge of describing the photon. Said Arthur
Zajonc, in his lead article titled “Light Reconsidered”:
Light is an obvious feature of everyday life, and yet light’s true nature has eluded us
for centuries. Near the end of his life, Albert Einstein wrote, “All the 50 years of con-
scious brooding have brought me no closer to the answer to the question: What are
light quanta?” We are today in the same state of “learned ignorance” with respect to
light as was Einstein.
1
“The Nature of Light: What is a Photon?” OPN Trends, Vol 3., No. 1, October 2003.
1
2 Chapter 1 Nature of Light
1 A BRIEF HISTORY2
2
A more in-depth historical account may be found, for example, in Vasco Ronchi, The Nature
of Light (Cambridge: Harvard University Press, 1970).
Nature of Light 3
by the experimental fact that when two beams of light intersected, they
emerged unmodified, just as in the case of two water or sound waves. Adopt-
ing a wave theory, Huygens was able to derive the laws of reflection and re-
fraction and to explain double refraction in calcite as well.
Within two years of the centenary of the publication of Newton’s
Optics, the Englishman Thomas Young performed a decisive experiment that
seemed to demand a wave interpretation, turning the tide of support to the
wave theory of light. It was the double-slit experiment, in which an opaque
screen with two small, closely spaced openings was illuminated by monochro-
matic light from a small source. The “shadows” observed formed a complex
interference pattern like those produced with water waves.
Victories for the wave theory continued up to the twentieth century. In
the mood of scientific confidence that characterized the latter part of the
nineteenth century, there was little doubt that light, like most other classical
areas of physics, was well understood. In 1821, Augustin Fresnel published re-
sults of his experiments and analysis, which required that light be a transverse
wave. On this basis, double refraction in calcite could be understood as a phe-
nomenon involving polarized light. It had been assumed that light waves in an
ether were necessarily longitudinal, like sound waves in a fluid, which cannot
support transverse vibrations. For each of the two components of polarized
light, Fresnel developed the Fresnel equations, which give the amplitude of
light reflected and transmitted at a plane interface separating two optical
media.
Working in the field of electricity and magnetism, James Clerk
Maxwell synthesized known principles in his set of four Maxwell equations.
The equations yielded a prediction for the speed of an electromagnetic wave
in the ether that turned out to be the measured speed of light, suggesting its
electromagnetic character. From then on, light was viewed as a particular
region of the electromagnetic spectrum of radiation. The experiment (1887)
of Albert Michelson and Edward Morley, which attempted to detect opti-
cally the earth’s motion through the ether, and the special theory of relativ-
ity (1905) of Albert Einstein were of monumental importance. Together
they led inevitably to the conclusion that the assumption of an ether was su-
perfluous. The problems associated with transverse vibrations of a wave in a
fluid thus vanished.
If the nineteenth century served to place the wave theory of light on a
firm foundation, that foundation was to crumble as the century came to an end.
The wave-particle controversy was resumed with vigor. Again, we mention
only briefly some of the key events along the way. Difficulties in the wave the-
ory seemed to show up in situations that involved the interaction of light with
matter. In 1900, at the very dawn of the twentieth century, Max Planck an-
nounced at a meeting of the German Physical Society that he was able to de-
rive the correct blackbody radiation spectrum only by making the curious
assumption that atoms emitted light in discrete energy chunks rather than in a
continuous manner. Thus quanta and quantum mechanics were born. Accord-
ing to Planck, the energy E of a quantum of electromagnetic radiation is pro-
portional to the frequency n of the radiation:
E = hn (1)
where the constant of proportionality h, Planck’s constant, has the very small
value of 6.63 * 10-34 J-s. Five years later, in the same year that he published
his theory of special relativity, Albert Einstein offered an explanation of the
photoelectric effect, the emission of electrons from a metal surface when irradi-
ated with light. Central to his explanation was the conception of light as a stream
of light quanta whose energy is related to frequency by Planck’s equation (1).
Then in 1913, the Danish physicist Niels Bohr once more incorporated the
4 Chapter 1 Nature of Light
h
l = (2)
p
Photons and electrons that behaved both as particles and as waves seemed at
first an impossible contradiction, since particles and waves are very different en-
tities indeed. Gradually it became clear, to a large extent through the reflections
of Niels Bohr and especially in his principle of complementarity, that photons
and electrons were neither waves nor particles, but something more complex
than either.
In attempting to explain physical phenomena, it is natural that we ap-
peal to well-known physical models like waves and particles. As it turns out,
however, the complete nature of a photon or an electron is not exhausted by
either model. In certain situations, wavelike attributes may predominate; in
other situations, particle-like attributes stand out. We know of no simpler
physical model that is adequate to handle all cases.
Quantum mechanics describes both light and matter and, together with
special relativity, predicts that the momentum, p, wavelength, l, and speed, y,
for both material particles and photons are given by the same general equations:
2E2 - m2c4
p = (3)
c
h hc
l = = (4)
p 2E - m2c4
2
pc2 m2c4
y = = c 1 - (5)
E B E2
Nature of Light 5
In these equations, m is the rest mass and E is the total energy, the sum of the
rest-mass energy mc2 and kinetic energy EK , that is, the work done to acceler-
ate the particle from rest to its measured speed. The proper expression for ki-
netic energy is no longer simply EK = 12 my2, but rather is EK = mc21g - 12,
where g = 1> 21 - 1y2>c22. This relativistic expression3 for kinetic energy EK
approaches 12 my2 for y V c.
A crucial difference between particles like electrons and neutrons and
particles like photons is that the latter have zero rest mass. Equations (3) to
(5) then take the simpler forms for photons:
E
p = (6)
c
h hc
l = = (7)
p E
pc2
y = = c (8)
E
Thus, while nonzero rest-mass particles like electrons have a limiting speed of
c, Eq. (8) shows that zero rest-mass particles like photons must travel with the
constant speed c. The energy of a photon is not a function of its speed but
rather of its frequency, as expressed in Eq. (1) or in Eqs. (6) and (7), taken to-
gether. Notice that for a photon, because of its zero rest mass, there is no dis-
tinction between its total energy and its kinetic energy. The following
example helps clarify the differences in the momentum, wavelength, and
speed of electrons and photons of the same total energy.
Example 1
An electron is accelerated to a kinetic energy EK of 2.5 MeV. (a) Determine
its relativistic momentum, de Broglie wavelength, and speed. (b) Determine
the same properties for a photon having the same total energy as the electron.
Solution
The electron’s total energy E must be the sum of its rest mass energy mc2
and its kinetic energy EK . The rest mass energy is
mc2 = 19.11 * 10-31 kg213 * 108 m>s22 = 8.19 * 10-14 J.
Since 1 eV = 1.6 * 10-19 J, we have mc2 = 5.11 * 105 eV = 0.511 MeV.
Thus,
or
The other quantities are then calculated in order. Working with SI units we ob-
tain, from Eq. (3):
2E2 - 1mc22 214.82 * 10-13 J22 - 18.19 * 10-14 J22
p = =
c 3 * 108 m>s
-21 #
= 1.58 * 10 kg m>s
3
This discussion is not meant to be a condensed tutorial on relativistic mechanics, but, with the
help of Eqs. (3) to (8), a summary of some basic relations that unify particles of matter and light.
6 Chapter 1 Nature of Light
h 6.626 * 10-34 J # s
l = = = 4.19 * 10-13 m = 0.419 pm
p 1.58 * 10-21 kg # m>s
and from Eq. (5):
pc2 11.58 * 10-21 kg # m>s213 * 108 m>s22
y = =
E 4.82 * 10-13 J
4.82 * 10-13 J
= 1.61 * 10-21 kg # m>s
E
p = =
c 3 * 108 m>s
h 6.626 * 10-34 J # s
l = = = 0.412 pm
p 1.61 * 10-21 kg # m>s
(m) (Hz)
10!16 3 " 1024
GAMMA
10!14 3 " 1022
RAYS
X-RAYS
1Å 10!10 3 " 1018
380 nm
1 nm V
(Vacuum)
10!8 3 " 1016 B
UV
(Near) G
Visible
1 mm 10!6 3 " 1014 spectrum
(Near) Y
OPTICAL
IR O
10!4 3 " 1012
R
(Far) 770 nm
c = ln (9)
Ultraviolet
On the short-wavelength side of visible light, this electromagnetic region
spans wavelengths ranging from 380 nm down to 10 nm. Ultraviolet light is
sometimes subdivided into three categories: UV-A refers to the wavelength
range 380–315 nm, UV-B to the range 315–280 nm and UV-C to the range
280–10 nm. The sun emits significant amounts of electromagnetic radiation in
all three UV bands but, due to absorption in the ozone layer of the earth’s at-
mosphere, roughly 99% of the UV radiation that reaches the earth’s surface
is in the UV-A band. UV radiation from the sun is linked to a variety of
health risks. UV-A radiation, generally regarded as the least harmful of the
three UV bands, can contribute to skin aging and is possibly linked to some
forms of skin cancer. UV-A radiation does not contribute to sunburns. UV-B
radiation has been linked to a variety of skin cancers and contributes to the
sunburning process. The link between UV-B radiation and skin cancer is a
primary reason for the concern related to ozone depletion, which is believed
to be in part caused by human use of so-called chlorofluorocarbon (CFC)
compounds. Ozone 1O32 is formed when UV-C radiation reacts with oxygen
in the stratosphere and, as mentioned, plays an important role in the filtering
of UV-B and UV-C from the electromagnetic radiation that reaches the
earth’s surface. CFC compounds can participate in chemical processes that lead
to the conversion of ozone into “ordinary” oxygen 1O22. The concern that CFCs
and other similar chemicals may contribute to the depletion of the ozone layer
and thus increase the risk of skin cancer led to protocols calling for the reduc-
tion of the use of refrigerants, aerosol sprays, and other products that release
these chemicals into the atmosphere. Sunblock and sunscreen lotions are in-
tended in part to block harmful UV-B radiation. On the other hand, UV radia-
tion has the beneficial effect of inducing vitamin D production in the skin.
X-rays
X-rays are EM waves with wavelengths in the 10 nm to 10-4 nm range. These
can be produced when high-energy electrons strike a metal target and are
Nature of Light 9
Gamma Rays
This type of EM radiation has its origin in nuclear radioactive decay and cer-
tain other nuclear reactions. Gamma rays have very short wavelengths in the
range from 0.1 nm to 10-14 nm. Like X-rays, penetrating gamma rays find use
in the medical area, often in the treatment of localized cancers.
Infrared Radiation
On the long-wavelength side of the visible spectrum, infrared (IR) radiation
has wavelengths spanning the region from 770 nm to 1 mm. Objects in thermal
equilibrium at terrestrial temperatures emit radiation that has its energy out-
put peak in the IR range. Consequently, infrared radiation is sometimes
termed “heat radiation” and finds application in nightvision scopes that detect
the IR emitted from objects in absolute “darkness” and in infrared photogra-
phy wherein objects at different temperatures (and so with different peak
wavelengths of emitted radiation) are imaged as areas of contrasting bright-
ness. Images such as these can be used to map the temperature variation
across the surface of the earth, for example. Infrared radiation is used as a
treatment for sore muscles and joints, and, more recently, lasers that emit IR
radiation have been used to treat the eye for vision abnormalities. Infrared ra-
diation is also used in optical fiber communication systems and in a variety of
remote control devices.
Microwaves
Beyond infrared radiation we find microwaves, with wavelengths from 1 mm
to 30 cm or so. Microwave ovens, which have become a common kitchen ap-
pliance, use microwaves to heat food. In addition, microwaves play an important
role in radar systems both on the ground and in the air, in telecommunications,
and in spectroscopy.
Radio Waves
Radio waves are long-wavelength EM radiations produced, for example,
by electrons oscillating in conductors that form antennas of various shapes.
Radio waves have wavelengths ranging from meters to thousands of me-
ters. Used commonly in radio and television broadcasts, they include the
AM radio band (540–1600 kHz) with wavelengths ranging from 188 to 556 m
as well as the FM radio band (88–108 MHz) with wavelengths from 2.78 to
3.41 m.
We have indicated that EM waves may lose and gain energy only in dis-
crete amounts that are multiples of the energy associated with the energy
quanta that have come to be called photons. Equation (1) gives the energy of
a photon as hn. When EM wave energy is detected, the detector can record
only energies that are multiples of a photon’s energy. As the following exam-
ple indicates, for macroscopic light sources, the energy of a photon is typical-
ly far less than the total detected energy, and so, in such a case, the restriction
that the detected energy must be only a multiple of a photon’s energy goes
unnoticed. Since the energy per photon decreases with increased wavelength,
for a given total energy, the energy “graininess” is less for long-wavelength
radiation than for short-wavelength radiation. To understand the interaction
of light with individual atoms and molecules, it is important to keep in mind
that EM waves gain and lose energy in discrete amounts proportional to the
frequency of the radiation. Consider the following example.
10 Chapter 1 Nature of Light
Example 2
A certain sensitive radar receiver detects an electromagnetic signal of fre-
quency 100 MHz and power (energy/time) 6.63 * 10-16 J>s.
Solution
3 * 108 m>s
a. l = c>n = = 3m
100 * 106 Hz
1 eV
E = 6.63 * 10-26 J a b = 4.14 * 10-7 eV
1.6 * 10-19 J
So each photon contributes but one part in 10 billion of the total power
in the radar wave even for this very weak signal. In such a case, the
“graininess” of the power in the signal is likely to go undetected.
In this example we have introduced the notion of a detector and the en-
ergy and power carried by an electromagnetic wave. The energy carried by an
EM wave can be specified in many related ways: the power, power per unit
area, and power per unit solid angle, for example. To quantify these charac-
teristics of EM waves, we turn to the topic of radiometry.
4 RADIOMETRY
d£
Ie = (10)
dv
r dA
Central ray
Figure 2 The radiant intensity is the flux
S through the cross section dA per unit of
dv solid angle. Here the solid angle dv = dA>r2.
4
The introduction “in the abstract” of so many new units, some rarely used and others mis-
used, is not very palatable pedagogically. Table 1 is meant to serve as a convenient summary that can
be referred to when needed.
12 Chapter 1 Nature of Light
d£ e £e 4pIe Ie
Ee = = = 2
= 2 , point source (11)
dA A 4pr r
dIe d2 £ e
Le = = (12)
dA cos u dv1dA cos u2
A
A
A
A
A A
A
r A
A
A
A
A A
A
Figure 3 Illustration of the inverse-square
law. The flux leaving a point source within any
solid angle is distributed over increasingly r
larger areas, producing an irradiance that 2r
decreases inversely with the square of the
3r
distance.
Direction
of viewing
Radiating
surface A
Ap
u r Normal
u
Projected
Figure 4 Radiant flux collected along a surface A cos u
direction making an angle u with the normal
to the radiating surface. The projected area of
the surface 1A cos u2 is shown by the dashed
rectangle.
Nature of Light 13
intensity I(0) is observed. As the aperture is moved along the circle of radius
r, thereby increasing the angle u, the cross section of radiation presented by
the surface decreases in such a way that
Thus, when a radiating (or reflecting) surface has a radiance that is indepen-
dent of the viewing angle, the surface is said to be perfectly diffuse, or a
Lambertian surface.
We show next that the radiance has the same value at any point along a ray
propagating in a uniform, nonabsorbing medium. Figure 5 pictures a narrow
beam of radiation in such a medium, including a central ray and a small bun-
dle of surrounding rays (not shown) that pass through the elemental areas
dA1 and dA2 situated at different points along the beam. The central ray
makes angles of u1 and u2 , respectively, relative to the area normals, as shown.
The solid angle dv1 = dA2 cos u2>r2, where dA2 cos u2 represents the projec-
tion of area dA2 normal to the central ray. According to Eq. (12), the radiance
L1 at dA1 is given by
d2 £ 1 d2 £ 1
L1 = = (15)
dv11dA1 cos u12 1dA2 cos u2>r221dA1 cos u12
By a similar argument, in which we reverse the roles of dA1 and dA2 in the
figure,
d2 £ 2 d2 £ 2
L2 = = (16)
dv21dA2 cos u22 1dA1 cos u1>r221dA2 cos u22
For a nonabsorbing medium, the power associated with the radiation passing
through the continuous bundle of rays remains constant, that is, d£ 1 = d£ 2 ,
so that we can conclude from Eqs. (15) and (16) that L1 = L2 . It follows that
the radiance of the beam is also the radiance of the source, at the initial point
of the beam, or L1 = L2 = L0 .
Suppose, referring to Figure 6, that we wish to know the quantity of
radiant power reaching an element of area dA2 on surface S2 due to the
source element dA1 on surface S1 . The line joining the elemental areas, of
l
ma
N
r
or
No
m
al
u1 u2
dv1 Central
dA1 ray
dA2 Figure 5 Geometry used to show the
r invariance of the radiance in a uniform,
lossless medium.
14 Chapter 1 Nature of Light
u1
dA1 r12
Figure 6 General case of the illumination dA2
of one surface by another radiating surface.
Each elemental radiating area dA1 con- u2
tributes to each elemental irradiated area S1
S2
dA2 .
length r12 , makes angles of u1 and u2 with the respective normals to the surfaces,
as shown. The radiant power is d2 £ 12 , a second-order differential because both
the source and receptor are elemental areas. By Eq. (15) or Eq. (16),
and the total radiant power at the entire second surface due to the entire first
surface is, by integration,
Example
Consider a £ e = 5 milliwatt Helium-Neon laser emitting a “pencil-like”
beam with a divergence angle a of 1.3 milliradians. The laser cavity is de-
signed so that the beam emerges from a surface area ¢AS = 2.5 * 10-3 cm2
at the output mirror. See the sketch shown in Figure 7.
a. Determine the solid angle ¢v in terms of R and a.
W
b. Determine the radiance of this laser light source in units of .
cm2 # sr
Solution
a. The solid angle ¢v is equal to ¢AT>R2 , where
¢A T = pr2T = p1R tan1a>2222 M pR21a>222,
since tan1a>22 M a>2 for small angles.
1pR2a22>4 pa2
Thus, ¢v = = , independent of the value of R. That is
R2 4
pa2 p10.001322
¢v = = sr = 1.33 * 10-6 sr
4 4
Mirror Output #v
mirror
rT
a
Area #AT
Radiating
area # AS R
Figure 7 Example 3.
Nature of Light 15
£e 5 * 10-3 W
¢Le = =
¢A S ¢v 12.5 * 10-3 cm2211.33 * 10-6 sr2
W
= 1.5 * 106
cm2 # sr
PROBLEMS
1 Calculate the de Broglie wavelength of (a) a golf ball of photons striking the surfaces per second for the two beams
mass 50 g moving at 20 m/s and (b) an electron with kinetic are in the same ratio as their wavelengths.
energy of 10 eV.
12 Calculate the band of frequencies of electromagnetic ra-
2 The threshold of sensitivity of the human eye is about 100 diation capable of producing a visual sensation in the nor-
photons per second. The eye is most sensitive at a wave- mal eye.
length of around 550 nm. For this wavelength, determine the
threshold in watts of power. 13 What is the length of a half-wave dipole antenna designed
to broadcast FM radio waves at 100 MHz?
3 What is the energy, in electron volts, of light photons at the
ends of the visible spectrum, that is, at wavelengths of 380 14 A so-called “rabbit-ears” TV antenna is made of a pair of
and 770 nm? adjustable rods that can spread apart at different angles. If
the rods are each adjusted to a quarter wavelength for a TV
4 Determine the wavelength and momentum of a photon
channel that has a middle frequency of 90 MHz, how long
whose energy equals the rest-mass energy of an electron.
are the rods?
5 Show that the rest-mass energy of an electron is 0.511
MeV. 15 A soprano’s voice is sent by radio waves to a listener in a
city 90 km away.
6 Show that the relativistic momentum of an electron, accel-
a. How long does it take for the soprano’s voice to reach
erated through a potential difference of 1 million volts, can
the listener?
be conveniently expressed as 1.422 MeV/c, where c is the
b. In the same time interval, how far from the soprano has
speed of light.
the sound wave in the auditorium traveled? Take the
7 Show that the wavelength of a photon, measured in speed of sound to be 340 m/s.
angstroms, can be found from its energy, measured in elec-
tron volts, by the convenient relation 16 A small, monochromatic light source, radiating at 500 nm, is
rated at 500 W.
12,400
l1Å2 = a. If the source radiates uniformly in all directions, deter-
E1eV2
mine its radiant intensity.
8 Show that the relativistic kinetic energy, b. If the surface area of the source is 5 cm2, determine the
radiant excitance.
EK = mc21g - 12 c. What is the irradiance on a screen situated 2 m from the
source, with its surface normal to the radiant flux?
reduces to the classical expression 12 my2, when y V c. d. If the receiving screen contains a hole with diameter 5 cm,
how much radiant flux gets through?
9 A proton is accelerated to a kinetic energy of 2 billion elec-
tron volts (2 GeV). Find (a) its momentum, (b) its de
17 A 1.5-mW helium-neon laser beam delivers a spot of light
Broglie wavelength, and (c) the wavelength of a photon with
5 mm in diameter across a room 15 m wide. The beam radi-
the same total energy.
ates from a small circular area of diameter 0.5 mm at the
10 Solar radiation is incident at the earth’s surface at an aver- output mirror of the laser. Assume that the beam irradiance
age of 1000 W>m2 on a surface normal to the rays. For a is constant across the diverging beam.
mean wavelength of 550 nm, calculate the number of pho-
a. What is the beam divergence angle of this laser?
tons falling on 1 cm2 of the surface each second.
b. Into what solid angle is the laser sending its beam?
11 Two parallel beams of electromagnetic radiation with differ- c. What is the irradiance at the spot on the wall 15 m from
ent wavelengths deliver the same power to equivalent sur- the laser?
face areas normal to the beams. Show that the numbers of d. What is the radiance of the laser?
ni ui ur
nt
ut
2 Geometrical Optics
INTRODUCTION
16
Geometrical Optics 17
Law of Reflection
When a ray of light is reflected at an interface dividing two optical media, the
reflected ray remains within the plane of incidence, and the angle of reflection
ur equals the angle of incidence ui . The plane of incidence is the plane contain-
ing the incident ray and the surface normal at the point of incidence.
1 HUYGENS’ PRINCIPLE
Surfa
c
norm e
al
Plane
incid of
ence
Plane contains surface
normal, incident,
In reflected, and
ci
ra den d refracted rays.
Incident y t cte
e fle y
medium ui R ra
ni ur
Refr Interface
a
medi cting
um n
t
ut
by the particles of the ether, an elastic medium filling all space. Consistent
with his conception, Huygens imagined each point of a propagating distur-
bance as capable of originating new pulses that contributed to the distur-
bance an instant later. To show how his model of light propagation implied
the laws of geometrical optics, he enunciated a fruitful principle that can be
stated as follows: Each point on the leading surface of a wave disturbance—
the wavefront—may be regarded as a secondary source of spherical waves
(or wavelets), which themselves progress with the speed of light in the medi-
um and whose envelope at a later time constitutes the new wavefront. Simple
applications of the principle are shown in Figure 2 for a plane and spherical
wave. In each case, AB forms the initial wave disturbance or wavefront, and
A¿B¿ is the new wavefront at a time t later. The radius of each wavelet is, ac-
cordingly, yt, where y is the speed of light in the medium. Notice that the new
wavefront is tangent to each wavelet at a single point. According to Huygens,
the remainder of each wavelet is to be disregarded in the application of the
principle. Indeed, were the remainder of the wavelet considered to be effec-
tive in propagating the light disturbance, Huygens could not have derived the
law of rectilinear propagation from his principle. To see this more clearly,
refer to Figure 3, which shows a spherical wave disturbance originating at O
and incident upon an aperture with an opening SS¿. According to the notion
A"
A A"
A
vt
vt
B B" B
A
P
S
S!
P!
Figure 3 Huygens’ construction for an ob- B
structed wavefront.
Geometrical Optics 19
of rectilinear propagation, the lines OA and OB form the sharp edges of the
shadow to the right of the aperture. Some of the wavelets that originate from
points of the wavefront (arc SS¿ ), however, overlap into the region of shadow.
According to Huygens, however, these are ignored and the new wavefront
ends abruptly at points P and P¿, precisely where the extreme wavelets origi-
nating at points S and S¿ are tangent to the new wavefront. In so disregarding
the effectiveness of the overlapping wavelets, Huygens avoided the possibili-
ty of diffraction of the light into the region of geometric shadow. Huygens
also ignored the wavefront formed by the back half of the wavelets, since
these wavefronts implied a light disturbance traveling in the opposite direc-
tion. Despite weaknesses in this model, remedied later by Fresnel and others,
Huygens was able to apply his principle to prove the laws of both reflection
and refraction, as we show in what follows.
Figure 4a illustrates the Huygens construction for a narrow, parallel
beam of light to prove the law of reflection. Huygens’ principle must be mod-
ified slightly to accommodate the case in which a wavefront, such as AC, en-
counters a plane interface, such as XY, at an angle. Here the angle of
incidence of the rays AD, BE, and CF relative to the perpendicular PD is ui .
Since points along the plane wavefront do not arrive at the interface simulta-
neously, allowance is made for these differences in constructing the wavelets
that determine the reflected wavefront. If the interface XY were not present,
the Huygens construction would produce the wavefront GI at the instant ray
CF reached the interface at I. The intrusion of the reflecting surface, howev-
er, means that during the same time interval required for ray CF to progress
from F to I, ray BE has progressed from E to J and then a distance equivalent
to JH after reflection. Thus, a wavelet of radius JN = JH centered at J is
drawn above the reflecting surface. Similarly, a wavelet of radius DG is drawn
centered at D to represent the propagation after reflection of the lower part of
the beam. The new wavefront, which must now be tangent to these wavelets at
points M and N, and include the point I, is shown as KI in the figure. A repre-
sentative reflected ray is DL, shown perpendicular to the reflected wavefront.
The normal PD drawn for this ray is used to define angles of incidence and re-
flection for the beam. The construction makes clear the equivalence between
the angles of incidence and reflection, as outlined in Figure 4a.
Similarly, in Figure 4b, a Huygens construction is shown that illustrates
the law of refraction. Here we must take into account a different speed of
light in the upper and lower media. If the speed of light in vacuum is c, we
express the speed in the upper medium by the ratio c>ni , where ni is a con-
stant that characterizes the medium and is referred to as the refractive index.
Similarly, the speed of light in the lower medium is c>nt . The points D, E, and
F on the incident wavefront arrive at points D, J, and I of the plane interface
XY at different times. In the absence of the refracting surface, the wavefront
GI is formed at the instant ray CF reaches I. During the progress of ray CF
from F to I in time t, however, the ray AD has entered the lower medium,
where its speed is, let us say, slower. Thus, if the distance DG is yit, a wavelet
of radius ytt is constructed with center at D. The radius DM can also be ex-
pressed as
DG ni
DM = ytt = yt a b = a bDG
yi nt
P
K ADX ! IDG
# DIG ! # DIM
IDM ! IDG
ui ! ur
ur L
C
ui M
B
F
N
A
E
X Y
D J I
(a)
F
B ui
A
E
ui J I ni
X Y
D ut nt
yit
ytt N H
G
ut M
K
DIM ! ut
IDF ! ui L
FI DM
sin ui ! and sin ut !
DI DI
sin ui FI DG nt
! ! ! ! constant
Figure 4 (a) Huygens’ construction to sin ut DM DM ni
prove the law of reflection. (b) Huygens’
construction to prove the law of refraction. (b)
2 FERMAT’S PRINCIPLE
The laws of geometrical optics can also be derived, perhaps more elegantly,
from a different fundamental hypothesis. The root idea had been introduced
by Hero of Alexandria, who lived in the second century B.C. According to
Hero, when light is propagated between two points, it takes the shortest path.
For propagation between two points in the same uniform medium, the path is
clearly the straight line joining the two points. When light from the first point
A, Figure 5, reaches the second point B after reflection from a plane surface,
however, the same principle predicts the law of reflection, as follows. Figure 5
Geometrical Optics 21
A B
ui ur
O C D E
shows three possible paths from A to B, including the correct one, ADB.
Consider, however, the arbitrary path ACB. If point A¿ is constructed on the
perpendicular AO such that AO = OA¿, the right triangles AOC and A¿OC
are equal. Thus, AC = A¿C and the distance traveled by the ray of light from
A to B via C is the same as the distance from A¿ to B via C. The shortest dis-
tance from A¿ to B is obviously the straight line A¿DB, so the path ADB is the
correct choice taken by the actual light ray. Elementary geometry shows that A ui
for this path, ui = ur . Note also that to maintain A¿DB as a single straight line,
the reflected ray must remain within the plane of incidence, that is, the plane a
of the page. O ni
The French mathematician Pierre de Fermat generalized Hero’s princi- nt
ple to prove the law of refraction. If the terminal point B lies below the surface x
of a second medium, as in Figure 6, the correct path is definitely not the short- b
est path or straight line AB, for that would make the angle of refraction equal
to the angle of incidence, in violation of the empirically established law of ut
c
refraction. Appealing to the “economy of nature,” Fermat supposed instead B
that the ray of light traveled the path of least time from A to B, a generaliza-
tion that included Hero’s principle as a special case. If light travels more slow- Figure 6 Construction to prove the law of
ly in the second medium, as assumed in Figure 6, light bends at the interface so refraction from Fermat’s principle.
as to take a path that favors a shorter time in the second medium, thereby
minimizing the overall transit time from A to B. Mathematically, we are
required to minimize the total time,
AO OB
t = +
yi yt
where yi and yt are the velocities of light in the incident and transmitting
media, respectively. Employing the Pythagorean theorem and the distances
defined in Figure 6, we have AO = 2a2 + x2 and OB = 2b2 + 1c - x22 ,
so that
Since other choices of path change the position of point O and therefore the
distance x, we can minimize the time by setting dt>dx = 0:
dt x c - x
= - = 0
dx 2
yi 2a + x2
yt 2b + 1c - x22
2
Again from Figure 6, in the two right triangles containing AO and OB,
respectively, the angles of incidence and refraction can be conveniently
22 Chapter 2 Geometrical Optics
x
introduced into the preceding condition, since sin ui = and
c - x 2a2
+ x2
sin ut = , giving
2b2 + 1c - x22
dt sin ui sin ut
= - = 0
dx yi yt
3 PRINCIPLE OF REVERSIBILITY
Refer again to the cases of reflection and refraction pictured in Figures 5 and 6. If
the roles of points A and B are interchanged, so that B is the source of light rays,
Fermat’s principle of least time must predict the same path as determined for
the original direction of light propagation. In general, then, any actual ray of
light in an optical system, if reversed in direction, will retrace the same path
backward. This principle of reversibility will be found very useful in various
applications to be dealt with later.
1
It is of interest to note here that a similar principle, called Hamilton’s principle of least action
in mechanics, that calls for a minimum of the definite integral of the Lagrangian function (the ki-
netic energy minus the potential energy), represents an alternative formulation of the laws of me-
chanics and indeed implies Newton’s laws of mechanics themselves.
Geometrical Optics 23
rays reflected in various directions and thus a diffuse scattering of the origi- z
nally parallel rays of light. Every plane surface will produce some such scat-
tering, since a perfectly smooth surface can only be approximated in practice. Q
The treatment that follows assumes the case of specular reflection.
Consider the specular reflection of a single light ray OP from the xy- rˆ 2
O
plane in Figure 7a. By the law of reflection, the reflected ray PQ remains
rˆ 1
within the plane of incidence, making equal angles with the normal at P. If y
the path OPQ is resolved into its x-, y-, and z-components, it is clear that the
direction of ray OP is altered by the reflection only along the z-direction, and P
then in such a way that its z-component is simply reversed. If the direction of
the incident ray is described by its unit vector, rN 1 = 1x, y, z2, then the reflec- x
tion causes (a)
z
rN 1 = 1x, y, z2 ¡ rN 2 = 1x, y, - z2
rˆ 1
It follows that if a ray is incident from such a direction as to reflect sequen-
tially from all three rectangular coordinate planes, as in the “corner reflector”
of Figure 7b,
rˆ 2
rN 1 = 1x, y, z2 ¡ rN 2 = 1-x, - y, - z2
y
and the ray returns precisely parallel to the line of its original approach. A
network of such corner reflectors ensures the exact return of a beam of
light—a headlight beam from highway reflectors, for example, or a laser
beam from a mirror on the moon. x
Image formation in a plane mirror is illustrated in Figure 8a. A point (b)
object S sends rays toward a plane mirror, which reflect as shown. The law of
reflection ensures that pairs of triangles like SNP and S¿NP are equal, so all Figure 7 Geometry of a ray reflected
from a plane.
reflected rays appear to originate at the image point S¿, which lies along the
normal line SN, and at such a depth that the image distance S¿N equals the
object distance SN. The eye sees a point image at S¿ in exactly the same way it
would see a real point object placed there. Since none of the actual rays of
light lies below the mirror surface, the image is said to be a virtual image. The
image S¿ cannot be projected on a screen as in the case of a real image. All
points of an extended object, such as the arrow in Figure 8b, are imaged by a
plane mirror in similar fashion: Each object point has its image point along its
normal to the mirror surface and as far below the reflecting surface as the
object point lies above the surface. Note that the image position does not
depend on the position of the eye. Further, the construction of Figure 8b
makes clear that the image size is identical with the object size, giving a mag-
nification of unity. In addition, the transverse orientation of object and image
are the same. A right-handed object, however, appears left-handed in its
image. In Figure 8c, where the mirror does not lie directly below the object,
the mirror plane may be extended to determine the position of the image as
seen by an eye positioned to receive reflected rays originating at the object.
Figure 8d illustrates multiple images of a point object O formed by two per-
pendicular mirrors. Images I1 and I2 result from single reflections in the two
mirrors, but a third image I3 results from sequential reflections from both
mirrors.
5 REFRACTION THROUGH
PLANE SURFACES
Consider light ray (1) in Figure 9a, incident at angle u1 at a plane interface
separating two transparent media characterized, in order, by refractive in-
dices n1 and n2 . Let the angle of refraction be the angle u2 . Snell’s law, which
now takes the form
requires an angle of refraction such that refracted rays bend away from the
S normal, as shown in Figure 9a, for rays 1 and 2, when n2 6 n1 . For n2 7 n1 ,
on the other hand, the refracted ray bends toward the normal. The law also
requires that ray 3, incident normal to the surface 1u1 = 02, be transmitted
without change of direction 1u2 = 02, regardless of the ratio of refractive
N P indices.
In Figure 9a, the three rays shown originate at a source point S below an
interface and emerge into an upper medium of lower refractive index, as in
S! the case of light emerging from water 1n1 = 1.332 into air 1n2 = 1.002.
(a) A unique image point is not determined by these rays because they have no
common intersection or virtual image point below the surface from which
they appear to originate after refraction, as shown by the dashed line extensions
of the refracted rays. For rays making a small angle with the normal to the sur-
face, however, a reasonably good image can be located. In this approximation,
where we allow only such paraxial rays2 to form the image, the angles of inci-
dence and refraction are both small, and the approximation
sin u ! tan u ! u 1in radians2
is valid. From Eq. (2), Snell’s law can be approximated by
(b)
n1 tan u1 ! n2 tan u2 (3)
n1 $ n2 (2) n1 $ n2
u2
O (3)
I1 (1)
u2 x
n2 n2
I3 I2 n1 u2 n1
s"
u1 s
u1
(d) S"
S u1
Figure 8 Image formation in a plane S
mirror.
(a) (b)
n1 $ n2
u2 u2
n2 90%
n1 uc
u1 u1 u1 u1
2
In general, a paraxial ray is one that remains near the central axis of the image-forming opti-
cal system, thus making small angles with the optical axis.
Geometrical Optics 25
The image point occurs at the vertical distance s¿ below the surface given by
n2
s¿ = a bs (4)
n1
or
n2
uc = sin-1 a b (5)
n1
For angles of incidence u1 7 uc , the incident ray experiences total internal re-
flection, as shown. For angle of incidence u1 6 uc both refraction and reflection
occur. The reflected rays for this case are not shown in Figure 9c. This phenome-
non is essential in the transmission of light along glass fibers by a series of total
internal reflections. Note that the phenomenon does not occur unless n1 7 n2 ,
so that uc can be determined from Eq. (5).
We return to the nature of images formed by refraction at a plane sur-
face when we deal with such refraction as a special case of refraction from a
spherical surface.
We discuss now what is meant by an image in general and indicate the practical
and theoretical factors that render an image less than perfect. In Figure 10, let
the region labeled “optical system” include any number of reflecting and/or re-
fracting surfaces, of any curvature, that may alter the direction of rays leaving
an object point O. This region may include any number of intervening media,
but we shall assume that each individual medium is homogeneous and isotrop-
ic, and so characterized by its own refractive index. Thus rays spread out radial-
ly in all directions from object point O, as shown, in real object space, which
precedes the first reflecting or refracting surface of the optical system. The fam-
ily of spherical surfaces normal to the rays are the wavefronts, the locus of
Optical
O I
system
points such that each ray contacting a wavefront represents the same transit
time of light from the source. In real object space the rays are diverging and
the spherical wavefronts are expanding. Suppose now that the optical system
redirects these rays in such a way that on leaving the optical system and en-
tering real image space, the wavefronts are contracting and the rays are con-
verging to a common point that we define to be the image point, I. In the
spirit of Fermat’s principle, we can say that since every such ray starts at O
and ends at I, every such ray requires the same transit time. These rays are
said to be isochronous. Further, by the principle of reversibility, if I is the ob-
ject point, each ray will reverse its direction but maintain its path through the
optical system, and O will be the corresponding image point. The points O
and I are said to be conjugate points for the optical system. In an ideal optical
system, every ray from O intercepted by the system—and only these rays—
also passes through I. To image an actual object, this requirement must hold
for every object point and its conjugate image point.
Nonideal images are formed in practice because of (1) light scattering,
(2) aberrations, and (3) diffraction. Some rays leaving O do not reach I due to
reflection losses at refracting surfaces, diffuse reflections from reflecting sur-
faces, and scattering by inhomogeneities in transparent media. Loss of rays by
such means merely diminishes the brightness of the image; however, some of
these rays are scattered through I from nonconjugate object points, degrad-
ing the image. When the optical system itself cannot produce the one-to-one
relationship between object and image rays required for perfect imaging of
all object points, we speak of system aberrations. Finally, since every optical
system intercepts only a portion of the wavefront emerging from the object,
the image cannot be perfectly sharp. Even if the image were otherwise per-
fect, the effect of using a limited portion of the wavefront leads to diffraction
and a blurred image, which is said to be diffraction limited. This source of
imperfect image formation, discussed further in the sections under diffrac-
O I tion, represents a fundamental limit to the sharpness of an image that cannot
be entirely overcome. This difficulty rises from the wave nature of light. Only
in the unattainable limit of geometrical optics, where l : 0, would diffrac-
tion effects disappear entirely.
Reflecting or refracting surfaces that form perfect images are called
(a) Ellipsoid
Cartesian surfaces. In the case of reflection, such surfaces are the conic sec-
tions, as shown in Figure 11. In each of these figures, the roles of object and
image points may be reversed by the principle of reversibility. Notice that in
Figure 11b, the image is virtual. In Figure 11c, the parallel reflected rays are
said to form an image “at infinity.” In each case, one can show that Fermat’s
I O
principle, requiring isochronous rays between object and image points, leads
to a condition that is equivalent to the geometric definition of the correspond-
ing conic section.
Cartesian surfaces that produce perfect imaging by refraction may be
more complicated. Let us ask for the equation of the appropriate refracting sur-
(b) Hyperboloid
face that images object point O at image point I, as illustrated in Figure 12.
There an arbitrary point P with coordinates (x, y) is on the required surface
©. The requirement is that every ray from O, like OPI, refracts and passes
through the image I. Another such ray is evidently OVI, normal to the surface
at its vertex point V. By Fermat’s principle, these are isochronous rays. Since
O the media on either side of the refracting surface are characterized by differ-
ent refractive indices, however, the isochronous rays are not equal in length.
The transit time of a ray through a medium of thickness x with refractive
index n is
(c) Paraboloid x nx
t = =
Figure 11 Cartesian reflecting surfaces
y c
showing conjugate object and image points.
Geometrical Optics 27
P(x, y)
do di
V I
O x
so si
no ni
Therefore, equal times imply equal values of the product nx, called the optical
path length. In the problem at hand, then, Fermat’s principle requires that
nodo + nidi = noso + nisi = constant (6)
where the distances are defined in Figure 12. In terms of the (x, y)-coordinates
of P, the first sum of Eq. (6) becomes
O I
n01x2 + y221>2 + ni[y2 + 1so + si - x22]1>2 = constant (7)
The constant in the equation is determined by the middle member of Eq. (6), (a)
noso + nisi , which can be calculated once the specific problem is defined.
Equation (7) describes the Cartesian ovoid of revolution shown in Figure 13a. no ni
In most cases, however, the image is desired in the same optical medium
O
as the object. This goal is achieved by a lens that refracts light rays twice, once
at each surface, producing a real image outside the lens. Thus it is of particu-
lar interest to determine the Cartesian surfaces that render every object ray
parallel after the first refraction. Such rays incident on the second surface can (b)
then be refracted again to form an image. The solutions to this problem are il- ni
lustrated in Figure 13b and c. Depending on the relative magnitudes of the no
refractive indices, the appropriate refracting surface is either a hyperboloid O
1ni 7 no2 or an ellipsoid 1no 7 ni2, as shown.
The first of these corresponds to the usual case of an object in air. A (c)
double hyperbolic lens then functions as shown in Figure 14. Note, however,
that the aberration-free imaging so achieved applies only to object point O Figure 13 Cartesian refracting surfaces.
(a) Cartesian ovoid images O at I by refrac-
at the correct distance from the lens and on axis. For nearby points, imaging tion. (b) Hyperbolic surface images object
is not perfect. The larger the actual object, the less precise is its image. point O at infinity when O is at one focus
Because images of actual objects are not free from aberrations and because and ni 7 no . (c) Ellipsoid surface images
hyperboloid surfaces are difficult to grind exactly, most optical surfaces are object point O at infinity when O is at one
spherical.3 The spherical aberrations so introduced are accepted as a com- focus and no 7 ni .
promise when weighed against the relative ease of fabricating spherical sur-
faces. In the remainder of this chapter, we examine, in detail, spherical
reflecting and refracting surfaces and, more briefly, cylindrical reflecting
and refracting surfaces. Note that a plane surface can be treated as a special
case of a cylindrical or a spherical surface in the limit that the radius of cur-
vature R of either type of surface tends to infinity. O I
3
The refinement of lens construction using injection molding technology has eased the pro-
duction of lenses with aspherical surfaces.
28 Chapter 2 Geometrical Optics
u
u P
R
h
a a" w
O V Q I C
s s"
rays of light originating at O are drawn, one normal to the spherical surface at its
vertex V and the other an arbitrary ray incident at P. The first ray reflects back
along itself; the second reflects at P as if from a plane tangent at P, satisfying the
law of reflection. The two reflected rays diverge as they leave the mirror. The in-
tersection of the two rays (extended backward) determines the image point I
conjugate to O. The image is virtual, located behind the mirror surface.
Object and image distances from the vertex are shown as s and s¿, respec-
tively. A perpendicular of height h is drawn from P to the axis at Q. We seek a
relationship between s and s¿ that depends only on the radius of curvature R of
the mirror. As we shall see, such a relation is possible only to first-order ap-
proximation of the sines and cosines of the angles made by the object and
image rays to the spherical surface. This means that in place of the expansions
w3 w5
sin w = w - + - Á
3! 5!
and
w2 w4
cos w = 1 - + + Á (8)
2! 4!
4
For example, for angles w around 10°, the approximation leads to errors around 1.5%.
Geometrical Optics 29
where we have also neglected the axial distance VQ, small when angle w is
small. Cancellation of h produces the desired relationship,
1 1 2
- = - (11)
s s¿ R
1 1 2
+ = - (12)
s s¿ R
1 1 1
+ = (14)
s s¿ f
s s"
(c)
5
Although this set of sign conventions is widely used, the student is cautioned that other schemes Figure 16 Location of focal points (a)
exist. No one with a continuing involvement in optics can hope to escape confronting other conventions, and (b) and construction to determine
nor should the matter be beyond the mental flexibility of the serious student to accommodate. magnification (c) of a spherical mirror.
30 Chapter 2 Geometrical Optics
The focal point F, located a focal length f from the vertex of the mirror, and
shown in Figure 16a and b, serves as an important construction point in graphi-
cal ray-tracing techniques, which we discuss following Example 1.
In Figure 16c, a construction is shown that allows the determination of
the transverse magnification. The object is an extended object of transverse
dimension ho . The image of the top of the object arrow is located by two
rays whose behavior on reflection is known. The ray incident at the vertex
must reflect to make equal angles with the axis. The other ray is directed
toward the center of curvature along a normal and so must reflect back
along itself. The intersection of the two reflected rays occurs behind the
mirror and locates a virtual image of dimension hi there. Because of the
equality of the three angles shown, it follows that
ho hi
=
s s¿
hi s¿
ƒmƒ = = (15)
ho s
Extending the sign convention to include magnification, we assign a 1 +2
magnification to the case where the image has the same orientation as the
object and a 1- 2 magnification where the image is inverted relative to the
object. To produce a 1+ 2 magnification in the construction of Figure 16c,
where s¿ must itself be negative, we modify Eq. (15) to give the general
form
s¿
m = - (16)
s
The following example illustrates the correct use of the sign convention.
Example 1
An object 3 cm high is placed 20 cm from (a) a convex and (b) a concave
spherical mirror, each of 10-cm focal length. Determine the position and na-
ture of the image in each case.
Solution
1 1 1 fs 1-1021202
+ = or s¿ = = = - 6.67 cm
s s¿ f s - f 1202 - 1-102
s¿ - 6.67 1
m = - = - = + 0.333 =
s 20 3
fs 11021202
s¿ = = = + 20 cm
s - f 20 - 10
s¿ 20
m = - = - = -1
s 20
Geometrical Optics 31
The location and nature of the image formed by a mirror can be de-
termined by graphical ray-trace techniques. Figure 17 illustrates how three
key rays—labeled 1, 2, and 3—each leaving a point P at the tip of an object,
can be drawn to locate the conjugate image point P¿. In fact, under the
conditions for which Eqs. (12) through (16) are valid, the paths of any two
rays leaving P are sufficient to locate the conjugate image point P¿. A third
ray serves as a convenient check on the accuracy of the first two chosen
rays. The three key rays discussed in connection with Figure 17 are chosen
as the basis of the graphical ray-trace technique because, once the mirror
center of curvature C, the focal point F, and vertex V are located along the
optical axis of a spherical mirror, these three rays can be drawn using only
a straightedge device. The conjugate image point P¿ marks the tip of the
image—the entire image then lies between P¿ and the point on the optical
axis directly above or below P¿.
Refer to Figure 17a, b, and c in connection with the following descrip-
tion of how the three key rays can be drawn. Note the difference in each ray
2" P"
3"
P 1 2
3
2 P
1
3 I 3"
V
O C F C F O V I
2" P"
1" 1"
(a) (b)
1"
3"
P 1
3
2" 2 P"
O V I F C
trace, depending on the object location before or after points C and F, and on
the geometry of the mirror surface, concave or convex.
• Ray 1. This ray leaves point P as a ray parallel to the optical axis, strikes
the mirror, reflects and passes through the focal point F of a concave
mirror—as in Figure 17a and b. Or, as in Figure 17c, it strikes a convex
mirror and reflects as if it came from the focal point F behind the mir-
ror. In each case, after reflection this ray is labeled 1¿.
• Ray 2. This ray leaves point P, passes through F, strikes a concave
mirror, and is reflected as a ray parallel to the optical axis, as in
Figure 17a. Or, as in Figure 17b, it leaves point P as if it is coming
from the point F to its left (dotted line), strikes the concave mirror,
and reflects as a parallel ray. Or, as in Figure 17c, for a convex mirror,
the ray leaves point P heading toward focal point F behind the
mirror, strikes the mirror, and reflects as a parallel ray. In each case,
after reflection, this ray is labeled 2¿.
• Ray 3. This ray leaves point P in Figure 17a, passes through point C for
the concave mirror, strikes the mirror, and reflects back along itself. Or,
as in Figure 17b—still for a concave mirror—ray 3 appears to come from
the point C to its left, strikes the mirror, and reflects back along itself.
Or, as in Figure 17c, for a convex mirror, it heads toward point C behind
the mirror, strikes the mirror, and reflects back along itself. In each case,
after reflection, this ray is labeled 3¿.
To understand how these rays locate the conjugate image point P¿ that
marks the tip of the image, it is useful to imagine that these three rays arrive
at the eye of one viewing the image. For the case shown in Figure 17a, the
three rays 1¿, 2¿, and 3¿ intersect at a real image point as they progress away
from the mirror and toward the viewer. For the arrangements shown in
Figure 17b and 17c, the rays 1¿, 2¿, and 3¿ appear to originate from a point
of intersection (a virtual image point) located behind the mirror. The real or
apparent point of intersection is interpreted as the emanation point of these
rays. That is, the viewer “sees” the tip of an image at point P¿.
8 REFRACTION AT A SPHERICAL
SURFACE
The two refracted rays appear to emerge from their common intersection, the
image point I. In triangle CPO, the exterior angle a = u1 + w. In triangle
CPI, the exterior angle a¿ = u2 + w. Approximating for paraxial rays and
substituting for u1 and u2 in Eq. (17), we have
Next, writing the tangents for the angles by inspection of Figure 18, where
again we may neglect the distance QV in the small angle approximation,
h h h h
n1 a - b = n2 a - b
s R s¿ R
Geometrical Optics 33
u2 al
orm
N
P
u1
u2 h
w a! a
C I O Q V
n1 n2
s
s!
or
n1 n2 n1 - n2
- = (19)
s s¿ R
Employing the same sign convention as introduced for mirrors (i.e., positive dis-
tances for real objects and images and negative distances for virtual objects and
images), the virtual image distance s¿ 6 0 and the radius of curvature R 6 0. If
these negative signs are understood to apply to these quantities for the case of
Figure 18, a general form of the refraction equation may be written as
n1 n2 n2 - n1
+ = (20)
s s¿ R
which holds equally well for convex surfaces. When R : q , the spherical
surface becomes a plane refracting surface, and
n2
s¿ = - a bs (21)
n1
hi n1s¿
m = = - (22)
ho n2s
n1 n2
ho
u1 C I
O V hi
u2
Figure 19 Construction to determine later-
al magnification at a spherical refracting s s"
surface.
Example 2
As an extended example of refraction by spherical surfaces, refer to Figure 20.
In (a), a real object is positioned in air, 30 cm from a convex spherical sur-
face of radius 5 cm. To the right of the interface, the refractive index is that
of water. Before constructing representative rays, we first find the image dis-
tance and lateral magnification of the image, using Eqs. (20) and (22).
Equation (20) becomes
1 1.33 1.33 - 1
+ =
30 s¿ 1 5
giving s¿ 1 = + 40 cm. The positive sign indicates that the image is real and
so is located to the right of the surface, where real rays of light are re-
fracted. Equation (22) becomes
1121 +402
m = - = -1
11.3321 +302
R !5
n1 ! 1 n2 ! 1.33
RI
RO1
30 40
(a)
n2 ! 1.33 n1 ! 1
n1 ! 1
s"2 ! 9
VO2
RI2 RI1
RO1
indicating an inverted image, equal in size to that of the object. Figure 20a
shows the image, as well as several rays, which are now determined. In
this example we have assumed that the medium to the right of the spher-
ical surface extends far enough so that the image is formed inside it,
without further refraction. Let us suppose now (Figure 20b) that the
second medium is only 10 cm thick, forming a thick lens, with a second,
concave spherical surface, also of radius 5 cm. The refraction by the first
surface is, of course, unaffected by this change. Inside the lens, therefore,
rays are directed as before to form an image 40 cm to the right of the
first surface. However, these rays are intercepted and refracted by the
second surface to produce a different image, as shown. Since the con-
vergence of the rays striking the second surface is determined by the po-
sition of the first image, its location now specifies the appropriate object
distance to be used for the second refraction. We call the real image formed
by surface (1) a virtual object for surface (2). Then, by the sign convention es-
tablished previously, we make the virtual object distance, relative to the sec-
ond surface, a negative quantity when using Eqs. (20) and (22). For the second
refraction, then, Eq. (20) becomes
1.33 1 1 - 1.33
+ =
-30 s¿ 2 -5
The final image is, then, 2>5 the lateral size of its (virtual) object and apears with
the same orientation. Relative to the original object, the final image is 2>5 as
large and inverted.
9 THIN LENSES
We have assumed that the lens faces the same medium of refractive index n1
on both sides. Now the second object distance, in general, is given by
s2 = t - s1œ (25)
36 Chapter 2 Geometrical Optics
where t is the thickness of the lens. Notice that this relationship produces the
correct sign of s2 , as in Figure 20, and also when the intermediate image falls
inside or to the left of the lens. In the thin-lens approximation, neglecting t,
s2 = - s1œ (26)
When this value of s2 is substituted into Eq. (24) and Eqs. (23) and (24) are
added, the terms n2>s1œ cancel and there results
n1 n1 1 1
+ œ = 1n2 - n12a - b
s1 s2 R1 R2
Now s1 is the original object distance and s2œ is the final image distance, so we
may drop their subscripts and write simply
1 1 n2 - n1 1 1
+ = a - b (27)
s s¿ n1 R1 R2
The focal length of the thin lens is defined as the image distance for an object
at infinity, or the object distance for an image at infinity, giving
1 n2 - n1 1 1
= a - b (28)
f n1 R1 R2
Equation (28) is called the lensmaker’s equation because it predicts the focal
length of a lens fabricated with a given refractive index and radii of curvature
and used in a medium of refractive index n1 . In most cases, the ambient
F
medium is air, and n1 = 1. The thin-lens equation, in terms of the focal
length, is then
1 1 1
+ = (29)
(a) s s¿ f
• Ray 1. A ray leaving the tip of the object, parallel to the optical axis, un-
dergoing refraction at the lens surfaces and passing through the right focal
point F of a converging lens, as in Figure 22a. Or, as in Figure 22b, a paral-
lel ray which refracts at the lens surfaces as if coming directly from the
left focal point F of a diverging lens.
• Ray 2. A ray leaving the tip of the object and passing through the left focal
point F of a converging lens, undergoing refraction at the lens surfaces,
Geometrical Optics 37
s s"
3
ho
2 RI
RO F F hi
(a)
1
2
ho 3
RO F VI F
s
s"
Figure 22 Ray diagrams for image forma-
tion by a convex lens (a) and a concave
(b) lens (b).
and emerging parallel to the axis as in Figure 22a. Or, as in Figure 22b, a
ray leaving the tip of the object, directed toward the right focal point F
of a diverging lens, undergoing refraction at the lens and emerging par-
allel to the axis.
• Ray 3. A ray leaving the tip of the object and passing directly through
the center of a converging or diverging lens, emerging unaltered, as in
Figure 22a or 22b.
The viewer, located at the far right in Figure 22a and 22b, receives these rays
as if they have come directly from an object and so “sees” the tip of the image
at the point where the backwards extensions of these rays either intersect or
appear to intersect. Any two rays are sufficient to locate the image; the third
ray may be drawn as a check on the accuracy of the graphical trace.
In constructing ray diagrams, as in Figure 22, observe that, except for the
central ray (ray 3), each ray refracted by a convex lens bends toward the axis
and each ray refracted by a concave lens bends away from the axis. From ei-
ther diagram, the angles subtended by object and image at the center of the
lens are seen to be equal. For either the real image RI in (a) or the virtual
image VI in (b), it follows that
ho hi
=
s s¿
and lateral magnification
hi s¿
ƒmƒ = ` ` = ` `
ho s
(1) (2)
RO2
F1 RI1 VI2 F2
RO1 F1 F2
(a)
(1) (2)
VO2
F2 RI2 RI1
RO1 F1 F1 F2
(b)
Figure 23 (a) Formation of a virtual image VI2 by a two-element train of a convex lens (1) and
concave lens (2). (b) Formation of a real image RI2 by a train of two convex lenses. The intermediate
image RI1 serves as a virtual object VO2 for the second lens.
Example 3
Find and describe the intermediate and final images produced by a two-lens
system such as the one sketched in Figure 23a. Let f1 = 15 cm, f2 = 15 cm,
and their separation be 60 cm. Let the object be 25 cm from the first lens, as
shown.
Solution
The first lens is convex: f1 = + 15 cm, s1 = 25 cm.
1 1 1 s1f 12521152
+ œ = or s1œ = = = + 37.5 cm
s1 s1 f s1 - f 25 - 15
s1œ 37.5
m1 = - = - = - 1.5
s1 25
Thus, the first image is real (because s1œ is positive), 37.5 cm to the right of the
first lens, inverted (because m is negative), and 1.5 times the size of the object.
Geometrical Optics 39
The second lens is concave: f2 = - 15 cm. Since real rays of light diverge
from the first real image, it serves as a real object for the second lens, with
s2 = 60 - 37.5 = + 22.5 cm to the left of the lens. Then,
s2f 122.521 -152
s2œ = = = - 9 cm
s2 - f 122.52 - 1 -152
s2œ -9
m2 = - = - = + 0.4
s2 22.5
Thus, the final image is virtual (because s2œ is negative), 9 cm to the left of
the seconds lens, erect with respect to its own object (because m is positive),
and 0.4 times its size. The overall magnification is given by m = m1m2
= 1 -1.5210.42 = - 0.6. Thus, the final image is inverted relative to the origi-
nal object and 6>10 its lateral size. All these features are exhibited qualita-
tively in the ray diagram of Figure 23a.
1 1 1
+ = (31)
s s¿ f
1 1 1 R
+ = ,f = - s¿ = -s
s s¿ f 2
s¿
Reflection m = - m = +1
s
Concave: f 7 0, R 6 0
Convex : f 6 0, R 7 0
n1 n2 n2 - n1 n2
+ = s¿ = - s
s s¿ R n1
n1s¿
Refraction Single surface m = - m = +1
n2s
Concave: R 6 0
Convex : R 7 0
1 1 1
+ =
s s¿ f
1 n2 - n1 1 1
Refraction Thin lens = a - b
f n1 R1 R2
s¿
m = -
s
Concave: f 6 0
Convex : f 7 0
40 Chapter 2 Geometrical Optics
VI3 VI3
RO1 RO2 RO3 #
# RO1 RO2 RO3
C F 2F F F 2F
RI1 RI1
#
# RI2
RI2
RO RO
# VI VI
F C F F
(a) (b)
Figure 24 Summary of image formation by (a) spherical mirrors and (b) thin lenses. The location, nature, magnifi-
cation, and orientation of the image are indicated or suggested. The letters R and V refer to real and virtual, O and I
to object and image. Changes in elevation of the horizontal lines suggest the magnification in the various regions.
notice that (1) the reciprocals of distances in the left member add to give the
reciprocal of the focal length and (2) the reciprocals of the object and image
distances describe the curvature of the wavefronts incident at the lens and
centered at the object and image positions O and I, respectively. A plane
wavefront, for example, has a curvature of zero. In Figure 25 spherical waves
expand from the object point O and attain a curvature, or vergence, V,
given by 1/s, when they intercept the thin lens. On the other hand, once
refracted by the lens, the wavefronts contract, in Figure 25a, and expand
further, in Figure 25b, to locate the real and virtual image points shown. The
curvature, or vergence, V¿, of the wavefronts as they emerge from the lens is
1>s¿. The change in curvature from object space to image space is due to the
refracting power P of the lens, given by 1/f. With these definitions, Eq. (31)
may be written
V + V¿ = P (32)
I
2F O F F 2F 3F O I
s s! s
s!
(a)
(b)
Figure 25 Change in curvature of wavefronts on refraction by a thin lens. (a) Convex lens. (b) Concave lens.
Geometrical Optics 41
The units of the terms in Eq. (32) are reciprocal lengths. When the lengths
are measured in meters, their reciprocals are said to have units of diopters
(D). Thus, the refracting power of a lens of focal length 20 cm is said to be
1
= 5 diopters. This alternative point of view emphasizes the degree of
0.2 m
wave curvature or ray convergence rather than object and image distances.
Accordingly, the degree of convergence V¿ of the image rays is determined
by the original degree of convergence V of the object rays and the refract-
ing power P of the lens, that is, the power to change incident wave curva-
ture. Eq. (32) can also be applied to the case of refraction at a single
surface, Eq. (20), in which case the refractive indices in object and image
space need not be 1. In this event, the power of the refracting surface is
1n2 - n12>R.
This approach is useful for another reason. When thin lenses are placed
together, in contact, the focal length f of the combination, treated as a single
thin lens, can be found in terms of the focal lengths f1 , f2 , Á of the individ-
ual lenses. For example, with two such lenses back-to-back, we write the lens
equations
1 1 1 1 1 1
+ œ = and + œ =
s1 s1 f1 s2 s2 f2
Since the image distance for the first lens plays the role of the object distance
for the second lens, we may write
s2 = - s1œ
1 1 1 1 1
+ œ = + =
s1 s2 f1 f2 f
The reciprocals of the individual focal lengths, therefore, add to give the reci-
procal of the overall focal length f of the pair. In general, for several thin lens-
es, in direct contact,
1 1 1 1
= + + + Á (33)
f f1 f2 f3
P = P1 + P2 + P3 + Á (34)
11 NEWTONIAN EQUATION
FOR THE THIN LENS
When object and image distances are measured relative to the focal points F
of a lens, as by the distances x and x¿ in Figure 26, an alternative form of the
thin-lens equation results, called the Newtonian form. In the figure, the two
rays shown determine two right triangles, joined by the focal point, on each
side of the lens. Since each pair constitutes similar triangles, we may set up
proportions between sides that represent the lateral magnification:
hi f hi x¿
= and =
ho x ho f
Introducing a negative sign for the magnification, due to the inverted image,
f x¿
m = - = - (35)
x f
The two parts of Eq. (35) also constitute the Newtonian form of the thin-lens
equation,
xx¿ = f2 (36)
This equation is somewhat simpler than Eq. (29) and is found to be more
convenient in certain applications.
12 CYLINDRICAL LENSES
Spherical lenses and mirrors with circular cross sections are far more com-
mon in optical systems than are cylindrical lenses. Nevertheless, cylindrical
lenses are important, for example, in the field of optometry for correcting the
visual defect known as astigmatism, as well as in novel visual displays where
it is useful to image points as lines. We close this chapter on geometrical op-
tics with a brief look at this special type of lens.
The optical axis for a spherical lens is an axis of symmetry since rotation
of the lens through an arbitrary angle about the optical axis leaves the lens
looking just as it did before the rotation. Because the orientation of the sur-
face curvature does not change in such a rotation, its optical behavior re-
mains unchanged. This rotational symmetry simplifies the analysis of the
imaging properties of such a spherical lens. On the other hand, a cylindrical
lens—shaped like a section of a soft drink can, sliced down the side from top
to bottom—lacks rotational symmetry about its optical axis. As a consequence,
a cylindrical lens has asymmetric focusing properties, as will be seen later in
greater detail. Whereas a spherical lens produces a point image of a point
s o"
f x"
ho ho
F F
x f hi hi
object, a cylindrical lens produces a line image of a point object. Because of Point Axis
these properties, a spherical lens is said to be stigmatic, and the cylindrical image
F
lens astigmatic.
Consider first a spherical lens, as shown in Figure 27a and b. Orthogo- Lens
nal vertical and horizontal axes are shown as solid diametrical lines through
the geometric lens center. Parallel rays of light passing through the vertical
axis (see Figure 27a) and through the horizontal axis (see Figure 27b) are
handled identically by the lens, converging them to a common focus at F.
The lens can be rotated through an arbitrary angle about its optical axis
with the same result. Thus, the focusing properties of a spherical lens are in-
variant to rotation about its central (optical) axis.
Next, consider the convex and concave cylindrical lenses shown in
Figure 28. One surface of the lens is cylindrical while the opposite is plane.6
Thus, the curved surface has a definite, finite radius of curvature, whereas the Vertical fan of rays
plane surface has an infinite radius of curvature. In Figure 29, two vertical (a)
slices or sections are shown perpendicular to the axis of a convex cylindrical
lens. Through each section, three representative rays are drawn. The opera- Point
tion of this lens is clearly asymmetric. Focusing occurs for rays along a verti- image Axis
F
cal section but not for rays along a horizontal section, where the lens presents
no curvature. Thus, rays 1, 2 and 3 focus to point A, and rays 4, 5 and 6 focus
Lens
to point B. However, there is no focusing of rays in a horizontal section, such
as the pairs of rays 1 and 4, 2 and 5, or 3 and 6. Other vertical sections would
produce other points along the focused line image AB in the same way. No-
tice that the line image AB so formed is always parallel to the cylinder axis.
This important feature is also shown in Figure 30, where the line image is real
for a cylindrical convex lens and virtual for a cylindrical concave lens. From
these figures, it is evident that the length of the line image is equal to the axial
length of the lens, assuming that rays of light parallel to the optical axis enter
along the entire extent of the lens. If an aperture is placed in front of the lens
to limit the bundle of light rays through the lens, the height of the line image Horizontal fan of rays
is just the aperture dimension along the cylinder axis, or the effective height of (b)
the lens.
In Figure 29, the line image formed is the result of an object point “at Figure 27 Parallel rays of light focused by a
infinity,” which produces parallel rays at the lens. In Figure 31, the object spherical lens. Because of its axis of symme-
try relative to rotation about an axis through
point O is near the lens, producing diverging rays of light incident on the its center, the lens treats (a) vertical and (b)
lens. Still, if the lens is thin, focusing occurs along the vertical sections, as horizontal fans of rays similarly, producing in
shown. Rays 1 and 3, in the left vertical section of Figure 31, focus at A; each case a point image at the same location.
rays 2 and 4 in the right vertical section focus at B. However, no focusing Each ray refracts twice through the lens, once
occurs for rays 1, 5, and 2 along the horizontal section. Because of the at each surface. For simplicity, only one re-
fraction is shown.
divergence of the rays entering the lens, however, the length of the focused
line image AB is no longer equal to the effective length CL of the lens. The
divergence of the extreme rays at each end of the lens now determines an
image that is longer than the length of the lens. The image length AB can
be found from a simple, geometrical argument that is apparent in Figure 32a,
a view of the central horizontal section in Figure 31 as seen from above. If
the effective length of the cylindrical lens is CL, then by similar triangles it
follows that
AB s + s¿
=
CL s
6
To be more precise, we are speaking of a plano-convex or plano-concave cylindrical lens as
shown in Figure 28. Generally speaking, both surfaces of the lens might be cylindrical. In such a case,
the behavior of the lens as a whole, due to the addition of the powers of the two surfaces, may not re-
duce to that of the simple plano-convex or plano-concave lens described here.
44 Chapter 2 Geometrical Optics
or
s + s¿
AB = a b CL (37)
s
Subject to our sign convention, this relation is a general form for the plano-
cylindrical lens that handles all cases, with s and s¿ object and image distances,
respectively, and AB always positive.
Example 4
Convex A thin plano-cylindrical lens in air has a radius of curvature of 10 cm, a re-
fractive index of 1.50, and an axial length of 5 cm. Light from a point object is
(a)
incident on the convex cylindrical surface from a distance of 25 cm to the left
of the lens. Find the position and length of the line image formed by the lens.
Solution
As given, s = 25 cm, R = 10 cm, n1lens2 = 1.50 and CL = 5 cm. Using the
spherical surface refraction equation (see Table 1),
n1 n2 n2 - n1
+ =
s s¿ R
and
s + s¿
AB = a bCL
Concave s
(b)
together with the sign convention—positive for real objects and images, neg-
Figure 28 Cylindrical lenses shown as sec- ative for virtual objects and images, positive R for convex surface.
tions of a solid and hollow cylindrical rod. 1
Entering values, we have for the first convex surface at entry, +
25
1.50 1.50 - 1.00
= , which gives s¿ = 150 cm, real. And for the second
s¿ 10
Line
A image B
Cylinder
axis
Virtual
vertical
line
image
B A
Real
vertical
line
image
A
(a) (b)
Figure 30 Formation of line images by cylindrical lenses for light incident from a distant object. In (a) the convex lens forms a real
image. In (b), the concave lens forms a virtual image. In either case, the line image is parallel to the cylinder axis.
I Line
image
B
3 C
5
1 L Figure 31 Formation of a line image AB by a
2 convex cylindrical lens when the object is a
point O at a finite distance from the lens. In
Parallel to
this case, the line image AB is longer than the
O 4 cylinder axis
axial length of the lens, CL.
1.50 1.0
(plane) surface at exit, we obtain + = 0, which gives s¿ = 100 cm.
- 150 s¿
25 + 100
Then with Eq. (37), AB = a b5 cm = 25 cm.
25
Thus, the line image is parallel to the cylindrical axis, enlarged to 25 cm, and
located 1 m from the lens. If the lens is rotated about its optical axis, the line
image also rotates, remaining always parallel to the cylindrical axis.
Looking again at Figure 31, imagine a screen placed on the exit side of
the lens so as to capture the light from the lens. We have argued that when the
screen is at the distance s¿ from the lens, one sees a focused line image AB on
the screen, in this case with a horizontal orientation. As the screen is moved
46 Chapter 2 Geometrical Optics
s s! further than s¿ from the lens, one sees an unfocused blur that has the general
A shape of the aperture—either of the rectangular cross section of the lens
C alone or of the lens with an aperture placed against it. Further, it should be
evident from Figure 31 that, as a screen, initially positioned just behind the
I
lens, moves toward the line image AB, the horizontal dimension (width) of
O
the blur increases and its vertical dimension (height) decreases. As the screen
L
moves beyond the line image, its width continues to increase but its height
B now also increases due to the divergence of the rays after focusing. If an aper-
(a) ture placed in front of the lens is circular, these blur images are elliptical in
shape, with changing major and minor axes formed by the width and height
of the blur. If the aperture is square, the blurs are rectangular in shape.
I Widths and heights of the blur pattern can be found at any position of the
O screen using the geometry apparent in Figure 32a and b, respectively. This
behavior can be observed easily in the laboratory.
s s!
Up to this point we have been dealing with a cylindrical lens whose
(b)
axis is either horizontal or vertical. Of course, the cylinder axis can be ori-
Figure 32 (a) Light rays in a top view of ented at any angle. An astigmatic eye, for example, while it possesses pre-
the horizontal (nonfocusing) section of the dominantly spherical optics, might have a cylindrical axis component whose
lens in Figure 31. (b) Light rays in a side
view of the vertical (focusing) section of the
axis could be horizontal, vertical, or some angle in between. To deal with
lens in Figure 31. cylindrical lenses and astigmatism in a general way, then, we must be able to
determine the effect of combining cylindrical lenses having arbitrary orien-
tations with each other and with spherical lenses. It turns out that two cylin-
drical lenses can produce the same effect as a sphero-cylindrical lens. Lens
prescriptions for vision correction are, in fact, expressed in terms of combi-
nations of spherical and cylindrical lenses. This subject is treated further
elsewhere.7
PROBLEMS
1 Derive an expression for the transit time of a ray of light 4 Determine the minimum height of a wall mirror that will
that travels a distance x1 through a medium of index n1 , a permit a 6-ft person to view his or her entire height.
distance x2 through a medium of index n2 , Á , and a dis- Sketch rays from the top and bottom of the person, and
tance xm through a medium of index nm . Use a summation determine the proper placement of the mirror such that
to express your result. the full image is seen, regardless of the person’s distance
2 Deduce the Cartesian oval for perfect imaging by a refract- from the mirror.
ing surface when the object point is on the optical x-axis 20 5 A ray of light makes an angle of incidence of 45° at the cen-
cm from the surface vertex and its conjugate image point ter of the top surface of a transparent cube of index 1.414.
lies 10 cm inside the second medium. Assume the refracting Trace the ray through the cube.
medium to have an index of 1.50 and the outer medium to
be air. Find the equation of the intersection of the oval with 6 To determine the refractive index of a transparent plate of
the xy-plane, where the origin of the coordinates is at the glass, a microscope is first focused on a tiny scratch in the
object point. Generate a table of (x, y)-coordinates for the upper surface, and the barrel position is recorded. Upon
surface and plot, together with sample rays. further lowering the microscope barrel by 1.87 mm, a fo-
cused image of the scratch is seen again. The plate thickness
3 A double convex lens has a diameter of 5 cm and zero thick-
is 1.50 mm. What is the reason for the second image, and
ness at its edges. A point object on an axis through the cen-
what is the refractive index of the glass?
ter of the lens produces a real image on the opposite side.
Both object and image distances are 30 cm, measured from 7 A small source of light at the bottom face of a rectangular
a plane bisecting the lens. The lens has a refractive index of glass slab 2.25 cm thick is viewed from above. Rays of light
1.52. Using the equivalence of optical paths through the totally internally reflected at the top surface outline a circle
center and edge of the lens, determine the thickness of the of 7.60 cm in diameter on the bottom surface. Determine
lens at its center. the refractive index of the glass.
7
See F. L. Pedrotti and L. S. Pedrotti, Optics and Vision (Upper Saddle River, N. J.: Prentice
Hall, Inc., 1998).
Geometrical Optics 47
2.25 cm
uc
Light
7.6 cm
Figure 33 Problem 7
8 Show that the lateral displacement s of a ray of light pene- 13 a. At what position in front of a spherical refracting surface
trating a rectangular plate of thickness t is given by must an object be placed so that the refraction produces
t sin1u1 - u22 parallel rays of light? In other words, what is the focal length
s = of a single refracting surface?
cos u2
b. Since real object distances are positive, what does your
where u1 and u2 are the angles of incidence and refraction, re- result imply for the cases n2 7 n1 and n2 6 n1?
spectively. Find the displacement when t = 3 cm, n = 1.50, 14 A small goldfish is viewed through a spherical glass fish-
and u1 = 50°. bowl 30 cm in diameter. Determine the apparent position
9 A meter stick lies along the optical axis of a convex mirror and magnification of the fish’s eye when its actual position
of focal length 40 cm, with its nearer end 60 cm from the mir- is (a) at the center of the bowl and (b) nearer to the oberver,
ror surface. How long is the image of the meter stick? halfway from center to glass, along the line of sight. As-
sume that the glass is thin enough so that its effect on the
10 A glass hemisphere is silvered over its curved surface. A
refraction may be neglected.
small air bubble in the glass is located on the central axis
through the hemisphere 5 cm from the plane surface. The 15 A small object faces the convex spherical glass window of a
radius of curvature of the spherical surface is 7.5 cm, and small water tank. The radius of curvature of the window is 5
the glass has an index of 1.50. Looking along the axis into cm. The inner back side of the tank is a plane mirror, 25 cm
the plane surface, one sees two images of the bubble. How from the window. If the object is 30 cm outside the window,
do they arise and where do they appear? determine the nature of its final image, neglecting any re-
fraction due to the thin glass window itself.
Window
cm R ! 5 cm
5
7. O Tank
!
R n ! 4/3
5 cm
30 cm 25 cm
Figure 34 Problem 10
16 A plano-convex lens having a focal length of 25.0 cm is to
be made with glass of refractive index 1.520. Calculate the
radius of curvature of the grinding and polishing tools to be
used in making this lens.
11 A concave mirror forms an image on a screen twice as large
as the object. Both object and screen are then moved to 17 Calculate the focal length of a thin meniscus lens whose spher-
produce an image on the screen that is three times the size ical surfaces have radii of curvature of magnitude 5 and 10 cm.
of the object. If the screen is moved 75 cm in the process, The glass is of index 1.50. Sketch both positive and negative
how far is the object moved? What is the focal length of the versions of the lens.
mirror?
18 One side of a fish tank is built using a large-aperture thin lens
12 A sphere 5 cm in diameter has a small scratch on its surface. made of glass 1n = 1.502. The lens is equiconvex, with radii of
When the scratch is viewed through the glass from a position curvature 30 cm. A small fish in the tank is 20 cm from the
directly opposite, where does the scratch appear and what is lens. Where does the fish appear when viewed through the
its magnification? Assume n = 1.50 for the glass. lens? What is its magnification?
48 Chapter 2 Geometrical Optics
19 Two thin lenses have focal lengths of - 5 and +20 cm. Deter- the diverging lens, and the mirror is placed a distance 3f on
mine their equivalent focal lengths when (a) cemented to- the other side of the lens. Using Gaussian optics, deter-
gether and (b) separated by 10 cm. mine the final image of the system, after two refractions (a)
by a three-ray diagram and (b) by calculation.
20 Two identical, thin, plano-convex lenses with radii of curva-
ture of 15 cm are situated with their curved surfaces in con- 23 A small object is placed 20 cm from the first of a train of
tact at their centers. The intervening space is filled with oil three lenses with focal lengths, in order, of 10, 15, and 20 cm.
of refractive index 1.65. The index of the glass is 1.50. Deter- The first two lenses are separated by 30 cm and the last two
mine the focal length of the combination. (Hint: Think of by 20 cm. Calculate the final image position relative to the
the oil layer as an intermediate thin lens.) last lens and its linear magnification relative to the original
object when (a) all three lenses are positive, (b) the middle
Oil n ! 1.65 lens is negative, (c) the first and last lenses are negative.
Provide ray diagrams for each case.
24 A convex thin lens with refractive index of 1.50 has a focal
n ! 1.5 n ! 1.5 length of 30 cm in air. When immersed in a certain trans-
parent liquid, it becomes a negative lens with a focal
length of 188 cm. Determine the refractive index of the
liquid.
25 It is desired to project onto a screen an image that is four
times the size of a brightly illuminated object. A plano-
|R| ! 15 cm convex lens with n = 1.50 and R = 60 cm is to be used.
Employing the Newtonian form of the lens equations, deter-
Figure 36 Problem 20 mine the appropriate distance of the object and screen from
the lens. Is the image erect or inverted? Check your results
21 An eyepiece is made of two thin lenses each of + 20-mm using the ordinary lens equations.
focal length, separated by a distance of 16 mm.
26 Three thin lenses of focal lengths 10 cm, 20 cm, and
a. Where must a small object be positioned so that light -40 cm are placed in contact to form a single compound
from the object is rendered parallel by the combination? lens.
b. Does the eye see an image erect relative to the object?
a. Determine the powers of the individual lenses and that
Is it magnified? Use a ray diagram to answer these ques-
of the unit, in diopters.
tions by inspection.
b. Determine the vergence of an object point 12 cm from
22 A diverging thin lens and a concave mirror have focal the unit and that of the resulting image. Convert the re-
lengths of equal magnitude. An object is placed (3/2)f from sult to an image distance in centimeters.
27 A lens is moved along the optical axis between a fixed
f f object and a fixed image screen. The object and image
positions are separated by a distance L that is more than
four times the focal length of the lens. Two positions of
the lens are found for which an image is in focus on the
screen, magnified in one case and reduced in the other. If
the two lens positions differ by distance D, show that the
3f
3f focal length of the lens is given by f = 1L2 - D22>4L.
2
This is Bessel’s method for finding the focal length of a
Figure 37 Problem 22 lens.
f f
Figure 38 Problem 27
Geometrical Optics 49
28 An image of an object is formed on a screen by a lens. Leav- 30 Determine the ratio of focal lengths for two identical, thin,
ing the lens fixed, the object is moved to a new position and plano-convex lenses when one is silvered on its flat side and
the image screen moved until it again receives a focused the other on its curved side. Light is incident on the unsil-
image. If the two object positions are S1 and S2 and if the vered side.
transverse magnifications of the image are M1 and M2 , re-
31 Show that the minimum distance between an object and its
spectively, show that the focal length of the lens is given by
image, formed by a thin lens, is 4f. When does this occur?
1S2 - S12
f = 32 A ray of light traverses successively a series of plane inter-
1 1 faces, all parallel to one another and separating regions of
a - b
M1 M2 differing thickness and refractive index.
This is Abbe’s method for finding the focal length of a lens. a. Show that Snell’s law holds between the first and last re-
29 Derive the law of reflection from Fermat’s principle by min- gions, as if the intervening regions did not exist.
imizing the distance of an arbitrary (hypothetical) ray from b. Calculate the net lateral displacement of the ray from
a given source point to a given receiving point. point of incidence to point of emergence.
uf
t1 t2
n0 n1 n2 nf
uo
Figure 39 Problem 32
33 A parallel beam of light is incident on a plano-convex lens 37 A plano-cylindrical lens in air has a curvature of 15 cm and an
that is 4 cm thick. The radius of curvature of the spherical axial length of 2.5 cm. The refractive index of the lens is
side is also 4 cm. The lens has a refractive index of 1.50 and 1.52. Find the position and length of the line image formed
is used in air. Determine where the light is focused for light by the lens for a point object 20 cm from the lens. Light from
incident on each side. the object is incident on the convex cylindrical surface of
the lens.
34 A spherical interface, with radius of curvature 10 cm, sepa-
rates media of refractive index 1 and 43 . The center of curva- 38 A plano-cylindrical lens in air has a radius of curvature of
ture is located on the side of the higher index. Find the focal 10 cm, a refractive index of 1.50, and an axial length of 5 cm.
lengths for light incident from each side. How do the results Light from a point object is incident on the concave, cylin-
differ when the two refractive indices are interchanged? drical surface from a distance of 25 cm to the left of the lens.
35 An airplane is used in aerial surveying to make a map of Find the position and length of the image formed by the
ground detail. If the scale of the map is to be 1:50,000 and the lens.
camera used has a focal length of 6 in., determine the proper 39 A plano-concave cylindrical lens is used to form an image
altitude for the photograph. of a point object 20 cm from the lens. The lens has a refrac-
36 Light rays emanating in air from a point object on axis tive index of 1.50, a radius of curvature of 20 cm, and an
strike a plano-cylindrical lens with its convex surface facing axial length of 2 cm. Describe as completely as possible the
the object. Describe the line image by length and location if line image of the point.
the lens has a radius of curvature of 5 cm, a refractive index 40 Consider the plano-convex cylindrical lens in problem 36. If
of 1.60, and an axial length of 7 cm. The point object is 15 cm the point object is only 6 cm from the lens, describe the line
from the lens. image.
L
P
3 Optical Instrumentation
INTRODUCTION
You should be familiar with ways to trace rays through an optical system
using the step-by-step application of Gaussian formulas and ray tracing.
However, not every light ray from an object point, directed toward or into an
optical system, reaches the final image. Depending on the location of the
object point and the ray angle, many of these rays are blocked by the limiting
apertures of lenses and mirrors or by physical apertures intentionally insert-
ed into the optical system. An aperture, in its broadest sense, is an opening
defined by a geometrical boundary. In this section, we wish to concentrate on
the effects of such spatial limitations of light beams in an optical system.
The apertures dealt with are often purposely inserted into an optical
system to achieve various practical purposes. Apertures can be used to
modify the effects of spherical aberration, astigmatism, and distortion. In
other applications, apertures may be introduced to produce a sharp border
to the image, like the sharp outline we see looking into the eyepiece of an
optical instrument. Apertures may also be used to shield the image from
50
Optical Instrumentation 51
AS Lens
L!
O! M!
Chie f ray
f ray F I Chie
O
I!
L
EnP ExP
(a)
AS
Lens
L!
O! N!
Chief ra
y M!
O Chief F
ray
M
L N
ExP
EnP
(b)
AS
Lens
O!
Chie
f ray I
O F
Chief
ray
I!
EnP
ExP
(c)
sees the lens directly but sees the AS through the lens. In other words, the
effective aperture due to the AS is its image formed by the lens, the dashed
line marked EnP. Since rays from O, directed toward this virtual aperture,
make a smaller angle 1N¿ON2 than rays directed toward the lens edge
1L¿OL2, this virtual aperture serves as the effective entrance pupil for the
system. Notice that rays from O, directed toward the edges of EnP, are in fact
first refracted by the lens so as to pass through the edges of the real aperture
stop. This must be the case, since AS and EnP are, by definition, conjugate
Optical Instrumentation 53
planes: The edges of EnP are the images of the edges of AS. This example il-
lustrates the general rule: The entrance pupil EnP is the image of the control-
ling aperture stop formed by the imaging elements preceding it.1 When the
controlling aperture stop is the first such element (a front stop), it serves itself
as the entrance pupil.
Another example, in which an aperture placed in front of the lens func-
tions as the AS for the system, is shown in Figure 1c. It is different from
Figure 1a in that the aperture is placed inside the focal length of the lens. Nev-
ertheless, the aperture is the AS for the system because it, not the lens, limits
the system rays to their smallest angle with the axis. Furthermore, it is the EnP
of the system because it is the first element encountered by the light from the
object.
Exit Pupil (E xP) We have described the EnP of an optical system as
the image of the AS one sees by looking into the optical system from the ob-
ject. If one looks into the optical system from the image, another image of the
AS can be seen that appears to limit the output beam size. This image is called
the exit pupil of the optical system. Thus, the exit pupil is the image of the con-
trolling aperture stop formed by the imaging elements following it (or to the
right of it in our figures). The rear stop in Figure 1b is automatically the ExP
for the system because it is the last optical component. According to our def-
inition of the ExP, the exit pupil is the optical conjugate of the AS; the ExP
and AS are conjugate planes. It follows that the ExP is also conjugate with
the EnP. In Figure 1a, the ExP is the real image of the EnP; in Figure 1c, it is
the virtual image. Notice that in each case, rays intersecting the edges of the
entrance pupil also (actually, or when extended) intersect the edges of the
exit pupil.
In a system like that of Figure 1a, a screen held at the position of the exit
pupil receives a sharp image of the circular opening of the aperture stop. If
the system represents the eyepiece of some optical instrument, the exit pupil
is matched in position and diameter to the pupil of the eye. Notice further
that if the screen is moved closer to the lens, it intercepts a sharp image II¿ of
the object OO¿. The exit pupil is seen to limit the solid angle of rays forming
each point of the image and therefore determines the image brightness, point
by point.
Chief Ray The chief, or principal, ray is a ray from an object point that
passes through the axial point, in the plane of the entrance pupil. Given the
conjugate nature of the entrance pupil with both the aperture stop and the exit
pupil, this ray must also pass (actually or when extended) through their axial
points. The chief ray in the cone of rays leaving object point O¿ is shown in all
three systems of Figure 1. The chief ray in the cone of rays leaving the axial
point O always coincides with the optical axis.
Before adding to our collection of new concepts that arise from a con-
sideration of apertures in optical systems, we consider a system slightly more
complex than those of Figure 1. In Figure 2, we specify a particular optical
system consisting of two lenses, L1 and L2, with an aperture A placed be-
tween them, as shown. The first question to be answered is: Which element
serves as the effective AS for the whole system? The answer to this question
is not always obvious. It can always be answered, however, by determining
which of the actual elements in the given system—in this case, A, L1, or L2—
has an entrance pupil that confines rays to their smallest angle with the axis, as
seen from the object point. To decide which candidate presents the limiting
aperture, it is necessary to find the entrance pupil for each by imaging each
one through that part of the optical system lying to its left:
1
“Preceding” is used in the sense that light must pass through those imaging elements first. If we
always use light rays directed from left to right, we can simply say, “by all imaging elements to its left.”
54 Chapter 3 Optical Instrumentation
A!1 A!2
A, AS
L2! L1 L2
O! b b!
Chief ray
a! ray
I Chief
O
a! I! b
a
b!
EnP ExP
Lens
T!
O O!
T
Object
EnP Image
plane
plane
(a)
Lens
A b
Chief ray
y
f ra
Chie U!
T!
O O!
T
U
V EnP Image
Object plane
plane
(b)
irradiance near the center. One can then also define the angular field of view
as twice the angle b made by the chief ray with the axis at the center of the
opening represented by the entrance pupil. The lens itself, in this case, acts
both as the field stop and the entrance window. In other, more complex, sys-
tems, the relevant quantities may be described as follows.
Field Stop (FS) The field stop is the aperture that controls the field of
view by limiting the solid angle formed by chief rays. As seen from the center
of the entrance pupil, the field stop (or its image) subtends the smallest angle.
When the edge of the field of view is to be sharply delineated, the field stop
should be placed in an image plane so that it is sharply focused along with the
final image. A simple example of such a field stop is the opening directly in
front of the film that outlines the final image in a camera. Intentional limita-
tion of the field of view using an aperture is desirable when either far-off axis
imaging is of unacceptable quality due to aberrations or when vignetting se-
verely reduces the illumination in the outer portions of the image.
Entrance Window (E nW) The entrance window is the image of the field
stop formed by all optical elements preceding it. The entrance window delin-
eates the lateral dimensions of the object to be viewed, as in the viewfinder of
a camera, and its angular diameter determines the angular field of view. When
the field stop is located in an image plane, the entrance window lies in the con-
jugate object plane, where it outlines directly the lateral dimensions of the ob-
ject field imaged by the optical system.
Exit Window (E xW) The exit window is the image of the field stop
formed by all optical elements following it. To an observer in image space, the
exit window seems to limit the area of the image in the same way as an out-
door scene appears limited by the window of a room.
In Figure 3c, the locations of the stops, pupils, and windows are shown in
a more complex optical system consisting of two lenses and two apertures.
The first aperture is the AS of the system and, as we have seen, is related to an
entrance pupil, its image in L1, and an exit pupil, its image in L2. The second
aperture is the field stop, FS, with its corresponding images through the lens-
es: the entrance window to the left and the exit window to the right. The field
of view in object space can then be described by a, the angle subtended by
the entrance window at the center of the entrance pupil. Similarly, the field of
view in image space can be described by a¿, the angle subtended by the exit
window at the center of the exit pupil. We see that the size of the field imaged
by the optical system is effectively determined by the entrance window and,
actually, by the size of the field stop. Notice that since EnW and ExW are both
images of the FS, they are conjugate planes. Thus, the same bundle of rays
that fills the entrance window also fills the field stop and the exit window.
The Summary of Terms that follows is provided as a convenient refer-
ence for a subject that requires patience, and experience with many exam-
ples, to master.
SUMMARY OF TERMS
Brightness
Aperture stop AS: The real element in an optical system that
limits the size of the cone of rays accepted by
the system from an axial object point.
Entrance pupil EnP: The image of the aperture stop formed by
the optical elements (if any) that precede it.
Exit pupil ExP: The image of the aperture stop formed by
the optical elements (if any) that follow it.
Field of view
Field stop FS: The real element that limits the angular field
of view formed by an optical system.
Optical Instrumentation 57
Entrance window EnW: The image of the field stop formed by the op-
tical elements (if any) that precede it.
Exit window ExW: The image of the field stop formed by the op-
tical elements (if any) that follow it.
The following example will provide practice with the thin-lens formula
1>s + 1>s¿ = 1>f and procedures for determining stops and pupils for a spe-
cific optical system. As you follow through the steps in the solution, be sure
to verify the correct use of the sign convention related to the object distance
s, image distance s¿, focal length f, and image magnification m = - s¿>s for
each calculation.
Example 1
An optical system (see sketch in Figure 4 below) is made up of a positive thin
lens L1 of diameter 6 cm and focal length f1 = 6 cm, a negative thin lens L2 of
diameter 6 cm and focal length f2 = - 10 cm, and an aperture A of diameter
3 cm. The aperture A is located 3 cm in front of lens L1 , which is located 4 cm
in front of lens L2 . An object OP, 3 cm high is located 18 cm to the left of L1 .
Problem
a. Determine which element (A, L1 , or L2) serves as the aperture stop AS.
b. Determine size and location of the entrance and exit pupils.
c. Determine the location and size of the intermediate image of OP
formed by L1 and the final image formed by the system.
d. Using a scale of 1 cm = 14 in., draw a diagram of the optical system and
locate to scale, on the drawing, the two pupils, intermediate image O¿P¿
and final image O–P–.
e. Draw the chief ray from object point P to its conjugate in the final
image, P–.
Solution
a. Determine first which element (A, L1 , or the image of L2 in L1) sub-
tends the smallest half-angle from rim to point O.
Elements A and L1 have no “optics” to their left, so each subtends a
half-angle directly:
For element A: uA M 1.5>15 = 0.1 rad
For element L1: uL1 M 3>18 = 0.17 rad
Optical System A L1 L2
P
f 1! 6 cm f 2! "10 cm
3 cm
1.5 cm
O F1, F2 F1 F2
3 cm 4 cm
18 cm
L!2
A
AS, EnP ExP
L1 L2
P
Chief f1 # 6 cm f 2# $10 cm
3 cm ray 3 cm
M N 1.5 cm O! O"
O F1, F2 F1 F2
P!
P"
3 cm 4 cm
12 cm
18 cm 10 cm
9 cm
rays from object points striking spherical surfaces at angles that exceed those
set by the paraxial approximation.
A brief description of spherical and chromatic aberration will serve us
here as a useful background for the treatment of optical elements and instru-
ments that conclude this chapter.
Spherical Aberration
In the paraxial ray approximation, all rays emanating from an object point,
after reflecting from a spherical mirror or passing through a lens with
spherical surfaces, either intersect at, or to the viewer appear to intersect at,
a common image point. In fact, rays emanating from an object point that
are incident on an optical element (spherical mirror or lens) at different
Mirror P
distances from the optical axis, after reflection or refraction, either inter-
sect, or to the viewer appear to intersect, at different positions. The result is Q
that point objects are not imaged as points but rather as small blurred lines.
The effect of spherical aberration for a concave spherical mirror that im-
M N
ages an axial object point is shown in Figure 6a. Note that rays incident on O
the mirror at symmetrically placed points Q converge at axial point M, Q
whereas rays incident on the mirror at points P converge at axial point N.
Consequently, light rays from a single point O form a blurred image along P
the line segment containing M and N. Figure 6b illustrates the effect of (a)
spherical aberration in a thin-lens system that forms a blurred line image of
P
an axial point object. In the case shown, rays from axial point object O that
encounter the lens at symmetrically placed points P converge at axial point Q
N, while rays from O that encounter the lens at points Q converge at axial M
point M. As in the spherical mirror case, the result is a blurred line image O N
Q
along the line segment containing M and N. The constructions shown in Fig-
ure 6a and 6b suggest that “stopping down” the optical system, so that only P
nearly paraxial rays get through the entrance pupil, will limit the effect of (b)
spherical aberration. This remedy for spherical aberration is, of course, ac-
Figure 6 Spherical aberration (exagger-
companied by a reduction in image brightness. Another remedy for spheri-
ated) in (a) a spherical mirror and (b) a thin
cal aberration is achieved by combining positive and negative lenses in an lens. For both arrangements, the rays from
arrangement such that the spherical aberration from one tends to cancel object point O fail to converge at a single
that from the other. image point.
60 Chapter 3 Optical Instrumentation
V R
V
R
fV TCA
fR
LCA
(a) (b)
Figure 7 Chromatic aberration (exaggerated) for (a) parallel rays of white light inci-
dent on a thin lens and (b) white light incident on a thin lens from an off-axis point P.
Chromatic Aberration
Chromatic aberration results because the index of refraction of a material
differs for different wavelengths. Since the focal length of a lens is depen-
dent on the index of refraction of the lens material, focal lengths and
image positions differ for different wavelength components of the light
used in the optical system. Thus, polychromatic light from a point object
images not as a point but as a series of points, one for each distinct wave-
length.
Figure 7 illustrates two related aspects of chromatic aberration. In
Figure 7a, parallel rays of light focus nearer the lens, at point V, for violet light
and further from the lens, at point R, for red light. Thus, white light coming
from a single distant object point fails to image as a single point. Rather, the
different wavelength components refract to form image points between the
focal lengths fV and fR , as indicated in Figure 7a. Figure 7b shows a slightly
different view of chromatic aberration evident in the behavior of white light
incident on a lens from an off-axis point P. The violet and red components of
the white light leaving point P refract differently at the lens and so converge
at different image points, labeled V and R, respectively. The amount of chro-
matic aberration, in such a case, can be described by two distances, shown in
Figure 7b, one called the longitudinal chromatic aberration (LCA) and the
other the lateral or transverse chromatic aberration (TCA).
Chromatic aberration in lenses can be effectively reduced by using mul-
tiple refractive elements of opposite powers. Of course, images formed in
mirrors do not suffer from chromatic aberration since the focal length of a
mirror is independent of wavelength.
3 PRISMS
d1 d2 u2
u1 u#2
u#1 B
due to the action of the prism as a whole is the sum of the angular deviations d1
and d2 at the first and second faces, respectively. Snell’s law at each prism face
requires that
Inspection will show that the following geometrical relations must hold be-
tween the angles:
d1 = u1 - u1œ (3)
d2 = u2 - u2œ (4)
B = 180 - u1œ - u2œ = 180 - A (5)
A = u1œ + u2œ (6)
The two members of Eq. (5) follow because the sum of the angles of a trian-
gle is 180° and because the sum of the angles of a quadrilateral must be 360°.
Notice that the angles A and B and the two right angles formed by the nor-
mals with the prism sides constitute such a quadrilateral.
Using Eqs. (1) through (6), a programmable calculator or computer
may easily be programmed to perform the sequential operations that finally
determine the angle of deviation, d. Assuming that the prism angle A and re-
fractive index n are given, then the stepwise calculation for a ray incident at
an angle u1 is as follows:
sin u1
u1œ = sin-1 a b (7)
n
d1 = u1 - u1œ (8)
u2œ = A - u1œ (9)
u2 = sin-11n sin u2œ 2 (10)
d = u1 + u2 - u1œ - u2œ (11)
The variation of deviation with angle of incidence for A = 30° and n = 1.50
is shown in Figure 10. Notice that a minimum deviation occurs for u1 = 23°.
Refraction by a prism under the condition of minimum deviation is most
often utilized in practice. We may argue rather neatly that when minimum de-
viation occurs, the ray of light passes symmetrically through the prism, mak-
ing it unnecessary to subscript angles, as shown in Figure 11. Suppose this
were not the case, and minimum deviation occurred for a nonsymmetrical case,
as in Figure 9. Then if the ray were reversed, following the same path back-
ward, it would have the same total deviation as the forward ray, which we
62 Chapter 3 Optical Instrumentation
40
35
25
20
dmin
15
23
Figure 10 Graph of total deviation versus
angle of incidence for a light ray through a
prism with A = 30° and n = 1.50. Minimum 20 40 60 80
deviation occurs for an angle of 23°. Angle of incidence (deg)
d
d/2
d/2 u
u
u! u!
B
Figure 11 Progress of a ray through a
prism under the condition of minimum
deviation.
d = 2u - 2u¿ (12)
and from Eq. (6),
A = 2u¿ (13)
A d + A
u¿ = and u = (14)
2 2
A + d A
sin a b = n sin a b
2 2
or
sin[1A + d2>2]
n = (15)
sin1A>22
Optical Instrumentation 63
1A + d2>2
n!
A>2
or
Dispersion
200 1600
The minimum deviation of a monochromatic beam through a prism is
l (nm)
given implicitly by Eq. (15) in terms of the refractive index. The refractive
index, however, depends on the wavelength, so that it would be better to
write nl for this quantity. As a result, the total deviation is itself wave-
length dependent, which means that various wavelength components of
the incident light are separated on refraction from the prism. A typical Red
White
normal dispersion curve and the nature of the resulting color separation Green
are shown in Figure 12. Notice that shorter wavelengths have larger re- Violet
fractive indices and, therefore, smaller speeds in the prism. Consequently, Figure 12 Typical normal dispersion curve
violet light is deviated most in refraction through the prism. The disper- and consequent color separation for white
sion indicated in the graph n versus l of Figure 12 is called “normal” dis- light refracted through a prism.
persion. When the refracting medium has characteristic excitations that
absorb light of wavelengths within the range of the dispersion curve, the
curve is monotonically decreasing, as shown, but has a positive slope in the
wavelength region of the absorption. When this occurs, the term anomalous
dispersion is used, although there is nothing anomalous about it. The nor-
mal dispersion curve shown is typical but varies somewhat for different
materials. An empirical relation that approximates the curve, introduced by
Augustin Cauchy, is
B C d
nl = A + 2
+ 4 + Á (17)
l l
sodium atoms in the sun’s outer atmosphere.2 Using the thin prism at min-
imum deviation for the D line, for example, the ratio of angular spread of the
F and C wavelengths to the deviation of the D wavelength, as suggested in
Figure 13, is
! nF - nC
=
d nD - 1
nF - nC
¢ = (18)
nD - 1
L
P
2
Because the yellow sodium D line is a doublet (589.0 and 589.6 nm), the more monochromatic
d line of helium at 587.56 nm is often preferred to characterize the center of the visible spectrum.
The green line of mercury at 546.07, which lies nearer to the center of the luminosity curve, is also
used.
Optical Instrumentation 65
prism in place. When the instrument is used for visual observations without
the capability of measuring the angular displacement of the spectral “lines,”
it is called a spectroscope. If means are provided for recording the spec-
trum, for example, with a photographic film in the focal plane of the tele-
scope objective, the instrument is called a spectrograph. When the prism is
made of some type of glass, its wavelength range is limited by the absorp-
tion of glass outside the visible region. To extend the usefulness of the spec-
trograph farther into the ultraviolet, for example, prisms made from quartz
1SiO22 and fluorite 1CaF22 have been used. Wavelengths extending further
into the infrared can be handled by prisms made of salt (NaCl, KCl) and
sapphire 1Al2O32.
Chromatic Resolving Power
If the wavelength difference between two components of the light incident
on a prism is allowed to decrease, the ability of the prism to resolve them
will ultimately fail. The resolving power of a prism spectrograph thus repre-
sents an important performance parameter, which we shall evaluate in this
section. Imagine two spectral lines formed on a photographic film in a
prism spectrograph. The lines are images of the slit, so that for precise
wavelength measurements the entrance slit should be kept as narrow as
possible consistent with the requirement of adequate illumination of the
film. Even with the narrowest of slit widths, however, the spectral line image
is found to possess a width, directly traceable to the limitation that the
edges of the collimating lens or prism face impose on the light beam. The
phenomenon is therefore due to the diffraction of light, treated later. Since
the line images have an irreducible width due to diffraction, as ¢l decreas-
es and the lines approach one another, a point is reached where the two
lines appear as one, and the limit of resolution of the instrument is realized.
No amount of magnification of the images can produce a higher resolution
or enhancement of the ability to discriminate between two such closely
spaced spectral lines.
Consider Figure 15a, in which a monochromatic parallel beam of light is
incident on a prism, such that it fills the prism face. Employing Fermat’s prin-
ciple, the ray FTW is isochronous with ray GX, since they begin and end on
the same plane wavefronts, GF and XW, respectively. Their optical paths can
be equated to give
FT + TW = nb
where b is the base of the prism and n is the refractive index of the prism, cor-
responding to the wavelength l. If a second neighboring wavelength compo-
nent l¿ is now also present in the incident beam, such that l¿ - l = ¢l,
the component l¿ will be associated with a different refractive index, n¿ =
n - ¢n. For normal dispersion, ¢n will be a small, positive quantity. The
T T %s
F W F W!
W %a
d d d %a
l& %l
G b X b l
(a) (b)
Figure 15 Constructions used to determine chromatic resolving power of a prism. (a) Refraction of
monochromatic light. (b) Refraction of two wavelength components separated by ¢l.
66 Chapter 3 Optical Instrumentation
emerging wavefronts for the two components, shown in Figure 15b, are thus
separated by a small angular difference, ¢a, and are accordingly focused at
different points in the focal plane of the telescope objective. Fermat’s princi-
ple, applied to the second component l¿, gives
FT + TW - ¢s = 1n - ¢n2b
¢s = b ¢n (19)
dn
¢s = b a b ¢l (20)
dl
Equation (20) now relates the path difference ¢s to the wavelength differ-
ence ¢l. The angular difference ¢a can also be introduced, using
¢s b dn
¢a = = a b a b ¢l (21)
d d dl
where d is the beam width. We appeal now to Rayleigh’s criterion, which de-
termines the limit of resolution of the diffraction-limited line images. This cri-
terion is explained and used in the later treatment of diffraction, where it is
shown that the minimum separation ¢a of the two wavefronts, such that the
images formed are just barely resolvable, is given by
l
¢a = (22)
d
l b dn
= a b a b ¢l
d d dl
l
1¢l2min = (23)
b1dn>dl2
l dn
" = = b (24)
1¢l2min dl
where we have incorporated Eq. (23). Since dispersion is limited by the glass,
prism resolving power might be improved by increasing the base b. However,
this technique soon requires impractically large and heavy prisms. The dis-
persion dn>dl may be calculated, for example, from the Cauchy formula for
the prism material, using Eq. (17).
Optical Instrumentation 67
Example 2
Determine the resolving power and minimum resolvable wavelength differ-
ence for a prism made from flint glass with a base of 5 cm.
Solution
With the help of Table 1 we can calculate an approximate average value of
¢n
the dispersion for l = 550 nm as
¢l
¢n nF - nD 1.7328 - 1.7205
= = = - 1.19 * 10-4 nm-1
¢l lF - lD 486 - 589
dn
" = ba b = 10.05 * 109 nm211.19 * 10-4 nm-12 = 5971
dl
The minimum resolvable wavelength difference in the region around 550 nm
is, then,
l 5550 Å
1¢l2min = = ! 1 Å
" 5971
Although grating spectrographs achieve higher resolving powers, they
are generally more wasteful of light. Furthermore, they produce higher-order
images of the same wavelength component, which can be confusing when in-
terpreting spectral records. These instruments are discussed later.
l1 l1 $ l
l
l2 $ l1
l2 % l
30
45
30
E
45
C
90
D
L
enters the prism at face AB and departs at face AD, making an angle of 90°
with the incident direction. The dashed lines are merely added to assist in ana-
lyzing the operation of the prism, a single structure. Of the incident wave-
lengths, only one will refract at the precise angle that conforms to the case of
minimum deviation, as shown, with the light rays parallel to the prism base
AE. At face BC, total internal reflection occurs to direct the light beam into
the prism section ACD, where it again traverses under the condition of mini-
mum deviation. Since the prism section BEC serves only as a mirror, the beam
passes effectively with minimum deviation through sections ABE and ACD,
which together constitute a prism of 60° apex angle. In use, the spectral line is
observed or recorded at F, the focal point of lens L. Thus, an observing tele-
scope may be rigidly mounted. Instead, the prism is rotated on its prism table
(or about an axis normal to the page), and as it rotates, various wavelengths in
the incident beam successively meet the condition of incidence angle for min-
imum deviation, producing the path indicated, with focus at F. The prism rota-
tion may be calibrated in terms of angle, or better, in terms of wavelength.
Reflecting Prisms
Total internally-reflecting prisms are frequently used in optical systems, both to
alter the direction of the optical axis and to change the orientation of an image.
Of course, prisms alone cannot produce images. When used in conjunction with
image-forming elements, the light incident on the prism is first collimated and
rendered normal to the prism face to avoid prismatic aberrations in the image.
Plane mirrors may substitute for the reflecting prisms, but the prism’s reflecting
faces are easier to keep free of contamination, and the process of total internal
reflection is capable of higher reflectivity. The stability in the angular relation-
ship of prism faces may also be an important advantage in some applications.
Some examples of reflecting prisms in use are illustrated in Figure 18. The
Porro prism, Figure 18d, consists of two right-angle prisms, oriented in such a
way that the face of one prism is partially revealed to receive the incident light
and the face of the second prism is partially revealed to output the refracted
light. The prism halves are separated in the figure to clarify its action. Images
are inverted in both vertical and horizontal directions by the pair, so that the
Porro prism is commonly used in binoculars to produce erect images.
Optical Instrumentation 69
(a) (b)
112.5 112.5
112.5 112.5
90
(c) (d)
Figure 18 Image manipulation by reflecting prisms. (a) Right-angle prism. (b) Dove
prism. (c) Penta prism; pentagonal cross section. (d) Porro prism.
4 THE CAMERA
The simplest type of camera is the pinhole camera, illustrated in Figure 19a.
Light rays from an object are admitted into a light-tight box and onto a pho-
tographic film through a tiny pinhole, which may be provided with any simple
means of shuttering, such as a piece of black tape. An image of the object is
projected on the back wall of the box, which is lined with a piece of film.
As stated earlier, an image point is determined ideally when every ray
from a given object point, each processed by the optical system, intersects at
the corresponding image point. A pinhole does no focusing and actually
blocks out most of the rays from each object point. Because of the smallness
of the pinhole, however, every point in the image is reached only by rays that
originate at approximately the same point of the object, as in Figure 19b.
Alternatively, every object point sends a bundle of rays to the screen, which
are limited by the small pinhole and so form a small circle of light on the (a)
screen, as in Figure 19a. The overlapping of these circles of light due to each
object point maps out an image whose sharpness depends on the diameter of
the individual circles. If they are too large, the image is blurred. Thus, as the
pinhole is reduced in size, the image improves in clarity, until a certain pin-
hole size is reached. As the pinhole is reduced further, the images of each ob-
ject point actually grow larger again due to diffraction, with consequent
degradation of the image. Experimentally, one finds that the optimum pin-
hole size is around 0.5 mm when the pinhole-to-film distance is around 25 cm.
The pinhole itself must be accurately formed in as thin an aperture as possi-
ble. A pinhole in aluminum foil, supported by a larger aperture, works well. (b)
The primary advantage of a pinhole camera (other than its elegant simplicity!)
is that, since there is no focusing involved, all objects near and far are in focus Figure 19 Imaging by a pinhole camera.
70 Chapter 3 Optical Instrumentation
Lens
on the screen. In other words, the depth of field of the camera is unlimited.
The primary disadvantage is that, since the pinhole admits so little of the
available light, exposure times must be long. The pinhole camera is not use-
ful in freezing the action of moving objects. The pinhole-to-film distance,
while not critical, does affect the sharpness of the image and the field of
view. As this distance is reduced, the angular aperture seen by the film is
larger, so that more of the scene is recorded, with corresponding decrease in
size of any feature of the scene. Also, the image circles decrease in size, pro-
ducing a clearer image.
If the pinhole aperture is opened sufficiently to accommodate a converg-
ing lens, we have the basic elements of the ordinary camera (Figure 20). The
most immediate benefits of this modification are (1) an increase in the
brightness of the image due to the focusing of all the rays of light from each
object point onto its conjugate image point and (2) an increase in sharpness
of the image, also due to the focusing power of the lens. The lens-to-film dis-
tance is now critical and depends on the object distance and lens focal
length. For distant objects, the film must be situated in the focal plane of the
lens. For closer objects, the focus falls beyond the film. Since the film plane is
fixed, a focused image is procured by allowing the lens to be moved farther
from the film, that is, by “focusing” the camera. The extreme possible posi-
tion of the lens determines the nearest distance of objects that can be han-
dled by the camera. “Close-ups” can be managed by changing to a lens with
shorter focal length. Thus, the focal length of the lens determines the subject
area received by the film and the corresponding image size. In general,
image size is proportional to focal length. A wide-angle lens is a short focal-
length lens with a large field of view. A telephoto lens is a long focal-length
lens, providing magnification at the expense of subject area. The telephoto
lens avoids a correspondingly “long” camera by using a positive lens, sepa-
rated from a second negative lens of shorter focal length, such that the com-
bination remains positive.
Also important to the operation of the camera is the size of its aperture,
which admits light to the film. In most cameras, the aperture is variable and is
coordinated with the exposure time (shutter speed) to determine the total ex-
posure of the film to light from the scene. The light power incident at the
image plane (irradiance Ee in watts per square meter) depends directly on
(1) the area of the aperture and inversely on (2) the size of the image. If, as in
Figure 21, the aperture is circular with diameter D and the energy of the light
is assumed to be distributed uniformly over a corresponding image circle of
diameter d, then
area of aperture D2
Ee r = 2 (25)
area of image d
Optical Instrumentation 71
D
d
The image size, as in Figure 21, is proportional to the focal length of the lens,
so we can write
D 2
Ee r a b (26)
f
The quantity f/D is the relative aperture of the lens (also called f-number or
f/stop), which we symbolize by the letter A,
f
A K (27)
D
but is, unfortunately, usually identified by the symbol f/A. For example, a lens
of 4-cm focal length that is stopped down to an aperture of 0.5 cm has a rela-
tive aperture of A = 4>0.5 = 8. This aperture is usually referred to by pho-
tographers as f/8. The irradiance is now
1
Ee r (28)
A2
Most cameras provide selectable apertures that sequentially change the irradi-
ance at each step by a factor of 2. The corresponding f-numbers, then, form a
geometric series with ratio 22 , as in Table 2. Larger aperture numbers corre-
spond to smaller exposures. Since the total exposure 1J>m22 of the film depends
on the product of irradiance A m2J # s B and time (s), a desirable total exposure may
be met in a variety of ways. Accordingly, if a particular film (whose speed is de-
scribed by an ISO number) is perfectly exposed by light from a particular scene
1
with a shutter speed of 50 s and a relative aperture of f/8, it will also be perfectly
exposed by any other combination that gives the same total exposure, for ex-
1
ample, by choosing a shutter speed of 100 s and an aperture of f/5.6. The change
in shutter speed cuts the total exposure in half, but opening the aperture to the
next f/stop doubles the exposure, leaving no change in net exposure.
A = f-number 1A = f-number22 Ee
1 1 E0
1.4 2 E0>2
2 4 E0>4
2.8 8 E0>8
4 16 E0>16
5.6 32 E0>32
8 64 E0>64
11 128 E0>128
16 256 E0>256
22 512 E0>512
72 Chapter 3 Optical Instrumentation
D
a
M!
M O N O! N! d
s0 s!0
s1
x x
Figure 22 Construction illustrating depth s2
of field MN. Object and image spaces are
not shown to the same scale.
D d
tan a ! and tan a !
s0œ x
so that
ds0œ
x! (29)
D
It is then required to find, from the lens equation, the object distance s1 cor-
responding to image distance s0œ + x and the object distance s2 corresponding
to image distance s0œ - x. After a moderate amount of algebra, one finds
s0f1f + Ad2
s1 = (30)
f2 + Ads0
s0f1f - Ad2
s2 = (31)
f2 - Ads0
Optical Instrumentation 73
2Ads01s0 - f2f2
depth of field = (32)
f4 - A2d2s20
Acceptable values of the circle diameter d depend on the quality of the pho-
tograph desired. A slide that will be projected or a negative that will be en-
larged requires better original detail and hence a smaller value for d. For
most photographic work, d is of the order of thousandths of an inch.
Example 3
A 5 cm focal length lens with an f/16 aperture is used to image an object 9 ft
away. The blurring diameter in the image is chosen to be d = 0.04 mm.
Problem
Determine the location of the near point 1s12, far point 1s22, and the depth
of field.
Solution
Based on Figure 22 and the given data, we have:
s0 = 9 ft M 275 cm d = 0.004 cm
f = 5 cm A = 16
Thus, the depth of field, MN in Figure 22, is about 25 ft for a 5-cm focal
length lens imaging an object 9 ft away. In effect this lens will image all ob-
jects from 5 ft to 30 ft with an acceptable sharpness.
Most cameras are equipped with a depth-of-field scale from which values of s1
and s2 can be read, once the object distance and aperture are selected. Ac-
cording to Eq. (32), depth of field is greater for smaller apertures (larger
f-numbers), shorter focal lengths, and longer shooting distances.
The camera lens is called upon to perform a prodigious task. It must pro-
vide a large field of view, in the range of 35° to 65° for normal lenses and as
large as 120° or more for wide-angle lenses. The image must be in focus and
reasonably free from aberrations over the entire area of the film in the focal
plane. The aberrations that must be reduced to an acceptable degree are, in ad-
dition to chromatic aberration, the five monochromatic aberrations: spherical
aberration, coma, astigmatism, curvature of field, and distortion. Since a correc-
tive measure for one type of aberration often causes greater degradation in the
image due to another type of aberration, the optical solution represents one of
many possible compromise lens designs. The labor involved in the design of a
suitable lens that meets particular specifications within acceptable limits has
been reduced considerably with the help of computer programming. Human
ingenuity is nevertheless an essential component in the design task, since there
74 Chapter 3 Optical Instrumentation
is more than one optical solution to a given set of specifications. The demands
made upon a photographic lens cannot all be met using a single element. Vari-
ous stages in solving the lens design problem are illustrated in Figure 23a, from
the single-element meniscus lens, which may still be found in simple cameras, to
the four-element Tessar lens. The use of a symmetrical placement of lenses, or
groups of lenses, with respect to the aperture is often a distinctive feature of such
lens designs. In such placements, one group may reverse the aberrations
introduced by the other, reducing overall image degradation due to factors
Cooke triplet
Tessar
Petzval
(a)
The simple magnifier is essentially a positive lens used to read small print, in
which case it is often called a reading glass, or to assist the eye in examining
small detail in a real object. It is often a simple convex lens but may be a dou-
blet or a triplet, thereby providing for higher-quality images.
Figure 24 illustrates the working principle of the simple magnifier. A small
object of dimension h, when examined by the unaided eye, is assumed to be held
at the near point of the normal eye—nearest position of distinct vision—at posi-
tion (a), 25 cm from the eye. At this position the object subtends an angle a0 at
the eye. To project a larger image on the retina, the simple magnifier is inserted
and the object is moved physically closer to position (b), where it is at or just
inside the focal point of the lens. In this position, the lens forms a virtual image
subtending a larger angle aM at the eye. The angular magnification3 of the sim-
ple magnifier is defined to be the ratio aM>a0 . In the paraxial approximation,
the angles may be represented by their tangents, giving
aM h>s 25
= =
a0 h>25 s
At the other extreme, if the virtual image is viewed at the nearpoint of the
eye, then s¿ = - 25 cm, and from the thin-lens equation,
25f
s =
25 + f
giving a magnification of
25
M = + 1 image at normal near point (34)
f
aM
h a0 h
(a) (b)
s
25 cm
Figure 24 Operation of a simple magnifier.
3
When viewing virtual images with optical instruments, the images may be at great distances,
even “at infinity,” when rays entering the eye are parallel. In such cases, lateral magnifications also ap-
proach infinity and are not very useful. The more convenient angular magnification is clearly a mea-
sure of the image size formed on the retina and is used to describe magnification when eyepieces are
involved, as in microscopes and telescopes.
76 Chapter 3 Optical Instrumentation
1 1 1 L
= + - (35)
f f1 f2 f1f2
where f1 and f2 represent the individual focal lengths of the pair. By the lens-
maker’s formula, for lenses made of the same glass,
1 1 1
= 1n - 12a - b = 1n - 12K1 (36)
f1 R11 R12
and
1 1 1
= 1n - 12a - b = 1n - 12K2 (37)
f2 R21 R22
1
= 1n - 12K1 + 1n - 12K2 - L1n - 122K1K2 (38)
f
d11>f2
= 0
dn
From Eq. (38),
d11>f2
= K1 + K2 - 2LK1K21n - 12 = 0
dn
4
Some longitudinal chromatic aberration remains because the principal planes of the system do not
coincide.
Optical Instrumentation 77
This condition is met, therefore, when the lenses are separated by the distance
1 1 1
L = c + d
2 K11n - 12 K21n - 12
This condition is valid independent of the lens shapes, leaving the choice of
shapes as latitude for compensating other aberrations.
Both the Huygens and Ramsden eyepieces, Figures 25 and 26, incorpo-
rate the design feature required by Eq. (39); that is, plano-convex lenses are
separated by half the sum of their focal lengths. In the diagram of Figure 25, the
focal length of the field lens, FL, is approximately 1.7 times the focal length of
the eye lens, or ocular, EL. The primary image “observed” by the eyepiece is
in this case a virtual object (VO) for the field lens, since its virtual position
falls between the lenses. The field lens then forms a real image (RI) that is
viewed by the eye lens. When the real image falls in the focal plane of the eye
lens, the magnified image is viewed at infinity by the eye located at the exit
pupil. Note that the Huygens eyepiece cannot be used as an ordinary magnifi-
er. If crosshairs or a reticle with a scale is used with the eyepiece to make pos-
sible quantitative measurements, then to be in focus with the image formed by
the ocular EL, the crosshairs must be placed in the focal plane of RI, conve- Huygenian eyepiece
niently attached to the field or aperture stop placed there (Figure 27). The Eye lens
image of the crosshairs does not share in the image quality provided by the
eyepiece as a whole, however, because the eye lens alone is involved in form-
ing the image. This is not a problem in the Ramsden eyepiece, Figure 26, in
Reticle
Retaining
RI VO ring
F2
EL Field lens
Exit
FL Field pupil Ramsden eyepiece
stop
Eye lens
Figure 25 Huygens eyepiece.
RO Field lens
Reticle
Retaining
EL ring
Field Exit
FL pupil
stop
Figure 27 Construction of Huygens and
Figure 26 Ramsden eyepiece. Ramsden eyepieces.
78 Chapter 3 Optical Instrumentation
which both the primary and intermediate images are located just in front of
the field lens. In this eyepiece, the lenses have the same focal length f and, ac-
cording to Eq. (39), are separated by f. Ideally, when the real object, RO, falls
at the position of the first lens, rays emerge from the eyepiece parallel to one
another, giving a virtual magnified image at infinity. Thus a reticle, the field
stop, and field lens are all essentially in the “same plane.” A disadvantage of
this arrangement is that the surface of the lens is then also in focus, including
dust and smudges. By using a lens separation slightly smaller than f, the reticle
is in focus at a position slightly in front of the lens, as shown in the ray diagram
of Figure 26 and in Figure 27. With a lens separation somewhat less than f,
however, the requirement on L that corrects for transverse chromatic aberra-
tion is somewhat violated. A modification of the Ramsden eyepiece that al-
most eliminates chromatic defects is the Kellner eyepiece, which replaces the
Ramsden eye lens with an achromatic doublet. Other eyepieces have also
been designed to achieve higher magnifications and wider fields.
Example 4
A Huygens eyepiece uses two lenses having focal lengths of 6.25 cm and 2.50
cm, respectively. Determine their optimum separation in reducing chromatic
aberration, their equivalent focal length, and their angular magnification
when viewing an image at infinity.
Solution
The optimum separation is given by
1 1 1 L 1 1 4.375
= + - = + -
f f1 f2 f1f2 6.25 2.50 16.25212.502
25 25
M = = = 7
f 3.57
In designing eyepieces, one usually desires an exit pupil that is not much
greater than the size of the pupil of the eye, so that radiance is not lost. Recall
that, in this instance, the exit pupil is an image of the entrance pupil as formed
by the ocular and that the ratio of entrance to exit pupil diameters equals the
magnification. Since the entrance pupil is determined by preceding optical el-
ements in the optical system (the diameter of the objective lens, in a simple
telescope), this requirement places a limit on the magnifying power of the
eyepiece and, thus, a lower limit on its focal length.
The important specifications of an eyepiece, assuming its aberrations
are within acceptable limits for a particular application, include the following:
6 MICROSCOPES
25
M = (40)
feff
where feff (in cm) is the effective focal length of the two lenses, separated by
a distance d, and given by Eq. (35).
1 1 1 d
= + - (41)
feff fo fe fofe
Eyepiece
Objective
O
I
ExP
EnP
f0 f0 L fe fe
s0 s#0
251fe + fo - d2
M = (42)
fofe
soœ d - fe - fo
= (43)
so fo
where we have used the fact that soœ = d - fe , evident in the diagram. Incor-
porating Eq. (43) into Eq. (42),
soœ 25
M = -a ba b (44)
so fe
showing that the total magnification is just the product of the linear magnifi-
cation of the objective 1soœ >so2 multiplied by the angular magnification of the
eyepiece 125>fe2 when viewing the final image at infinity. The negative sign in-
dicates an inverted image. Comparing Figure 28 with the geometry associated
with Newton’s equation for a thin lens, we see that the magnitude of the later-
al magnification is given by
hi soœ x¿ L
ƒmƒ = ` ` = ` ` = = (45)
ho so fo fo
since x¿ = L is the distance between the objective image and its second focal
point, as shown. The magnification of the microscope may then be expressed,
perhaps more conveniently, as
25 L
M = -a ba b (46)
fe fo
Example 5
A microscope has an objective of 3.8-cm focal length and an eyepiece of
5-cm focal length. If the distance between the lenses is 16.4 cm, find the mag-
nification of the microscope.
Solution
25 L 25 7.6
M = -a ba b = -a ba b = - 10, a magnification of 10*
fe fo 5 3.8
Numerical Aperture
To collect more light and produce brighter images, cones of rays from the
object, intercepted by the objective lens (usually the aperture stop), should
Optical Instrumentation 81
Oil Air
no a!o a!a
ao aa
ng Cover glass Figure 29 Microscope objective, illustrating
the increased light-gathering power of an oil-
O immersion lens.
N. A. = n sin a (47)
The numerical aperture is an invariant in object space, due to Snell’s law. That
is, in the case of air,
N. A. = ng sin aa = sin aaœ
and when an oil-immersion objective is used,
The maximum value of the numerical aperture when air is used is unity, but
when object space is filled with a fluid of index n, the maximum numerical
aperture may be increased up to the value of n. In practice, the limit is around
1.6. The numerical aperture is an alternative means of defining a relative
aperture or of describing how “fast” a lens is. As shown previously, image
brightness is inversely proportional to the square of the f-number. Here also,
image brightness is proportional to the square of the numerical aperture. The
numerical aperture is an important design parameter also because it limits
the resolving power and the depth of focus of the lens. The resolving power is
proportional to the numerical aperture, whereas the depth of focus is inverse-
ly proportional to the square of the numerical aperture. Most microscopes use
objectives with numerical apertures in the approximate range of 0.08 to 1.30.
Biological specimens are covered with a cover glass of 0.17- or 0.18-mm
thickness. For objectives with numerical apertures over 0.30, the cover glass has
increasing influence on the image quality, since it introduces a large degree of
spherical aberration when oil immersion is not involved. Thus, a biological
82 Chapter 3 Optical Instrumentation
7 TELESCOPES
Refracting Telescopes
Figures 31 and 32 show two refracting telescope types, producing, respective-
ly, inverted and erect images. The Keplerian telescope in Figure 31 is often
referred to as an astronomical telescope since inversion of astronomical
objects in the images produced creates no difficulties. The Galilean telescope,
illustrated in Figure 32, produces an erect image by means of an eyepiece of
negative focal length. In either case, nearly parallel rays of light from a distant
object are collected by a positive objective lens, which forms a real image in
its focal plane. The objective lens, being larger in diameter than the pupil of
the eye, permits the collection of more light and makes visible those point
sources such as stars that might otherwise not be detected. The objective lens
is usually a doublet, corrected for chromatic aberration. The real image
formed by the objective is observed with an eyepiece, represented in the fig-
ures as a single lens. This intermediate image, located at or near the focal
point of the ocular, serves as a real object (RO) for the ocular in the astro-
nomical telescope and a virtual object (VO) in the case of the Galilean tele-
scope. In either case, the light is refracted by the eyepiece in order to produce
parallel, or nearly parallel, light rays. An eye placed near the ocular views an
image at infinity but with an angular magnification given by the ratio of the
Optical Instrumentation 83
Final image
Exit pupil
(eyepoint)
Real intermediate
image
Exit pupil
of objective
Specimen
Condenser diaphragm
Field diaphragm
Light source
angles a¿>a, as shown. The object subtends the angle a at the unaided eye and
the angle a¿ at the eyepiece.
From the two right triangles formed by the intermediate image and the
optical axis, it is evident that the angular magnification is
a¿ fo
M = = - (48)
a fe
84 Chapter 3 Optical Instrumentation
Objective
Ocular
fe
fo
a RO
a a!
ExP
EnP
Objective
Ocular
ExP
VO
a a a!
fe
fo
E nP
The negative sign is introduced, as usual, to indicate that the image is invert-
ed in Figure 31, where fe 7 0, and is erect in Figure 32, where fe 6 0. In
either case, the length L of the telescope is given by
L = fo + fe (49)
Dex
me = (50)
Dobj
f fe
me = - = - (51)
x f0
where x is the distance of the object (objective lens) from the focal point of
the eyepiece, or f0 . Combining Eqs. (48), (50), and (51),
1 Dex
me = =
M Dobj
so that
Dobj
Dex = (52)
M
Thus, the diameter of the bundle of parallel rays filling the objective lens is
greater by a factor of M than the diameter of the bundle of rays that pass
through the exit pupil. It should be pointed out that the image is not, there-
fore, brighter by the same proportion, however, because the apparent size of
the image increases by the same factor M. The brightness of the image cannot
be greater than the brightness of the object; in fact, it is less bright due to in-
evitable light losses due to reflections from lens surfaces.
Binoculars (Figure 33) afford more comfortable telescopic viewing,
allowing both eyes to remain active. In addition, the use of Porro or other
prisms to produce erect final images also permits the distance between objec-
tive lenses to be greater than the interpupillary distance, enhancing the
stereoscopic effect produced by ordinary binocular vision. The designation
“6 * 30” for binoculars means that the angular magnification M produced is
6 * and the diameter of the objective lens is 30 mm. Using Eq. (52), we con-
clude that the exit pupil for this pair of binoculars is 5 mm, a good match for
the normal pupil diameter. For night viewing, when the pupils are somewhat
larger, a rating of 7 * 50, producing an exit pupil diameter of 7 mm, would be
preferable.
86 Chapter 3 Optical Instrumentation
Example 6
Determine the eye relief and field of view for the 6 * 30 binoculars just de-
scribed. Assume an objective focal length of 15 cm and a field lens (eyepiece)
diameter of 1.50 cm.
Solution
The focal length of the ocular is found from
fo 15
fe = - = - = 2.5 cm
M -6
The eye relief is the distance of the exit pupil from the eyepiece. Since the
exit pupil is the image of the objective formed by the eyepiece, the eye re-
lief is the image distance s¿, given by
sf Lfe 1fo + fe2fe 115 + 2.5212.52
s¿ = = = = = 2.92 cm
s - f L - fe 1fo + fe2 - fe 15
The angular field of view from the objective subtends both the object on
one side and the field lens of the eyepiece on the other. Thus, for objects at
a standard distance of 1000 yd,
h Df
u = =
s L
or
sDf 13000 ft211.502
h = su = = = 257 ft at 1000 yd
L 15 + 2.5
Reflection Telescopes
Larger-aperture objective lenses provide greater light-gathering power and
resolution. Large homogeneous lenses are difficult to produce without optical
defects, and their weight is difficult to support. These problems, as well as the
Optical Instrumentation 87
fs
fp fp fs
(a) (b)
fp fs
(c)
Aperture
(a)
Aperture
(b)
Schmidt
correcting plate
Aperture
(c) Figure 36 The Schmidt optical system.
PROBLEMS
1 An object measures 2 cm high above the axis of an optical the lens. Determine the position and size of the entrance
system consisting of a 2-cm aperture stop and a thin con- and exit pupils, as well as the image. Sketch the chief ray
vex lens of 5-cm focal length and 5-cm aperture. The object and the two extreme rays through the optical system,
is 10 cm in front of the lens and the stop is 2 cm in front of from the top of the object to its conjugate image point.
90 Chapter 3 Optical Instrumentation
f ! 5 cm
2 cm
2 cm
2 cm 5 cm
Object
AS
10 cm
Figure 37 Problem 1.
2 Repeat problem 1 for an object 4 cm high, with a 2-cm 6 Plot a curve of total deviation angle versus entrance angle
aperture stop and a thin convex lens of 6-cm focal length and for a prism of apex angle 60° and refractive index 1.52.
5-cm aperture. The object is 14 cm in front of the lens and the
7 A parallel beam of white light is refracted by a 60° glass
stop is 2.50 cm behind the lens.
prism in a position of minimum deviation. What is the angu-
3 Repeat problem 1 for an object 2 cm high, with a 2-cm aper- lar separation of emerging red 1n = 1.5252 and blue (1.535)
ture stop and a thin convex lens of 6-cm focal length and 5-cm light?
aperture. The object is 14 cm in front of the lens and the stop is
4 cm in front of the lens. 8 a. Approximate the Cauchy constants A and B for crown
and flint glasses, using data for the C and F Fraun-
4 An optical system, centered on an optical axis, consists of
hofer lines from Table 1. Using these constants and
(left to right)
the Cauchy relation approximated by two terms, cal-
1. Source plane culate the refractive index of the D Fraunhofer line
2. Thin lens L1 at 40 cm from the source plane for each case. Compare your answers with the values
3. Aperture A at 20 cm farther from L1 given in the table.
4. Thin lens L2 at 10 cm farther from A b. Calculate the dispersion in the vicinity of the Fraunhofer D
5. Image plane line for each glass, using the Cauchy relation.
Lens L1 has a focal length of 40/3 cm and a diameter of 2 cm; c. Calculate the chromatic resolving power of crown and
lens L2 has a focal length of 20/3 cm and a diameter of flint prisms in the vicinity of the Fraunhofer D line, if each
2 cm; aperture A has a centered circular opening of 0.5-cm prism base is 75 mm in length. Also calculate the mini-
diameter. mum resolvable wavelength interval in this region.
a. Sketch the system. 9 An equilateral prism of dense barium crown glass is used in
b. Find the location of the image plane. a spectroscope. Its refractive index varies with wavelength,
c. Locate the aperture stop and entrance pupil. as given in the table:
d. Locate the exit pupil.
e. Locate the field stop, the entrance window, and the exit nm n
window.
f. Determine the angular field of view. 656.3 1.63461
587.6 1.63810
5 Refer back to the extended example in the text, involving 486.1 1.64611
both a positive and a negative lens, of focal lengths 6 cm and
-10 cm, respectively. For the identical optical system, al- a. Determine the minimum angle of deviation for sodium
ready partially analyzed, light of 589.3 nm.
a. Determine the location and size of the field stop, FS. b. Determine the dispersive power of the prism.
b. Determine the location and size of the entrance and exit c. Determine the Cauchy constants A and B in the long
windows. wavelength region; from the Cauchy relation, find the
c. Using the chief ray from object point P to image point dispersion of the prism at 656.3 nm.
P– as shown in the example, draw the two marginal rays d. Determine the minimum base length of the prism if it is to
from P to P–, which, with the chief ray, define the cone resolve the hydrogen doublet at 656.2716- and 656.2852-
of light that successfully gets through the optical system. nm wavelengths. Is the project practical?
Optical Instrumentation 91
10 A prism of 60° refracting angle gives the following an- points in the first and third rows is to be kept smaller
gles of minimum deviation when measured on a spec- than a typical silver grain of the emulsion, say 1 mm. At
trometer: C line, 38°20¿; D line, 38°33¿; F line, 39°12¿. what object distance nearer and farther than the middle
Determine the dispersive power of the prism. row does an unacceptable blur occur if the camera has
a focal length of 50 mm and is stopped down to an f/4
11 The refractive indices for certain crown and flint glasses are
setting?
Crown: nC = 1.527, nD = 1.530, nF = 1.536 16 A telephoto lens consists of a combination of two thin lens-
Flint: nC = 1.630, nD = 1.635, nF = 1.648 es having focal lengths of +20 cm and -8 cm, respectively.
The two glasses are to be combined in a double prism that is The lenses are separated by a distance of 15 cm. Determine
a direct-vision prism for the D wavelength. The refracting the focal length of the combination, distance from negative
angle of the flint prism is 5°. Determine the required angle lens to film plane, and image size of a distant object sub-
of the crown prism and the resulting angle of dispersion be- tending an angle of 2° at the camera.
tween the C and the F rays. Assume that the prisms are thin 17 A 5-cm focal length camera lens with f/4 aperture is focused
and the condition of minimum deviation is satisfied. on an object 6 ft away. If the maximum diameter of the cir-
12 An achromatic thin prism for the C and F Fraunhofer lines is cle of confusion is taken to be 0.05 mm, determine the
to be made using the crown and flint glasses described in depth of field of the photograph.
Table 1. If the crown glass prism has a prism angle of 15°, de- 18 The sun subtends an angle of 0.5° at the earth’s surface,
termine (a) the required prism angle for the flint glass and where the irradiance is about 1000 W>m2 at normal inci-
(b) the resulting “mean” deviation for the D line. dence. What is the irradiance of an image of the sun formed
13 A perfectly diffuse, or Lambertian, surface has the form of a by a lens with diameter 5 cm and focal length 50 cm?
square, 5 cm on a side. This object radiates a total power of 19 a. A camera uses a convex lens of focal length 15 cm. How
25 W into the forward directions that constitute half the total large an image is formed on the film of a 6-ft-tall person
solid angle of 4p. A camera with a 4-cm focal length lens and 100 ft away?
stopped down to f/8 is used to photograph the object when it b. The convex lens is replaced by a telephoto combination
is placed 1 m from the lens. consisting of a 12-cm focal length convex lens and a con-
a. Determine the radiant exitance, radiant intensity, and ra- cave lens. The concave lens is situated in the position of
diance of the object. the original lens, and the convex lens is 8 cm in front of it.
b. Determine the radiant flux delivered to the film. What is the required focal length of the concave lens
c. Determine the irradiance at the film. such that distant objects form focused images on the
same film plane? How much larger is the image of the
14 Investigate the behavior of Eq. (32), giving the dependence person using this telephoto lens?
of the depth of field on aperture, focal length, and object dis-
tance. With the help of a calculator or computer program, 20 The lens on a 35-mm camera is marked “50 mm, 1:1.8.”
generate curves showing each dependence.
a. What is the maximum aperture diameter?
15 A camera is used to photograph three rows of students at b. Starting with the maximum aperture setting, supply the
a distance 6 m away, focusing on the middle row. Suppose next three f-numbers that would allow the irradiance to
that the image defocusing or blur circles due to object be reduced to 13 the preceding at each successive stop.
Lens
Sun dL ! 5 cm
Image
fL ! 50 cm
c. What aperture diameters correspond to these f-numbers? distant object. If the telescope is focused on a telephone
1
d. If a picture is taken at maximum aperture and at 100 s, pole 30 m away, how much of the post falls between mil-
what exposure time at each of the other openings pro- limeter marks on the graticule? The focal length of the ob-
vides equivalent total exposures? jective is 20 cm.
27 A pair of binoculars is marked “7 * 35.” The focal length
21 The magnification given by Eq. (33) is also valid for a dou-
of the objective is 14 cm, and the diameter of the field lens
ble-lens eyepiece if the equivalent focal length given by
of the eyepiece is 1.8 cm. Determine (a) the angular magni-
Eq. (35) is used. Show that the magnification of a double-
fication of a distant object, (b) the focal length of the ocular,
lens eyepiece, designed to satisfy the condition for the elim-
(c) the diameter of the exit pupil, (d) the eye relief, and
ination of chromatic aberration, is, for an image at infinity,
(e) the field of view in terms of feet at 1000 yd.
1 1 28 a. Show that when the final image is not viewed at infinity,
M = 12.5a + b
f1 f2 the angular magnification of an astronomical telescope
may be expressed by
22 A magnifier is made of two thin plano-convex lenses, each of
3-cm focal length and spaced 2.8 cm apart. Find (a) the equiva-
mocfobj
lent focal length and (b) the magnifying power for an image M = -
formed at the near point of the eye. s–
23 The objective of a microscope has a focal length of 0.5 cm where moc is the linear magnification of the ocular and
and forms the intermediate image 16 cm from its second s– is the distance from the ocular to the final image.
focal point. b. For such a telescope using two converging lenses with
a. What is the overall magnification of the microscope focal lengths of 30 cm and 4 cm, find the angular magni-
when an eyepiece rated at 10* is used? fication when the image is viewed at infinity and when
b. At what distance from the objective is a point object the image is viewed at a near point of 25 cm.
viewed by the microscope?
29 The moon subtends an angle of 0.5° at the objective lens of
24 A homemade compound microscope has, as objective and an astronomical telescope. The focal lengths of the objective
eyepiece, thin lenses of focal lengths 1 cm and 3 cm, respec- and ocular lenses are 20 cm and 5 cm, respectively. Find the
tively. An object is situated at a distance of 1.20 cm from the diameter of the image of the moon viewed through the tele-
objective. If the virtual image produced by the eyepiece is scope at near point of 25 cm.
25 cm from the eye, compute (a) the magnifying power of 30 An opera glass uses an objective and eyepiece with focal
the microscope and (b) the separation of the lenses. lengths of + 12 cm and -4.0 cm, respectively. Determine the
25 Two thin convex lenses, when placed 25 cm apart, form a length (lens separation) of the instrument and its magnify-
compound microscope whose apparent magnification is ing power for a viewer whose eyes are focused (a) for infin-
20. If the focal length of the lens representing the eyepiece ity and (b) for a near point of 30 cm.
is 4 cm, determine the focal length of the other. 31 An astronomical telescope is used to project a real image of
26 A level telescope contains a graticule—a circular glass on the moon onto a screen 25 cm from an ocular of 5-cm focal
which a scale has been etched—in the common focal plane length. How far must the ocular be moved from its normal
of objective and eyepiece so that it is seen in focus with a position?
Screen
f # 5 cm
Obj Ocular 25 cm
32 a. The Ramsden eyepiece of a telescope is made of two pos- focal lengths, as it is for a refracting telescope when the
itive lenses of focal length 2 cm each and also separated image is formed at infinity.
by 2 cm. Calculate its magnifying power when viewing an 34 The primary mirror of a Cassegrain reflecting telescope has
image at infinity. a focal length of 12 ft. The secondary mirror, which is con-
b. The objective of the telescope is a 30-cm positive lens, with vex, is 10 ft from the primary mirror along the principal
a diameter of 4.50 cm. Calculate the overall magnification axis and forms an image of a distant object at the vertex of
of the telescope. the primary mirror. A hole in the primary mirror permits
c. What is the position and diameter of the exit pupil? viewing the image with an eyepiece of 4-in. focal length,
d. The diameter of the eyepiece field lens is 2 cm. Deter- placed just behind this mirror. Calculate the focal length of
mine the angle defining the field of view of the telescope. the secondary convex mirror and the angular magnification
of the instrument.
33 Show that the angular magnification of a Newtonian re-
flecting telescope is given by the ratio of objective to ocular
Lens
Mirror
am h ao
C Fo , Fe
ao
fe fo
Object
M1(f # 12!)
M2
f # 4"
2! 10!
12!