chapter · next room. you can easily hear the voices from the room because the sound waves bend...

26
Wave Optics W hy is the sky blue? What causes the beautiful colors in a soap bubble or an oil film? Why are clouds and ocean surf white, though both are formed by tiny drops of clear, colorless water? Why do Polaroid sunglasses reduce reflected glare? In this chapter we shall answer these questions. Additionally, we shall see how two light beams can combine to produce darkness, we shall show how to measure the wavelength of light using a meter stick, and we shall see why the magnification of any optical microscope is limited by the wave properties of light. We begin by giving a brief qualitative introduction to the wave phenomena of polar- ization, diffraction, and interference before returning to a more quantitative discussion of each. In our description of various experiments we shall often use a laser as our light source because of its wonderfully simple properties. 731 C . HAPTER 26 Return to Table of Contents Colors produced by a thin layer of oil on the surface of water result from constructive and destructive interference of light.

Upload: doankhuong

Post on 21-Aug-2018

213 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: CHAPTER · next room. You can easily hear the voices from the room because the sound waves bend around the doorway. ... 1 * CHAPTER 26 Wave Optics *,

Wave Optics

W hy is the sky blue? What causes the beautiful colors in a soap bubble or an oil

film? Why are clouds and ocean surf white, though both are formed by tiny

drops of clear, colorless water? Why do Polaroid sunglasses reduce reflected glare? In this

chapter we shall answer these questions. Additionally, we shall see how two light beams

can combine to produce darkness, we shall show how to measure the wavelength of

light using a meter stick, and we shall see why the magnification of any optical microscope

is limited by the wave properties of light.

We begin by giving a brief qualitative introduction to the wave phenomena of polar-

ization, diffraction, and interference before returning to a more quantitative discussion of

each. In our description of various experiments we shall often use a laser as our light

source because of its wonderfully simple properties.

731

C.HAPTER26Return to Table of Contents

Colors produced by a thin layer ofoil on the surface of water resultfrom constructive and destructiveinterference of light.

Page 2: CHAPTER · next room. You can easily hear the voices from the room because the sound waves bend around the doorway. ... 1 * CHAPTER 26 Wave Optics *,

Fig. 26–2 A polarized laser beam.

Fig. 26–3 Cross section of an unpolar-ized light beam.

732 CHAPTER 26 Wave Optics

Fig. 26–1 The electric field in a linearly polarized, plane, monochromatic wave.

Wave Properties of LightPolarizationThe simplest kind of light wave is a plane, monochromatic wave, which is linearlypolarized (Fig. 26–1). Linear polarization means that the electric field vector isalways directed parallel to a single line (the y-axis in the figure). Fig. 26–1 showsonly the electric field. An associated magnetic field is parallel to the z-axis and oscil-lates in phase with the electric field. Values of the electric field are shown at a partic-ular instant of time for various points along the x-axis. In a plane wave the value of theelectric field is the same along any plane perpendicular to the direction of the wave’smotion. The figure shows plane wavefronts, along which the electric field is maximum.Rays show the direction of motion of the wave—along the x-axis.The plane wave described in Fig. 26–1 is approximated by a section of a polarized

laser beam (Fig. 26–2). Within the beam, rays are approximately parallel, and wave-fronts are approximately cross sections of the beam.* By turning the laser on its side,rotating it 90�, we can produce a wave linearly polarized in the horizontal direction,rather than in the vertical direction. By rotating the laser through some other angle wecan get polarization in any direction perpendicular to the beam.Most natural light sources and many lasers have random polarization. This means

that at a given instant the electric field at any point in the wave is just as likely to bedirected along any line perpendicular to the direction of motion. The light is then saidto be unpolarized, and we indicate this state as shown in Fig. 26–3.

Frequency BandwidthsAny real source of light is not exactly monochromatic; that is, there is never just oneprecise value of frequency. Instead there is a range or band of frequencies, which maybe wide or narrow. The narrower the band, the more nearly the wave approximates amonochromatic wave. Laser light is nearly monochromatic. A common helium-neonlaser emits light at a frequency of 4.74 � 1014 Hz with a bandwidth of about 108 Hz.This means that the frequency range is less than 1 part in 106. More expensive,frequency-stabilized lasers have bandwidths as low as 104 Hz. Some lasers have evenachieved a stabilized frequency range of less than 100 Hz.

*There is some slight spreading of the beam, and hence the beam is not exactly a plane wave. A plane waveis also approximated by a small section of a spherical wave from a distant point source.

26–1

Page 3: CHAPTER · next room. You can easily hear the voices from the room because the sound waves bend around the doorway. ... 1 * CHAPTER 26 Wave Optics *,

Frequency band widthstypical of various kinds of light

Fig. 26–4 What phase of the cycleoccurs at a time �t after t 0?

Light �f (Hz)

Stabilized He-Ne laserCommon He-Ne laserSpectral lineWhite light

104

108

109

1014

Table 26–1

73326–1 Wave Properties of Light

By way of comparison, spectral lines emitted by various gas discharge tubes typicallyhave bandwidths of roughly 109 Hz, and white light, ranging in frequency from 4 �1014 Hz to 7 � 1014 Hz, has a bandwidth of 3 � 1014 Hz. The order of magnitude ofthese bandwidths is summarized in Table 26–1.

CoherenceIt is often important to be able to predict the relationship between the phase of a lightwave at two different times at the same point in space. For example, suppose that atsome instant t0, at one point in a laser beam, the electric field vector has its maximumvalue; that is, you are at a peak in the wave. At a time �t later, will the wave again haveits peak value or will it have some other value (Fig. 26–4)? If the laser light were trulymonochromatic, the solution would be easy. We could simply determine the exactnumber of cycles elapsed in a given time interval �t by multiplying the frequency (thenumber of cycles per second) by the time interval �t. (If the result were a wholenumber, the wave would again be at a peak. Or if the result were a whole number plus*1

4*, the electric field would be zero at that instant.) However, even laser light is notexactly monochromatic. There is always some frequency range f to f �f. Thenumber of cycles per second during any particular time interval can be anywhere inthis range. So the number of cycles completed during a time interval �t is somewherein the range f �t to ( f �f ) �t. If the time interval �t is small enough, the product �f�t will be much less than 1 cycle, and there will be little uncertainty in the number ofcycles completed, or in the final phase of the cycle. We can then predict the final phasefrom the initial phase, and we say that the wave is coherent over the time interval �t.This means that there is a definite, predictable phase relationship. The condition forcoherence then is that the time interval be small enough that

�f �t �� 1

or that

�t �� (condition for coherence) (26–1)1*�f

EXAMPLE 1 Coherence of Light Sources

Will you have coherence over a time interval of 10�6 s for lightfrom (a) a gas discharge tube; (b) a stabilized He-Ne laser?

SOLUTION (a) For a single line from a gas discharge tube,we see from Table 26–1 that even for a single spectral line�f � 109 Hz. Thus 1/�f � 10�9 s, and the time interval �t �10�6

s is much too long to satisfy Eq. 26–1, since 10�6 s is certainlynot less than 10�9 s. Thus this kind of light is not coher ent oversuch a time interval. The number of cycles completed is uncer-tain by �f �t � (109 s�1)(10�6 s) � 103 cycles, and so there is noability to predict the phase over such a time interval.

(b) For light from a stabilized He-Ne laser, �f � 104 Hz. Thusthe condition that must be satisfied for such light is

�t ��

or

�t �� 10�4 s

A time interval of 10�6 s satisfies this condition. Thus the laserlight is coherent over this time interval. Notice that �f�t �

(104 s�1)(10�6 s) � 10�2, and so we know the number of elapsedcycles with an uncertainty of only one hundredth of a cycle.

1��f

This book is licensed for single-copy use only. It is prohibited by law to distribute copies of this book in any form.

Page 4: CHAPTER · next room. You can easily hear the voices from the room because the sound waves bend around the doorway. ... 1 * CHAPTER 26 Wave Optics *,

Fig. 26–5 During a time interval �t a wave peak advances a distance �x c �t. Thus com par -ing the phases at two points a distance �x apart at a fixed time is the same as comparing thephases at the same point in space over a time interval �t.

Each wavefront in a light wave advances at the speed of light. Therefore, we canrelate the coherence of light at a fixed point in a plane wave at two different times tocoherence at two different points in a plane wave at the same time. As illustrated in Fig.26–5, during a time interval �t a wavefront advances a distance �x c �t, and socomparison of phases at two points �x apart is equivalent to comparing the phases ata fixed point over a time interval �t. Since �t�� 1/� f is the condition for coherence,two points in a wavefront will be coherent if

�x ��

The distance c /� f is called the coherence length, denoted by xc.

xc (26–2)

The condition for coherence may be expressed in terms of xc:

�x �� xc (coherence condition) (26–3)

For a stabilized He-Ne laser with a frequency range of 104 Hz, we find

xc 3 � 104 m

Two points in the laser beam have a predictable phase relationship, as long as they aremuch less than 30,000 m apart! For a common laboratory He-Ne laser, the bandwidthis of the order of 108 Hz. Thus

xc 3 m

The two points must be much closer than 3 m. Certainly two points a few cm apart arecoherent. For white light,

xc 3 � 10–6 m

Thus points in a beam of white light must be considerably less than a thousandth of amillimeter apart to be coherent.

c*�f

3 � 108 m/s**1014 Hz

c*�f

3 � 108 m/s**108 Hz

c*�f

3 � 108 m/s**104 Hz

c*�f

c*�f

734 CHAPTER 26 Wave Optics

Page 5: CHAPTER · next room. You can easily hear the voices from the room because the sound waves bend around the doorway. ... 1 * CHAPTER 26 Wave Optics *,

(a)

(b)

Fig. 26–6 (a) Little diffraction isproduced by a slit wider than 1 mm.(b) Considerable diffraction is producedby a slit less than 0.1 mm wide.

Fig. 26–7 Cross section of a sphericalwave, resulting from diffraction of lightby a circular hole with a diameter muchsmaller than the wavelength of the light.

Fig. 26–8 Diffraction of water waves.

26–1 Wave Properties of Light 735

DiffractionSuppose you are standing behind an open doorway, listening to a conversation in thenext room. You can easily hear the voices from the room because the sound wavesbend around the doorway. This phenomenon is called diffraction. It is a propertycommon to all waves to bend or diffract around an obstacle. However, the amount ofbending depends on the wavelength of the wave and the dimensions of the obstacle. Ingeneral, the longer the wavelength, the greater is the diffraction. Light, with its rela-tively short wavelength, bends or diffracts very little around an open doorway, butsound waves with their much longer wavelengths, diffract a great deal. Thus you canhear the conversation though you cannot see those who are talking.Seeing diffraction of light requires careful observation. Suppose we pass an intense

beam of light through a narrow slit in an opaque screen and project it onto a whitescreen (Fig. 26–6). If the slit is relatively wide, (say, at least a millimeter), we get animage of the slit on the screen (Fig. 26–6a). As predicted by geometrical optics, therays passing through the slit travel straight to the screen. But if we make the slit verynarrow (say, less than about 0.1 mm), the image on the screen actually gets wider (Fig.26–6b), violating the prediction of geometrical optics. We find that the narrower wemake the slit, the more the light bends outward. Of course if we make the slit muchless than 0.1 mm, there will be too little light to be seen, even if the light source illu-minating the slit is very intense. But if we could make a slit with a width much lessthan the wavelength of light and still have enough light intensity to see the smallamount of light passing through, we would see the light spread out in all directions,forming a cylindrical wave. Or if we replaced the slit by a tiny circular hole, with adiameter much less than the wavelength of light, we would produce a spherical wave.This result, illustrated in Fig. 26–7, is the essence of the Huygens-Fresnel prin-ciple, according to which each section of wavefront in the diffracting aperture isthe source of a spherical wave. Fig. 26–8 shows a photograph, illustrating this prin-ciple for water waves. The waves in Fig. 26–8 are incident on an aperture muchsmaller than the wavelength.Diffraction of light was first observed and recorded by the Jesuit priest Francesco

Grimaldi, a contemporary of Newton. Grimaldi observed the spreading of a narrowbeam of sunlight entering a darkened room. Geometrical optics, which was then basedon a picture of light consisting of particles,* could not account for this phenomenon, andso Grimaldi proposed that light is a wave. Grimaldi’s idea was rejected by Newton, whobelieved that if light were a wave, the diffraction effect would be much greater thanobserved. It must have seemed unlikely to Newton that light could have the incrediblysmall wavelength necessary to explain such a small amount of diffraction. Newton’sauthority was so great that the wave theory was not accepted for another 200 years.

InterferenceLike sound waves or waves on a string, light waves can interfere constructively ordestructively. Constructive interference occurs when two light waves are in phase, anddestructive interference occurs when two light waves are 180� out of phase. The colorsin the photograph of the oil slick at the beginning of this chapter result from interferenceof the light reflected from the top and bottom surfaces of the thin film of oil. The thick -ness of the film varies, and, as a result, different colors of light interfere constructively.

*In the twentieth century it was discovered that light does indeed have some particle-like properties.However this modern idea of a photon as a “particle” of light refers to emission or absorption of light, notto the way it propagates. Light travels as a wave, not as a bunch of particles, contrary to Newton’s belief. Thedual character of light as wavelike (in transmission) and particle-like (in absorption and emission) is at theheart of modern quantum physics, to be discussed in Chapter 28.

Page 6: CHAPTER · next room. You can easily hear the voices from the room because the sound waves bend around the doorway. ... 1 * CHAPTER 26 Wave Optics *,

(a) Constructive interference

(b) Destructive interference

Fig. 26–9 Two light waves caninterfere either (a) constructivelyor (b) destructively.

Fig. 26–11 An interference pattern.

736 CHAPTER 26 Wave Optics

For example, where the film appears blue the thickness is such that blue light reflectedfrom the two surfaces interferes constructively, while red light interferes destruc-tively. In the sections that follow we shall investigate interference and diffractionphenomena quantitatively.

InterferenceFig. 26–9a illustrates constructive interference of light, which occurs when twolight waves are in phase at a certain point in space over a period of time. Fig. 26–9bshows destructive interference, which occurs at a point in space where the waves are180� out of phase over a period of time. If the two waves are of equal amplitude and180� out of phase, the presence of two sources of light actually produces darkness! Fig.26–10 illustrates the intensity of light that is seen at a given point in space where twowaves of equal amplitude interfere either constructively or destructively. Since theintensity of a wave is proportional to the square of its amplitude, constructive inter-ference of two equal-amplitude waves produces in the resultant wave 2 times theamplitude or 4 times the intensity of the individual waves. Whether the interference isconstructive or destructive at a given point in space depends on the position of the pointrelative to the sources of light. Interference of light typically produces a pattern of lightand dark areas (Fig. 26–11).

(a) (b) (c)

Fig. 26–10 Light seen on a screen at a point in space where: (a) There is one light waveof intensity I; (b) There are two light waves, each of intensity I, interfering constructively andproducing a total intensity 4I; (c) There are two light waves, each of intensity I, interferingdestructively and producing no light.

In order for the eye to perceive interference of light, there must be a definite phaserelationship between the two waves over a time interval that the eye can detect. The eyehas a response time on the order of of a second. Thus interference effects must bestable for at least this long to be visible. This is longer than the coherence time of eventhe most monochromatic sources available today. The relative phases of two inde-pendent light sources (say, two different lasers) will vary randomly over time intervalsgreater than the coherence time. Thus, if we illuminate an area with two differentsources, the interference of their light waves at a given point will rapidly oscillate fromconstructive to destructive, and so no interference pattern is visible. All one sees is auniform illumination equal to the sum of the two intensities. For example, two equal-amplitude waves from separate sources produce instantaneous intensities rapidlyoscillating between 0 and 4 times each wave’s intensity I, at each point in space. Thusone sees only the time average of the instantaneous intensity, which at all points is thesame: twice the intensity of each source’s wave, since the average of 0 and 4I equals2 I. With the present state of technology, it is impossible to see interference of lightfrom two independent light sources.*

*It is possible to detect electronically interference of two independent sources, as demonstrated by Brownand Twiss in 1952.

26–2

1*20

Page 7: CHAPTER · next room. You can easily hear the voices from the room because the sound waves bend around the doorway. ... 1 * CHAPTER 26 Wave Optics *,

(a)

(b)

Fig. 26–12 (a) Overlapping wavefrontsinterfere constructively at certain pointsand destructively at others. (Figure isnot drawn to scale.) (b) Photographof interference fringes from double slitsilluminated with a He-Ne laser.

73726–2 Interference

Interference effects are easily observable when a single wavefront is divided intotwo separate parts, which then follow separate paths to the point where interference isobserved. Since the two waves arise from a common wavefront, changes in phase arecommon to each, and the interference pattern remains constant.

Young’s Double SlitOne of the simplest ways to produce interfering light is to use a double-slit arrange-ment like the one studied by the English physician Thomas Young in 1801.* If a mono-chromatic plane wave is incident on a pair of thin, closely spaced slits, the two slitsserve as sources of coherent light. The slits must be narrow enough and close enoughthat there is a significant amount of diffraction and overlap of the two wavefronts. Asillustrated in Fig. 26–12, what is seen on a screen in front of the slits is a pattern ofalternating light and dark fringes. Fig. 26–13 shows how the location of the fringesrelates to the distance to each of the slits. Point P is equidistant from the two slits, andso the two waves are exactly in phase at this point, and therefore point P is at the centerof an interference maximum—a bright fringe. The first dark fringe above the centralbright fringe is at a point Q, which is one-half wavelength farther from slit 2 than fromslit 1. The next bright fringe is at point R, which is 1 wavelength farther from slit 2than from slit 1.

Fig. 26–13 Constructive interference occurs at P and R . Destructive interference occurs at Q.(Figures are not drawn to scale.)

*In 1801 Young used his double-slit experiment to measure the wavelength of light and to provide supportfor his belief that light is a wave. His ideas gained wide acceptance only after many years.

Page 8: CHAPTER · next room. You can easily hear the voices from the room because the sound waves bend around the doorway. ... 1 * CHAPTER 26 Wave Optics *,

738 CHAPTER 26 Wave Optics

Fig. 26–14 A point S on a screen is located a distance r1 from slit 1 and a distance r2 from slit 2.(This figure is not drawn to scale.)

The location of the fringes can be determined with the aid of Fig. 26–14. The figureshows an arbitrary point S some distance y from the center of the interference patternat point P. The angular displacement of point S is measured by the angle '. Thedifference in the path lengths from S to each of the slits is related to the same angle '.As shown in the figure, this path-length difference is d sin '. Constructive interferenceoccurs when this distance equals a whole number of wavelengths:

d sin ' m m 0, 1, 2, … (constructive interference) (26–4)

Destructive interference occurs when the difference in path lengths equals a wholenumber of wavelengths plus wavelength:

d sin ' (m ) m 0, 1, 2, … (destructive interference) (26–5)1*2

1*2

EXAMPLE 2 Measuring Light’s Wavelength With a Meter Stick

Light from a He-Ne laser illuminates two narrow slits, 0.20 mmapart, producing interference fringes on a wall 6.67 m from theslits (Fig. 26–12). The centers of the bright fringes are 2.1 cmapart. (a) Determine the wavelength of the laser light. (b)Whatwould the fringe separation be if the slits were illuminatedwith violet light of wavelength 400 nm?

SOLUTION (a) From Fig. 26–14, we see that the distance yfrom the center of the interference pattern to any point S isrelated to the angle � and the distance � from slits to screen:

tan � �

Any point in the interference pattern is at a very small angle �,for which sin � and tan � are very nearly identical. Thus

sin � �

Inserting this equation into Eq. 26–4 yields an expression forthe distance ym to the mth fringe

d � m� m � 0, 1, 2, …

or

ym �m���

d

ym��

y��

y��

Page 9: CHAPTER · next room. You can easily hear the voices from the room because the sound waves bend around the doorway. ... 1 * CHAPTER 26 Wave Optics *,

Fig. 26–15 The top and bottom sur -faces of a thin film reflect light upward.

Fig. 26–16 An interference pattern isformed by reflection from the thin filmof air between two microscope slides.

73926–2 Interference

Thin FilmsThere is another simple way to split a single light wave into separate, coherent waves,which can then interfere. When light is incident on a partially reflecting surface (forexample, a glass surface), part of the incident light is reflected and part is transmittedinto the second medium. If the two light waves again come together, after havingfollowed paths of somewhat different length, they will interfere. The difference in pathlengths, however, must be less than the coherence length of the light, or the two waveswill be totally incoherent. The Michelson interferometer, described in Problem 21, usesa half-silvered mirror to equally divide the incident wavefront from a monochro-matic source.A thin transparent film can also serve to produce two coherent light waves, one

reflected from the top and one from the bottom of the film (Fig. 26–15). If the film isvery thin, you can see the interference, even with a white light source, as indicated bythe photo of the colored oil film at the beginning of this chapter.In Fig. 26–15 the second light wave travels a distance greater than the first. The

difference in the path lengths may cause a phase difference between the two waves. Inaddition, there may also be phase changes produced by the reflections. When light isincident on the surface of a medium with a higher refractive index than that ofthe incident medium, reflected light experiences a 180� phase change. If thesecond medium has a lower index than the first, reflection causes no phase change. Thesituation is the same as that of a wave pulse on a rope, partially reflected because of achange in density of the rope. As illustrated in Chapter 16, Fig. 16–11, the reflectedpulse is inverted if the second section of rope has higher density than the first. If thesecond section is of lower density than the first, there is no inversion of the reflectedpulse, that is, no change in phase.The microscope slides in Fig. 26–16 show a pattern of interference fringes produced

by light reflected on either side of the thin film of air trapped between the slides. Theair film varies in thickness, and hence the interference alternates between constructive(bright) and destructive (dark).

EXAMPLE 2 Measuring Light’s Wavelength With a Meter Stick—Continued

Thus

y0 � 0, y1 � y2 � , …

The fringes are equally spaced, separated by a distance

�y � (26–6)

Solving for �, we find

� � �

� 6.3 �10�7 m

� 630 nm

Thus, using measurements in cm and m, we indirectly measurethe wavelength of the laser light, a length less than a thousandthof a mm.

(b) Applying Eq. 26–6, we find that for violet light of wave-length 400 nm, the fringe spacing changes to

�y � � �1.3 �10�2 m

�1.3 cm

Because of its shorter wavelength, violet light produces inter-ference fringes that are closer together.

(400 �10–9 m)(6.67 m)���������

2.0 �10–4 m���d

(2.0 �10–4 m)(2.1 �10–2 m)����

6.67 md �y�

���d

2���

d

���d

Page 10: CHAPTER · next room. You can easily hear the voices from the room because the sound waves bend around the doorway. ... 1 * CHAPTER 26 Wave Optics *,

740 CHAPTER 26 Wave Optics

EXAMPLE 3 Colors On An Oil Slick

What color will be brightest when white light is reflected atnormal incidence from a film of oil 250 nm thick on the surfaceof a puddle of water? The oil has a refractive index of 1.4.

SOLUTION The light reflected from the upper surface ofthe oil film undergoes a 180° phase change, since oil’s refractiveindex is higher than air’s (Fig. 26–17). The light reflected fromthe lower surface of the oil experiences no phase change, sincewater’s index is lower than oil’s. Thus, if light reflected from thelower surface is to arrive at the upper surface in phase with thelight reflected from the top surface, the extra distance traveledmust give it a net 180° phase change. This means that the addi-tional path length, equal to twice the film’s thickness t fornormal incidence, must equal a whole number of wavelengthsplus wavelength.

2t � (m � )�n where m � 0, 1, 2, … (26–7)

The wavelength �n is of course the wavelength in the oil. Thisis related to the vacuum wavelength �0 by Eq. 23–10:

�n �

where n is the refractive index of the oil. Inserting this expres-sion into Eq. 26–7 and solving for �0, we find

�0 � for m � 0, 1, 2, … (26–8)

Fig. 26–17

Trying possible values of m, we find

m � 0: �0 � � �1400 nm

m �1: �0 � � � 470 nm

m � 2: �0 � � � 280 nm

Only for m �1 do we find a wavelength in the visible range (400to 700 nm). Blue light of wavelength 470 nm will interfere con -structively, and therefore blue will be the color most strongly re -flected by the film. The film will appear blue. One can show thatred light will experience destructive interference (Problem 14).

1400 nm�3

2tn�

�5

2�

2(250 nm)(1.4)��

�1

2�

1400 nm�3

2tn�

�3

2�

2tn�

�1

2�

2tn�m � �

1

2�

�0�n

1�2

1�2

EXAMPLE 4 Nonreflective Glass Coating

Uncoated glass reflects 4% of the light incident on its surface atnormal incidence. Sometimes glass is coated with a thin layer ofa transparent material so that the intensity of the reflected lightis reduced. Find the minimum thickness of a coating of magne-sium fluoride, MgF2 (n �1.38), which will produce destructiveinterference at a wavelength in the middle of the visible spec-trum (550 nm).

SOLUTION Both reflected waves experience a 180° phasechange, since both are reflected from a medium with a higherindex than that of the incident medium (Fig. 26–18). The onlyrelative change in phase results from a difference in path length.Destructive interference occurs for a minimum path differenceof wavelength.

2t � �n �

or

t � � � 99.6 nm

Fig. 26–18

550 nm�4(1.38)

�0�4n

�0�2n

1�2

1�2

Page 11: CHAPTER · next room. You can easily hear the voices from the room because the sound waves bend around the doorway. ... 1 * CHAPTER 26 Wave Optics *,

Fig. 26–20 Diffraction pattern of apenny. Constructive interference of lightdiffracted around the edge of the pennyproduces a bright spot at the center ofthe shadow.

74126–3 Diffraction

DiffractionHistorical BackgroundThe quantitative study of diffraction was of great historical importance in establishingthe wave nature of light. Although Young’s double-slit experiment supported the wavetheory, many nineteenth-century scientists clung to Newton’s particle theory of light.Full acceptance of the wave theory followed careful quantitative studies of diffractionby various scientists, especially Fresnel and Arago. In evaluating the wave theory ofdiffraction proposed by Fresnel, Poisson objected that it led to a rather strange and, toPoisson, an obviously false prediction: that at the center of the shadow of a roundobject would be a bright spot. Poisson argued that waves diffracted around the edgeswould travel an equal distance to the center and therefore interfere constructivelythere, if the wave theory were correct. Poisson presented this argument as a proof thatthe wave theory was wrong. Arago promptly performed the crucial experiment andfound the predicted bright spot at the center of the shadow (Fig. 26–20). Based on suchresults, the wave theory of light was strongly established by 1820.

26–3

EXAMPLE 4 Nonreflective Glass Coating—Continued

This is the minimum thickness of MgF2 that will producedestructive interference for �0 � 550 nm. Such a layer willnot completely eliminate reflection at this wavelength, since theamplitudes of the two interfering waves are not equal, and sothere is only partial cancellation of the waves, as in Fig. 26–9b.

One might wonder whether this coating will produce areduction in reflected intensity at other wavelengths or whetherit could perhaps enhance reflection by constructive interfer-ence at some wavelengths. The condition for constructive inter-ference here is that the path-length difference equals a wholenumber of wavelengths.

2t � m�n � mor

�0 � where m �1, 2, 3, …

m �1: �0 � 2tn � 2(99.6 nm)(1.38) � 275 nm

m � 2: �0 � �138 nm

Other values of m yield smaller wavelengths. Thus no value ofm gives a wavelength in the visible range. No visible light

interferes constructively. A more detailed analysis, which takesinto account the intensities of the interfering waves, shows thatthe effect of the coating is to reduce reflection fairly uniformlyacross the visible spectrum to an average of about 1% of incidentintensity, compared to a 4% reflection for uncoated glass. How -ever, there is a slight enhancement of reflected intensity in theblue part of the spectrum. Coating glass with several thin layersof different materials, carefully selected for index and thickness,can provide a further reduction in intensity. One of the mostimportant applications of such coatings is for lenses in opticalinstruments, which use a large number of lenses that wouldotherwise produce much unwanted reflected light. Fig. 26–19shows eyeglasses that have a nonreflective coating on one lens.

Fig. 26–19

2tn�2

2tn�m

�0�n

Page 12: CHAPTER · next room. You can easily hear the voices from the room because the sound waves bend around the doorway. ... 1 * CHAPTER 26 Wave Optics *,

Fig. 26–21 Diffraction by a straight edgeand the corresponding graph of intensityversus position. The dashed line indicatesthe intensity predicted by geometricaloptics, implying a sharp-edged shadow.

Fig. 26–22 Single-slit diffraction. Lightdiffracted by the slit forms a series oflight and dark bands on a distant screen.

742 CHAPTER 26 Wave Optics

Fraunhofer and Fresnel DiffractionWhen an object producing diffraction is illuminated by a plane wave and the resultingdiffraction pattern is viewed on a screen at a large enough distance from the object,detailed analysis of the diffraction pattern is greatly simplified. This case is referred toas “Fraunhofer diffraction.” When the illuminating source is not a plane wave or thescreen is not far enough away, the diffraction is called “Fresnel diffraction.” Analysisof Fresnel diffraction is more complicated than analysis of Fraunhofer diffraction, andthe Fresnel diffraction pattern itself looks quite different from the Fraunhofer diffrac-tion pattern of the same object. Figs. 26–20 and 26–21 are examples of Fresnel dif -fraction. We shall analyze only Fraunhofer diffraction because of its relative simplicity.

Single SlitA particularly simple case, Fraunhofer diffraction by a single slit, is shown in Fig.26–22. In Section 26–1 we described diffraction of light around the edges of the slit(Fig. 26–6b). Actually the phenomenon is somewhat more complicated because lightcoming from various parts of the slit interferes to form a series of light and dark bands,as shown in Fig. 26–22.

Fig. 26–23 Rays emanate from points in the opening, a thin slit of width a. All rays in the 'direction strike the screen at point P. (This figure is not drawn to scale.)

Page 13: CHAPTER · next room. You can easily hear the voices from the room because the sound waves bend around the doorway. ... 1 * CHAPTER 26 Wave Optics *,

(a)

(b)

(c)

Fig. 26–25 Locating the (a) first,(b) second, and (c) third diffractionminima for a single slit of width a.

74326–3 Diffraction

We can use the Huygens-Fresnel principle to analyze single-slit diffraction.According to this principle, each section of a wavefront in the diffracting apertureis the source of a spherical wavelet. The amplitude of the light wave at any pointbeyond the aperture is the superposition of all these wavelets. Fig. 26–23 shows aplane wave incident on a narrow slit and a distant screen for viewing the resultingdiffraction pattern. An enlarged section of the figure shows the spherical wavelets andassociated rays emanating from several points in the opening. Rays striking the screenat point P interfere constructively or destructively, depending on the relative phases ofthe waves. Since P is at a great distance, the rays reaching P are nearly parallel. Theserays form an angle ' with a line drawn to point O, which is directly opposite the slit.Fig. 26–24 shows rays directed at the center of the diffraction pattern (' 0). Since

these parallel rays travel equal distances to the screen, interference is constructive, andthe center of the pattern is therefore a diffraction maximum.

Fig. 26–24 Rays in the ' 0 direction strike the screen at point O, directly in front of the slit.

Next we shall locate the diffraction minima produced by a single slit. We can useFig. 26–25 to find the values of ' corresponding to diffraction minima. The firstminimum is indicated in Fig. 26–25a, where the slit of width a has been divided intotwo halves. We compare any ray from the top half with a ray a distance a/2 below it.Let the difference in path length of two such rays be /2, so that they interfere destruc-tively. As shown in the figure this occurs for rays at an angle ', where

a sin '

Since all rays from the top half of the slit can be paired with a cancelling ray from thebottom half, this angle corresponds to a diffraction minimum. A similar argument canbe applied to find other diffraction minima. In Fig. 26–25b the slit is divided into 4 seg -ments, and it is shown that another diffraction minimum occurs at an angle ', such that

a sin ' 2

In Fig. 26–25c, the slit is divided into 6 segments, and a diffraction minimum isfound at an angle ', where

a sin ' 3

In general, single-slit diffraction minima occur at any angle ' satisfying the equation

a sin ' m m 1, 2, 3, … (diffraction minima) (26–9)

Page 14: CHAPTER · next room. You can easily hear the voices from the room because the sound waves bend around the doorway. ... 1 * CHAPTER 26 Wave Optics *,

Fig. 26–26 Fraunhofer diffraction by acircular aperture. The central maximumis called the “Airy disk,” which contains84% of the light in the pattern. Sincethe diameter of the Airy disk is inverselyproportional to the aperture diameterD, as the diameter of the aperturedecreases, the disk gets bigger.

744 CHAPTER 26 Wave Optics

Circular ApertureFraunhofer diffraction by a circular aperture is of particular importance because of itsapplication to the eye and to optical instruments, which generally have circular aper-tures. However, quantitative analysis of the circular aperture is considerably morecomplicated than analysis of the slit. Therefore we shall simply state without proof theone most important result of that analysis: the first diffraction minimum of an apertureof diameter D is at an angle ', where

D sin ' 1.22

Notice that this equation is similar to the equation for the first minimum of a slit ofwidth a (Eq. 26–9 with m 1: a sin ' ). Since sin ' is very nearly equal to ' inradians, for the angles we normally encounter, we can express this result

' 1.22 (first minimum; circular aperture) (26–10)

Fig. 26–26 shows the Fraunhofer diffraction pattern of a circular aperture and thecorresponding graph of intensity versus '. The circular aperture was illuminated by aplane wave. One can think of this wave as originating from a distant point source.Since most of the light energy is concentrated in the central disk, called the “Airy disk,”for simplicity we can regard this disk as the image of the point.

*D

EXAMPLE 5 A Wide Image of a Thin Slit

Find the width of the central diffraction maximum of a slit ofwidth 0.100 mm, as seen on a screen 2.00 m from the slit. Theslit is illuminated by a He-Ne laser beam (� � 633 nm).

SOLUTION The edge of the central diffraction maximumwill correspond to the first minimum (m � 1). Applying Eq.26–9, we find the angle � corresponding to this point.

a sin � � m� � �or

sin � � � � 6.33 �10�3

� � 6.33 �10�3 rad (or 0.363°)

Next we find the linear distance y from the center of the patternto the point on the screen corresponding to this angle. Since the

angle is quite small, we may approximate sin� by y/�, where yand � are shown in Fig. 26–23. Thus

� sin � � 6.33 �10�3

or

y � (6.33 �10�3)� � (6.33 �10�3)(2.00 m)

�1.27 �10�2 m �1.27 cm

Since the central maximum extends an equal distance below themidpoint, the width of the maximum is double this value, or2.54 cm. Thus the central maximum has a width 254 times theslit width (0.0100 cm). In other words, the width of the slit’scentral image is 254 times greater than the width of the imagepredicted by geometrical optics.

y��

633 nm��1.00 �10�4 m

��a

Page 15: CHAPTER · next room. You can easily hear the voices from the room because the sound waves bend around the doorway. ... 1 * CHAPTER 26 Wave Optics *,

Fig. 26–27 Resolution of two pointsources of light diffracted by a circularaperture.

(c) Pointsunre-solved

(b) Pointsarebarelyresolved

(a) Pointsareclearlyresolved

74526–3 Diffraction

Rayleigh Criterion for ResolutionWhen a point source of light is imaged by an optical system with a circular aperture,the image is an Airy disk. For example, the image of a star formed by a telescope issuch a disk. If two points are very close, their Airy disks will overlap, and you may notbe able to distinguish separate images. Fig. 26–27 shows the image of two points thatare (a) clearly resolved, (b) barely resolved, or (c) unresolved. As a quantitativemeasure of the resolution of two points, Lord Rayleigh proposed the following crite-rion— the Rayleigh criterion: two points are barely resolved when the center ofone’s Airy disk is at the edge of the other’s Airy disk.Fig. 26–28 illustrates the formation of the image of two points that are barely

resolved according to Rayleigh’s criterion. Notice that the angular separation 'min of thetwo points P and Q is the angle from the center of an Airy disk to the first minimum,expressed by Eq. 26–10. Thus two points are resolved only if they subtend an angle atleast as big as this minimum value 'min, where

'min 1.22 (Rayleigh’s criterion for resolution) (26–11)

Any image formed by an optical system consists of a set of Airy disks, each of which isthe image of a single point on the object. The size of these disks determines the resolution.

Fig. 26–28 Rayleigh’s criterion for resolution.

*D

EXAMPLE 6 Diffraction of Light by the Eye

(a) Find the prediction of diffraction theory for the minimumangle subtended by two points that are barely resolved by the eye.Assume a pupil diameter of 2.0 mm, and use a wavelength atthe center of the visible spectrum. (b) Find the distance betweenthe two points if they are 25 cm from the eye, at its near point.

SOLUTION (a) We apply Eq. 26–11, using for � the wave-length inside the eye, where the refractive index n �1.34. At thecenter of the visible spectrum the vacuum wavelength �0 � 550nm, and

� �

Thus

�min �1.22 �1.22 �1.22

� 2.5 �10�4 rad

(b) From Fig. 26–28 we see that this angle equals the separationd between the points divided by the distance of 25 cm.

� 2.5 �10�4 rad

d � (25 cm)(2.5 �10�4 rad) � 6.3 �10�3 cm

The eye should be unable to resolve points closer than about0.06 mm.

The minimum angle we have calculated is fairly close to themeasured minimum angle between points barely resolved bythe normal eye. The factors other than diffraction that affect thisminimum angle are discussed in Chapter 25, Section 25–5(Factors Limiting Visual Acuity).

d�25 cm

5.5 �10–7 m���(1.34)(2.0 �10–3 m)

�0�nD

��D

�0�n

Page 16: CHAPTER · next room. You can easily hear the voices from the room because the sound waves bend around the doorway. ... 1 * CHAPTER 26 Wave Optics *,

746 CHAPTER 26 Wave Optics

Diffraction GratingsA diffraction grating consists of thousands of very narrow, closely spaced slits, madeby etching precisely spaced grooves on a glass plate (Fig. 26–30). The slits are thetransparent spaces between the grooves. Typically a grating consists of thousands ortens of thousands of transparent lines per cm. It follows from our discussion of single-slit diffraction that each line, because of its extremely narrow width, producesdiffracted light spread over a considerable angle—perhaps 20� or 30�. Of course, thediffraction pattern of one such line alone would not produce enough intensity to beseen by itself. But when diffracted light from thousands of lines interfere, bright, sharpdiffraction maxima are produced, (Figs. 26–31 and 26–32).

EXAMPLE 7 Diffraction Limit of a Microscope

Find an expression for the minimum separation between twopoints that are barely resolved by a microscope with an object -ive of diameter D and focal length f.

SOLUTION Using Fig. 26–29 and applying Eq. 26–11, wefind

d � f�min �1.22 (26–12)

To make d as small as possible we need to minimize the ratiof /D. But it is not possible to make f less than about D/2, theradius of the lens.* Setting f � D/2 in the expression above, weobtain

d � 0.61�

*For a simple symmetrical lens made of glass with a refractive index of 1.5,the focal length equals the radius of curvature of the first surface, as shownin Problem 24–37. But the radius of curvature can be no smaller than theradius of the lens itself. Therefore the focal length is always less than theradius D/2.

Fig. 26–29

Thus the minimum distance between two points that can beresolved by any microscope equals roughly half the wave-length of the light used to illuminate the points. For example,using the value � � 550 nm for the center of the visible spec-trum, we find

d � (0.61)(550 �10�9 m)

� 3.4 �10�7 m

f��D

Fig. 26–32 The diffraction pattern produced by two per -pendicular diffraction gratings illuminated by a He-Ne laser.

Fig. 26–31 The diffraction pattern produced by agrating illuminated by a He-Ne laser.

Fig. 26–30 A diffraction gratingconsists of thousands of narrow,closely spaced slits.

Page 17: CHAPTER · next room. You can easily hear the voices from the room because the sound waves bend around the doorway. ... 1 * CHAPTER 26 Wave Optics *,

Fig. 26–33 Rays from a diffractiongrating.

Fig. 26–35 X-ray diffraction patternof DNA. The double helix structureof DNA was revealed by this historicphotograph taken by Rosalind Franklin.

74726–3 Diffraction

Fig. 26–33 can be used to locate the diffraction maxima of a grating with spacingd between adjacent slits. As indicated in the figure, for light in the ' direction, thedifference in path length of rays from adjacent slits is d sin '. These adjacent rays willinterfere constructively if the difference in path length is an integral multiple of , thatis, 0, , 2 , 3 , and so forth. Indeed rays from all the slits will interfere constructivelyif they are directed at an angle ', such that

d sin ' m m 0, 1, 2, … (26–13)

Unlike the maxima produced by a double slit, diffraction grating maxima are verynarrow and sharp, as shown in Fig. 26–31. This can be understood when we considerhow a slight change in the angle ' away from a value satisfying Eq. 26–13 will affectthe intensity. Suppose that the change in angle is so slight that light from adjacent slitsis still nearly in phase. If there were only two slits interfering, such an angle would stillgive an intensity close to the maximum value. However, with the thousands of slits ina grating, there can be destructive interference in many ways. For example, lightfrom slits 100 spacings apart might interfere destructively. If all pairs of slits 100 spac-ings apart interfere destructively, there will be no light in that particular direction.Diffraction gratings can be used to measure the wavelength of light, as illustrated

in the following example.

X-ray diffraction is a technique that utilizes the small spacing between the atoms ina crystal as a three-dimensional diffraction grating. The atomic spacing is on the sameorder as wavelengths in the X-ray portion of the electromagnetic spectrum. So X rays,rather than visible light, are diffracted by a crystal. And the resulting diffraction patterncan be used to discover the crystal structure. X-ray diffraction of DNA was used byWatson and Crick in discovering the structure of DNA in 1951 (Fig. 26–35).

EXAMPLE 8 Separating the Sodium Doublet

Find the first-order (m � 1) diffraction angles for the sodiumdoublet, using a grating with 106 lines/m. The sodium doubletconsists of two yellow lines in the spectrum of sodium, withnearly identical wavelengths: 589.00 nm and 589.59 nm.

SOLUTION Applying Eq. 26–13 to each of the wavelengths,with d � 10�6 m and m � 1, we find

sin � �

sin �1 � � 0.58900

�1 � 36.09°

Fig. 26–34

sin �2 � � 0.58959

�2 � 36.13°

Thus the angular separation between the lines is 0.04°, or 7 �10�4 rad. Viewed at a distance of 1 m, the lines are 0.7 mm apart.

(1)(589.59 � 10–9 m)���

10–6 m

m��

d

(1)(589.00 � 10–9 m)���

10–6 m

This book is licensed for single-copy use only. It is prohibited by law to distribute copies of this book in any form.

Page 18: CHAPTER · next room. You can easily hear the voices from the room because the sound waves bend around the doorway. ... 1 * CHAPTER 26 Wave Optics *,

748 CHAPTER 26 Wave Optics

PolarizationPolarization by AbsorptionMost light sources produce unpolarized light (Fig. 26–3), as opposed to the polarizedlight produced by some lasers (Figs. 26–1 and 26–2). However, there are ways topolarize light that is initially unpolarized, or to change the direction of polarization ofpolarized light. One way is to pass the light through a Polaroid sheet, a synthetic mate-rial first produced by Edwin Land in 1928 when he was an undergraduate.There is a direction along each Polaroid sheet called its “transmission axis.” Light

linearly polarized along this axis passes through the sheet (Fig. 26–36a), whereas lightpolarized in the perpendicular direction is completely absorbed (Fig. 26–36b). If theincident light is linearly polarized at some angle ' relative to the transmission axis, thelight will be partially absorbed and partially transmitted. As illustrated in Fig. 26–36c,the component of the electric field parallel to the axis is transmitted. The light thatemerges is thus polarized along the direction of the transmission axis and has anamplitude E related to the incident amplitude E0 by the equation

E E0 cos '

Fig. 26–36 The effect of a Polaroid sheet on initially polarized light depends on the directionof initial polarization relative to the direction of the sheet’s transmission axis.

The intensity of light is proportional to the square of its amplitude (Eq. 23–5: Iav �0cE0

2). Squaring the equation above, we obtain

E 2 E02 cos2 '

Multiplying both sides of this equation by the appropriate constant ( �0c), we obtain a re -lationship between the average transmitted intensity I and the average incident intensity I0:

I I0 cos2 ' (26–14)

26–4

(a) Incident lightpolar ized along trans-mission axis

(b) Incident lightpolarized perpendicu -lar to transmission axis

(c) Incident lightpolarized at an angle 'with transmission axis

1*2

1*2

Page 19: CHAPTER · next room. You can easily hear the voices from the room because the sound waves bend around the doorway. ... 1 * CHAPTER 26 Wave Optics *,

(a)

(b)

(c)

Fig. 26–37 Equivalent representationsof unpolarized light.

74926–4 Polarization

This result is known as the law of Malus. The intensity of the transmitted light has itsmaximum value, I I0, when ' 0, and has its minimum value, I 0, when ' 90�.Unpolarized light consists of a superposition of linearly polarized waves, with vary -

ing directions of polarization, as illustrated in Fig. 26–37a. The electric field associatedwith each of these waves can be resolved into x and y components, relative to an arbi-trary coordinate system. Since the direction of polarization is random, the resultant xand y components are equal. We can replace the many randomly directed linearlypolarized waves by just two linearly polarized waves of equal intensity, with mutuallyperpendicular polarization directions, as illustrated in Figs. 26–37b and 26–37c.When unpolarized light is incident on a Polaroid sheet, only the component along

the transmission axis is transmitted. Since the two components have equal intensity inunpolarized light, this means that the intensity of the transmitted light is half theintensity of the incident light (Fig. 26–38).

I I0 (initially unpolarized light) (26–15)

Fig. 26–38 A Polaroid sheet polarizes initially unpolarized light.

1*2

EXAMPLE 9 Light Passing Through Two Polarizers

An unpolarized laser beam of intensity 1000 W/m2 is incident ona Polaroid sheet with a vertical transmission axis. The lightpassing through this sheet strikes a second Polaroid sheet, witha transmission axis at an angle of 30.0° from the vertical (Fig.26–39). Find the polarization and the intensity of the lightemerging from the second sheet.

SOLUTION The light transmitted by the first Polaroid sheetis vertically polarized and, according to Eq. 26–15, has anintensity equal to half the incident intensity.

I � I0 � (1000 W/m2) � 500 W/m2

Fig. 26–39

The light incident on the second sheet is polarized at an angle of30° relative to this sheet’s transmission axis and, according tothe law of Malus (Eq. 26–14), has intensity

I� � I0� cos2 � � (500 W/m2)(cos2 30.0°) � 375 W/m2

1�2

1�2

Page 20: CHAPTER · next room. You can easily hear the voices from the room because the sound waves bend around the doorway. ... 1 * CHAPTER 26 Wave Optics *,

(a)

(b)

Fig. 26–40 (a) Light polarized parallelto the reflecting surface is more stronglyreflected than (b) light polarized in aperpendicular direction.

750 CHAPTER 26 Wave Optics

Polarization by ReflectionWhen light is reflected from the surface of a dielectric, such as water or glass, theintensity of the reflected light depends on the angle of incidence and on the polariza-tion of the incident light. Light polarized parallel to the reflecting surface (Fig.26–40a) is always more strongly reflected than light polarized in a perpendiculardirection (Fig. 26–40b). Unpolarized light can be thought of as consisting of two equal-intensity polarized waves—one polarized parallel to the surface and a second polarizedperpendicular to the first. After reflection, the component polarized parallel to thesurface is more intense than the other component. Fig. 26–41 shows the intensities ofthe two components in a beam of initially unpolarized light reflected by water, forseveral angles of incidence. The intensities are predicted by Maxwell’s equations.Notice that both components are more strongly reflected for very large angles ofincidence. Any smooth dielectric surface becomes mirror-like as the angle of incidenceapproaches 90�. You can observe this effect by holding up a smooth sheet of paper sothat rays from a light source are reflected at glancing incidence.The angle of incidence at which the reflected light is 100% polarized is known as

“Brewster’s angle,” denoted by 'B. Brewster’s angle has a value of 53� for reflection bywater, as indicated in Fig. 26–41c. Maxwell’s equations can be used to derive an ex pres -sion for Brewster’s angle, relating it to the refractive index n of the incident mediumand the index n� of the reflecting medium. We present the result here without proof:

tan 'B (Brewster’s angle) (26–16)n�*n

(e)(d)

(c)(b)(a)

Fig. 26–41 Initially unpolar-ized light of total intensity200 W/m2 is reflected bywater at various angles ofincidence.

Page 21: CHAPTER · next room. You can easily hear the voices from the room because the sound waves bend around the doorway. ... 1 * CHAPTER 26 Wave Optics *,

75126–4 Polarization

Polaroid sunglasses are effective at reducing reflected glare from the surface of abody of water or from other surfaces (Fig. 26–42). The lenses are made of Polaroidsheets with vertical transmission axes. The reflected light consists mainly of horizon-tally polarized light, and such light is completely absorbed by the lenses.

Fig. 26–42 An early ad for Polaroid sunglasses.

EXAMPLE 10 Brewster’s Angle for Glass

Calculate Brewster’s angle for light incident from air onto aglass surface if the glass has a refractive index of 1.5.

SOLUTION Applying Eq. 26–16, using the refractive indicesfor glass and air, we find

tan �B � � �1.5

�B � 56°

1.5�1.0

n��n

EXAMPLE 11 A Sunset Seen Through Polaroid Sunglasses

The setting sun is reflected from the surface of a lake at anangle of incidence of 75°. The intensity of the incident light is200 W/m2, as in Fig. 26–41e. Find the intensity of the reflectedlight reaching the eye of an observer wearing Polaroid sun -glasses (Fig. 26–43).

SOLUTION From Fig. 26–41e we see that the intensity ofthe reflected light having a polarization along the transmissionaxis of the sunglasses is 11 W/m2. Therefore this is the intensityof the reflected light reaching the eye.

Without sunglasses, the observer would see reflected light ofboth polarizations. From Fig. 26–41e, the intensity of reflectedlight seen by the observer would be 11 W/m2 � 31 W/m2 �

42 W/m2.

Fig. 26–43

Non-Polaroid sunglasses that produce the same darkening asthese Polaroid sunglasses would absorb 50% of all incidentlight, and so they would transmit to the observer reflected lightof intensity 21 W/m2.

Page 22: CHAPTER · next room. You can easily hear the voices from the room because the sound waves bend around the doorway. ... 1 * CHAPTER 26 Wave Optics *,

Polarization by ScatteringWhen an electromagnetic wave is incident on an atom, the atom’s electrons oscillatein response to the oscillating electric field. The electrons behave like tiny antennas;they emit their own radiation with the same frequency as the incident electromagneticwave, but scatter the radiation in various directions. The intensity of this scattered radi-ation depends on the light’s frequency. Blue light is scattered much more effectivelythan red light. Fig. 26–44 shows how scattering of sunlight by the earth’s atmospheregives us blue skies and red sunsets. Observers A and B both see blue sky as a result ofthe blue part of the sun’s spectrum being scattered toward their eyes by the atmosphere.Meanwhile, observer C sees an orange or red sun as a result of the blue part of thespectrum’s having been scattered out of the beam of direct sunlight. This kind of scat-tering, called “Rayleigh scattering,” is the result of independent, incoherent radiationby many atoms.

Fig. 26–44 Scattering of sunlight by the earth’s atmosphere results in blue skies (seen by A andB) and red sunsets (seen by C).

Scattering is also the basic mechanism at the heart of reflection and refraction by asolid or a liquid. But the higher density and relative immobility of the atoms in theliquid or solid mean that light scattered by neighboring atoms is coherent and cantherefore interfere constructively or destructively. The result of this interference isremarkably simple. The scattered waves interfere destructively in all directions, exceptthose corresponding to the reflected and refracted waves, for which the interference isconstructive. So we see only a reflected wave and a refracted wave.The particles of water in a cloud or in ocean surf also “scatter” sunlight, but in this

case the scattered light is white, in contrast to the blue sky. The tiny water dropletssimply reflect and refract incident light. The result of multiple reflections and refrac-tions by a very large number of droplets is to redirect or scatter the incident white lightin all directions.

CHAPTER 26 Wave Optics752

Page 23: CHAPTER · next room. You can easily hear the voices from the room because the sound waves bend around the doorway. ... 1 * CHAPTER 26 Wave Optics *,

Scattering of sunlight by the atmosphere tends to polarize the light. Fig. 26–45shows how this polarization arises. A beam of unpolarized sunlight, incident on theatmosphere, travels along the x-axis. This transverse electromagnetic wave has an elec-tric field that oscillates in the yz plane; there is no component along the x-axis, thedirection of propagation of the wave. Electrons within atoms in the atmosphere oscil-late in the yz plane, in response to the incident wave, and scatter light in variousdirections. The nature of radiation produced by any source of radiation is that there canbe a component of the electric field in a given direction only if there is a component ofmotion of the radiating source parallel to that direction. Therefore there can be no x-component of the electric field in scattered radiation, since there is none in the incidentwave. This implies that radiation scattered along the yz plane, perpendicular to theincident beam, must be polarized. For radiation scattered in any such directionthere is only a single line perpendicular to the direction of propagation, along whichthe electric field vector can oscillate, as indicated in Fig. 26–45. Light scattered in otherdirections is partially polarized.

Fig. 26–45 Polarization by scattering.

26–4 Polarization 753

Page 24: CHAPTER · next room. You can easily hear the voices from the room because the sound waves bend around the doorway. ... 1 * CHAPTER 26 Wave Optics *,

Mirage! Looming! Mountain specter!What do you think of when you hear theseterms? Illusions or ghosts, perhaps. Thenames are suggestive of the fear and be -wilderment that these phenomena havearoused over the centuries. It may surpriseyou to learn, however, that the namesdescribe real optical effects in the atmos-phere. You already know enough about thebasic principles of optics to understandsuch phenomena.

Mirages

You have heard of people lost in the desertwho imagined that they saw tempting poolsof blue water just beyond the next sand

dune and who dragged themselves forwardin hopeless pursuit of them. You may thinkthat such mirages are simply hallucinationscaused by heat and thirst. Actually, theyare almost always real phenomena—reallight is behaving in a way that creates anillusion in which anyone, hot and thirsty ornot, can share (Fig. 26–A).Mirages are refraction phenomena. Light

rays are bent by layers of air at dif ferenttemperatures. Warm air has lower densitythan cool air and has a lower refractiveindex. As a ray coming from a cooler layerenters a warmer layer—which is whathappens when the ray is moving downward

toward a searingly hot desert surface—the ray is refracted away from the normal.The ray can be bent so much that it curvesback upward (Fig. 26–B). When it reachesthe viewer’s eye, it is automatically tracedback by the brain as if it came from a sourcedirectly in line with its final segment. Thatis, it is seen as if it were ahead and below,rather than ahead and above. Light comingfrom a clear blue sky produces the illusionof a bright blue pool on the ground ahead.A mirage of the sort seen in a hot desert

is called an “inferior mirage” because itappears below the light source (the sky).There is another kind of mirage, called a“superior mirage,” which appears abovethe source of light (Fig. 26–C). A superiormirage is typically caused when light movesupward from a layer of cool, dense air intowarmer, less dense layers. The rays are bentaway from the normal, as before, but thistime they turn down rather than upward(Fig. 26–D). When they strike the eye, theyare traced up to a mirage seen at theirapparent point of origin. Under such condi-tions, a ship moving on the water belowthe horizon can look like a ghost ship sail -ing through the sky! The appearance ofa superior mirage is sometimes calledlooming, for obvious reasons: the miragelooms above its source.Under special conditions, looming can

produce truly uncanny effects. One par -ticular type of looming is called fatamorgana, after Morgan le Fay, the fairy-enchantress of the King Arthur legendswho lived on a magical island. Fata morganais most often seen in the Strait of Messina,a waterway that separates Sicily from Italyand that was long dreaded for its deadlycurrents, rocks, and whirlpools. The mirageis caused by irregular layerings of air ofvarious densities, which produce multiplerefractions and multiple overlapping images.The result is an apparent vertical elonga-tion of the source object, sometimes toenormous proportions. For example, when

A Closer Look

Magic in the Sky

Fig. 26–A A fairly common kind of mirage: the dry surface of a road appears to be wet.

Fig. 26–B Formation of an inferior mirage.

Page 25: CHAPTER · next room. You can easily hear the voices from the room because the sound waves bend around the doorway. ... 1 * CHAPTER 26 Wave Optics *,

seen from a ship in the strait, objects suchas trees or hills on the shore can look likehuge, weirdly shaped figures that can bedisorienting and dangerous to unsuspectingmariners (Fig. 26–E).

Coronas and anticoronas

Interference or diffraction effects can occurwhen light rays pass near the edges of tinyobjects in the atmosphere. If the diffractionappears around the light source, the effectis usually a ring, called a corona.Coronas can appear around the sun or

moon when light rays pass near the edgesof water droplets in the atmosphere. Ringsof different colors can be seen becausedifferent wavelengths are diffracted tovarying degrees.

A Closer Look

Fig. 26–C A superior mirage of a ferry,which appears to be vertically elongated.

Fig. 26–D Formation of a superior mirage.

Fig. 26–E What appears in this picture to be icy castles in the sky is actually an extremely rare,mirage-like effect known as fata morgana.

Page 26: CHAPTER · next room. You can easily hear the voices from the room because the sound waves bend around the doorway. ... 1 * CHAPTER 26 Wave Optics *,

A Closer Look

A related phenomenon caused by thepresence of hexagonal ice crystals is calleda halo (Fig. 26–F). If the ice crystals line upin just the right way on a sunny day, theycan produce two separate bright spots, oneto either side of the sun. These ghost sunsare called parhelia or more popularly,sundogs.When diffraction occurs around a

shadow area, the effect is called an anti-corona. Anticoronas are also known bythe names glory (Fig. 26–G) and moun-tain specter.These are rare and awesomephenomena.The glory is seen as a halo surrounding

the shadow of the observer’s head. Be -cause of the extreme directional depend-ence of this effect, no one else sees thehalo around the observer’s head. Twoobservers might each see a halo aroundhis or her own head, but each observerwill not be able to see the other’s halo.

One particularly famous example of ananticorona occurs occasionally near theBrocken, a mountain in the Harz range ofcentral Germany. Because it is known inGerman legend as the site of the WalpurgisNight witchcraft rituals, the Brocken is an

appropriate location for the anticorona,which is there given the name “Brockenspecter.” The specter can be seen at twi -light on sunny days by observers who standnear the foot of the mountain when thereare misty banks of fog or cloud just abovethem that do not reach as high as themountain top. The low-lying sun then castsa huge shadow of the mountain peak ontothe upper surface of the mist. This silhou-ette of the peak appears surrounded byrings of colored light, as rays passing aroundthe edge of the peak are bent and sep -arated out according to their wavelength(Fig. 26–H).

Fig. 26–H Viewing the Brocken Specter.

Fig. 26–F A street light blocks the sun’sdirect rays, allowing a halo around the sun tobe clearly seen.

Fig. 26–G The bright circle around theplane’s shadow is called a glory.