Course on Waves

Embed Size (px)

Citation preview

  • 7/27/2019 Course on Waves

    1/632

  • 7/27/2019 Course on Waves

    2/632

    Plancks constant h

    Reduced Plancks constant = h / i n

    Electron rest meProton rest mp

    G r a v i t a t i o n a l c o n s t a i f ^ ^ ^ ^ ^ ^ ^ ^ ^ ^ ^ ^ ^ S G

    Acceleration of gravity at sea level g

    Bohr radius a

    Avogadros No

    B o l t z m a n n s c o n s t a n l ^^^^^^^^^^^^^^S k

    Standard temperatur^^^^^^^^^^^^^^^BTb

    Standard p r e s s u r a ^ ^ ^ ^ ^ ^ ^ ^ ^ ^ ^ ^ ^ ^ H poMolar volume at Vo

    Thermal energy k Tat kT0

    Density of air at po

    Speed of sound in air at v0

    Sound impedan ce of air at Z0

    Standard sound i n t e n s i t ) ^^ ^ ^^ ^^ ^^ ^^ ^| l o

    Factor of ten in intensity

    One fermi (F)One angstrom unit ( A )One micron (ju)

    One hertz (Hz)

    Wavelength of one-electron-volt photon

    One electron volt (ev)

    One watt (W)

    One coulomb (coul)

    One volt (V)One ohm (fi)

    Thirty ohms

    Impedance per square of vacuum for

    electromagnetic waves

    One farad (F)

    One henry (H)

    Useful Constants2.997925 X 1010 cm/sec = 3 X 1010 cm/sec

    4.8 X 10~10 statcoulomb

    1.6 X 10~19 coulomb

    6 .6 X 10 ~27 erg-sec

    1.0 X 10 27 erg-sec

    0.9 X 10~27 gm1.7 X 10-2 4 gm

    6.7 X lO- 8 CGS units

    980 cm /sec2

    0.5 X 10-8 cm

    6.0 X 1023 mole-1

    1.4 X 10~16 erg/deg Kelvin

    273 deg Kelvin

    1 atm = 1.01 X 106 dyne/cm 222 .4 x 103 cm3/mole

    3.8 X 10-1 4 erg ^ ev

    1.3 X 10-3 gm/cm 3

    3.32 X 104 cm/sec

    42.8 (dyne/cm 2) /(cm/sec)

    1 jnwatt/cm2

    1 bel = 10 db

    10-13 cm 10-8 cm

    10~4 cm

    1 cycle per second (cps)

    1.24 X 104 cm ^ 1 23 45 A1.6 X 10-1 2 erg

    1 joule/sec = 107 erg/sec

    3 X 109 statcoul = c/ 1 0 s tat coul0

    s f ostatvolt =; 108/ c statvolt1/(9 X 10n ) statohm = 109/ c 2 statohm

    1/ c statohm

    At t/c statohm = 377 ohm

    9 X 1011 statfarad = c2/1 0 9 statfarad

    1/ (9 X 1011) stathenry = 109/ c 2 stathenry

    In converting from practical units to electrostatic units we have approximated the velocity of light as 3.00 X 1010 cm/sec. Wherever a 3

    appears, a more accurate conversion factor can be obtained by using the more accurate value of c. Similarly wherever 9 appears, it is moreaccurately (2.998)2.

  • 7/27/2019 Course on Waves

    3/632

    Recommended Unit Prefi xes

    M u l t i p l es and

    Submu l t i p l es Pref i xes Symbols

    1012 tera T109 giga G106 mega M103 kilo k

    102 hecto h10 deka dalO"1 deci dio-2 centi clO"3 milli m10-6 microIO"9 nano nio-*2 pico P

    Useful I dent i t i es

    cos x +cos y= [2 cos (x t/)] cos \ (x+ y )

    cos x cos y [2 sin %(x y)] sin (x + y)

    sin x+ sin y= [2 cos (x y)] sin (x + y)

    sin x sin y =[2 sin (x t/)] cos (x + y)

    cos (x y)= cos x cos yqp sin x sin t/sin ( x y ) =sin x cos y sin ycos x

    cos 2 x = cos2 x sin2 x

    sin 2x = 2 sin x cos x

    cos2 x = { 1 + cos 2 x)

    sin2x = (1 cos 2x)

    sin x = x c 3 +

    cos x = 1 |x2 +

    (1 + x)n = 1 + nx + n(nl)x2 + cos Oi+ cos (#i + y) + cos (0i + 2 y) +

    ; x2 < 1 .+ cos [0i + (N l)y] = cos [01 + %(N l)y]

    sin %Ny

    sin i y

  • 7/27/2019 Course on Waves

    4/632

    waves

  • 7/27/2019 Course on Waves

    5/632

    N e w Yo r k St . Lo u is

    Sa n Fr a n c i s co

    To r o n t o L o n d o n

    S y d n e y

    mcgraw -h i l l book com pany

  • 7/27/2019 Course on Waves

    6/632

    wavesb e r k e l e y p h y s i c s c o u r s e vo l u m e 3

    Th e prepara t io n of th i s cour se teas suppor ted, hi j a grant

    f r om the Na t i onal Sci ence Founda t i on t o Educa t i on D e-

    ve l opmen t C en t er

    Fra nk S. Craw ford,Jr .Pro f esso r o f P hys icsUn i ver si t y o f Ca l i f o r n i a , Be r kel e y

  • 7/27/2019 Course on Waves

    7/632

    COVER DESIGN

    Photograph ic adapta t ion by Fel i x Cooper

    f ro m an or i g ina l by John Severson

    WAVES

    Copyr i gh t 1965, 1966, 1968 by Educat i on

    D eve lop m ent Cent er , I nc . (successor by m erger

    to Educat io nal Servi ces In corporat ed). A l l

    Ri ghts Reserv ed. Pr i n ted in the U ni t ed Sta teso f Am er i ca . Th i s book , o r pa r t s thereo f , may

    no t be r ep r odu ced i n an y f o rm w i t h ou t t h e

    w r i t t en perm i s si o n o f Educa t i o n D evel o pmen t

    Center , In c ., N ew ton, M assachuset t s .

    L i b ra r y o f Cong ress Ca ta log Card Number

    6466016

    04860

    7890 BPBP 7654

  • 7/27/2019 Course on Waves

    8/632

    Preface to the Berkeley Physics Course

    This is a two-year elementary college physics course for students majoringin science and engineering. The intention of the writers has been to pre

    sent elementary physics as far as possible in the way in which it is used by

    physicists working on the forefront of their field. We have sought to make

    a course which would vigorously emphasize the foundations of physics.

    Our specific objectives were to introduce coherently into an elementary

    curriculum the ideas of special relativity, of quantum physics, and of sta

    tistical physics.

    This course is intended for any student who has had a physics course inhigh school. A mathe matics course including the calculus should be taken

    at the same time as this course.

    There are several new college physics courses under development in the

    United States at this time. The idea of making a new course has come to

    many physicists, affected by the needs both of the advancement of science

    and engineering and of the increasing emphasis on science in elementary

    schools and in high schools. Our own course was conceived in a conversa

    tion between Philip Morrison of Cornell University and C. Kittel late in

    196 1. We wer e encourage d by John Mays and his colleagues of the

    National Science Foundation and by Walter C. Michels, then the Chair

    man of the Commission on College Physics. An informal comm ittee was

    formed to guide the course through the initial stages. The comm ittee con

    sisted originally of Luis Alvarez, William B. Fretter, Charles Kittel, Walter

    D. Knight, Philip Morrison, Edward M. Purcell, Malvin A. Ruderman, and

    Jerrold R. Zacharias. The committee met first in May 196 2, in Berkeley;

    at that time it drew up a provisional outline of an entirely new physics

    course. Bec ause of heav y obligations of several of the original members,

    the committee was partially reconstituted in January 1964, and now con

    sists of the undersigned. Contributions of others are acknowledged in the

    prefaces to the individual volumes.

    The provisional outline and its associated spirit were a powerful influence

    on the course materia l finally produ ced. The outline covered in detail the

    topics and attitudes which we believed should and could be taught to

    beginning college students of scien ce and engineering. It was never our

    intention to develop a course limited to honors students or to students with adva nce d standing. We have sought to present the principles of physics

    from fresh and unified viewpoints, and parts of the course may therefore

    seem almost as new to the instructor as to the students.

    The five volumes of the course as planned will include:

    I. Mechanics (Kittel, Knight, Ruderm an) III. Waves (Crawford)

    II. Electricity and Magnetism (Purcell) IV. Quantum Physics (Wichm ann)

    V. Statistical Physics (Reif)

    The authors of each volume have been free to choose that style and method

    of presentation which seemed to them appropriate to their subject.

  • 7/27/2019 Course on Waves

    9/632

    The initial course activity led Alan M. Portis to devise a new elementary

    physics laboratory, now known as the Berkeley Physics Laboratory. Because the course emphasizes the principles of physics, some teachers

    may feel that it does not deal sufficiently with experimental physics. The

    laboratory is rich in important experiments, and is designed to balance the

    course.

    The financial support of the course development was provided by the

    National Science Foundation, with considerable indirect support by the

    University of California. The funds wer e admin istered by Educational

    Services Incorporated, a nonprofit organization established to administer

    curriculum improv ement programs. We are particularly indebted to

    Gilbert Oakley, James Aldrich, and William Jones, all of ESI, for their

    symp athetic and vigorous support. ES I established in Berkeley an office

    under the very competent direction of Mrs. Mary B. Maloney to assist the

    development of the course and the laboratory. The U niversity of Califor

    nia has no official connection with our program, but it has aided us in im

    portan t ways. Fo r this help we thank in particula r two successive chair

    men of the Department of Physics, August C. Helmholz and Burton J.

    Moyer; the faculty and nonacademic staff of the Department; Donald

    Con ey, and many others in the University. Abraham Olshen gave much

    help with the early organizational problems.

    A Fu r t h er No t e Volumes I, II, and V were published in final form in the period from Janu

    ary 19 65 to June 19 67. During the preparation of Volumes III and IV for

    final publication some organizational changes occurred. Education Devel

    opment Center succeeded Educational Services Incorporated as the

    administering organization. The re were also some changes in the com

    mittee itself and some redistribution of responsibilities. The com mittee isparticularly grateful to those of our colleagues who have tried this course

    in the classroom and who, on the basis of their experience, have offered

    criticism and suggestions for improvements.

    As with the previously published volumes, your corrections and sugges

    tions will always be welcome.

    January, 1965

    Eugene D. Commins

    Frank S. Crawford, Jr.

    Walter D. Knight

    Philip Morrison

    Alan M. Portis

    Edward M. Purcell

    Frederick Beif

    Malvin A. Buderman

    Eyvind H. Wichmann

    Charles Kittel, Cha i rman

    June, 1968 Berkeley, California

    Frank S. Crawford, Jr.

    Charles Kittel

    Walter D. Knight Alan M. Portis

    Frederick Beif

    Malvin A. Buderman

    Eyvind H. WichmannA. Carl Helmholz 1

    Edward M. PurcellJCha i rmen

  • 7/27/2019 Course on Waves

    10/632

    Preface to Volume I I I

    This volume is devo ted to the study of waves. Tha t is a broad subject.

    Eve ryon e knows many natural phenome na that involve waves there are

    water waves, sound waves, hght waves, radio waves, seismic waves,

    de Broglie waves, as well as other waves. Fur therm ore, perusal of the

    shelves of any physics library reveals that the study of a single facet of wave

    phenomenafor example, superson ic sound w aves in w a termay occupy

    whole books or periodicals and may even absorb the complete attention of individual scientists. Amazingly, a professional specialist in one of these

    narrow fields of study can usually communicate fairly easily with other

    supposedly narrow specialists in other supposedly unrelated fields. He has

    first to learn their slang, their units (like what a parsec is), and what num

    bers are important. Indeed, when he experiences a change of interest, he

    ma y becom e a narrow specialist in a new field surprisingly quickly. This

    is possible because scientists share a common language due to the remark

    able fact that many entirely different and apparently unrelated physical

    phenomena can be described in terms of a common set of concepts. Many

    of these shared concepts are implicit in the word w ave.

    The principal objective of this book is to develop an understanding of

    basic wave concepts and of their relations with one another. To that end

    the book is organized in terms of these concepts rather than in terms of

    such observable natural phenomena as sound, light, and so on.

    A complementary goal is to acquire familiarity with many interesting and

    important examples of waves, and thus to arrive at a concrete realization

    of the wide applicability and generality of the concep ts. After each new

    concept is introduced, therefore, it is illustrated by immediate application

    to many different physical systems: strings, slinkies, transmission lines, mail

    ing tubes, light beams, and so forth. This may be contrasted with the dif

    ferent approach of first developing the useful concepts using one simple

    example (the stretched string) and then considering other interesting

    physical systems.

    By choosing illustrative examples having geometric similitude with one

    another I hope to encourage the student to search for similarities and analogies betw een different wave phenomena. I also hope to stimulate him

    to develop the courage to usesuch analogies in hazarding a guess when

    confron ted with new phenomena. The use of analogy has well-known

    dangers and pitfalls, but so does everything. (The guess that light waves

    might be just like mechanical waves, in a sort of jelly-like ether was

    very fruitful; it helped guide Maxwell in his attempts to guess his famous

    equations. It yielded interesting predictions. Wh en experiments espe

    cially those of Michelson and Morley indicated that this mechanical model

  • 7/27/2019 Course on Waves

    11/632

    could not be entirely correct, Einstein showed how to discard the model yet

    keep Maxwells equations. Einstein preferred to guess the equationsdirectly what might be called pure guesswork. Nowadays, although

    most physicists still use analogies and models to help them guess new equa

    tions, they usually publish only the equations.)

    The home experiments form an importan t part of this volume. They can

    provide pleasure and insight of a kind not to be ac quired through the

    ordinary lecture demonstrations and laboratory experiments, important as

    these are. The home experiments are all of the kitchen physics type, re

    quiring little or no special equipment. (An optics kit is provided. Tuningforks, slinkies, and mailing tubes are not provided, but they are cheap and

    thus not special. ) These experiments really aremeant to be done at

    home, not at the lab. Many would be better termed demonstra t ionsrather

    than experiments.

    Every major concept discussed in the text is demonstrated in at least one

    home experiment. Besides illustrating con cepts, the home experiments give

    the student a chance to experience close personal contact with phe

    nomena. Because of the home aspect of the experiments, the contact is

    intimate and leisurely. This is importan t. There is no lab partn er who

    may pick up the ball and run with it while you are still reading the rules of

    the game (or sit on it when you want to pick it up); no instructor, explain

    ing the meaning of hi sdemonstration, when what you really need is to per

    form you rdemonstration, with your own hands, at your own speed, and as

    often as you wish.

    A very valuable feature of the home experiment is that, upon discovering

    at 10 p .m . that one has misunderstood an experiment done last week,

    by 10:15 p .m . one can have set it up once again and rep eate d it. This is

    important. Fo r one thing, in real experimental work no one ever gets it

    right the first time. Afterthoughts are a secre t of success. (There are

    others.) Nothing is mo re frustrating or more inhibiting to learning than

    inability to pursue an experimental afterthought because the equipment is

    torn down, or it is after 5 p .m . , or some other stupid reason.

    Finally, through the home experiments I hope to nurture what I may call

    an appreciation of phenom ena. I would like to beguile the student into

    creating with his own hands a scene that simultaneously surprises and delights his eyes, his ears, and his brain . . .

    Clear-colored stones

    are vibrating in the brook-bed . . .

    or the water is.

    S O S E K l f

    t Reprinted from The Four Seasons (tr. Peter Beilenson), copyright 1958, by The Peter

    Pauper Press, Mount Vernon, N.Y., and used by permission of the publisher.

  • 7/27/2019 Course on Waves

    12/632

    Acknow ledgments

    In its preliminary versions, Vol. Ill was used in several classes at Berkeley.

    Valuable criticisms and comments on the preliminary editions came from

    Berkeley students; from Berkeley professors L. Alvarez, S. Parker, A. Portis,

    and especially from C. Kittel; from J. C. Thompson and his students at the

    University of Texas; and from W . Walker and his students at the University

    of California at Santa Barbara. Extre m ely useful specific criticism was pro

    vided by S. Pasternacks attentive reading of the preliminary edition.

    Of particular help and influence were the detailed criticisms of W. Walker,

    who read the almost-final version.

    Luis Alvarez also contributed his first published experiment, A Simpli

    fied Method for Determination of the Wavelength of Light, School Sci ence

    and M a t h ema t i c s 32, 89 (1932), which is the basis for Home Exp. 9.10.

    I am especially grateful to Joseph Doyle, who read the entire final

    manuscript. His considered criticisms and suggestions led to many impor

    tant changes. He also introduced me to the Japanese haiku that ends the

    preface. He and another graduate student, Robert Fisher, contributed

    many fine ideas for hom e experimen ts. My daughter Sarah (age 4^) and

    son Matthew (2) not only contributed their slinkies but also demon

    strated that systems may have degrees of freedom nobody ever thought of.

    My wife Bevalyn contributed her kitchen and very much more.

    Publication of early preliminary versions was supervised by Mrs. Mary R.

    Maloney. Mrs. Lila Lowell supervised the last preliminary edition and

    typed most of the final manuscrip t. The illustrations owe their final form

    to Felix Cooper.

    I acknowledge gratefully the contributions others have made, but final

    responsibility for the manuscrip t rests with me. I shall welcom e any fur

    ther corrections, complaints, compliments, suggestions for revision, andideas for new home experiments, which may be sent to me at the Physics

    De partm ent, University of California, Berkeley, California, 94 720 . Any

    home experiment used in the next edition will show the contributors

    name, even though it may first have been done by Lord Rayleigh or

    somebody.

    F. S. Crawford, Jr.

  • 7/27/2019 Course on Waves

    13/632

  • 7/27/2019 Course on Waves

    14/632

    Teachi ng Notes

    Traveling waves have great aesthetic appeal, and it would be tempting to begin with

    them. In spite of their aesthet ic and mathem atical beauty, however, waves are physi

    cally rather complicated because they involve interactions between large numbers of

    particles. Since I want to emphasize physical systems rather than mathem atics, I

    begin with the simplest physical system, rather than with the simplest wave.

    Chap t er 1 Free O sci l la t i ons o f Sim p le Sys tems : W e first review the free oscillations

    of a one-dimensional harmonic oscillator, emphasizing the physical aspects of inertia

    and return force, th e physical meaning of

  • 7/27/2019 Course on Waves

    15/632

    W ha t to om i t : Sec. 2. 3 is optional especially if the students already know some

    Fourier analysis. Example 5 (Sec. 2.4 ) is a linear array of coupled pendulums, the

    simplest system having a low-fre quen cy cutoff. They are used later to help explain

    the behavior of other systems that have a low-frequency cutoff. A teach er who does

    not intend to discuss at a later time systems driven below cutoff (waveguide, iono

    sphere, total reflection of hght in glass, barrier penetration of de Broglie waves, high-

    pass filters, etc.) need not consider Example 5.

    Chap t er 3 Forced Osci l la t i ons: Chapters 1 and 2 started with free oscillations of a

    harmonic oscillator and ended with free standing waves of closed systems. In Chaps.

    3 and 4 we consider forced oscillations, first of closedsytems (Chap. 3) where we

    find resonances, and then in opensystems (Chap. 4) where we find traveling waves.

    In Sec. 3.2 we review the damped driven one-dimensional oscillator, considering its

    transient behavior as well as its stead y-state behavior. Then w e go to two or more

    degrees of freedom, and discover that there is a resonance corresponding to every

    mode of free oscillation. We also consider closed systems driven below their lowest

    (or above their highest) mode frequency and discover exponential waves and filtering

    action.

    W ha t t o om i t : Transients (in Sec. 3.2) can be omitted. Some teach ers may also

    wish to omit everything about systems driven beyond cutoff.

    H om e experi ments: Home Exps. 3.8 (Forced oscillations in a system of two coupled

    cans of soup) and 3.16 (Mechanical bandpass filter) require phonograph turntables.

    They make excellent class demonstrations, especially of exponential waves for systems

    driven beyond cutoff.

    Chap t er 4 Trave l ing W aves : Here we introduce t rave l ingwaves resulting from

    forced oscillations of an opensystem (contrasted with the s t a n d i n g waves resulting

    from forced oscillations of a closedsystem that we found in Chap. 3). The remainder

    of Chap. 4 is devoted to studying phase velocity (including dispersion) and impedance in traveling waves. We con trast the two traveling wave concepts, pha se ve loc i ty

    and impedance, with the standing wave concepts, i n e r t i a and retu r n fo rce, and also

    contrast the fundamental difference in phase relationships for standing versus traveling

    waves.

    H ome exper i ments : We recommen d Home Exp. 4.12 (Wate r prism). This is the

    first optics kit experiment; it uses the purple filter (which passes red and blue but cuts

    out green). We strongly recomm end Home Exp . 4.18 (Measuring the solar constant

    at the earths surface) with your face as detector.

    Chap t er 5 Re f lec t io n : By the end of Chap. 4 we have at our disposal both stand

    ing and traveling waves (in one dimension). In Chap. 5 we consider gen eral super

    positions of standing and traveling waves. In deriving reflection coefficients we make

    a very physical use of the superposition principle, rather than emphasizing bound

    ary conditions. (Use of boundary conditions is emphasized in the problems.)

    W ha t to om i t : There are many examples, involving sound, transmission lines, and

    light. Don t do them all! Cha pter 5 is essentially the application of what we have

    acquired in Chaps. 1 -4 . Any or all of it can be omitted.

    H ome exper i ments : Everyone should do Home Exp. 5.3 (Transitory standing waveson a slinky). Home Exps. 5 .17 and 5.18 are especially interesting.

  • 7/27/2019 Course on Waves

    16/632

    Chap t er 6 M odu la t ions , Pu lses, and W ave Packets : In Chaps. 1-5 we work mainly

    with a single frequency w (exc ept for Sec. 2.3 on Fourie r analysis). In Chap. 6 we

    consider superpositions, involving different frequencies, to form pulses and wave

    packets and to extend the concepts of Fourier analysis (developed in Chap. 2 for periodic functions) so as to include nonperiodic functions.

    W ha t t o om i t : Most of the physics is in the first three sections. A teach er who

    has omitted Fourier analysis in Sec. 2.3 will undoubtedly want to omit Secs. 6.4 and

    6.5, where Fourier integrals are introduced and applied.

    H om e experim ent s: No one believes in group velocity until they have watched water

    wave packets (see Home Exp. 6. 11). Everyone should also do Home Exps. 6.1 2 and

    6.13.

    Problems: Frequency and phase modulation are discussed in problems rather than

    in the text. So are such interesting recen t developments as Mode-locking of a laser(Prob. 6.23), Frequency multiplexing (Prob. 6.32), and Multiplex Interferometric

    Fourier Spectroscopy (Prob. 6.33).

    Chap te r7 W aves in Tw o and Three D im ens ions : In Chaps. 1-6 the waves are all

    one-dimensional. In Chap. 7 we go to three dimensions. The propagation vecto r k

    is introduced. Elec trom agn etic waves are studied using Maxwells equations as the

    starting point. (In earlier chapters there are many examples of electrom agnet ic waves

    in transmission lines, evolving from the LC -circu it example.) W ate r waves are alsostudied.

    W ha t t o om i t : Sec. 7.3 (Water Waves) can be omitted, but we recommend the

    home experiments on wat er waves wheth er or not Sec. 7.3 is studied. A teacher mainly

    interested in optics could actually start his course at Sec. 7.4 (Electromagnetic Waves)

    and continue on through Chaps. 7, 8, and 9.

    Chap t er 8 Po la r i z a t i o n : This chapte r is devoted to study of polarization of electro

    magnetic waves and of waves on slinkies, with emphasis on the physical relation between partial polarization and coherence.

    H ome exper i ments : Everyone should do at least Home Exps. 8.12, 8.14, 8.16, and

    8.18 (Exp. 8.14 requiring slinky; the others, the optics kit).

    Chap t er 9 In t er f e rence and D i f f r a c t i o n : Here we consider superpositions of waves

    that have traveled different paths from source to detec tor. We emphasize the physical

    meaning of coherence. Geometrical optics is treated as a wave phenomen on the

    behavior of a diffraction-limited beam impinging on various reflecting and refracting

    surfaces.

    H ome exper i men ts : Everyone should do at least one each of the many home ex

    periments on interference, diffraction, coherence, and geometrical optics. We also

    strongly recommend 9.50 (Quadrupole radiation from a tuning fork.)

    Problems: Some topics are developed in the problems: Stellar interferometers, in

    cluding the recently developed long-base-line interferometry (Prob. 9.57); the

    analogy between the phase-contrast microscope and the conversion of AM radio

    waves to FM is discussed in Prob. 9.59.

  • 7/27/2019 Course on Waves

    17/632

    Genera l r emar ks : At least one home experimen t should be assigned per week. For

    your convenience we list here all experiments involving water waves, waves in

    slinkies, and sound waves. We also late r describe the optics kit.

    W ater wav es : Discussed in Chap. 7; in addition they form a recurring theme devel-

    oped in the following series of easy home experiments:

    1.24 Sloshing modes in pan of water

    1.25 Seiches

    2.31 Sawtooth shallow-water standing waves

    2.33 Surface tension modes

    3.33 Sawtooth shallow-water standing waves

    3.34 Rectangular two-dimensional standing surface waves on water3.35 Standing waves in water

    6.11 Water wave packets

    6.12 Shallow-water wave packetstidal waves

    6.19 Phase and group velocities for deep-water waves

    6.25 Resonance in tidal waves

    7.11 Dispersion law for water waves

    9.29 Diffraction of water waves.

    Sl ink ies: Ev ery student should have a Slinky (about $1 in any toy store). Fo ur of

    the following experiments require a record-player turntable and are therefore outside

    the kitchen physics cost range. However, many students already have record

    players. (The experiments involving recor d players make good lectu re demonstrations.)

    1.8 Coupled cans of soup

    2.1 Slinkydependence of frequency on length

    2.2 Slinky as a continuous system

    2.4 Tone quality of a slinky

    3.7 Resonance in a damped slinky (turntable required)

    3.8 Forced oscillations in a system of two coupled cans of soup (turntable required)

    3.16 Mechanical bandpass filter (turntable required)3.23 Exponential penetration into reactive region (turntable required)

    4.4 Phase velocity for waves on a slinky

    5.3 Transitory standing waves on a slinky

    8.14 Slinky polarization

    Sound : Many home experiments on sound involve use of two identical tuning forks,

    preferably C52 3.3 or A440. The cheapest kind (about $1.2 5 each), which are

    perfectly adequ ate, are available in music stores. Mailing tubes can be purchased for

    about 25 0 ea ch in stationery or art-supply stores. The following home experimentsinvolve sound:

    1.4 Measuring the frequency of vibrations

    1.7 Coupled hacksaw blades

    1.12 Beats from two tuning forks

    1.13 Nonlinearities in your earcombination tones

    1.18 Beats between weakly coupled nonidentical guitar strings

    2.4 Tone quality of a slinky

    2.5 Piano as Fourier-analyzing machineinsensitivity of ear to phase

    2.6 Piano harmonicsequal-temperament scale3.27 Resonant frequency width of a mailing tube

  • 7/27/2019 Course on Waves

    18/632

    4.6 Measuring the velocity of sound with wave packets

    4.15 Whiskey-bottle resonator (Helmholtz resonator)

    4.16 Sound velocity in air, helium, and natural gas

    4.26 Sound impedance5.15 Effective length of open-ended tube for standing waves

    5.16 Resonance in cardboard tubes

    5.17 Is your sound-detecting system (eardrum, nerves, brain) a phase-sensitive detector?

    5.18 Measuring the relative phase at the two ends of an open tube

    5.19 Overtones in tuning fork

    5.31 Resonances in toy balloons

    6.13 Musical trills and bandwidth

    9.50 Radiation pattern of tuning forkquadrupole radiation

    Componen ts : Four linear polarizers, a circular polarizer, a quarter-wave plate, a

    half-wave plate, a diffraction grating, and four color filters (red, green, blue, and

    purple). The comp onents are described in the text (linear polarizer on p. 41 1; c ircu

    lar polarizer, p. 433; quarter- and half-wave retardation plates, p. 435; diffraction

    grating, p. 49 6). Some experiments also require microscop e slides, a showcase-lamp

    line source, or a flashlight-bulb point source as described in Home Exp. 4.12, p. 217.

    Aside from Exp. 4.12, all experiments requiring the optics kit are in Chaps. 8 and 9.They are too numerous to list here.

    The first experiment involving the complete optics kit should be identification of all

    the components by the student. (Compon ents are listed on the envelope container

    glued to the inside back cove r.) Label the components in some way for future refer

    ence. Fo r example, use scissors to round off slightly the four corne rs of the circular

    polarizer, and then scratch IN near one edge of the input face or stick a tiny piece

    of tape on that face. Clip onecorner of the one-quarter-wave retarder; clip tw ocorners

    of the two-q uarter- (half-) wave retarder. Scratch a line along the axis of easy transmission on the linear polarizers. (This axis is parallel to one of the edges of the polarizer.)

    We should rem ark that the quarter-wave plate gives a spatial retardation of

    1400 2 0 0 A, nearly independent of wavelength (for visible light). Thus the wave

    length for which it is a quarter-wav e retarde r is 560 0 800 A. The 2 0 0 A is

    the manufacturers tolerance. A manufactured batch that gives retardation 140 0 A

    is a quarter-wave retarder for green (5600 A), but it retards by less than one quarter-

    wave for longer wavelengths (red) and more for shorter (blue). Another batch that

    happens to give retardation 1400 + 2 0 0 = 1600 A is a quarter-wave retarder only for

    red (6400 A). One that retards by 1400 2 0 0 is a quarter-wave plate only for blue

    (4800 A). Similar remarks apply to the circular polarizer, since it consists of a sand

    wich of quarter-wave plate and linear polarizer at 45 deg, and the quarter-wave plate

    is a retarder of 1400 200 A. Thus there may be slightly distracting color effects

    when using white light. The student must be warned that in any experim ent where

    he is supposed to get black, i.e., extinction, he will always have some non-extin-

    guished light of the w rong color leaking through. Fo r example, I was naive when

    I wrote Home Exp. 8 .12. You should perhaps strike out everything after the word

    band in the sentences Do you see the dark band at green? That is the color of5600 A!

    Op t i cs K i t

    Home experiment

  • 7/27/2019 Course on Waves

    19/632

    Use o f Compl ex N um bers Complex numbers simplify algebra when sinusoidal oscillations or waves are to be

    superposed. They may also obscure the physics. Fo r that reason I have avoidedtheir use, especially in the first part of the book. All the trigonom etric identities that

    are needed will be found inside the front cover. In Chap. 6 1 do make use of the complex

    representation exp ioot, so as to use the well-known graphical or phasor diagram

    method of superposing vibrations. In Chap. 8 (Polarization) I use complex quantities

    extensively. In Chap. 9 (Interfere nce and Diffraction) I do not make much use of com

    plex quantities, even though it would sometimes simplify the algebra. Many teachers

    may wish to make much more extensive use of complex numbers than I do, especially

    in Chap. 9. In the sections on Fourier series (Sec. 2.3 ) and Fourie r integrals (Secs.

    6.4 and 6.5), I make no use of complex quantities. (I especially wanted to avoid Fourier

    integrals involving negative frequencies!)

    A Not e on the M K S Sys tem o f E lect r i ca l U n i t sf

    t Reprinted from Berkeley Physics Course,

    Vol. II, Electricity and Magnetism, by Ed

    ward M. Purcell, 1963, 1964, 1965 by

    Education Development Center, Inc., suc

    cessor by merger to Educational Services

    Incorporated.

    Most textbooks in electrical engineering, and many elementary physics

    texts, use a system of electrical units called the ra t i ona l i zed M K S system.

    This system employs the MKS mechanical units based on the meter, the

    k i l og ram, and the second. The MKS unit of force is the new t on , defined

    as the force which causes a 1-kilogram mass to accelerate at a rate of

    1 m ete r/se c2. Thus a newton is equivalent to exactly 1 05 dynes. The

    corresponding unit of energy, the newton-meter, orj oul e, is equivalent to

    107 ergs.

    The electrical units in the MKS system include our familiar practical

    units coulomb, volt, ampere, and ohm along with some new ones.

    Someone noticed that it was possible to assimilate the long-used practicalunits into a com plete system devised as follows. Write Coulom bs law as

    we did in Eq. 1.1:

    F = k3132

    *21(1)

    Instead of setting kequal to 1, give it a value such that F2 will be given in

    newtons if q1 and 92 are expressed in coulombs and r21 in meters. Know

    ing the relation between the newton and the dyne, between the coulomb

    and the esu, and between the meter and the centimeter, you can easily

    calculate that k must have the value 0.8 98 8 X 1010. (Two 1-coulomb

    charges a meter apart produce quite a forc e around a million tons!) It

    makes no difference if we write 1 /( 4 weo) instead of k , where the constant

    c0 is a number such that l / ( 477c0) = k= 0 .89 88 X 1 010. Coulombs law

    now reads:

    1 qiq2

    ^ 477o r2

    with the constant c0 specified as

    Co = 8.8 54 X 10 -1 2 coulomb2/n ew to n-m 2

    (2)

  • 7/27/2019 Course on Waves

    20/632

    Separating out a factor l/4 w was an arbitrary move, which will have the

    effect of removing the 477 that would appear in many of the electrical

    formulas, at the price of introducing it into some others, as here in

    Cou lomb s law. That is all that rationalized means. The constant e0

    is called the dielectric constant (or permittivity) of free space.

    Electric potential is to be measured in volts, and electric field strength E

    in volts/m eter. The force on a charge q, in field E, is:

    F (newtons) = r/E (coulombs X volts/meter) (4)

    An ampere is 1 cou lom b/se c, of course. The force per me ter of length

    on each of two parallel wires, r meters apart, carrying current Imeasured

    in amperes, is:

    f (ne wton s/m eter) = ( - g ) M (5 )

    Recalling our CGS formula for the same situation,

    / ( dy nes /cm ) = ^ ( esu^ )Z21 (6 )r c2 (cm:y se c2)

    we compute that (jiio/477) must have the value 10~7. Thus the constant jUo,called the permeability of free space, must be:

    ju,0 = 477 X 10~7 newtons/amp2 (exactly) (7)

    The magn etic field B is defined by writing the L ore ntz force law as follows:

    F (newtons) = c/E + qvX B (8 )

    where v is the velocity of a particle in me ters/s ec, qits charge in coulombs.

    This requires a new unit for B. The unit is called a tesla, or a w eber /m2.One tesla is equivalent to precisely 104 gauss. In this system the auxiliary

    field H is expressed in different units, and is related to B, in free space, in

    this way:

    B = ju0H (in free space) (9)

    The relation of H to the free current is

    J H ds = /free (10)

    /free being the free current, in amperes, enclosed by the looparound which

    the line integral on the left is taken. Since dsis to be measured in meters,

    the unit for H is called simply, ampere /me te r .

    Maxwells equations for the fields in free space look like this, in the

    rationalized MKS system:

    div E = p curl E

    div B = 0 curl B = Moo

    8B

    31

    SE

    01+ M0J

  • 7/27/2019 Course on Waves

    21/632

    If you will com pare this with our G aussian CGS version, in which c appears

    out in the open, you will see that Eqs. 11 imply a wave velocity l/\/oMo

    (in me ters/ sec , of course). That is:

    oMo = \ (12)c

    In our Gaussian CGS system the unit of charge, esu, was established by

    Coulombs law, with k= 1. In this MKS system the coulom b is defined,

    basically, not by Eq. 1 but by Eq. 5, that is, by the force between currents

    rather than the force betw een charges. Fo r in Eq . 5 we have ju,o= 4 7 7 x 10~7.

    In other words, if a new experimental measurement of the speed of lightwere to change the accepted value of c, we should have to revise the value

    of the constant o, not that of ju

  • 7/27/2019 Course on Waves

    22/632

    Contents Pr eface to th e Berkeley Phy sics Course vA Fu r t h er No t e vi

    Preface to Volu me I I I vii

    Acknow ledgmen ts ixTeach ing N otes xi

    A Not e on the M K S Sys tem o f E lec t r i ca l Un i ts xvi

    Chap t er 1 Fr ee O sc i l l a t i onso f Sim p le Sys tems 1

    1.1 Introduction 2

    1.2 Free Oscillations of Systemswith One Degreeof Freedom 3

    1.3 Linearity and the Superposition Principle 12

    1.4 Fre e Oscillations of Systems with Two Degrees of Freedom 16

    1.5 Beats 28

    Problems and Hom e Experiments 36

    Chap te r 2 Fr ee O sc i l l a t i ons o f Sys tems w i t h M any D egrees o f

    Freedom 47

    2.1 Introduction 48

    2.2 Transverse Modes of Continuous String 50

    2.3 General Motion of Continuous String and Fou rier Analysis 59

    2.4 Modes of a Noncontinuous System with NDegrees of Freedom 72

    Problems and Home Experiments 90

    Chap t er 3 For ced O sc i l l a t i ons 101

    3.1 Introduction 102

    3.2 Dam ped Driven One-dimensional Harm onic Oscillator 102

    3.3 Besonances in System with Two Degrees of Freedom 116

    3.4 Filters 122

    3.5 Forced Oscillations of Closed System with Many Degrees of

    Freedom 130

    Problems and Home Experiments 146

    Chap t er 4 Trav el i ng W aves 155

    4.1 Introduction 156

    4.2 Harmonic Traveling Waves in One Dimension and Phase

    Velocity 157

    4.3 Index of Befraction and Dispersion 176

    4.4 Impedance and Energy Flux 191

    Problems and Home Experim ents 214

  • 7/27/2019 Course on Waves

    23/632

    Chap te r 5 Re f lec t i on 225

    5.1 Introduction 226

    5.2 Perfect Termination 22 6

    5.3 Reflection and Transmission 23 3

    5.4 Impedan ce Matching between Two Transparent Media 24 5

    5.5 Reflection in Thin Films 24 9

    Problems and Home Experiments 25 2

    Chap te r6 M odu la t i ons , Pu lses , and W ave Packets 267

    6.1 Introduction 268

    6.2 Group Velocity 26 86.3 Pulses 279

    6.4 Four ier Analysis of Pulses 29 5

    6.5 Fourier Analysis of Traveling Wave Packet 308

    Problems and Home Experimen ts 311

    Chap te r7 W aves in Two and Thr ee D im ens ions 331

    7.1 Introduction 3327.2 Harmonic Plane Waves and the Propagation Vector 332

    7.3 Water Waves 346

    7.4 Electromagnetic Waves 355

    7.5 Radiation from a Point Charge 36 6

    Problems and Home Experimen ts 381

    Chapter8 Po la r i za t i on 393

    8.1 Introduction 394

    8.2 Description of Polarization States 39 5

    8.3 Production of Polarized Transverse Waves 40 7

    8.4 Double Refraction 41 9

    8.5 Randwidth, Cohe rence Time, and Polarization 42 7

    Problems and Home Experiments 437

    Chap ter 9 I n t er f e r ence and D i f f r a c t i o n 451

    9.1 Introduction 453

    9.2 Interference between Two Coherent Point Sources 454

    9.3 Interference between Two Independent Sources 466

    9.4 How Large Can a Point Light Source Re? 470

    9.5 Angular Width of a Ream of Traveling Waves 473

    9.6 Diffraction and Huygens Principle 47 8

    9.7 Geometrical Optics 498

    Problems and Home Experiments 519

  • 7/27/2019 Course on Waves

    24/632

    Supp l em ent a ry Top i cs 545

    1 M icroscopic Exam ples of Weakly Coupled Identical

    Oscillators 546

    2 Dispersion Relation for de Broglie Waves 54 8

    3 Penetration of a Particle into a Classically Forbidden Region ofSpace 552

    4 Phase and Group Velocities for de Broglie Wav es 55 5

    5 Wave Equations for de Broglie Waves 556

    6 Electrom agne tic Radiation from a One-dimensional Atom 557

    7 Time Coherence and Optical Beats 558

    8 W hy Is the Sky Bright? 55 9

    9 Electro ma gne tic Wave s in Material Media 563

    A ppend i c es 585

    Supp l ementa r y Read i ng 59 1

    I ndex 593

    Opt i cs K i t , Tables o f U ni ts , Val ues, an d U sefu l Constants and

    I dent i t ies I ns ide covers

    O pt i ca l Spect ra fo l l ow ing page528

  • 7/27/2019 Course on Waves

    25/632

  • 7/27/2019 Course on Waves

    26/632

    waves

  • 7/27/2019 Course on Waves

    27/632

  • 7/27/2019 Course on Waves

    28/632

    Chapt er 1

    Free Osci l lat ions o f Simple Systems

    1.1 Introduction 2

    1.2 Fre e Oscillations of Systems with One Degree of Freedom

    Nomenclature 3

    Return force and inertia

    Oscillatory behavior 4

    Physical meaning of u 2 4

    Damped oscillations 5Example 1: Pendulum 5

    Example 2: Mass and springslongitudinal oscillations

    Example 3: Mass and springstransverse oscillations 7

    Slinky approximation 9

    Small-oscillations approximation 9

    Example 4: LC circuit 11

    1.3 Linearity and the Superposition Principle 12

    Linear homogeneous equations 13

    Superposition of initial conditions 14Linear inhomogeneous equations

    Example 5: Spherical pendulum

    14

    15

    1.4 Fre e Oscillations of Systems with Two Degrees of Freed om 16

    Properties of a mode 16

    Example 6: Simple spherica l pendulum 17Example 7: Two-dimensional harmonic oscillator

    Normal coordinates 19

    17

    Systematic solution for modes 20

    Example 8: Longitudinal oscillations o f two coupled m assesExample 9: Transverse oscillations o f two coupled masses

    Example 10: Two coupled LC circuits 27

    2125

    1.5 Beats 28

    Modulation 29

    Almost harmonic oscillation 29

    Example 11: Beats produced by two tuning forksSquare-law detector 30

    30

    Example 12: Beats between two sources o f visible light 31

    Example 13: Beats between the two normal modes o f two weakly

    coupled identical oscillators 32

    Esoteric examples 36

    Problems and Hom e Experiments 36

  • 7/27/2019 Course on Waves

    29/632

    Chapt er 1 Free Osci l l ati ons o f Simpl e Systems

    1 .1 In t roduc t i on

    The world is full of things that move. Their motions can be broadly cate

    gorized into two classes, according to whether the thing that is moving

    stays near one place or travels from one place to another. Examples of the

    first class are an oscillating pendulum, a vibrating violin string, water slosh

    ing back and forth in a cup, electrons vibrating (or whatever they do)

    in atoms, hght bouncing back and forth between the mirrors of a laser.

    Parallel examples of traveling motion are a sliding hockey puck, a pulse

    traveling down a long stretched rope plucked at one end, ocean waves roll

    ing toward the beach, the electron beam of a TV tube; a ray of hght

    emitted at a star and detec ted at your eye. Sometimes the same phenom

    enon exhibits one or the other class of motion (i.e., standing still on the

    average, or traveling) depending on your point of view: the ocean waves

    travel toward the beach, but the water (and the duck sitting on the surface)

    goes up and down (and also forward and backward) without traveling.

    The displacement pulse travels down the rope, but the material of the rope

    vibrates without travehng.

    We begin by studying things that stay in one vicinity and oscillate or vi

    brate about an ave rage position. In Chaps. 1 and 2 we shall study many

    examples of the motion of a closed system that has been given an initial

    excitation (by some external disturbance) and is thereafter allowed to oscil

    late freely without further influence. Such oscillations are called f r e e or

    na tu ra l osci l l a t i ons. In Chap. 1 study of these simple systems having one or two moving parts will form the basis for our understanding of the free

    oscillations of systems with many moving parts in Chap. 2. There we shall

    find that the motion of a complicated system having many moving parts

    may always be regarded as compounded from simpler motions, called

    modes, all going on at once. No matter how comp licated the system, we

    shall find that each one of its modes has properties very similar to those of

    a simple harmon ic oscillator. Thus for motion of any system in a single

    one of its modes, we shall find that each moving part experiences the samereturn force per unit mass per unit displacement and that all moving parts

    oscillate with the same time dependence cos (u t +

  • 7/27/2019 Course on Waves

    30/632

    tion. The displacement is described by a vect or \ j/(x,y,x,t). Sometimes we

    call this vector function of x, y , z, ta w ave f u n c t i o n . (It is only a contin

    uous function of x, y , and zwhen we can use the continuous approxima

    tion, i.e., when near neighbors have essentially the same motion.) In some

    of the electrical examples, the physical quantity may be the current in a

    coil or the charge on a capac itor. In others, it may be the electric field

    E (x,y,z,t)or the magnetic field B(x,y,z,t) . In the latter cases, the waves are

    called electromagnetic waves.

    1 .2 Free O sc i l l a t i ons o f Sys tems w i t h O ne D egree o f Freedom

    W e shall begin with things that stay in one vicinity , oscillating or vibratingabout an average position. Such simple systems as a pendulum oscillating

    in a plane, a mass on a spring, and an LC circuit, whose configuration at

    any time can be completely specified by giving a single quantity, are said

    to have one degree of freedomloosely speaking, one moving part (see

    Fig. 1.1). Fo r example, the swinging pendulum can be described by the

    angle that the string makes with the vertical, and the L C circuit by the

    charge on the capac itor. (A pendulum free to swing in any direction, like

    a bob on a string, has not one but two degrees of freedom; it takes two coordinates to specify the position of the bob. The pendulum on a grand

    father clock is constrained to swing in a plane, and thus has only one de

    gree of freedom.)

    For all these systems with one degree of freedom, we shall find that the

    displacement of the moving part from its equilibrium value has the

    same simple time dependence (called harm on ic osci l l a t i on ),

    \ p(t)= A cos (a t +

  • 7/27/2019 Course on Waves

    31/632

    second, or hertz (abbreviated cps, or Hz). The inverse of vis called the

    per i od T, which is given in seconds per cycle:

    T = X . (2)

    Th e pha se constant

  • 7/27/2019 Course on Waves

    32/632

    D am ped osci l l a t i o n s . If left undisturbed, an oscillating system will con

    tinue to oscillate forever in accord anc e with Eq. (1). However, in any real

    physical situation, there are frictional, or resistive, processes which damp the motion. Thus a more realistic description of an oscillating

    system is given by a damped oscillation. If the system is excited into

    oscillation at t = 0 (by giving it a bump or closing a switch or something),

    we find (see Vol. I, Chap. 7, page 209)

    xp(t) = A e~t /2rcos (ttf +

  • 7/27/2019 Course on Waves

    33/632

    F ig . 1 .3 L ong i t ud i na l osci l l a t ions.

    (a) Spr ings re laxed and unattached.

    (b ) Spr ings a t t ached , M a t equ i l i b r i um

    posi t ion , (c ) Genera l conf igu ra t ion .

    If we retain only the first term in Eq. (6 ), then Eq. (5) takes on the form

    ! r = (7)where

    2 = f (8)

    The general solution of Eq. (7) is the harmonic oscillation given by

    \ p(f) A cos ( t +

  • 7/27/2019 Course on Waves

    34/632

    total force Fzin the + zdirection is the superposition (sum) of these two

    forces:

    F* K (z ao) + K (2a z ao)

    = 2 K (z a).

    Newtons second law then gives

    ^ = Fz - 2 K ( z a). (9)

    The displacement from equilibrium is z a. We designate this by \ p(t):

    44*) = z(t) ~ a.then

    d 2xp _ d2z

    d t2 dt2

    Now we can write Eq. (9) in the form

    = - < * . (10)

    with

    2 =2K

    M(11)

    The general solution of Eq. (10) is again the harmonic oscillation

    = A co s (u t+ (p). Note that Eq. (11) has the form to2 = force per unit

    displacement per unit mass, since the return force is 2 K\ pfor a displace

    ment xp.

    Example 3: Mass and springstransverse oscillations

    The system is shown in Fig. 1.4. Mass M is suspended betw een rigid sup

    ports by means of two identical springs. The springs each have zero mass,

    spring constant K , and unstretched length ao They each have length aat

    the equilibrium position of M. W e neg lect the effect of gravity. (Gravity

    does not produce any return force in this problem. It does cause the sys

    tem to sa g, but tha t does not affect the results in the order of approximation that we are interested in.) Mass M now has three degrees of freedom:

    It can move in the zdirection (along the axis of the springs) to give longi

    tudinal oscillation. Tha t is the motion we considered above, and we need

    not repea t those considerations. It can also move in the xdirection or in

    the ydirection to give transv erse oscillations. Fo r simplicity, let us con

    sider only motion along x. We may imagine that there is some frictionless

    constraint that allows complete freedom of motion in the transverse x di

    rection but prevents motion along either yor z. (For example, we could

  • 7/27/2019 Course on Waves

    35/632

    Fig. 1.4 Transverse oscil lat i ons.

    (a) Equil ibr ium configurati on, (b) Gen-

    eral configurati on (for moti on along x).

    drill a hole through M and arrange a frictionless rod passing through the

    hole, rigidly attached to the walls, and oriented along x. However, you

    can easily convin ce yourself that such a constraint is unnecessary. From

    the symmetry of Fig. 1.4, you can see that if at a given time the system is

    oscillating along x, there is no tendency for it to acquire any motion along

    yor z. The same circumstance holds true for each of the other two de

    grees of freedom: no unbalanced force along xor y is developed due tooscillation alongz, nor alongxor zdue to oscillation along y .)

    At equilibrium (Fig. 1.4a), each of the springs has length aand exerts a

    tension T0, given by

    T0 = K (a a0). (12)

    In the general configuration (Fig. 1.4b), each spring has length Iand tension

    T = K ( l a0). (13)

    This tension is exerte d along the axis of the spring. Taking the x compo

    nent of this force, we see that each spring contributes a return force Tsin 6

    in the x direction. Using Newtons second law and the fact that sin 6 is

    x / l , we find

    M d?x _ F x _ _ 2 Tsin 8a t z

  • 7/27/2019 Course on Waves

    36/632

    Equation (14) is exact, under our assumptions (including the assumption,

    expressed by Eq. (13), that the spring is a linear or Hookes law spring).

    Notice that the spring length Iwhich appears on the right side of Eq. (14)

    is a function of x. Therefore Eq. (14) is not exactly of the form that gives

    rise to harmonic oscillations, because the return force on M is not exactly

    hnearly proportional to the displacement from equilibrium, x.

    Sl i n k y a pp r o x im a t i o n . There are two interesting ways in which we can

    obtain an approximate equation with a linear restoring force. The first way

    we shall call the sl i nky approx ima t ion , in which we neglect ao /acompared

    to unity. Henc e, since Iis always greater than a, we neglect ao / lin Eq. (14).

    [A shnky is a helical spring with relaxed length ao about 3 inches. It canbe stretched to a length aof about 15 feet without exceeding its elastic

    limit. That would give ao /a< 1 /6 0 in Eq. (14).] Using this approxima

    tion, we can write Eq. (14) in the form

    This has the solution x= A cos (cot + cp), i.e., harmonic oscillation. Notice

    that there is no restriction on the amplitude A. We can have large oscil

    lations and still have perfect linearity of the return force. Notice also that

    the frequency for transverse oscillations, as given by Eq. (16), is the same

    as tha t for longitudinal oscillations, as given by Eq. (11). Tha t is not true

    in general. It holds only in the shnky approximation, where we effectively

    take ao= 0 .

    Sma l l osci l l a t i ons approx im a t i on . If aocannot be neglected with respect

    to a(as is the case, for example, with a rubber rope under the conditions

    ordinarily met in lecture demonstrations), the shnky approximation does

    not apply. Then Fxin Eq. (14) is not hnear in x. However, we shall show

    that if the displacements xare small compared with the length a, then I

    differs from aonly by a quantity of order a(x /a )2 . In the smal l osc i l la t ions

    app rox ima t i on , we neglect the terms in Fxwhich are nonlinear in x / a . Let

    us now do the algebra: We want to express Iin Eq. (14) as I= a + some

    thing, where something vanishes when x 0. Since Iis larger than a,

    whether xis positive or negative, something must be an even function

    of x. In fact we have from Fig. 1.4

    (15)

    with

    (16)

    I2 = a2 + x2

    a2( l + c),

  • 7/27/2019 Course on Waves

    37/632

    Thus

    1 = 1(1 +l a

    = H 1 - ( r ) + ( H * + a( f ) 3 + - -

    Discarding the cubic and higher-order terms, we obtain

    d2x_ 2K 2 T0x . . . .

    Therefore x(t)is given by the harmonic oscillation

    x(t)= A cos (cot -)-

  • 7/27/2019 Course on Waves

    38/632

    'mass is M. Thus the retu rn force :per unit displacem ent pe r unit mass is

    ~2To(x/a)/xM .

    Notice that the frequency for transverse oscillations is given by to2 =

    2T o /M a for both the case of the shnky approximation (a 0 = 0) and the

    small-oscillations approximation ( x / a

  • 7/27/2019 Course on Waves

    39/632

    Because of Eqs. (25) and (26), there is only one degree of freedom. We

    can describe the instantaneous configuration of the system by giving @i, or

    Q2, or 7. The cur ren t Iwill be most convenient in our later work (whenwe go to systems having more than one degree of freedom), and we shall

    use it here. W e first use Eq . (25) to eliminate Qi from Eq . (24); the n we

    differentiate with respect to tand use Eq. (26) to eliminate >2:

    Lft=C_1

  • 7/27/2019 Course on Waves

    40/632

    nonlinear, as we can see from the expansion of sin t[/given by Eq. (6 ). Only

    when we neglect the higher powers of xpdo we obtain a linear equation.

    Nonlinear equations are generally difficult to solve. (The nonlinear pen

    dulum equation is solved exactly in Volume I, pp. 22 5 ff.) Fortunately, there are many interesting physical situations for which linear equations

    give a very good approximation. We shall deal almost entirely with linear

    equations.

    L i n ea r h om ogeneous equa t i ons . Linear homogeneous differential equa

    tions have the following very interesting and important property: Th e sum

    of any tw o so lu t i ons is i t sel f a so lu t i on . Nonlinear equations do not have

    that prop erty. The sum of two solutions of a nonlinear equation is not it

    self a solution of the equation.

    We shall prove these statem ents for both cases (linear and nonlinear) at

    once. Suppose that we have found the differential equation of motion of

    a system with one degree of freedom to be of the form

    = CxP + o p + W + y r + ( 2 8 )

    as we found, for example, for the pendulum [Eqs. (5) and ( 6 )] or for the

    transverse oscillations of a mass suspended by springs [Eq. (19)]. If the

    constants a , (3, y , etc. are all zero or can be taken to be zero as a sufficiently

    good approximation, then Eq. (28) is linear and homogeneous. Otherwise,

    it is nonlinear. Now suppose that ipi(t) is a solution of Eq. (28) and that

    \ ^2(t )is a different solution. For example, ipimay be the solution corre

    sponding to a particular initial displacement and initial velocity of a pendu

    lum bob, and xp2 may correspond to different initial displacement and velocity. By hypothesis \ piand xp2 each satisfy Eq. (28). Thus we have

    CxPi + a\ pi2 + A f c3 + y 4 + -, (29)

    C\ p2 + axp22 + P^23 + yxp24 + (30)

    The question of interest to us is whether or not the supe rpos i t i onof and

    ip2, defined as the sum \ p(t) = \ pi(t)+ xp2( t) , satisfies the same equation of

    motion, Eq. (28). Do we have

    ^'^2~ = + 4*2)+ ('/'l + >/'2)2 + + ife)3 + ? (31)

    The question (31) has the answer yes if and only if the constants a, j},

    etc. are zero. That is easily shown as follows. Add Eqs. (29) and (30).

    d^h_

    d t2

    and

    d2\ p2 _

  • 7/27/2019 Course on Waves

    41/632

    The sum gives Eq. (31) if and only if all the following conditions are

    satisfied:

    Equations (32) and (33) are both true. Equations (34) and (35) are not true

    unless aand f iare zero. Thus we see that the superposition of two solu

    tions is itself a solution if and only if the equation is linear.

    The property that a superposition of solutions is itself a solution is unique

    to homogeneous linear equations. Oscillations that obey such equations are

    said to obey the superpos i t ion pr inc i p l e. We shall not study any other kind.

    Superposi t i o n o f i n i t i a l cond i t i o n s. As an example of the applications of

    the concept of superposition, consider the motion of a simple pendulum

    under small oscillations. Suppose that one has found a solution corre

    sponding to a certain set of initial conditions (displacement and velocity)

    and another solution \ p2 corresponding to a different set of initial conditions.

    Now suppose we prescribe a third set of initial conditions as follows: We

    superpose the in i t i a l cond i t i onscorresponding to \ piand ip2That means

    that we give the bob an initial displacement that is the algebraic sum of the

    initial displacement corresponding to the motion \ pi(t)and that correspond

    ing to \p2(t), and we give the bob an initial velocity that is the algebraic sum

    of the two initial velocities corresponding to and \ p2 Then there is no

    need to do any m ore work to find the new motion, described by ip3(t). The solution \p3 is just the superposition ip1 + xp2 We let you finish the proof.

    This result holds on l yif the pendulum oscillations are sufficiently small so

    that we can neglect the nonlinear terms in the return force.

    L i nea r inhom ogeneous equa t i ons . Linear inhomogeneous equations (i.e.,

    equations containing terms independent of \ p)also give rise to a superposi

    tion principle, though of a slightly different sort. There are many physical

    situations analogous to a driven harmonic oscillator, which satisfies theequation

    where F(t)is an external driving force that is independent of \ p(t). The

    corresponding superposition principle is as follows: Suppose a driving force

    F\ (t ) produces an oscillation \ pi( t )(when F\ is the only driving force), and

    suppose another driving force F2(t) produces an oscillation xp2(t) [when F^t)

    dt2 dt2 dt2

    C\ p1 C\ p2 = C(\ p1 + \ p2),

    a\ p12 + a\ p22 = a(\ pi+ \ f2)2,

    /Sxpi3 + /i\ p 23 + M3, etc-

    d 2xpi d2\p2 _ d2(xpi+ Xp2)(32)

    (33)

    (34)

    (35)

    (36)

    i t b it lf] Th if b th d i i f t i lt l

  • 7/27/2019 Course on Waves

    42/632

    is present by itself]. Then, if both driving forces are present simultaneously

    [so that the total driving force is the superposition F i ( t )+ F2(t)], the corre

    sponding oscillation [i.e., corresponding solution of Eq. (36)] is given by the

    superposition \ p(t) = 'pii.t)+

    - - * * PS)

    These two equations are uncoupled, by which we mean that the xcom

    ponent of force depends only on x, not on y , and vice versa. Thus Eq . (37)

    does not contain y , and similarly Eq. (38) does not contain x. Equations

    (37) and (38) can be solved independently to give

    x(t) =A i cos (cot +

  • 7/27/2019 Course on Waves

    43/632

    Fig. 1.6 Systems w i t h tw o degrees o f

    f r eedo m . (T he masses are const ra in ed to

    r ema i n i n t he p l ane o f t he f i gu r e .)

    1 .4 Free O sc i l l a t ions o f Sys tems w i th Tw o D egrees o f Freedom

    In nature there are many fascinating examples of systems having two

    degrees of freedom . The mo st beautiful examples involve molecules and

    elementary particles (the neutral Kmesons especially); to study them re

    quires quantum m echanics. Some simpler examples are a double pendu

    lum (one pendulum attached to the ceiling, the second attached to the bob

    of the first); two pendulums coupled by a spring; a string with two beads;

    and two coupled L Ccircuits. (See Fig. 1.6.) It takes two variables to de

    scribe the configuration of such a system, say \paand \ pj,. For example, in

    the case of a simple pendulum free to swing in any direction, the moving

    parts tpaand \ pj, would be the positions of the pendulum in the two perpendicular horizontal directions; in the case of coupled pendulums, the

    moving parts xpaand \pbwould be the positions of the pendulums; in the

    case of two coupled L Ccircuits, the moving parts xpaand \ pbwould be

    the charges on the two capacitors or the currents in the circuits.

    The general motion of a system with two degrees of freedom can have a

    very complicated appearance; no part moves with simple harmonic motion.

    However, we will show that for two degrees of freedom and for linear

    equations of motion the most general motion is a supe rpos i t ionof two independent simple harm onic motions, both going on simultaneously. These

    two simple harmonic motions (described below) are called no rm a l m odes

    or simply modes. By suitable starting conditions (suitable initial, values of

    xpa, M dxpa/d t , and dipb/dt) , we can get the system to oscillate in only one

    mode or the other. Thus the modes are uncou pled, even though the

    moving parts are not.

    Prope r t ies o f a mode . When only one mode is present, each moving part

    undergoes simple harmonic motion. All parts oscillate with the same fre

    quen cy. All parts pass through their equilibrium positions (where \pis zero)

    simultaneously. Thus , for example, one never has in a single mode,

    \pa(t) = A cosoof and M t ) = B sin cot (different phase constants) or

    \pa(t) = A cos u\ t and xph(t ) = Bcos co2t (different frequ encies). Instead

    one has, for one mode (which we call mode 1),

    M t ) = M cos (i t+

  • 7/27/2019 Course on Waves

    44/632

    Each mode has its own characteristic frequency: toi for mode 1, t02 for

    mode 2. In each mode the system also has a characteristic configuration

    or shape, given by the ratio of the amplitudes of motion of the moving

    parts: A 1/B .1 for mode 1 and A 2/ B2 for mode 2. Note that in a mode theratio is constan t, independent of time. It is given by the appro

    priate ratio A 1/ B1 or A 2/B 2, which ca n be either positive or negative.

    The most general motion of the system is (as we will show) simply a

    superposition with both modes oscillating at once:

    \pa(t) = A t cos ( u i t +

  • 7/27/2019 Course on Waves

    45/632

    F ig . 1 .7 Tw od imensiona l harm on ic

    osci l la t or , (a) Equi l i br i um , (b) General

    con f i gu r a t i on .

    M ~ = - 2 K i X , and \ = - 2 K2y , (45)

    which have the solutions

    2Kx Ai cos (wit +

  • 7/27/2019 Course on Waves

    46/632

    general, is still not as general in app eara nce as Eqs. (43). Tha t is because

    we were lucky! Our natural choic e for xand yalong the springs gave us

    the uncoupled equations (45), each of which corresponds to one of the

    modes. In terms of Eq . (43), we cam e out with \paluckily chosen so that

    A 2 came out identically zero and with \ pbchosen so that Bi came out iden

    tically zero. Our fortunate choic e of coordinates gave us what are called

    no rma l coo rd i na t es ;in this example the normal coordinates are xand y.

    Suppose we had not been so lucky or so wise. Suppose we had used a

    coordinate system x' and y ' related to x and y by a rotation through

    angle a , as shown in Fig. 1.8. By inspection of the figure we see that the

    normal c oord inate X; is a linear co mb ination of the coordina tes x'and i f ,

    as is the other normal coordinate, y. If we had used the dumb coordi

    nates x!and y 'instead of the smart coordinates xand y , we would have

    obtained two coupled differential equations, with both x'and y 'appear

    ing in each equation, rather than the uncoupled equations (5).

    In most problems involving two degrees of freedom it is not easy to find

    the normal coordinates by inspection, as we did in the present example.

    Thus the equations of motion of the different degrees of'freedom are

    usually coupled equations. One method of solving these two coupled

    differential equations is to search, for new variables th at are linear combi

    nations of the original dumb coordinates such that the new variables

    satisfy uncoupled equations of motion. The new variables are then called

    normal coordin ates. In the present example we know how to find the F ig . 1 .8 Ro t a t i on o f coo rd ina t es .

    normal coordinates, given the dumb coordinates x'and y ' . Simply rotate

  • 7/27/2019 Course on Waves

    47/632

    o a coo d ates, g e t e du b coo d ates a d y S p y otate

    the coordinate system so as to obtain x and y , each of which is a linear

    combination of x'and y ' . In a more general problem, we would have to

    use a more general linear transformation of coordinates than can beobtained by a simple rotation. Th at would be the case if, for example, the

    pairs of springs in Fig. 1.7 were not orthogonal.

    Sys tema t ic so lu t i on fo r modes . Without considering any specific physical

    system, we assume that we have found two coupled first-order linear

    homogeneous equations in the dumb coordinates xand y:

    ^ = a u x - a 12y ( 4 7 )

    - ^ jr = a 2i x - a 22 y . ( 4 8 )

    Now we simply assumethat we have oscillation in a single normal mode.

    That means weassume that both degrees of freedom, namely xand y , os

    cillate with harm onic motion with the same f r equency andsam e phase

    constant . Thus we assumewe have

    x A cos (cot +

  • 7/27/2019 Course on Waves

    48/632

    of the linear homogeneous equations (51) and (52) must vanish:

    an w2 012021 a22 w2 = { ( I I I 2) (f l2 2 CO2) 021012 = 0. (56)

    Equation (55) or (56) is a quadratic equation in the variable co2. It has

    two solutions, which we call coi2 and CO22. Thus we have found that if we

    assume we have oscillation in a single mode, there are exactly two ways

    that that assumption can be realized. Fre qu en cy coi is the frequency of

    mode 1; C02 is tha t of mode 2. The shape or configuration of xand yin

    mode 1 is obtained by substituting co2 = coi2 back into either one of

    Eqs. (53) and (54). [They are equivalent, because of Eq. (56).] Thus

    = ( A ) = - g L = co!2 - a n . (57fl)

    \X/mode 1 VA /mode 1 A i #12

    Similarly,

    (M.) = ( A ) = J * L = c22 - B l L . (5 7b)\ X/ mode 2 V A /mode 2 A 2 d i2

    Once we have found the mode frequencies coi and C02 and the amphtude

    ratios B1/A 1 and B2/ A 2, we can write down the most general superpositionof the two modes as follows:

    x(t) Xi(t) + x2(t) = Ai cos (coi* +

  • 7/27/2019 Course on Waves

    49/632

    F ig . 1 .9 Lo ng i t ud i na l osci l l a t i ons,

    ( a) Equ i l i b r i um , ( b ) Gener a l c on f i g u r a-

    t ion .

    modes, since there are-tw o degrees of freedom . In a mode, each moving

    part (each mass) oscillates- with harm onic motion. This means that each moving part oscillates with the same frequency, and thus t he retu rn fo rce

    pe r un i t d isp l acemen t pe r un i t m ass is t he same fo r bo t h m asses. (We

    learned in Sec. 1.2 that oo2 is the return force per unit displacement per

    unit mass. Th at holds for each moving part, w hether it is a single isolated

    system with one degre e of freedom or is part of a larger system. The

    only requirement is that the motion be harmonic motion with a single

    frequency.)

    In the present example the masses are equal. We need therefore only

    search for configurations that have the same return force per unit displace

    ment for both masses. Let us guess that the displacements may be the

    same, and see if that works: Suppose we start at the equilibrium position

    and then displace both masses by the same amount to the right. Is the re

    turn force the same on each mass? Notice that the central spring has the

    same length as it had at equilibrium, so that it' exerts no force on either

    mass. The left-hand mass is pulled to the left because the left-hand spring

    is extended. The right-hand mass is pushed to the left with the sameforce,

    because the right-hand spring is comp ressed by the same amount. Wehave therefore discovered one mode!

    Mode 1: \pa{t ) = \pb(t ), i 2 = (60)M

    The frequency i 2 = K /M in Eq. (60) follows from the fact that each mass

    oscillates just as it would if the central spring were removed.

    Now let us try to guess the second mode. Fro m the symm etry, we guess

    h if d 6 i l h d If di

  • 7/27/2019 Course on Waves

    50/632

    that if aand 6 move oppositely we may have a mode. If amoves a distance

    \pato the right and 6 moves an equal distance to the left, each has the same

    return force. Thus the second mode has ipb = \pa. The frequencycan be found by considering a single mass and finding its return force per

    unit displacement per unit mass. Consider the left-hand mass a. It is

    pulled to the left by the left-hand spring with a force Fz = K\ pa. It is

    pushed to the left by the middle spring with a force Fz 2K\ pa. (The

    factor of two occurs because the central spring is compressed by an amount

    2 ipa.) Thus the net force for a displacement \pais 3K\ pa, and the return

    force per unit displacement per unit mass is 3 K /M :

    Mode 2: i//a = \pb, w22 (61)

    The modes are shown in Fig. 1.10.

    F ig . 1 .10 N o rma l modes o f l ong i t u d i na l

    osci l l a t i o n , (a ) M ode w i t h l ow er f r e-

    quency . (b) M ode w i t h h i gher f r equency .

    We shall solve this problem on ce more, using the method of searching

    for normal coordinates, i.e., sm art coordinates. The sm art coordinates

    are always a linear combination of ordinary dumb coordinates, such that

    instead of two coupled linear equations, one obtains two uncoupled equations. From Fig. 1.9b , we easily see that the equations of motion for a

    general configuration are

    By inspection of these equations of motion, we see that alternately adding

  • 7/27/2019 Course on Waves

    51/632

    y p q , y g

    and subtracting these equations will produce the desired uncoupled equa

    tions. AddingEqs. (62) and (63), we obtain

    M ~ ( i a + i b) = K ( 4 a + U (64)

    Subtracting Eq. (63) from Eq. (62), we obtain

    M d2('Pad ~ ^ = - 3 K ( 4a - U (65)

    Equations (64) and (65) are uncoupled equations in the variables 4a + 4b

    and if/a 4b They have the solutions

    4a+ 4b= 4 l ( t ) =Ai cos (toit + z), C022 = , (67)

    M

    where Ai and

  • 7/27/2019 Course on Waves

    52/632

    The system is shown in Fig. 1.11 . The oscillations are assumed to be con

    fined to the plane of the paper. There fore there are just two degrees of

    freedom. The three identical massless springs have a relaxed length aothatis less than the equihbrium spacing aof the masses. Thus they are all

    stretch ed. When the system is at its equilibrium configuration (Fig. 1.11a ),

    the springs have tension To.

    Because of the symmetry of the system, the modes are easy to guess.

    They are shown in Fig . 1.11. The lower mode (the one with the lower

    frequency, i.e., the one with the smaller return force per unit displacement

    per unit mass for each of the masses) has a shape (Fig. 1.11c) such that the

    cen ter spring is never compressed or extended. The frequency is thus obtained by considering either one of the masses separately, with the return

    force provided only by the spring that conn ects it to the wall. Fo r either

    the slinky approximation (unstretched spring length of zero) or the small-

    oscillations approximation (displacements very small compared with the

    spacing a), we shall show presently that a displacement \paof the left-hand

    Fig. 1.11 Transverse osci ll ati ons,

    (a) Equi li bri um, (b) General configura-

    ti on. (c) Mode w it h low er fr equency,

    (d) M ode w it h hi gher fr equency.

    mass causes the left-hand spring to ex ert a return force of To(\ p/a). Hence,

  • 7/27/2019 Course on Waves

    53/632

    p g ( p ) ,

    in this mode the return force per unit displacement per unit mass, coi2, is

    given by

    M o d el: Wl2 = - 5 - , ^ = + 1 . (70)M a xpa

    We see this as follows. Firs t consider the shnky app roximation (Sec. 1.2).

    In this approximation, the tension T is larger than T0 by the factor l / a ,

    where Iis the spring length and ais the length at equilibrium (Fig. 1.11a).

    The spring exerts a transverse return force equal to the tension Ttimes the

    sine of the angle between the spring and the equilibrium axis of the springs,i.e., the return force is T(\ pa/ l ). But T T0( l /a ) . Thus the return force is

    Ti)(\ pa/a ), and this gives Eq . (70). Nex t consid er the small-oscillations ap

    proximation (Sec. 1.2). In tha t approxim ation, the increase in length of

    the spring is neglected, because it differs from the equilibrium length a only

    by a quantity of order a(xpa/a ) 2, and therefore the increase in tension also

    is neglected . The tension is thus T0 when the displacement is xpa. The

    return force is equal to the tension To times the sine of the angle between

    the spring and the equilibrium axis. This angle may be taken to be asmall angle, since the oscillations are small. Then the angle (in radians)

    and its sine are equal, and both are equal to xpa/a . Thus the return force

    is To(xpa/a ) . This gives Eq. (70).

    Similarly, we can obtain the frequency for mode 2 (Fig. 1.11 d)as follows:

    Consider the left-hand mass. The left-hand spring contributes a return

    force per unit displacement per unit mass of To /M a , as we have just seen

    in considering mode 1. In mode 2 the cent er spring is helping the left-

    hand spring, and in fact it is providing twice as great a return force as is

    the left-han d spring. This is easily seen in the small-oscillations approxi

    mation: The spring tension is Tofor both springs, but the center spring

    makes twice as large an angle with the axis as does the end spring, so that

    it gives twice as large a transverse force com ponent. The total return

    force per unit displacem ent per unit mass, CO22, is thus given by

    Mode 2:

  • 7/27/2019 Course on Waves

    54/632

    y g , ; qy

    motion motion of the charges in this case. Th e electromotive^ force

    (emf) across the left-hand inductance is L d l a/ d t . A positive charge Qi onthe left-hand capacitor gives an emf C~XQ\ that tends to increase l a(with

    ou r sign conventions). A positive charg e Q2 on the middle capacitor gives

    am emf'" G-1 @2 that tends to d e c r e a s e ' T h u s we have for the complete

    contribution to L d l a/ d t

    Similarly,

    Ldlg

    dt

    L d l b

    dt

    = c -'Q i -

    = c - i< ? 2 - c - i< ? 3.

    (72)

    (73)

    As in Sec. 1.2, we will:express the configuration of the system in terms of

    curre nts rather than charge s. To do this, we differentiate Eqs. (72) and

    (73) with respect to time and use conservation of charge. Differentiating

    gives

    t d2I a _ x dQ i j dQ'2

    d t 2 d t d t

    ^ d2h _ dQz _ -i_i dQ z

    d t2 d t d t

    (74)

    (75)

    Charge conservation gives

    dQ1 _ T dQ2 _ r ,

    d t ~ d t ~ a ~

    dQ z

    dth. (76)

    Substituting Eqs. ( 76) into Eqs. (74) and (75 ), ,We obtain the coupled equations of motion

    d2I aL

    L

    d t2

    d2h

    d t2

    = - c ~ n a + c~\ h - ia)

    = - ~ G ~ H h ~ h ) - C U b.

    (77)

    (78)

    F ig . 1 .12 Tw o coup l ed L C ci rcu i ts .

    G enera l con f igu ra t ion o f cha rges and

    currents . T he arrow s giv e s ign conven-

    t i o n s fo r posi t i v e cu r ren t s .

    h h

    ^ n n n n p r L L

    Qi Qz Qs

    ~Q i - q2 ~Q s

    Now that we have the two equations of motion we want to find the two

    l d h b f d b h f l d

  • 7/27/2019 Course on Waves

    55/632

    normal modes. These can be found by searching for normal coordinates,

    by guessing, or by the systematic method (see Prob. 1.21 ). One finds

    c1Mode 1: l a h , wi2 r (79)q r ' - i ' '

    Mode 2: I a = h ,

  • 7/27/2019 Course on Waves

    56/632

    frequency (0mod:

    wav = + w2)> Wm0(j EE -((Oi (02). (82)

    The sum and difference of these give

    toi - (0av + wmod> *02 = Wav COmod. (83)

    Then we may write Eq. (81) in terms of (oav and comod:

    1p = A COS Wit + A COS (02f

    = A COS ((0avf + tomodf) + A COS ((0avf tOrnodf)

    = [2A cos wmodt] cos wavt,

    i.e.,

    Amod(t) COS C0av, (84)

    where

    Am od(^ ) - A CO S COm od t. (8 5 )

    We can think of Eqs. (84) and (85) as representing an oscillation at angular

    frequ enc y coav, with an am plitud e Amod that is not consta nt bu t rather varies

    with time accordin g to Eq . (85). Equa tions (84) and (85) are exact. However, it is most useful to write the superposition, Eq. (81), in the form of

    Eqs. (84) and (85) when coi and 2 are of com parable magnitude. Then

    the modulation frequ ency is small in magnitude com pared with the average

    frequency:

    (Oi ^

  • 7/27/2019 Course on Waves

    57/632

    are not exactly constant, but only almost constan t. Their variation is

    negligible during one cycle of oscillation at the average fast frequency

    coav, provided that the frequency range or bandwidth of the component

    harm onic oscillations is small compared with coav. (W e shall prove these

    remarks in Chap. 6 .)

    Some physical examples of beats follow:

    Example 11: Beats produced by two tuning forks

    Whe n a sound wave reaches your ear, it produces a varia tion in air pres

    sure at the eardrum. Le t ipi and \ p2 represent the respective contributionsto the gauge pressure produced outside your eardrum by two tuning forks,

    numbe red 1 and 2. (The gauge pressure is just the pressure on the outer

    surface of your eardrum minus the pressure on the inner surface; the pres

    sure on the inner surface is normal atmo spher ic pressure. This pressure

    difference provides the driving force to drive the eardrum.)

    If both forks are struck equally hard at the same time and are held at

    the same distance from the eardrum, the amplitudes and phase constants

    for the gauge pressuresj p iand xp2 are the same, and thus Eq. (80) correctly represents the two pressure contributions. The total pressure (which gives

    the total force on the drum) is the superposition xp = \ pi+ \p2 of the con

    tributions from the two forks. It is given either by Eq. (81) or by Eqs. (84)

    and (85). If the frequencies of the two forks, i>iand v2, differ by more

    than about 6 % of their average value, then your ear and brain ordinarily

    prefer Eq. (81). Tha t is, you hear the total sound as two separate notes

    with slightly different pitches. Fo r example, if v2 is f times v i , you hear

    two notes with an interval of a major third. If v2 is 1.06i'i, you hear v2as a note one half -tone higher in-, pitch than v\ . However, if and v2

    differ by less than about 10 cps, your ear (plus brain) no longer easily

    recogn izes them as different notes. (A musicians trained ear may do

    much better.) Then a superposition of the two is not heard as a chord

    made up of the two notes v\ and v2, but rather as a single pitch of

    frequency vav with a slowly varying amplitude A mod, just as given by Eqs.

    (84) and (85).

    Square law detector . The modulation amplitude Amod oscillates at the

    >modulation angular frequen cy comod. Whenever thas increased by an

    amoun t 2 w (radians of phase), the am plitude A modhas-gone through one

    complete cycle of oscillation (i.e., the slow oscillation at the modulation

    frequency) and has returned to its original value. At two times during one

    cycle, Amod is zero. At those times, the ear doesnt hear anything there

    is no sound. In between the silences, you hea r a sound at the average

    pitch. Since; cos .comodtgoes from zero to + 1 , to zero, to 1, to zero, to

    + 1, etc ., we see that A mod has opposite signs at successive loud times.

    Nevertheless, your ear does not recognize two kinds of loud times, as

    ill di if f h i i h i f k Th

  • 7/27/2019 Course on Waves

    58/632

    you will discover if you perform the experim ent with tuning forks. Thus

    your ear (plus brain) does not distinguish positive from negative values ofAmod- It only distinguishes w hether the magnitude of A mod is large

    (loud) or small (soft), that is, whether the squa reof A mod is large or

    small. Fo r that reason, you r ear (plus brain) is sometimes said to be a

    squa re la w det ec to r . Since A mod2 has t w omaxima for every modulation

    cycle (during which wmodtincreases by 2 t t), the repetition rate for the se

    quence loud, soft, loud, soft, loud, soft, . . . is twice the modulation fre

    quenc y. This repetition rate of large values of Amod2 is called the beat

    f r e quency :

    k beat 2(0mod COi 02. (86)

    We can see thisalgebraically as follows:

    A m o d ( f) 2 A CO S Wm od*.

    [Amod(f)]2 = 4 A 2 COS2 COmodt;but

    cos26 = ^[cos26 + sin26 + cos29 sin26} J [1 + cos 20],

    Thus

    [Amod(i)]2 = 2A2[1 + cos 2wm0dt ],

    i.e.,

    (Amod)2 = 2A2[1 + COS Wheat*]- (87)

    ThusAmod2 oscillates about its average valueat twice the modulation fre

    quency, i.e., at the beat frequency, coi w2-

    The superposition of two harmonic oscillations with nearly equal fre

    quencies to produce beats is illustrated in Fig. 1.13.

    Example 12: Beats between two sources of visible light

    In 1955, Forrester, Gudmundsen, and Johnson performed a beautiful ex

    periment showing beats between two independent sources of visible light

    with nearly the same frequency, f The light sources were gas discharge

    tubes containing freely decaying mercury atoms with an average frequency

    of j'av = 5.4 9 X 10 14 cps, corresponding to the bright green line of mer

    cury. The atoms were placed in a mag netic field. This caused the green

    radiation to split into two neighboring frequencies, with the frequency

    difference proportional to the mag netic field. The beat frequency was

    v\ v2 ~ 1010 cps. This is a typical rada r or microwav e frequency.

    t A. T. Forrester, R. A. Gudmundsen, and P. O. Johnson, Photoelectric mixing of incoherent

    light, Phys. Rev. 99, 1691 (1955).

  • 7/27/2019 Course on Waves

    59/632

    0 5 10 15 20 25 30 35F>'! v> " . V . ' . iAvfliWliW1111liVft

    Miimw0 5 10 15 20 25 30

    1 beat

    Fi g. 1.13 Beat s, xf/i a n d \ p 2 ar e the pres-sure va r i a t i ons a t yo u r ea r p rod uced by

    t w o t u n i n g f o r k s w i t h f r e q u en c y r at i o

    v \ /v i 10 /9 . The to ta l p r essu re is the

    superpos i t ion ip i+ 2, w h ich is an

    almost ha rm on i c osci l l a t i o n a t f r e-

    quencyz'avw i t h sl ow l y v a r y i n g amp l i t u d