Classical Electrodynamics
Classical Electrodynamics
ElectrDdynam ics
Julian Schwinger • Lester L. DeRaad, Jr.
A d v a n c e d Bo o k P r o g r a m
CRC Press
Taylor & Francis Group
Boca Raton London New York
CRC Press is an imprint of the Taylor & Francis Group, an informa business
This book contains information obtained from authentic and highly regarded sources.
Reasonable efforts have been made to publish reliable data and information, but the
author and publisher cannot assume responsibility for the validity o f all materials
or the consequences of their use. The authors and publishers have attempted to trace
the copyright holders of all material reproduced in this publication and apologize to
copyright holders if permission to publish in this form has not been obtained. If any
copyright material has not been acknowledged please write and let us know so we
may rectify in any future reprint.
Except as permitted under U.S. Copyright Law, no part of this book may be reprinted,
reproduced, transmitted, or utilized in any form by any electronic, mechanical,
or other means, now known or hereafter invented, including photocopying,
microfilming, and recording, or in any information storage or retrieval system,
without written permission from the publishers.
2 Magnetic Charge I IT
2.1 A Very Brief History of Magnetic C h a r g e ..................................... 19
2.2 Problems for Chapter 2 .................................................................... 19
3 Conservation Laws 21
3.1 Conservation of Energy .................................................................... 21
3.2 Conservation of M om en tu m .............................................................. 22
3.3 Conservation of Angular Momentum. Virial Theorem ................ 24
3.4 Conservation Laws and the Speed of L ig h t..................................... 25
3.5 Problems for Chapter 3 .................................................................... 29
4 Macroscopic Electrodynamics 33
4.1 Force on an Atom ............................................................................. 33
4.2 Force on a Macroscopic B o d y ........................................................... 38
4.3 Macroscopic Electrodynam ics..................... 40
4.4 Problems for Chapter 4 .................................................................... 42
v
vi CONTENTS
6.3 Paramagnetism.................................................................................... 67
6.4 Ferromagnetism.................................................................................... 69
6.5 Problems for Chapter 6 .................................................................... 72
22 Multipoles 257
22.1 Problems for Chapter 22 .................................................................... 261
26 Magnetostatics 313
26.1 Variational P rin cip le ............................................................................. 313
26.2 Boundary Conditions............................................................................. 315
26.3 Vector Potential ................................................................................... 316
26.4 Problems for Chapter 2 6 ....................................................................... 318
47 Diffraction I 491
47.1 Diffracted Electric F i e l d .......................................................................491
47.2 Diffraction by a Circular A p e r t u r e ................................................. 495
47.3 Diffraction by a Slit ........................................... 499
47.4 Diffraction by a Straight E dge............................................................. 503
47.5 Problems for Chapter 47...................................................................... 507
48 Diffraction II 509
48.1 Approximate S o lu t io n ..........................................................................510
48.2 Exact Solution for Current .................................................................512
48.3 Exact Diffraction Cross S e c tio n ................................................. . . . 5 1 6
48.4 Field Near Edge ................................................................................... 517
48.5 Problems for Chapter 48 .................................................................... 520
A Units 555
B Bibliography 561
Index 563
List of Figures
5.1 The real and imaginary parts of the conductivity, as givenin (5.14). 48
5.2 Singularities of the integrand in (5.33) in the complexa;plane. . 50
6.1 Solutions to (6.58) for temperatures above and below the Curie
temperature. The solid curves are the plots of the right side of
(6.58) for T < T C and T > Tc , respectively, while the dashed line
is the left side of that equation........................................................... 71
xiii
XIV LIST OF FIGURES
22.1 Geometry of field point and source point for a bounded charge
distribution............................................................................................... 257
22.2 Two interacting charge distributions.................................................... 260
23.1 Sphere bisected by plane. Shown are the locations of the physical
charge at r' and the image charges at r', r^, and ....................... 275
23.2 Geometry of dielectric sphere with source point outside................ 276
32.1 Electric and magnetic fields for a wave propagating in the direc
tion n ................................................................................................... . 353
LIST OF FIGURES xv
35.1 Lorentzian line shape for the energy radiated per unit frequency
range...................................................................................................... 382
1J. Schwinger and D. Saxon, Discontinuity in Waveguides (Gordon and Breach, New York,
1968).
2 “ On Radiation by Electrons in a Betatron,” unpublished (1945) [transcribed by M. A.
Furman, LBNL-39088, July 1996]; “On the Classical Radiation of Accelerated Electrons,”
Phys. Rev. 75, 1912 (1949).
XVII
have retained our original organization of the manuscript, and not the few, very
long chapters of Schwinger’s later revision, because we feel that this makes the
contents more accessible to the student and teacher. We have tried to retain
Schwinger’s inimitable lecturing style in this book, in which everything flows
inexorably from what has gone before. However, as an aid to the reader, we
have included a Reader’s Guide, which identifies major themes in each chapter,
suggests a possible path through the book, and identifies topics for inclusion or
exclusion from a given course, depending upon circumstances.
We dedicate this book to the memory of Julian Schwinger. We are indebted
to Clarice Schwinger for her gracious permission to pursue this project, and
for her encouragement. We are also grateful to Professor Michael Strauss, who
attended Schwinger’s 1983 course, and made available his notes, and the corre
sponding book manuscript at that time. Further materials have been obtained
from the Julian Schwinger Papers (Collection 371), Department of Special Col
lections, University Research Library, University of California, Los Angeles.
Many students at the University of Oklahoma contributed by finding innumer
able errors in the typescript, so we thank them as well.
Finally, we acknowledge the Alfred P. Sloan Foundation and the U. S. Na
tional Science Foundation for partial financial support during the early stages
of this project, and the U. S. Department of Energy for partial support of KAM
during the completion of this project.
X V lll
Reader’s Guide
8+ 4(/ ; 8M
ther included, in its original incarnation, two chapters, now Chapters 51 and
52, which were based on lectures Schwinger gave, around the same time, on the
energy loss by electrically and magnetically charged particles when they pass
through matter. This latter work was evidently based on the excitement gener
ated by the purported discovery of magnetic charge in cosmic ray experiments,3
evidence that was subsequently interpreted in terms of a conventional but rare
nuclear fragmentation event.4 The manuscript prepared during the 1977-79
period remains the core of the present volume.
However, as noted, Schwinger undertook extensive revisions beginning in
1979, and continuing at least through 1983. Indeed, the UCLA archives contain
an extensive revision of the book through electrostatics, corresponding to the
material in the present Chapters 1-25. In the interest of preserving the more
spontaneous flavor of the original lectures, we have retained most of the origi
nal material, but incorporated new material at the appropriate points. A few
chapters were not in the original version and are based closely on the 1979-83
revision; they are Chapter 10, “Einsteinian Relativity,” Chapter 18, “Modi
fied Bessel Functions,” Chapter 19, “Cylindrical Conductors,” and Chapter 25,
“Modes and Variations.” Schwinger ceased work on the project in 1984, but
he did present new material on radiation in his UCLA lectures in the previous
year, and that new material now appears in Chapter 44, “Waveguides,” and
Chapter 46, “ Partial-Wave Analysis of Scattering.” Of course, the present au
thors have incorporated numerous improvements and additional material. Most
significant is the inclusion of a great many problems, of a wide range of diffi
culty, some of which had been prepared by us in the late 1970s, but many of
which are new. The problems constitute an integral part of the text; indeed
many important concepts appear only in the problems, and an instructor might
feel it appropriate to base lectures on those topics. Examples include covariant
notation, introduced in Chapter 10, and vector spherical harmonics, introduced
in Chapter 50.
In view o f this long history, the reader will not be surprised to detect a certain
variation in the level of the material. This seems to us entirely appropriate.
3P. B. Price, E. K. Shirk, W . Z. Osborne, and L. S. Pinsky, Phys. Rev. Lett 35, 487 (1975).
4L. W . Alvarez, in Proceedings of the 1975 International Symposium on Lepton and Photon
Interactions at High Energies, Stanford, August 21-27, 1975, ed. W . T. Kirk, p. 967.
XIX
Some o f the topics are quite elementary, but there are many which are more
advanced, and which therefore could be omitted in a first course. Indeed, it is
the authors’ experience that there is far more material here than can be covered
in a typical two-semester graduate electrodynamics course. It is hoped that the
more advanced material will be of utility to practicing researchers, and indeed,
the authors have found the material contained herein extremely valuable in their
own research in electromagnetic theory and quantum field theory.
The balance o f this Guide will sketch major subjects covered in each chapter,
and make suggestions for material that is essential, and inessential, for a first
course or reading.
C h a p ter 1. The central point of this chapter is a hueristic derivation of
Maxwell’s equations starting from Coulomb’s law together with the imposition
of Galilean invariance. This is not to be taken as a rigorous derivation, especially
because the correct relativity group is that of Einstein. Essential.
C h a p ter 2. Here we show how Maxwell’s equations are modified if magnetic
charge is present. Inessential.
C h a p ter 3. We discuss the conservation of energy, momentum, and an
gular momentum, thereby establishing the consistency of Maxwell’s equations.
Essential.
C h a p ter 4. This chapter is devoted to the inference of the macroscopic
Maxwell equations. Essential.
C h a p ter 5. Here we present simple classical models for the electric perme
ability and conductivity, and discuss the Clausius-Mossotti equation. Essential.
C h a p ter 6. We give models for the magnetic properties of matter here, and
introduce the vector potential. Essential.
C h a p ter 7. We treat the somewhat subtle issues of how to construct the
energy and momentum for macroscopic electrodynamics. Essential.
C h a p ter 8. Mechanical action principles are the subject here. Although
these should be familiar, the viewpoint is somewhat novel and crucial in the
following. Essential.
C h a p ter 9. Here we present the action principle for electrodynamics. Es
sential, except for Section 9.6.
C h a p ter 10. We discuss the modification of the particle Lagrangian re
quired by Einstein’s relativity, as well as the usual relativistic kinematics and
field transformations. Probably only Section 10.1 is essential.
C h a p ter 11. Here the study of electrostatics begins. We show that the
general action principle implies the stationary principles for electrostatics. Es
sential.
C h a p ter 12. We introduce Green’s functions. Essential.
C h a p ter 13. We discuss Green’s function for free space. Essential.
C h a p ter 14. We derive Green’s function for a semi-infinite dielectric. Only
Sections 14.1 and 14.4 are essential.
C h a p ter 15. Here, we use the previous Green’s function to compute the
force between a point charge and a dielectric slab. Inessential.
C h a p ter 16. Now, we introduce Bessel functions. As they are used through
out the following, this is essential.
xx
C h a p ter 17. We give the derivation of Green’s function for parallel con
ducting plates, along with an extensive discussion of image charges in this case.
Only Section 17.1 is essential.
C h a p ter 18. Now, we introduce modified Bessel functions. This material
is very useful, but perhaps inessential for a first reading.
C h a p ter 19. We discuss cylindrical conductors, of rectangular, triangular,
and circular cross section, and derive useful mathematical results. Inessential,
C h a p ter 20. Now we introduce spherical harmonics. Essential,
C h a p ter 21. Legendre’s polynomials and spherical Bessel functions are the
principal topics here. Only the beginning through Section 21.1, and perhaps
Section 21.3, are essential.
C h a p te r 22. The general multipole expansion for the energy of interac
tion of two arbitrary bounded charge distributions is the subject. At least the
beginning of this chapter is essential.
C h a p ter 23. We consider Green’s functions for conducting and dielectric
spheres. The first parts of Sections 23.1 and 23.4 seem essential.
C h a p ter 24. We give a general treatment of electrostatics in the presence of
dielectrics and conductors . We prove Thomson’s theorem, and give the general
expression for capacitance. At least the latter, given in Section 24.6, is essential.
C h a p ter 25. Variational methods for estimating eigenvalues are the sub
jects treated here. Elegant, but inessential.
C h a p ter 26. We now introduce the variational principle for magnetostatics.
Essential.
C h a p ter 27. The energy of interaction of steady currents, and inductance,
are the subjects here. Essential.
C h a p ter 28. Here the topic in magnetic dipoles. Essential.
C h a p ter 29. We introduce the magnetic scalar potential. Inessential.
C h a p ter 30. Here, the subject is the “string” required for the definition of
the vector potential when magnetic charge is present. Inessential.
C h a p ter 31. Now the treatment of radiation begins, with the retarded
Green’s function and the Lienard-Wiechert potentials. Essential.
C h a p ter 32. We discuss the asymptotic fields, and dipole radiation. Es
sential.
C h a p ter 33. Here we study radiation from the point of view of energy
transferred from the source, and derive the Darwin-Breit Hamiltonian for the
charges. Inessential.
C h a p ter 34. We discuss radiation from simple models of antennas. Inessen
tial.
C h a p ter 35. We derive the general formula for the spectral distribution of
radiation. Essential.
C h a p ter 36. Now, we obtain a formula for the power spectrum, and apply
it to Cerenkov radiation. Essential.
C h a p ter 37. We discuss radiation in the extreme situations of constant
acceleration, and of impulsive scattering. In the problems, we give general
formulas for the energy radiated by a particle undergoing arbitrary acceleration
for a finite amount of time. Inessential.
xxi
C h a p ter 38. Synchrotron radiation is the subject here. Essential.
C h a p ter 39. We derive the power spectrum for the different polarization
states in synchrotron radiation. Inessential.
C h a p ter 40. The high energy regime for synchrotron radiation is the sub
ject here. Sections 40.1 and 40.5 are essential.
C h a p ter 41. We derive Snell’s law, and the reflection and transmission
coefficients for light incident on a plane dielectric interface. Mostly essential.
C h a p ter 42. We treat reflection by a conductor having finite conductivity
here. Inessential.
C h a p ter 43. Here we perform the 2 + 1 dimensional break up of Green’s
function for Helmholtz’ equation leading to Hankel functions. Inessential.
C h a p ter 44. Here we give a general treatment of cylindrical waveguides,
harking back to Chapters 19 and 25. Inessential.
C h a p ter 45. We present simple models of scattering, including Thomson
scattering. Essential.
C h a p ter 46. We give here a partial-wave approach to scattering. Although
this is very useful for quantum mechanics, it is here inessential.
C h a p ter 47. The elementary theory of diffraction is the subject. At least
Sections 47.1 and 47.2 are essential.
C h a p ter 48. Here we show that the diffraction of a normally incident plane
wave on a straight edge is exactly solvable. Inessential.
C h a p ter 49. We derive Babinet’s principle in the case of a plane with
an aperture or slit to give the diffraction for the opposite polarization state.
Inessential.
C h a p ter 50. The general formulation of scattering, the Born approxima
tion, and the optical theorem are the subjects. Essential.
C h a p ter 51. We derive general dispersion relations for the dielectric con
stant from causality requirements. Inessential.
C h a p ter 52. We give a general treatment of energy loss by an electrically
or magnetically charged particle traversing matter. Inessential.
XXII
Chapter 1
Maxwell’s Equations
1.1 Electrostatics
Our intention is to move toward the general picture as quickly as possible, start
ing with a review of electrostatics. We take for granted the phenomenology of
electric charge, including the Coulomb law of force between charges of dimen
sions that are small in comparison with their separation. This is expressed by
the interaction energy, E ) of a system of such charges in otherwise empty space,
a vacuum:
&ct&b
E I v a-
Tab
a*b
where ea is the charge o f the ath particle while
is the separation between the ath and 6th particles. (Throughout this book
we use the Gaussian system of units. Connection with the SI units will be
given in Appendix A.) As we shall see, this starting point, the Coulomb energy
(1.1), summarizes all the experimental facts of electrostatics. The energy of
interaction of an individual charge with the rest of the system can be emphasized
1
2 CHAPTER 1. M AXW ELL’S EQUATIONS
by rewriting (1.1) as
(1.3)
a bjia a
where we have introduced the electrostatic potential at the location of the ath
charge that is due to all the other charges,
where we now treat every charge on an equal footing, which means that in (1.5)
we sum over all charges e¿. In terms of this potential, which is different from
<j)a, the energy E can be written as
16
( - )
a a
where F a is the force acting on the ath point charge. Comparing this with the
energy expression (1.1) we find the force on the ath particle to be
Fa = - V a J 2 — = -V «e „ E — = -V «e.*(r«). (1-8)
fe r‘ 4 f e r“
In the last form, we have substituted <£(ra) for <j>ay so it would appear that an
extra self-action contribution has been introduced. To see that this is not true,
we first argue physically that the difference between <£(ra) and <f>a is independent
1.1. ELECTROSTATICS 3
Can we make sense of this? We could define the limit here by arbitrarily adding
a displacement vector e of fixed direction to r a and letting its length approach
zero:
r = ra + c, € —►0, (1.10)
but at the cost o f picking out a particular direction. In order to remove the
most blatant aspect of this directional dependence, let us also approach r a from
the opposite direction,
r = ra - c , € —►0, (1.11)
and average over the two possibilities, so that the additional term (1.9) becomes
21 ( 1 .12)
a2
More elaborate limiting procedures, such as an average over all directions, can
be used, but the simple procedure of (1.12) suffices. Therefore, we can employ
^(r) in (1.8), with the implicit use of the two-sided limit, (1.12), to calculate
the force.
With the force given in terms of the gradient of a field (the potential), the
electric field E can now be defined by
E (r) = - V * ( r ) , (1.13)
Fa = eaE(ra). (1.14)
Besides the curl, the other elementary differential operation that can be applied
to a vector field is the divergence. To find V •E, we consider a related integral
statement. The integral of the normal component of E over a closed surface S
bounding a volume V is the electric flux (see Fig. 1.1):
r — Tb 1
j ) d S .E (r ) = ] C C4/ d S ‘ | r - r 6||r — rj|2 p dilb- (1.17)
a) b)
ej = f (dr)p(r). (1.21)
6iny Jv
[Throughout this book we use the following notation for the element of volume:
(dr) = d x d y d z ] (1.22)
For point charges, the charge density is zero except at the location of the charges,
f , , v c/ v f 0 if Yb is outside V
r —r . ) = ( l i f r i i g i n s l d e V (1.24)
by use of the divergence theorem relating surface and volume integrals. (See
Problem 1.2.) Since (1.25) is true for an arbitrary volume V , the integrands of
the volume integrals must be equal, so we obtain the equation satisfied by the
divergence of E,
V •E (r) == 4irp(r). (1.26)
These differential equations for the curl and divergence of E, (1.16) and
(1.26), respectively, completely characterize E when appropriate boundary con-
ditions are imposed. It is evident from (1.15) that, for a localized charge dis
tribution, the magnitude of the electric field becomes vanishingly small with
increasing distance from the collection of charges:
One can also specify how rapidly this occurs. But it is remarkable that the weak
boundary condition (1.27) already implies a unique solution to the differential
equations (1.16) and (1.26). To show this, we suppose that Ei and E 2 are two
such solutions. The difference, E = Ei — E 2 satisfies
V * £ = 0, V x£ = 0 everywhere, (1.28)
\E\ —►0 as r —* 00, (1.29)
1.1. ELECTROSTATICS 7
V x ( V x f ) = V ( V - f ) - V 2f , (1.30)
- V 2£ = 0. (1.31)
Let the single function £ (r) be any Cartesian component of the vector field
£; it obeys
—V 2£ (r) = 0. (1.32)
We present this as the everywhere valid statement
0 = - £ V 2£ = - V . (£ V £ ) + ( V £ ) 2, (1.33)
or
( V £ ) 2 — V 2-^£2 = 0. (1.34)
z
Now we integrate this over the interior volume V (R ) of a sphere of radius R
centered about an arbitrary point, which we take as the origin. The integral of
the second term in (1.34) is turned into an integral over the surface S(R) of the
sphere by means of the divergence theorem,
Using the relation between an element of area and an element of solid angle,
dS = R 2dily we can present this surface integral in terms of the average value
o f £ 2 over the surface o f the sphere,
(1.36)
(1.37)
The decisive step now is to divide by the area 4irR2, and then integrate (1.37)
over R from 0 to oo:
which finally incorporates the boundary condition (1.29), that £ vanishes at all
infinitely remote points. Everything on the left side of (1.38) is non-negative, yet
it all adds up to zero. Accordingly, every individual contribution must be zero.
This tells us quite explicitly that £ = 0 at the origin, which is anywhere, and,
consistently, that V £ = 0 everywhere, or, that £ is a constant, which is required
8 CHAPTER 1. M AXW ELL’S EQUATIONS
V * E = 4np, ^ p = 0,
Q
V x E = 0, — E = 0, (1.39)
where the time independence has been made explicit. We are now going to
remove the restriction to static conditions by letting the charges move in a
particularly simple way. The equations of electromagnetism that emerge from
this discussion will then be accepted as applicable to more general motions, as
justified by various tests of internal consistency.
(1.40)
so, in going from the static system to the uniformly moving system, we make
the replacement
d d d f .
S ~ 5 = « + v -V ' (1'41)
The equation for the constancy of the charge density in (1.39) becomes, in
the moving system
We recognize here a particular example of the charge flux vector or the (electric)
current density j,
j = pv. (1.44)
1.2. INFERENCE OF M AXW ELL’S EQUATIONS 9
Since V is arbitrary, the local conservation law, (1.45), follows. We also note
that the expression for the current density, (1.44), continues to be valid even
when v is dependent upon position, v —> v (r ,t). (See Problem 1.4.)
We can perform a similar transformation on the equation for the electric
field d'E/dt = 0; namely,
V x ( v x E ) = v ( V - E ) - ( v . V )E (1.48)
= v 47rp — (v •V )E
■= 4 7 r j - ( v V ) E , (1.49)
0 = ^ + 4*j-Vx(vxE). (1.50)
[Notice that by taking the divergence of (1.50) we recover the local charge conser
vation equation (1.45), so that the conservation of charge is not an independent
statement.] The quantity v x E represents a new phenomenon combining the
effects of motion with those of electric charge. To describe this new, induced
effect, we define the magnetic induction1 B by
v x E = cB, (1.51)
where c is a constant having the dimensions of velocity (which will turn out to
be the speed of light). Expressed in terms of the magnetic field, (1.50) becomes
an equation determining the curl of B,
^ 1 d„ 4x . ^
V X B = - — E + — j. (1.52)
COt C
Next, we naturally ask for the divergence of B . According to the definition,
(1.51), we have
V •B = V ♦ Q x e ) = - ( ^ x v ) . E = ~ . ( V x E ) = 0, (1.53)
1We will usually call B the magnetic field, but see Chapter 4.
10 CHAPTER 1. M AXW ELL’S EQUATIONS
or
V •B = 0. (1.54)
Moreover, in the co-moving coordinate system where the charges are at rest—
static— the magnetic field should also not change in time:
S B = ¿ B + ( W ) B = 0, (1.55)
¥ = V x (v x B), (1.56)
consistent with V •B = 0.
What do we do now? We need one experimental fact. Light is an electro
magnetic oscillation. The evidence for this is overwhelming. As examples, we
note that electric and magnetic fields are known to influence the emission, prop
agation, and absorption of light; and that radio and infrared waves, which differ
only in wavelength from visible light, are emitted by electric charge oscillations.
What must be done so that this fact is built into the equations we are inferring?
The existence of electromagnetic waves means that the equations determining
the electric field have solutions of the form
E ~ i ( z — ct), (1.57)
where c is the speed of the waves. Such waves, propagating in the z direction,
satisfy the second-order differential equation
<92
(1.58)
d z ^ ~ C2 5 i2 E;
for an arbitrary direction of propagation, the corresponding wave equation is
V’ E = i ^ E . (1.59)
More precisely, we require that this equation should hold far from the charges
that produce the field. The left side of this equation can be written as [cf. (1.30)]
V 2E = - V x ( V x E), (1.60)
E = - - xB. (1.62)
c
1.2. INFERENCE OF M AXW ELL’S EQUATIONS 11
But this cannot be a completely correct statement, since then v —> 0 would
require E —* 0. No electrostatics! However, all that is really necessary is that
the curl of this tentative identification be valid:
V x E = —V X i (1.63)
_ _ 1 9_ 47r .
V xB = E + — j, V •E = 47rp,
c at c
- V x E = — B, V « B = 0. (1.65)
c at
These equations of electromagnetism, as local, differential field equations, are
no longer restricted to the initial assumption of a common small velocity for all
charges.
To complete the dynamical picture we ask: What replaces (1.14) to describe
the force on an electric charge, when that charge moves with some velocity v
in given electric and magnetic fields E and B ? We consider two coordinate
systems, one in which the particle is at rest (co-moving coordinate system) and
one in which it moves at velocity v. Suppose in the latter coordinate system,
the electric and magnetic fields are given by E and B, respectively. In the
co-moving frame, the force on the particle is
F = eEeff, (1.66)
where Eeff is the electric field in this frame. In transforming to the co-moving
frame, all the other charges— those responsible for E and B — have been given
an additional counter velocity —v. We then infer from (1.62) that (v /c ) x B
has the character of an additional electric field in the co-moving frame. Hence,
the suggested Eeff is
Eeff = E H— XB, (1.67)
c
leading to the force law, due to Hendrick Antoon Lorentz (1835-1928),
F = e (E + ^ x B ) . (1.68)
These results, Maxwell’s equations, (1.65), and the Lorentz force law, (1.68),
have not been derived, but inferred from a special circumstance. We will adopt
12 CHAPTER 1. M A XW ELL’S EQUATIONS
these equations as describing the electromagnetic fields produced by, and acting
on, charges possessing arbitrary velocities, although the above discussion does
allow room for additional terms if v/c is no longer small. The fact that no
such terms are actually required is part of the implication of the special theory
o f relativity (see problem 1.6). We will prefer, instead, to show the physical
consistency of the equations as they stand (see Chapter 3).
1.3 Discussion
We have arrived at the Maxwell-Lorentz electrodynamics by combining three
ingredients: the laws of electrostatics; the Galileo-Newton principle of relativity
(charges at rest, and charges with a common velocity viewed by a co-moving
observer, are physically indistinguishable); and the existence of electromagnetic
waves that travel in a vacuum at the speed c. The historical line of development
was otherwise. Until the beginning of the nineteenth century, electricity and
magnetism were unrelated phenomena. The discovery in 1820 by Hans Christian
Oersted (1777-1851) that an electric current influences a magnet— creates a
magnetic field— is formulated, for stationary currents, in the field equation
^ « 4tt.
V x B = — j. (1.69)
c
The symbol c that appears in this equation is the ratio of electromagnetic and
electrostatic units of electricity (see Appendix A). Then, in 1831, Michael Fara
day (1791-1867) discovered that relative motion of a wire and a magnet induces
a voltage in the wire— creates an electric field. Such is the content of
-V X E = Í ¿ B, (1.70)
V - B = 0, (1.71)
that expresses the empirical absence of single magnetic poles. Finally, in 1864,
James Clerk Maxwell (1831-1879) recognized that the restriction to stationary
currents in (1.69), as expressed by V •j = 0, was removed in
_ Air. 1 d_
V xB = t j + -s e, (1.72)
V •E = Airp. (1.73)
The deduction of the existence of electromagnetic waves that travel at the speed
c, in remarkable numerical agreement with the speed of light, was confirmed in
1867 by Heinrich Rudolf Hertz (1857-1894). It was the conflict between the
1.4. PROBLEMS FOR CHAPTER 1 13
existence of this absolute speed c and the relativity concept of Newtonian me
chanics that set the stage for Einsteinian relativity. Already at the age of 16,
Albert Einstein (1879-1955) had recognized this paradox: To a co-moving New
tonian observer, light waves should oscillate in space, but not move; however,
Maxwell’s equations admit no such solutions. Einsteinian relativity is an out
growth of Maxwellian electrodynamics, not the other way about. That is the
spirit in which electrodynamics is developed as a self-contained subject in this
book.
A x ( B x C ) + B x ( C x A ) + C x (AxB) = 0,
V x ( A x B ) = A x ( V x B ) - B x ( V x A ) - ( A x V ) x B + (BxV)xA,
V •(AAxB) = A(B • V x A - A * V x B ) + A x B * V A .
V.E
Í dS *E,
where V is the volume contained within the closed surface 5, dS being
the surface element in the direction of the outward normal, and Stokes’
theorem,
3. This question has to do with the uniqueness theorem which follows from
(1.37).
(a) Directly from that equation, what assumption about |£(r)|, |r| —►'oo,
will produce the conclusion that £ = 0 everywhere?
(b) How would it work out if one had integrated this equation from R = 0
to oo, without dividing by R21
(c) How fast would ( £ 2) r have to fall off with R so that we could conclude
£ = 0 everywhere by simply taking R —►oo in (1.37)?
14 CHAPTER 1. M AXW ELL’S EQUATIONS
4. For an arbitrarily moving charge, the charge and current densities are
where R (t) is the position vector of the charged particle. Verify the state
ment of conservation of charge,
0¡P(r’t) + V.j(r,t) = 0.
x'3 = x 3.
? = -, 7 = (1 - P 2)~ 1/2,
v being the relative velocity of the two coordinate frames. We can regard
the four quantities Xp, p = 0, 1, 2, 3, as forming a four-vector. The four-
current jo = c/9, ji, i = 1 , 2 , 3 , constructed from the electric charge
and current densities, transforms by the same law:
On the other hand, the electric and magnetic field vectors are components
of a four-tensor, and so they have a more complicated transformation law.
Consider a boost by an arbitrary velocity v. Then the components of the
electric and magnetic fields in the direction of v do not change, while the
components in directions perpendicular to v are entangled:
Ei| = E,|, E1 = T ( e + ^ x b ) x ,
B' x = 7 ( B ~ xe )x .
Magnetic Charge I
VxE = - V x ( ^ x B ) , (2.2)
B = —xE, (2.3)
c
with its implication that all magnetic fields are due to the motion of electric
charges? This is the (1820) hypothesis of André Marie Ampere (1775-1836).
But is it true? An affirmative response is conventional, but the mathemati
cal development allows a more general possibility. Again, all that was really
required in the above was the curl relation
V xB = V x (^ xe), (2.4)
V •B = 47rpm. (2.5)
The implication of (2.5) is that a further source of magnetic fields, other than
moving electric charges, could exist in magnetic charge. Whether this possibility
is realized in nature still awaits experimental confirmation.
17
18 CHAPTER 2. MAGNETIC CHARGE I
° = + (v •V )B = ^ B - V x ( v x B ) + v ( V . B). (2.6)
jm = PmV, (2.7)
—V X E = ~ B + — j m. (2.8)
c at c
d
+ V •jm = 0. (2.9)
Id 47T.
VxB = E + -Je, V * E = 47TPe,
c at c
—V x B = ~ B + - j m, V 'B = 4irpm , (2.10)
c at c
where we have consistently used the subscript e to denote densities for electric
charge. Observe that these equations are invariant in form under the replace
ments (duality transformation)
Pe * pm > E —» B, je ►j m,
Pm * ~Pe> B ¥ B, j m ►“ je* (2-11)
F = e(E + ^ x B )+ i(B -^ x E ), ( 2 .1 2 )
for a particle carrying both electric and magnetic charge, e and g ) respectively.
Although from time to time there have been spectacular reports of the dis
covery of magnetic charge (Price, 1975; Cabrera, 1982), these “discoveries” were
never replicated, and serious objections were raised in each instance. Neverthe
less, there are strong theoretical reasons to believe that magnetic charge exists
in nature, and may have played an important role in the development of the uni
verse. Searches for magnetic charge continue at the present time, emphasizing
that electromagnetism is very far from being a closed subject.
2.1. A VERY BRIEF HISTORY OF MAGNETIC CHARGE 19
F = E -f iB , i = \/~T,
and related combinations of charge and current. Verify that these equa
tions retain their form under the transformation illustrated by
F -* e -^ F ,
where <j>is an arbitrary constant. Express this as a transformation of E, B,
and the charge-current quantities. What is the geometric interpretation?
What is the particular form of this transformation when <j) = 7t/ 2 ?
2. Suppose every charged particle carried electric and magnetic charge in the
universal ratio gk/^k = A. Is there another way of looking at this situation
in which we would be unaware of magnetic charge?
Conservation Laws
In order to check the physical consistency of the above set of equations governing
Maxwell-Lorentz electrodynamics [(2.10) and (2.12) or (1.65) and (1.68)], we
examine the action of, and reaction on, the sources of the electromagnetic fields.
To be precise, we ask whether there is a correct balance in the exchange of
energy, momentum, and angular momentum between the charged particles and
the electromagnetic fields. As we shall see, the Maxwell-Lorentz system as it
stands implies the conservation of these mechanical properties, no matter how
rapidly the charges are moving.
(3.1)
where we have used the Lorentz force law, (2.12), and the expressions for the
currents, (1.44) and (2.7), for a point particle. We interpret this equation as
meaning, even for general current distributions, that j e •E + j m •B is the rate of
energy transfer from the field to the particles, per unit volume. Then through
elimination of the currents by use of Maxwell’s equations, (2.10), this rate can
be rewritten as
(3.2)
The general form of any local conservation law, (1.45) or (1.46), suggests the
following interpretations:
21
22 CHAPTER 3. CONSERVATION LAWS
E2 + B2
energy density = U = (3.4)
8 tt 5
-f V •S -|-je •E 4* jm •B = 0, (3.6)
The three terms here are identified, respectively, as the rate of change
of the electromagnetic field energy within the volume, the rate of flow of
electromagnetic energy out of the volume, and the rate of transfer of elec
tromagnetic energy to the charged particles. Thus, (3.6) gives a complete
description of energy conservation.
F = e ( e + ~ x b ) + 0 (b —~ x e )
= J {d t)f, (3.8)
f = ¿ [E (V .E ) + B (V .B )]
3.2. CONSERVATION OF MOMENTUM 23
+i P[(-ilE
wD
+VxB) xB+Ex(4 Ib- VXE)]
1
<39>
= + t -[-E x ( V x E) + E(V «E) — B x ( V x B ) + B ( V ♦B)].
ut 4irc 4ir
The quadratic structure in E occurring here is
E2
— E x ( V x E ) + E(V •E) = - V — + (E. V)E + E(V.E)
¿i
= V . ( - 4 + E e ), (3.10)
which introduces dyadic notation, including the unit dyadic 1 , with components
(3n>
where Ski is the Kronecker 6 symbol. (See Problem 3.1.) The analogous result
holds for B. Accordingly, the force density is
d ExB / E2+ B2 EE + B B \
(3.12)
dt 47re \ 87t 4w )
EkEi + BkBj
Tki = h\U - (3.16)
4ir
Notice that the stress tensor is symmetrical, Tki = T}fc, which, as we shall see
in the next section, is required in order to obtain a local conservation law for
angular momentum. The trace of T , the sum of the diagonal elements T**, is
simply the energy density, (3.4),
We also note that the Poynting vector, (3.5), is proportional to the momentum
density,
S = c2G, (3.18)
which has the structure of
This is the first indication of the relativistic connection between energy and
mass, E = me2.
When we now multiply (3.25) with e,-,-* and sum over repeated indices, we find
that the terms involving spatial derivatives can be written as a divergence:
This final step is justified only because Tki is symmetrical (thus this symmetry is
required for the existence of a local conservation law of angular momentum). We
therefore icKvn.áfy the following electromagnetic angular momentum quantities:
¿ ( r . G ) + V . ( T t ) —tf + r . f = 0, (3.30)
at
which we call the electromagnetic virial theorem, in analogy with the mechanical
virial theorem of Rudolf Clausius (1822-1888). (See Chapter 8 .)
4 p = / (* )¿G = - / ( * ) V . T = 0, (3.32)
at Jpulse 01 Jpulse
With an eye toward relativity, we consider the space and time moments of
( 3 .6 ) and (3.15), respectively, combined as a single vector statement:
outside the charge and current distributions. Exploiting the connection between
S and G [(3.18)], we can rewrite (3.34) as a local conservation law, much as the
equality of Tjk and Tkj lead to the conservation of angular momentum:
G = P, (3.37)
E= f (dr)U. (3.40)
Jpulse
Thus the motion of this energy-centroid vector is governed by
which is to say that the center of energy, (r)£ (t), moves with constant velocity,
|<r>j*(<) = v js , (3.42)
m = E/c2. (3.43)
£
E. (3.44)
dt
3.4. CONSERVATION LAWS AND THE SPEED OF LIGHT 27
(3.46)
«S.
II
which is constant in the direction of the momentum,
(3.47)
>
II
ft,
We combine (3.47) with (3.41) to yield
vP •v e = c2. (3.48)
If the flow of energy and momentum takes place in a single direction, it would
be reasonable to expect that these mechanical properties are being transported
with a common velocity,
V p = v P = v, (3.49)
which then has a definite magnitude,
v •v = c2, v = c, (3.50)
E = v P, P = 4 V> (3.51)
cl
so we learn that p
E - Pc, v = c—, (3.52)
(r 2 - cH2) = 0, (3.53)
Integrating this over all space and using the idea of energy and momentum
weighting to define averages, as before, we obtain
E2+ B2 ExB
(3.58)
87T 47r
where the volume integrations are extended over the pulse. Now, a sum of
vectors of given magnitudes is of maximum magnitude when all those vectors
are parallel, which is to say here that
where equality holds only when E x B everywhere points in the same direction,
that of the pulse’s total momentum or velocity. On the other hand, we note the
inequality,
(ExB)2 = £ 2B 2 -(E.B)2
2
/ E2 + B2
V 2
+ (E ♦B)2
Comparing (3.59) and (3.61), we see that both equalities must hold, so that
E •B = 0, E 2 = B 2, (3.62)
3.5. PROBLEMS FOR CHAPTER 3 29
Figure 3.1: Electric and magnetic fields for an electromagnetic pulse propagating
with velocity v.
1 = i i + j j + kk.
A •1 = A, 1 •A = A, 1-1 = 1.
V-(AAB), V-(AAxB).
where the direction o f Su> points in the direction of the rotation, and has
magnitude equal to the (infinitesimal) amount of the rotation. Check that
5(V2) = 0.
Í(BXC) = 6 B X C + B x 6 C = ^ x ( B x C ) .
(a)
T tT = Tkk = U,
0 >)
T rT 2 = TkiTik = W 2 - 2(cG )2 > U2,
and
(c)
detT = —U[U2 — (cG )2].
Here the summation convention is employed, and the trace and determi
nant refer to T thought of as a 3 x 3 matrix.
5. Show that the angular momentum conservation law derived in Section 3.3
can be written as
J p 7 + V-JC + rxf = 0,
J = rxG,
K = — Txr,
r(<) = r ( 0 ) + vt.
Square and average this position vector, and upon comparison with the
solution of (3.55), identify (r(0) •v) and v.
3.5. PROBLEMS FOR CHAPTER 3 31
F = E -H B , F* = E — ¿B.
8^ - F’
the vector
1
tF* XF,
&7TÍ
and the dyadic
— (FF* + F*F).
n t,
eg — —he.
y 2
Chapter 4
Macroscopic
Electrodynamics
33
34 CHAPTER 4. MACROSCOPIC ELECTRODYNAMICS
Since the system is small, all the charges are near the center of mass of the
charge distribution, which lies at the position R . (For the purposes of the fol
lowing expansion we could let R represent an arbitrary point inside the charge
distribution; the use of the center of mass allows us to separate intrinsic proper
ties from those due to the motion of the atom as a whole.) We can then expand
the electric and magnetic fields about this reference point,
and likewise for B, in which the subsequent terms are considered negligible.
Here V means the gradient with respect to R . Now, the total force on the
atom, (4.2), can be rewritten in terms of this expansion as
F = ^ e aj E ( R ) + ^ e 0 [(ra - R ) . V ] E ( R ) + xB(R)
+ X > « ^ x [ ( r a - R ) * V ] B ( R) + . . . . (4.4)
a
The first term here is zero because of the neutrality of the system, (4.1). In the
second term, we identify the electric dipole moment, d,
d= ^ ea(ra - R ) = ^ e° r a, (4.5)
a a
5 > v« = £ d - (4.6)
Momentarily setting aside the fourth term, we find the force on the system
to be
F = (d •V )E (R ) + i ( ^ d ) X B (R ) + . . . . (4.7)
For the second term here, we can transfer the time derivative,
where V = dR/dt. Using (1.64) for <9B/3t, and rewriting the resulting double
cross product according to
Recalling that force is the time rate of change of momentum, we see that
( l / c ) d x B introduces a redefinition of the momentum of the system.
We now return to the fourth term of (4.4), which would seem to correspond
to a small effect, since for atomic systems, va/c < 1. A rearrangement of it is
£ e« v x - R ) •V 1B (R ) = E 7 V x Kr <* - R ) •V !B (R )
a a
where, recalling the definition of the electric dipole moment (4.5), we can express
the first term on the right side as
i v x ( d - V )B (R ). (4.12)
Combining this contribution with the second term on the right side of (4.10),
and using (1.54), we obtain
-[(d * V )V — (V •V )d ]x B (R ) = i [ ( d x V ) x V ] x B ( R )
c c
Collecting the various results to this point, we can now rewrite the total force
on the atom, (4.4), as
F b = Jt £ ~ ( ra - R) x [(ra - R) •V ]B (R )
- £ T ( ro - R ) x - v )• B (R )
We must now recall the restricted nature of this description: The electric and
magnetic fields change only slightly over the dimensions of the system. The first
of the three terms on the right side of (4.16) is a small correction to what is
36 CHAPTER 4. MACROSCOPIC ELECTRODYNAMICS
is of the same order o f magnitude as the omitted terms in the expansion (4.4),
and is therefore also to be neglected. An average of the initial form of F # ,
(4.15), with the single remaining contribution of (4.16), the second line there,
now gives
Fb = \ £ 7 (v “ - v ) ;* Kr « - R ) •V 1R (R )
a
- 5 E 7 ( r a - R ) x [ ( Va- v ) . V ] B ( R )
a
What has finally emerged here is the magnetic dipole moment of the system,
**- i
^ = ¿ E e<*(ro - R ) x (v“ - v )’ (4-19>
a
so (4.18) is equal to (¿i is constant in space)
where we have used (1.54). (It is, of course, the similarity of this structure to
V [ d « E ( R ) ] that justifies the identification of as the magnetic analogue of
d.) We also recognize that a contribution of this form already appears in the
second term on the right side of (4.14), bearing the information that a moving
electric dipole also acts as a magnetic dipole. The comparison of the two effects,
characterized by £ d x V and \i, is that of the typical speeds of the relatively
heavy atoms, |V|, and of the light electrons, |va|, in the interior of atoms,
Accordingly, we neglect the motional effects of the atoms, and finally write
(4.14) as
F = V[d •E(R) + fi •B(R)] + ^ ( i d x B ( R ) J . (4.22)
In the absence of time variation, what remains is a force associated with the
respective potential energies o f a given electric dipole in an electric field,
-d*E (4.23)
The energy interpretation does more than supply the force components as
negative gradients with respect to position coordinates. It also produces torques
as negative gradients with respect to angles. Take the example of a magnetic
dipole ¡i in the presence of a magnetic field B. If 0 is the angle between /¿ and
B, the magnetic potential energy is
-|¿z||B|cos0. (4.25)
The implied internal torque, that is, the torque on this individual dipole, and
not the moment of the force on the dipole, is then
(the reference point of this torque is at the position of the dipole), which can
be represented by a vector perpendicular to the plane formed by fi and B,
r = pxB. (4.27)
We shall now derive this vectorial result directly, along with its electric
counterpart; for simplicity, additional time derivative terms are omitted. The
torque, the moment of the force about the center of the charge distribution at
R is
r = ^ ( r a - R ) X ( e aE (r J + i eav a X B ( r j ) . (4.28)
a ' '
The part proportional to the electric field is, when we neglect the variation of
E over the system, the electric torque
te = dxE(R), (4.29)
as expected in analogy with (4.27). In deriving the magnetic torque, we first
make the unimportant change, v a —►v a — V , using (4.21), and then transfer
the time derivative to get
tb = X > a - R ) x ( i Ca( va —V ) X B ( R ) )
- £ ( v‘ - v ) x - e a(ra - R ) x B ( R )
1
f « r<* " R ) x [ ( v « - V ) X B ( R ) ] - ( va - V ) x [ ( r „ - R ) x B ( R ) ] }
2 z—' c
a
= pxB(R) (4.30)
where in the second line we have omitted the —7 JjB = V x E contribution as
negligible in comparison with r E . [See Problem 4.2 for a justification of (4.30).]
In the third line, we averaged the two preceding forms, and then used the first
identity in Problem 1.1. Putting all this together, we find the torque on the
system is given by
t = d x E + v X B, (4.31)
so that, as with the force, the result can be expressed in terms of the electric
and magnetic dipole moments, d and f,1.
38 CHAPTER 4. MACROSCOPIC ELECTRODYNAMICS
4 O®) (4.32)
Notice that we have rewritten (4.22) with the aid of the identities
V ( d •E) = d x ( V x E) + (d •V )E , (4.33)
V ( m - B ) = m x ( V x B) + ( / x * V) B . (4.34)
First a word about d and \i in these expressions. In the single atom formula
(4.22), the derivatives act only on E and B, which is reflected in (4.32). For
a many-atom system, the dipole moments could well vary from one location to
another and so have macroscopic spatial dependence. Accordingly, d (r) and
fjt(r) are the average dipole moments at the point r. We now define the electric
polarization, P , and the magnetization, M , by
and
M ( r , i ) = n (r)fi(r,t), (4.36)
(Here, the distinction between | B and J¿B has been dropped, because the
difference is of order of the small atomic velocity V , which is averaged to zero
in any case.)
4.2. FORCE ON A MACROSCOPIC BODY 39
P x (V x E) + | ( i r x B ) = ( I | p ) x B , (4.38)
V ( M •B ). = M X ( V X B ) + (M •V )B + B x ( V X M ) + (B •V ) M , (4.39)
(P •V )E - ( V •P )E , (4.40)
The comparison of this with the microscopic description of the force on charge
and current densities, ( 3 .8 ) for zero magnetic charge, suggests the definition of
an effective charge density, peff, and an effective current density, j eff, as
Peff(r,t) = - V . p ( r , f ) , (4.43)
Notice that these effective densities satisfy the equation of charge conservation,
¿P e ff + V . j eff = 0. (4.45)
It is left to the reader to verify (Problem 4.3) that the total torque, r , on the
body, the sum over all atoms of the external torques:
+ I l ( d x B) (4.46)
40 CHAPTER 4. MACROSCOPIC ELECTRODYNAMICS
„ 1 .
T = P effE + ~ J e f f X B (4.48)
C
T T h = ñ +h, J f = A /, (4.50)
V/(r,<) = V / ( M ) . (4.51)
The microscopic charge distribution 's composed of two parts. That which
is confined to atoms is called bound charge. When the remaining, “free,” mi
croscopic charge distributions are appropriately averaged, we obtain the macro
scopic densities
P — Pivee J — Jfree* (4.52)
4.3. MACROSCOPIC ELECTRODYNAMICS 41
E B p - V .p J + ^P + cVxM
peff = - V . P , (4.53)
jes = ¿ P + c V x M . (4.54)
As we have seen in the preceding section, these densities are examples of macro-
scopically measured quantities, disclosed by slowly varying electric and magnetic
fields. The physical measurements necessary for the definitions of /?eff and j efr,
since they employ slowly varying fields, should correspond to the mathematical
process of averaging involved in the definitions of pbound and jb ou n d , so we have
the identifications
Abound — P efí
jb ou n d = Jeff- ( 4 . 55)
In view o f (4.45), these two forms of macroscopic charge are separately con
served.
The correspondence between microscopic and macroscopic quantities is given
by Table 4.1: The microscopic Maxwell equations now read
V xB = ( j + ^ P + cV x m ) , V •E = 4ir(p — V •P),
-V xE = -¿ B , V •B = 0, (4.57)
c ox
which can be cast into the form of the microscopic equations if we define the
displacement, D ,
D = E H- 47tP, (4.58)
42 CHAPTER 4. MACROSCOPIC ELECTRODYNAMICS
1 3 4?r,
V xH =
„ _ _ _
D H------J,
:
V * D = 47r/>,
- - t v
c ot c
1 r)
- V x E = - — B, V *B = 0. (4.60)
c at
Note that the macroscopic charge is conserved,
V . j + J ^ = °, (4.61)
which follows from the first pair of equations. As microscopically smooth distri
butions, the density and flux of free charge will serve to measure the macroscopic
fields E and B. That is exhibited in the expression for the force on a macroscopic
charge distribution,
F = J (dr) (p E + ^ J x b ) . (4.62)
p(r) = —d •V<S(r).
A = - i r xB , V x A = B.
z
r a = raX (e „E (r a) + e„ X B (ra)) ,
first, over the charges in an individual atom, and thereby obtaining ex
pressions in terms of d a and fjtai the dipole moments of the atom, and then
over the atoms making up a macroscopic body, obtain the result that
where the external and internal torques are given by (4.46) and (4.47),
respectively. Then, verify that the torque acting on a macroscopic object
in electric and magnetic fields is given in terms of peff and j eff according
to (4.48).
Chapter 5
5.1 Conductivity
We start by considering a simple model of a metal in which the current is linearly
related to the electric field. The model is to be considered as suggestive only
but it does lead to a qualitative understanding of the important phenomena of
conduction. Of course, an accurate description requires quantum mechanics.
First consider a free electric charge (an electron) moving under the influence
of an external electric field, and subject to collisions with the atoms of the
substance. The electric field accelerates the charge, and the collisions slow it
down. Our model represents the effects of the collisions by a frictional force
that is proportional— and opposed— to the velocity. The equation of motion for
the particle, having charge e and mass m, is
any initial velocity decreases exponentially in time, due to collisions with atoms,
with I / 7 supplying the characteristic decay time. The general solution to (5.2)
is found by first rewriting it as
Jt P * v ( 0 ] = ^ E ( i ) , (5-4)
45
46 CHAPTER 5. SIMPLE MODEL FOR CONSTITUTIVE RELATIONS
and then integrating from t* = —oo (a time before any field has been applied),
to t (the time o f observation),
(5.5)
J—00
Notice that the response of the system under the action of an external electric
field is nonlocal in time. (That is, the velocity at a given time depends on the
electric field at earlier times.) The main contribution to the integral comes from
the region of time differences that are of the order of I / 7 .
The current density for a single charge is proportional to its velocity. If n /
is the (constant) density of (free) conduction electrons, then J is
(5.6)
(5.7)
my
which is a statement of Ohm’s law (Georg Simon Ohm, 1787-1854), a being the
static conductivity. [This, of course, can be more directly obtained by looking
for the static solution of (5.2).] A more general situation arises when the electric
field exhibits harmonic variation (i.e., has a definite frequency),
(5.8)
— E (w )e-iu,t (5.10)
m 7 — iu>
Here displayed is a complex amplitude for J(t), J(u>). In terms of it, the complex
conductivity, <r(o;), is defined by
(5.12)
5.2. DIELECTRIC CONSTANT 47
are, respectively, even and odd functions of u;, as shown in Fig. 5.1. Finally, we
note that the integral of the conductivity over all frequencies is
rife*
r to * (» ) = ^ 2 fd a , , r °_ * L
7-00 rn Jo m V o 1 + ^2
_ rife*
(5.15)
m
which is called a “sum rule.” The significant feature of this sum rule is that the
right hand side is independent of the frictional force constant, so that we could
use it to determine n/ experimentally. What underlies this is the simplicity of
the response to an electric field pulse that is localized at time t = 0. Without
time to act, the frictional forces are effectively absent. (See Problem 5.1.)
w 1 + u «r + 7 J tr = ¿ Re (E (w)e_ib,t) • (5 1 7 )
This implies that the steady-state solution for the position vector will also ex
hibit harmonic time variation, that is,
0
co
Figure 5.1: The real and imaginary parts of the conductivity, as given in (5.14).
5.2. DIELECTRIC CONSTANT 49
Under the usual circumstance of 7 <C the amplitude of the induced oscilla
tion becomes very large for u> = u>o, the condition of resonance.
It is now immediate to calculate the polarization (4.35) in terms of the
induced electric dipole moment and the density of bound electrons, n&,
P = n^er, ( 5 .1 9 )
nbel
p (t) = Re
m —u>2 + u)q — iyu
= R e [ x e ( w ) E ( w ) e _ u *'t ], (5.20)
nbe¿ 1
Xe(w) = (5.21)
m —u)2 — iu>7 + coq ’
which satisfies
Xe(w ) = X e ( - w ) * . (5.22)
The static susceptibility, the real value when u> = 0, is
z>2
Xe = > 0. (5.23)
nbe2
(5.25)
mu)2
Since a typical value of the electron’s kinetic energy is of the same order of
magnitude as a typical Coulomb potential energy (this is the virial theorem—
see Chapter 8 ),
p2
mv 2 ~ —e , (5.26)
nbe2 3
0 ~ ribl , (5.27)
the number of bound electrons in an atomic volume, For dense matter, where
the atoms are tightly packed, we may have, therefore,
Xe ~ 1. (5.28)
50 CHAPTER 5. SIMPLE MODEL FOR CONSTITUTIVE RELATIONS
e = l + 47rxc>l, (5.32)
where, as noted, the excess over unity can be significant in dense substances.
For example, the static dielectric constant for mica is about 6 , and for water,
about 80.
We can also derive a sum rule for x c(u>), by considering the integral
rue- —VjJ
du(-iu)Xe(w) = - du>- "Ó2 = Wo - Q t ) •
r ■ f
(u + h ) 2 ~ u f
(5.33)
For the following discussion, we assume that y < 2u>o, so that u>02 > 0. (This
restriction is actually unnecessary— see Problem 5.2.) Now think of a; as a
complex variable, and note that the integrand has singularities at
w = - | 7 ± <*>(,, (5.34)
which lie in the lower half plane (see Fig. 5.2). Accordingly, the integrand is
everywhere regular in the upper half plane, and the integral over a path that is
closed by a large semicircle in the upper half plane equals zero. The integral of
5.3. PLASMA 51
interest, (5.33), is therefore the negative of that over the semicircle of arbitrarily
large radius, which is
n¿e2 f , —i e2
------ / duj-— = TTUb— . (5.35)
m Jn u m
Thus, the desired sum rule, independent of 7 and u)0, is
f du) (-i v ) X e ( v ) = 7T
m
(5.36)
In retrospect, the same method could have been applied to the conductivity
integral, (5.15), using (5.12), which integral is obtained from that in (5.33) by
putting (júq = 0. Physically, we should have expected this result on the basis of
the a- relation, (5.15), since the two quantities e and a are just parts of a whole.
The two phenomena being discussed are just the free electron and bound electron
contributions to the total current, which have the following form for a definite
frequency u>,
so that the sum rule corresponding to this total current is proportional to the
total electron density rib + n / ,
e2
/ du[a(u) - *u>Xe(w)] = (nt + nb)it— , (5.38)
J- o o m
which expresses the fact that, in the response to an electric pulse localized at
time zero, only the inertia of the electrons matter— frictional and binding forces
have no time in which to act.
5.3 Plasma
Let us combine the results of the preceding two sections by considering the mo
tion of free charge in a conducting dielectric material, for which the conduction
current is
J = crE = j D . (5.39)
J^> + V . j = 0, (5.40)
becomes
d <T
52 CHAPTER 5. SIMPLE MODEL FOR CONSTITUTIVE RELATIONS
d or
dtP + - 4vp = 0. (5.42)
The solution to this equation, corresponding to an initial charge density p(r, 0),
is
p(v,t) = p ( r , 0 ) e - ^ > (5.43)
implying that the charge disappears from the interior of the conducting body
at a rate measured by
47T0"
(5.44)
e
The charge in the interior of the volume eventually all migrates to the surface.
(See Problem 5.3.)
The use of the static conductivity here assumes that this decay rate, 7 ', is
small compared to the frictional constant 7 , 7 ' <C 7 , or, using (5.7), that
47r n fe 2
< 7 (5.45)
e m
This must be satisfied by some combination of small density of free charge and
a high coefficient of friction. But what of the opposite situation, where there
is a high density of free charge and little friction, as encountered in a plasma?
For that, it may be clearer to return to the equation of motion (5.2), presented
as the differential equation
^ í ! e = !V£!d (5.46)
m me
G H ¿ , + v . ( | + 7) a = o, (5.48)
( d2 d 2\
(5.49)
W + ', M + V P = 0-
2 47rn/e2
wp = (5.50)
me
5.4. POLAR MOLECULES 53
(Note that the usual form of the plasma frequency has e = 1.) In general,
the solutions to (5.49) are damped oscillations. If the rate of change is small
compared to the scale set by 7 , the equation (5.49) simplifies to
which will be recognized as the previous result (5.42), including the restriction
(5.45). This limit corresponds to exponential decay. But in the other limit,
where change occurs rapidly relative to 7 , it is the 7 term that can be approxi
mately neglected,
<5-52)
/> = - v . p , (5.53)
arising from the displacement r of the electrons relative to the oppositely charged
heavy ions,
P = ne r. (5.54)
d2r
(5.55)
m ~dt? ~ eE’
or
d2 ne“
^ 2 P = -----E
(5.56)
at1 m '
From this, and the field equation
V •E = 4 ir(—V •P ), (5.57)
With the exception of atomic hydrogen where the orbital motion respects a
preferred direction in space (as in the classical elliptical orbits), atomic electric
dipole moments change direction in space so rapidly in response to the fast
electronic motion that no average effect survives. But things are different with
molecules, specifically those of a polar nature. In the example of H+ C1“ , the
hydrogenic electron is transferred to form the chlorine ion, and a dipole moment
is associated with the relative motion of the heavy ions. Other examples of polar
molecules associated with familiar substances are H2 O, SO 2 , NH3 , and CH 3 CI.
For such molecules, in isolation, it is not misleading to think of a permanent
electric dipole moment that changes its spatial orientation only in response to
the slow rotation o f the molecule.
But molecules are not ordinarily isolated; they exist in an environment in
which other molecules collide with them at a rate determined by the temperature
of the substance. The effect of these collisions is to remove any particular spatial
orientation o f the dipole moments; it still requires an electric field to provide
a preferred direction. But now there is a competition between the organizing
effect of the electric field, with its preference for lower values of the energy,
and the disorganizing effect o f the ambient temperature T. For a static field,
the net balance of that competition is expressed by the Boltzmann factor, which
gives the probability of finding a configuration of energy E ,
e ~ E /kT ^ (5.59)
where Ar, the constant of Ludwig Boltzmann (1844-1906) has the value
In thermal equilibrium, the fraction of dipole moments that are directed within
the solid angle
dSl = sin 0 d0 d<j) (5.61)
l^ d -E /fcT (5.62)
Z 4ir
where the choice of the normalization constant, the so-called “partition func
tion”
Z = f — eA ' ^ kT (5.63)
J 4tt
ensures that the totality of such fractions equals unity. Consequently, the aver
age dipole moment is
5.4. POLAR MOLECULES 55
Except for very low temperatures or very high fields, the condition |d| |E|
kT holds true, so that we may expand the exponential Boltzmann factor in the
partition function Z:
In this uniformly weighted average over all directions, where d and —d appear
with equal weight, the average of d is zero, and that of the square of some
component, say d2 z , is the same as any other component, so that
d 1 d2 p
(5.67)
0E 3 (kT )2 ’
(d )T « (5-68)
What kind of limit does our weak field condition impose on the magnitude of
the electric field? If we recall that room temperature, T = 300K, corresponds
to kT « eV, and typical atomic dimensions set the scale for the dipole
moment, |d| 1 0 ” 8 e cm (where e represents the magnitude of the charge on
the electron), then |d| |E| C kT requires
which is indeed a large electric field. (For the average dipole moment when
this approximation is not valid, see Problem 5.4.) We note that one of the
largest observed moments, that of potassium chloride, (foci = 2.14 x 10~8 ecm,
is slightly more than twice the value that was used in the estimate of (5.69).
With a density of nmoi polar molecules per unit volume, the (weak field)
contribution to the polarization is
(5.70)
P - n m o l 3fcTE ’
and the complete static susceptibility becomes
d2
Xe — Xe,atom “h nmo\ (5.71)
where Xe,atom is the atomic susceptibility due to the induced dipole moments
of the atoms [see (5.23)]. This result is accurate for a “polar gas,” where the
densities are low. Note that the first term in (5.71) is independent of T, while
56 CHAPTER 5. SIMPLE MODEL FOR CONSTITUTIVE RELATIONS
(5.72)
is directly verified to be
' d2 7
(d )(t) = Re (5.74)
S kTy — iu
m
The implication for the static situation, u = 0, and for the high-frequency
limit, u) 7 , are as already described. In particular, (5.74) expresses the fact
that as uj increases, (d) decreases eventually like y/u. In general, the static
susceptibility contribution of the polar molecules is multiplied by
1
(5.75)
1 — iiOT ’
where
r = - (5.76)
7
is the relaxation time of the electric dipole moments. The molecular effect
disappears at relatively low frequencies, leaving only the atomic contribution,
since the atomic frequency, u>0, is much greater than the molecular frequency,
7-
5.5. CL A USIUS-MOSSOTTI EQUATION 57
Here the driving force has been taken to be eE where the macroscopic field is
E( r, f ) = e (r,¿). The same assumption for the driving field was made for the
alignment of polar molecules in Section 5.4. This is incorrect since e includes the
field of the atom (or molecule) itself, the effects of which are already represented
in the harmonic restoring force. The correct driving field is the field acting on the
electron due to all the other atoms (we use the word atom to stand, generically,
for either atom or molecule),
where now the overbar represents a spatial average over a volume V which
contains exactly one atom, that average volume per atom being the inverse of
the density,
n=y- (5-79)
[In (5.78) we assume that no significant contribution is produced by neighboring
atoms, so that Edriving does not differ appreciably from edriving? and also that
the average over the atom is already sufficiently representative of a macroscopic
average that the field E can be introduced. What we are doing should not be
expected to apply for a strongly polar liquid or solid where the forces produced
by neighboring molecules could be the dominant effect.] The field due to the
atom in which the electron is located is
e atom (5.80)
I*- r a l ’
where the summation extends over all charges in the atom. We can calculate
eatom by averaging over a sphere (of radius a and volume V = 47ra3 / 3 ) which
is large enough to include the atom; the negative of the average field can be
written as
(5.81)
-* = -b
An individual term here,
where R is the position vector of the center o f the sphere. The last form merely
uses (with subscript a) the fact that
V •(r — R ) = V •r = 3. (5.84)
And now the spherical symmetry of the situation, telling us that the electric
field at a point must be directed along the line from the center, immediately
yields (see also Problem 5.5)
4?r
e a(ra) = -g-nea(r 0 - R ), (5.85)
and then
_____ 47t 4tt _ 47
e atom = — - R ) = — nd = — P . (5.86)
We conclude that the correct driving field (subject to the caveats mentioned) is
47T
^driving =: E H P• (5.87)
To appreciate what effect this has on our earlier results, let use denote by
what was previously, and incorrectly, called the susceptibility. For static fields,
then,
d2
xé = nb + nmol (5.88)
rrnuÁ 3k T ’
and now we have
(5.89)
or
(5.90)
Xe , _ Xe
(5.91)
1 m Ú 2 .v ^ e’ ^e 1 _ i2_ / >
1 t" 3 1 3
it is clear that the earlier identification of Xe with x*e ls valid only when x e < l ,
as in substances of low density. Yet another way of presenting matters, one that
introduces the dielectric constant according to (5.32)
4irxe = e - 1 , (5.92)
namely
e — 1 _ 47T ,
(5.93)
7+2 - T Xe’
is known, from its historical origins, as the Clausius-Mossotti formula [O. F.
Mossotti (1850), R. Clausius (1879); the years cited indicate the dates of signif
icant publication].
5.6. PROBLEMS FOR CHAPTER 5 59
When one turns to fields of definite frequency, the names change, the relation
e(w) - 1 47T ,
(5.94)
Í H ! = T * '* " *
being called that of Lorenz and Lorentz [L. Lorenz (1869) and H. A. Lorentz
(1880)]. For the particular example of a nonpolar substance, where [(5.21)]
n&e2/m
X » (5.95)
u;$ —ÜJ2 — iujy ’
^ ^ 47rx'e(a;)
(5.96)
immediately gives
nje2/m
X e (w ) = (5.97)
WÍ?1 —
—a; — «20^7
>
47Tn&e2
Wi = wñ - = <^o> (5.98)
m 6+2
where e is the static dielectric constant given by 1 + 47TXe(0). Clearly the sum
rule (5.36) remains unchanged since it is independent of uo.
We conclude with an experimental example of a situation in which the modi
fication of the driving field is both significant and accurate. The static dielectric
constant of nonpolar nitrogen gas has been measured at low and high densities.
At the density 0.06604 g /cm 3, the observed value of e — 1 is 0.03109. If e — 1
were equal to 47r%g, which is proportional to the density, the value predicted for
a density of 0.5780 g /cm 3— slightly more than half the density of water— would
be e — 1 = 0.2721. The measured value is 9% higher: 0.29633. If, however, one
uses the Clausius-Mossotti relation (5.93), it is (e — l ) / ( e + 2) that is propor
tional to the density. Now the value of e — 1 predicted for the larger density is
0.2959, which falls short of the measurement by only 0.1%.
2. Evaluate the integral in (5.33) in the case that 7 > 2u>o, and, thereby,
establish the sum rule (5.36) in general.
3. Show that the charge that disappears from the interior of a conductor
according to (5.43) appears on the surface of that conductor.
60 CHAPTER 5. SIMPLE MODEL FOR CONSTITUTIVE RELATIONS
X (<fr)
k —r«|
= 47T
1
2a
2 1 2
6 r*
where the volume integral is over a sphere of radius a which contains the
point r a, and the origin is taken to be the center of the sphere. [Hint:
Take r a to lie in the z direction.] Upon taking the gradient in (5.81), the
result in (5.86) follows.
8. On the basis of the formula derived for the dielectric constant, (5.32)
and (5.23), estimate as accurately as you can, based on simple physical
arguments, the value of e for air and water, in the regime u>moi <
^atom- What about the regime probed by capacitive measurements, cv <C
u>moi, where, for water, e « 80? How closely do your estimates agree with
the observed values? Discuss possible sources of error in your estimate.
How are your results changed if the Clausius-Mossotti equation (5.93) is
used instead?
10 . Consider the Hall effect discussed in the previous problem. If the mag
netic field (perpendicular to the current flow) is in the 2 direction, the
diplacement current is neglected, and the resistivity l /<7 is very small,
show that there exist waves of the form exp[i(kz — ut)] in E and J, with
the dispersion relation
Be
U) = k2
2wne
These are helicon waves, which provide a means of measuring 1/ne.
5.6. PROBLEMS FOR CHAPTER 5 61
11. (a) Show that Ohm’s law for a neutral conducting fluid moving with
velocity v is
J = <r ( e + -^ x b ) ,
dB „ . c2
—— = V x ( v x B ) + -— V 2B.
at 7 47TCT
(Argue that the displacement current, ignored here, gives only v2/c2
corrections.)
(b) For a fluid at rest this means that B satisfies the diffusion equation
dB _ c 2
V 2B.
dt Aircr
If B varies over a characteristic distance L, what is the characteristic
time r for the decay of the field? Estimate r for the earth’s core,
where L ~ 106 m, <7 ~ 1017 esu.
(c) For times short compared to the diffusion time r, B satisfies
dB „ ,
= Vx(vxB ).
Show that this means that the magnetic flux through any closed loop
moving with the local fluid velocity is constant in time.
w + V .(pv) = 0
(conservation of fluid),
d 1
P q,-V + p(v •V ) v = - V p - — B X (V X B )
Í = V X (.X B )
B = B 0 + B i( r, t ) ,
P = po + p i ( r , t ) ,
v = v^r.f),
62 CHAPTER 5. SIMPLE MODEL FOR CONSTITUTIVE RELATIONS
d2
O P Vi - s 2V ( V •V i ) + VA X ( V X [ V X ( v i X v a )]) = 0,
s2
Bo
"VA ri
y jw p
13. Show that for a collisionless plasma, for which the charge density satisfies
Magnetic Properties of
M atter
( 6 . 1)
V - B = 0, ( 6. 2)
B = V x A. (6.3)
(6.4)
(6.5)
63
64 CHAPTER 6. MAGNETIC PROPERTIES OF MATTER
E + ~ A = -V<j>. ( 6 .6 )
c ot
d _ , e d . e . Ax
m —-v = -e V < ^ -----— A H— v X (V X A ), (6.7)
at cot c
d cd t-t ( i ^
( 6 .8 )
ms y = - . ; s A _ e V AJ'
v X (V X A ) = V ( v •A ) - (v •V ) A , (6.9)
and Q I
p = mv + - A , (6.11)
c
ip = -v (* -iv .A ). (6.12)
jt { l ™ 2) = eE•'v = - eé + + 4 (* - r A) ’ (613)
where we have used ( 6 .6 ), as well as (6.10) for <j> instead of A . [In (6.13), v is
an implicit function of t, so d/dt does not act on it.] Rewriting this equation as
é a = z H- <6'20)
where, in the case we are considering here, the variables are identified as
q -> r , p - + p, ( 6 .2 1 )
6.2 Diamagnetism
We now apply this formalism to a simple mutually interacting system (an atom)
immersed in a homogeneous magnetic field, B, for which the vector potential
can be taken to be
A = ^BXr. (6.22)
(6.23)
where the last term, U ) represents the potential energy of atomic forces that
keep the atom together. The energy can be rewritten as
( B x r 0f , (6.24)
where the first term is the ordinary expression for the energy of the atom. The
second term in (6.24) involves the intrinsic magnetic moment of the atom, /x0,
defined by
= X ^ 9 ^ r _ra><Po’ (6-25)
66 CHAPTER 6. MAGNETIC PROPERTIES OF MATTER
when r„ is the position of the ath particle relative to the center of mass of the
atom. This is very similar to the definition of the magnetic moment given in
(4.19),
v = Yl^ñ;raXVa’ (6-26)
but is not the same since p ^ mv. To see that m is the actual magnetic moment,
which is required to satisfy [see (4.24)]
dE
(6.27)
dB ~
we note that the change in the energy, 6&E, caused by a change in the magnetic
field, <SB, is
5b E = ^ Pa ~ i .
? Xro .lirax5B
ma 2c
= - ^C^ra><Va^*¿B = (6.28)
The actual field-dependent magnetic moment can then be read off from (6.28)
to be given by (6.26) or explicitly by
which can be derived from (6.26) by use of ( 6 . 1 1 ), the second term being an
induced effect. [Note that this argument provides a demonstration that the
intrinsic magnetic moment Mo is independent of the magnetic field B.]
We first consider an atom which has zero intrinsic magnetic moment,
Mo = 0 . (6.30)
This will be true for a spherically symmetric system or, since for a single particle
for a system with zero total angular momentum as long as all the particles
involved have the same e/m ratio (which is very nearly true for an atom). For
such a spherically symmetric system, which has no preferred direction, we see
that
(rr) = - r 2 l , (6.32)
and hence
( r X ( B x r ) ) = ([r2B — (r *B)r]) = - r 2B, (6.33)
o
so that the time average of m is
■ r2B (6.34)
**— E 6 m aC2 a
6.3. PARAMAGNETISM 67
Note that the induced magnetic moment is directed oppositely to the magnetic
field. If n is the density of such atoms, the magnetization (4.36) is
where the last step is justified since —Xm <C 1. The relative sizes of Xm and Xe>
from (5.23), are estimated by the ratio
(6.39)
if we use typical atomic dimensions r and speeds v. The effect we have discussed
here is called diamagnetism, which is universally exhibited by all matter.
That the induced magnetization in atoms is opposite in sign to the inducing
field is a microscopic example of what is called Lenz’s law [Heinrich Lenz (1804-
1865)]. Bodies for which this effect is the dominant one are called diamagnetic.
They are repelled by regions of strong magnetic field, where the energy is in
creased. (See Chapter 26.) The characterization of diamagnetism, /i < 1 , is seen
to be analogous to the usual situation for dielectrics, e > 1 , since the parallel
relationship between magnetic and electric quantities is H = ^-B and D = eE.
We close this section by remarking that the simple formula (6.38) suggests
that the ratio of susceptibility to density for a given substance is independent of
temperature. This is almost universally valid experimentally. In the example of
water, where the mass density is practically constant, the measured susceptibil
ity is close to —0.7 x 10"6. The major exception is the metal— with the largest
diamagnetic susceptibility— bismuth. Its susceptibility at room temperature is
about —1.3 x 10“ 6, but decreases significantly with rising temperature. Here
the quantum insights of the modern theory of metals are indispensable.
6.3 Paramagnetism
What happens if there is a permanent intrinsic moment,
Mo # 0? (6.40)
68 CHAPTER 6. MAGNETIC PROPERTIES OF M ATTER
M r = (6.41)
m = J£| b . (6.42)
3 kT
This is appropriate to the weak field circumstance
Xra,para mv
100 at room temperature, (6.45)
Xm,dia Hr
where mv2 is related to the magnitude of energies in the atom. The estimate in
(6.45) is in general agreement with the observation that paramagnetic gaseous
oxygen at standard pressure and room temperature has a positive susceptibility
about one fifth the susceptibility of water, although the molecular density of the
oxygen is less than a thousandth of that of water. The susceptibilities of para
magnetic substances are still so small compared with unity (for liquid oxygen,
Xm = 3 x 10~4) that the approximation of neglecting the distinction between
B and H in (6.44) is well justified. The inverse dependence on temperature
displayed there was discovered experimentally by Pierre Curie (1859-1906).
Again, we have persisted in an error. In the electric case (see Section 5.5),
the correct driving electric field, Edriving was obtained by removing the field
of the atom itself. Exactly the same arguments apply here. The effect is one
of no practical importance but is conceptually significant. Analogously to the
previous arguments, we define the driving magnetic field by
Recalling from Section 1.2 that the magnetic field, which is present in a coordi
nate system moving relative to a static electric field, is
B = ~ X E , ( 6 .4 7 )
t>atom := ^ ] ~~~~X e a, ( 6 .4 8 )
a
where now e a is an individual term in the second sum in (5.80). The average
value of this field is, according to (5.85) [e^ = —ea(ra) there],
47T ,
" 3
-'atom
= E t * - = x: t * V
4ir 1 r-A ea
<±7T 87r 87r ,
= ------- > — r aX v a = — n\i — — M, (6.49)
T V ^ c T T
where we recall (6.26). Hence, the driving magnetic field is
« 871V „ 47T
Bdiriving (6.50)
which is analogous to the electric case, (5.87). Since the magnetic susceptibility
is defined by
M = XmH, (6.51)
the correct driving field is negligibly different from B ,
/ 4 tt
* driving
=( 1 +T Xm) H « B , ( 6 .5 2 )
because %m < 1 .
6.4 Ferromagnetism
The history o f magnetism did not begin with the phenomena of paramagnetism
and diamagnetism, which were first recognized by Faraday in 1845. The ancients
were familiar with the remarkable properties of Magnesian stone, the iron oxide
Fe3 Ü4 . The term ferromagnetism refers to the property of such substances,
primarily members of the iron group, of exhibiting permanent magnetization.
A simple model of this effect was introduced by Pierre Weiss (1865-1940), who
effectively postulated that the driving magnetic field within ferromagnets is not
(6.50), but rather
B driving = H + A M , ( 6 .5 3 )
(but see Problem 6.3), it is simpler and more accurate quantum mechanically
to suppose that the atomic magnetic moment is either lined up parallel or
anti-parallel to Bdriving, which defines the 2 axis. Since the interaction energies,
for the two possibilities, are
(6.55)
* = + AM). (6.56)
(6.58)
in which
Tc ^ A . (6.59)
k
In Fig. 6.1 there is plotted the left side of this equation, and also the right side
with examples of the two situations T > Tc and T < T C. For T > T C the curves
corresponding to the two sides of (6.58) intersect only at M = 0; there is no
magnetization. But, for T < Tc, there is also an intersection at a positive value
of M/rtfio that is less than unity; a permanent magnetization can exist. The
critical temperature Tc above which no permanent magnetization is possible is
called the Curie temperature. Its value is of the order Tc ~ 1000 K. If T <C Tc
all the magnetic moments are lined up,
M = n/i0, (6.60)
M/n]^
Figure 6.1: Solutions to (6.58) for temperatures above and below the Curie
temperature. The solid curves are the plots of the right side of (6.58) for T < T C
and T > Tc, respectively, while the dashed line is the left side of that equation.
72 CHAPTER 6. MAGNETIC PROPERTIES OF MATTER
which is the inverse of the typical factor that relates magnetic energy to electric
energy. The clear suggestion is that the underlying mechanism of ferromag
netism is not magnetic, but electrical, in origin. The quantum theory of ferro
magnetism initiated by Werner Heisenberg (1901-1976) vindicates this conclu
sion.
(b) Show that for such a uniform field acting on a system of charged
particles {ea}, the electric dipole moment is given by the analogue of
(6.27),
dE
<9E “
where E is the energy.
(c) Taking this differential equation as defining d generally, find the E
dependence of E in the case that d = do + #E.
(a)
<j>= 0 , As — A z = 0, Ay = K x, (K — constant)
(b)
0 = o, Ax = Ay = 0, Az - Kt
6.5. PROBLEMS FOR CHAPTER 6 73
(c)
<j) = 0, A = a(a •r), (a = constant)
(d)
a •r
, A = 0; r > 0
(e)
= 0- r > 0.
J
that is,
( a*o ) t = ^ dSlti0 exp [//0 •( H + A M )/¿ r ],
Z= J do exp[A*o •( H + XM)/kT].
Repeat the analysis given in Section 6.4 for this distribution and discuss
similarities and differences, particularly in the low and high temperature
limits.
d2 2 dr d2 _ ( 1 dH _ 1 dr
m _ r = _ mWor_ m 7 _ _ m _ R + e ^E + _ _ x B + - - x B
For E = 0 , and R (t) a circular motion with constant angular velocity fi,
show that to first order in Q there is an effective electric field acting on the
charge. What would thus be the polarization of a dielectric body rotating
in an external constant magnetic field?
c 1 d '
J •E = E « (7.1)
4 ^ V X H — 4ñ d t B
J . E = - V . (¿E X H ) - ¿ ( E •¿ D + H •| B ) . (7.3)
[Recall that if there were free magnetic currents, (7.2) would represent the work
done on the magnetic charges.] Our aim is to write this result as a local energy
conservation law. We immediately identify, from the divergence term, the energy
flux or Poynting’s vector, S, to be
S = ¿E X H , (7.4)
which has the same form as that of the microscopic flux, (3.5), except that here
B is replaced by H. More intractable is the identification of the last term in
75
76 CHAPTER 7. MACROSCOPIC ENERGY AND MOMENTUM
— ¡J ± — ( E.^ D + H (7.5)
dt 4w v
— U + V - S + J . E = 0. (7.6)
at
f = />E + -J x B . (7.7)
c
0 = ¿ (V .B )H + ¿ ( v x E+ I | b) x D. (7.9)
G = — DxB, (7.11)
47TC 7
which is analogous to (3.13) except the correct electric field here is D. If a local
momentum conservation law holds, the last two terms of (7.10)should be the
divergence of a tensor, representing the flow of momentum,
where e and ¡jl are constants. This is not an unrealistic situation for many
substances over a sufficiently limited frequency range. For this case, the energy
density and the stress tensor exist, and have the following forms:
S = -^ -E x H , G = -^ -E xH = ^S . (7.16)
47r 47tc c¿
It is interesting that we can transform these expressions, as well as Maxwell’s
equations, to look like those in vacuum, by redefining the fields, the charges,
and the speed of light as follows:
(See Problem 7.1.) The ratio of c to cf is the index of refraction for the medium,
4 = n = ^/eji. (718)
c
By this transformation we see that the speed of propagation of electromagnetic
waves in the medium is c'; for propagation in a definite direction, the transcrip
tion from the vacuum statement (see Section 3.4) that Ex and B ' are mutually
perpendicular and equal in magnitude is
eE 2 = i i H V E •H = 0. (7.19)
In the usual situation at low frequencies, where / i « 1, c > 1, the speed of light
in the medium, c', is less than that in vacuum, c, and the electric field is smaller
than the magnetic field, by the same factor:
the constants e and // with frequency dependent functions, e(u>) and /i(u>), which
for the simple model leading to (5.97), are
47rne2
// = !, e(w) = 1 + (7.21)
m u>2 —a?2 — iu>7
e(w) (7.22)
where u>p is the plasma frequency, (5.50) with e = 1. Substituting this into
(7.18), we see that the speed of light in the medium, c', is larger than the speed
of light in the vacuum: c' = £ > c- If d is the speed of energy flow, as is
true for a nondispersive medium, this is impossible, since it manifestly violates
causality. (Causality is the concept that information cannot travel faster than
the speed of light in vacuum.)
We must now return to our starting point and carefully re-examine energy
flow. A realistic electromagnetic wave is characterized by a finite spatial extent,
so correspondingly contains a finite range of frequencies. This is relevant be
cause the identification of the energy density according to (7.5) involves a time
integration, the building up of the field. But a field with a definite frequency
has no such transient behavior; a range of frequencies is required. Since the
properties of the medium are frequency dependent, the spread in frequency has
very important consequences, no matter how small the spread is.
We begin again by representing the time behavior of the electric field as a
superposition of frequencies (Fourier transform)
/
oo
du> D (u )e~ iwt, (7.25)
-OO
where [see (5.31)]
D(ci) = c(w)E(w). (7.26)
Again, since D (t) is real, the dielectric constant obeys
just as we saw in the model (7.21)— see (5.30). For simplicity, we will assume
e(u>) is real (that is, we stay away from the absorption regions) so that it is an
7.3. DISPERSIVE MEDIUM 79
(7.30)
In the last form the symmetry u> <-►—u/ has been made manifest. This can be
easily seen to be a time derivative by inserting the factor (u> — u/)/(u> — u /) so
that (7.30) becomes
5_JL
[ é j E(o.) ■E (-u .'). (7.31)
dt 87t J w —u>*
[A check of this result is to note that if we let e(a>) = e, we recover the result
for the nondispersive energy density, (e/87r)(E(¿))2.] This provides a general
expression for the electric part of the energy density.
Now suppose that the turning on of the field takes place so slowly that only
a very small band of frequencies about a central frequency & occurs (the precise
relation between turning-on time and band width is not required here— but see
Problem 7.5). Now with u « uj' « U or —a7, we have, since c(—w) = c(u;),
On the other hand, for cj & the above quantity is simply e(a7). However,
this contribution to the energy density oscillates in time with the frequency 267
and will not contribute significantly when averaged over one or more periods.
In addition, the spatial dependence will also be rapidly oscillating, suppressing
the contribution in spatial averages. In the sense of such averages, then, the
final form for the energy density in this situation is inferred from (7.29) and
(7.31) to be
H2 E2 d
(we(w)) (7.33)
u = s ^ + 8;r duj
To find the speed at which energy is propagated in an electromagnetic pulse,
we proceed as in Section 3.4. In regions without charges, p, J = 0, local energy
80 CHAPTER 7. MACROSCOPIC ENERGY AND MOMENTUM
conservation implies
U + V * S = 0, (7.34)
3t
which, when integrated over the pulse, implies
f J ( d r ) r U = J (d r )S , (7.35)
E2
(dr) E x H , (7.36)
8ir
Since we are considering a nearly monochromatic pulse, with the effects of dis
persion adequately summarized in (7.32), the relations given at the end of Sec
tion 7.2 between E and H are still true, at frequency U:
IP H2
|ExH| = E H = E2 (7.38)
v?’
The speed o f propagation o f energy is less than the speed of light in vacuum, c,
in accord with the general idea that information cannot travel faster than c.
|ve| = c
3. Repeat the discussion in Section 7.3 when e(u>) = 1 and /¿(u>) ^ 1. What
happens if both e and \i are frequency dependent functions?
4. Use the model for the dielectric constant, (7.21), for |u/2 — u/2|>> tuy} so
that 6(u/) is real, to calculate the speed of energy flow, v # , from (7.39) in
the two cases oj < and w > uq. In both cases, show that the speed of
energy flow is less than c. (While this conclusion is not altered, the story
is more complicated when \u> — uq| ~ y. Treatment of this question by
Sommerfeld and Brillouin can be found in L. Brillouin, Wave Propagation
and Group Velocity, Academic Press, New York, 1960. See Problems 7
and 8 below.)
5. Imagine an electric field E(t) is slowly built up from a value of zero in the
distant past to a value Re (E(u/)e~Ut'i) in the present. Let T be the time
scale over which this turning on of the field occurs. Compute the band
width of frequencies, Ac*/, occuring in the Fourier transform of E (t).
u>2
Ren
vE = c ------ ,
where
1+
( W2- W2)2 + W V
9. (a) The Pauli exclusion principle asserts that only one electron can oc
cupy a given state. The number of states per element of phase space
is given by
0 (<fr)(dp)
ft3
where h is Planck’s constant, and the factor of 2 arises from the two
possible spin orientations of the electron. At zero temperature the
lowest energy states are completely filled. Show that the energy of
the highest occupied state, the Fermi energy, is
V 2<j>= A2<¿,
and compute A in terms of no. What value does A have for copper?
Show that the spherically symmetric solution of this equation is the
screened Coulomb potential,
e -A r
</Kr ) = « - —
r •
(c) From the analysis of (b), we see that placing an external charge dis
tribution pext in a metal leads to an induced electronic charge density
pe. Derive the following relation between the Fourier transforms,
7.4. PROBLEMS FOR CHAPTER 7 83
*ftext(k)
e(k) =
m
<!> being the total potential, and 0ext that due to the external charge
distribution. From (b) show that
4irpext(k) 4X/Jext(k)
@ex\ ( k ) — m =
k2 k 2 + A2 ’
from which the result follows.
Chapter 8
We have already mentioned the Hamiltonian formulation for the particle equa
tions of motion [see (6.17)-(6.19)]. We now want to show that the whole system
of particles and fields is a mechanical system derivable from a Hamilton action
principle.
We start by reviewing and generalizing the Lagrange-Hamilton principle
for a single particle. The action, W 12 , is defined as the time integral of the
Lagrangian, L, where the integration extends from an initial configuration or
state at time t% to a final state at time 1 1 :
ti
dt L ( 8 . 1)
2
The integral refers to any path, any line of time development, from the initial
to the final state, as shown in Fig. 8.1. The actual time evolution of the system
is selected by the principle of stationary action: In response to infinitesimal
variations of the integration path, the action W 12 is stationary— does not have
a corresponding infinitesimal change— for variations about the correct path,
provided the initial and final configurations are held fixed,
SW12 = 0. (8.2)
85
86 CHAPTER 8. REVIEW OF ACTION PRINCIPLES
This means that, if we allow infinitesimal changes at the initial and final times,
including alterations of those times, the only contribution to SW\2 then comes
from the endpoint variations, or
6W 12 = G i - G2, (8.3)
í=H£)í~nM)' (8-4)
Here, the independent variables are r and t } so that two kinds of variations can
be considered. First, a particular motion is altered infinitesimally, that is, the
path is changed by an amount <$r:
Second, the final and initial times can be altered infinitesimally, by St\ and St2 ,
respectively. It is more convenient, however, to think of these time displacements
as produced by a continuous variation of the time parameter, St(t),
) d< , ( 8 . 8)
± ( dSt\ d_
(8.9)
dt \ dt ) dt
Because of this redefinition of the time variable, the limits of integration in the
action,
(8.10)
8.2. HAMILTONIAN VIEWPOINT 87
are not changed, the time displacement being produced through 6t(t) subject
}
to (8.7). The resulting variation in the action is now
Sii v
dr d dSt
SW12 = ¿r.V F -
™dt * d tSr dt
r_ £ _ _ 1 (d r
+ Sr *
[ m*2r v v -I* 6t —m —
2 V dt
where, in the last form, we have shifted the time derivatives in order to isolate
<$r and St.
Because <$r and St are independent variations, the principle of stationary
action implies that the actual motion is governed by
d2
v = - w’ ( 8 . 12)
LmUj +yJ
A ri
dt 2
(d r y i
= -v
dt ’
(8.13)
while the total time derivative gives the change at the endpoints,
with
L = p -^ -H (r ,p ,t ), (8.16)
W i2 = j \ p ' d v - d t H ] . (8.18)
dc . dH . dv 8H dSt
6W 12 = j f dt [|>
= 1 -isr)
(8.19)
and the equation supplied by the stationary action principle for p variations,
(8.25), also guarantees that there is no contribution here to the results of r and
t variations.
i = p * ( < ^ ~ v) + = (8-27)
where
H ( r ,p ,v ,t ) = p . v - imt;2 + V(r ,t). (8.28)
X dr ^ d dH 8H 8H
,ñr - 3 í + p - dx
i
dt dt
dp 8H \
— (p •Sr — HSt) — Sr <
at dt ^ dr )
8H \ (dH_ dH
+ 5p ¿V • + St (8.29)
dp) ov V * dt
dp dH
= -W , (8.30)
dt dr
dv dH _
v, (8.31)
dt dp ’
0 _ _dH
_ = _ p+raV) (8.32)
dH dH
(8.33)
dt dt ’
G p * <5r — HSt. (8.34)
Notice that there is no equation of motion for v since dv/dt does not occur in
the Lagrangian, nor is it multiplied by a time derivative. Consequently, (8.32)
refers to a single time and is an equation of constraint.
From this third approach, we have the option of returning to either of the
other two viewpoints by imposing an appropriate restriction. Thus, if we write
90 CHAPTER 8. REVIEW OF ACTION PRINCIPLES
(8.28) as
H ( r , p , v , t ) = § ^ + V (r ’ t ) ~ 2 ~ ( P — m v)2> (8-35)
and we adopt
V = —p (8.36)
m
as the definition of v, we recover the Hamiltonian description, (8.16) and (8.17).
Alternatively, we can present the Lagrangian (8.27) as
_ m ( dr\2 , ( dr \ m ( dr \2
L = 2{di) + ■ (8'37)
v = ^ , p = mv, (8.38)
(p — m v) •— (Sr. (8.39)
dt
In the next chapter, where the action formulation of electrodynamics is con
sidered, we will see the advantage of adopting this third approach, which is
characterized by the introduction of additional variables, similar to v, for which
there are no equations of motion.
SW12 = 0, (8.40)
independently of the choice of initial and final times. We say that the action,
which is left unchanged, is invariant under this alteration of path. Then the
stationary action principle (8.3) asserts that
SW12 = Gi - G 2 = 0, (8.41)
or, there is a quantity G (t) that has the same value for any choice of time t\ it
is conserved in time. A differential statement of that is
j f G {t) = 0. (8.42)
8.4. INVARIANCE AND CONSERVATION LAWS 91
The G functions, which are usually referred to as generators, express the inter
relation between conservation laws and invariances of the system.
Invariance implies conservation, and vice versa. A more precise statement is
the following:
Gt = -H S t, (8.45)
(8.46)
(8.48)
(8.51)
i-3
h3
©
X
u
II
II
dL
— T = —rX W = - r X r~9V
— := 0, (8.52)
at dr
since V depends only on |r|.
Conservation of linear momentum appears analogously when there is invari
ance under a rigid translation. For a single particle, (8.21) tells us immediately
that p is conserved if V is a constant, say zero. Then, indeed, the action
(8.53)
Wu=i H '- f ? "
is invariant under the displacement
Sy = 6e = constant, (8.54)
and
Gse = p *8e (8.55)
is conserved. But the general principle acts just as easily for, say, a system of
two particles, a and 6, with Hamiltonian
H = j L + l L + v ( r a - v b). (8.56)
2m a 2mb
This Hamiltonian and the associated action
8.4. INVARIANCE AND CONSERVATION LAWS 93
P = Pa+p&, ^ P = 0. (8.60)
using the example of two particles. This gives each particle the common addi
tional velocity 6v, and therefore must also change their momenta,
The action is not invariant; its variation has end-point contributions. But there
is still a conservation law, not of G = P •Svt, but of N •Sv, where
„ mar a + mi r h
R = -------—-------, M = ma+nib, (8.65)
N = Pi - M R, ( 8 . 66)
namely
dN dR
0 = P - M (8.67)
dt dt ’
is the familiar fact that the center of mass of an isolated system moves at the
constant velocity given by the ratio of the total momentum to the total mass of
that system.
94 CHAPTER 8. REVIEW OF ACTION PRINCIPLES
8H = SX(—2T + r •W ) . (8-71)
Therefore we have, for an arbitrary time interval, for the variation of the action
(8.18),
/*! /»1 j
SW12 = j d t [ 6 X ( 2 T - r .V V ) ] = G 1 - G 2 = j dt— ( p . « A r ) (8.72)
^ ( r . p ) = 2T + K (8.75)
r . p ( t i ) - r » p ( < 2) iQ
r *p (8.76)
6r = SX-,
r
(8.78)
+ -* W . (8.80)
mró r
(8.81)
which, when applied to the Coulomb potential, gives the bound-state time av
erage relation
.(8.82)
determine the relationship between the velocity v = dr/dt and the momen
tum. Compute the energy in terms of the velocity. Write the Lagrangian
in terms o f v.
96 CHAPTER 8. REVIEW OF ACTION PRINCIPLES
V = arh.
L= p • + ^m v2 - e<f>+ A. (9.1)
v (9.6)
97
98 CHAPTER 9. ACTION PRINCIPLE FOR ELECTRODYNAMICS
and find
dr
(9.7)
p - d ¡ ~ H’
1 / e \2
(9.8)
* = 2s ( » - ; a) + e *-
Here we make contact with the energy considerations of Sec. 6.1; in particular,
H coincides with the form of the energy given in (6.16).
i « A = - E -V *, (9.11)
B = VxA. (9.12)
Thus, we recognize that A(r,t), E(r, t), in analogy with r(t), p(t), obey equa
tions of motion while <¡>(v, t), B(r,¿), as analogues of v(t), do not. There are
enough clues here to give the structure of £fieid> apart from an overall factor.
The anticipated complete Lagrangian for microscopic electrodynamics is
'dtra ^ , 1
L = - v a ) + - m av 2 - e oi6(ra) + ^ v a . A ( r a)
+ E•( ) - B - V x A + ^ (B 2 - E 2) .(9.13)
47T r ; f A - v *
The terms that are summed in (9.13) describe the behavior of charged parti
cles under the influence of the fields, while the terms that are integrated describe
the field behavior. The independent variables are
(9.15)
6L = 8va • [-P a + mav a + y A ( r j ] , (9.16)
Spa SL = S p ,- ( ^ - t. ) - (9.17)
^ = - e aV a (¿(ra) - ^ . A ( r a) ) , (9.18)
so that
The volume integrals extend over sufficiently large regions to contain all the
fields of interest. Consequently, we can integrate by parts and ignore the sur
face terms. The responses of the Lagrangian (9.13) to field variations, and the
corresponding equations of motion deduced from the action principle are
V *E = 47rp, (9.24)
8A : 6L - ÍA .E
4 ttc (
1 f /J \ CA i l9E 4jr- ~
(9.25)
100 CHAPTER 9. ACTION PRINCIPLE FOR ELECTRODYNAMICS
1 9_ 47T.
V xB = (9.26)
~cdiE + T J’
6L =
¿ /<*>“ •( (9.27)
1 d
E = (9.28)
- < I M A - V *-
SL = ¿ / ( * ) » • ( ■ - V x A + B), (9.29)
B = V XA. (9.30)
We therefore recover Maxwell’s equations, two of which are implicit in the con
struction of E and B in terms of potentials. By making a time variation of the
action [variations due to the time dependence of the fields vanish by virtue of
the stationary action principle— that is, they are already subsumed in (9.23)-
(9.30)],
St : (9.31)
which is a constant of the motion, dH/dt = 0. The generators are inferred from
the total time derivative terms in (9.15), (9.25), and (9.31),
to be
G = ^ i r a .pa E -6 A -H 6 t. (9.34)
a
9.3 Energy
Notice that the total Lagrangian (9.13) can be presented as
t =I> a
<935>
where the Hamiltonian is given by (9.32). The narrower, Hamiltonian, descrip
tion is reached by eliminating all variables that do not obey equations of motion,
and, correspondingly, do not appear in G. Those “superfluous” variables are
the v a and the fields (¡> and B, which are eliminated by using (9.19), (9.24), and
9.3. ENERGY 101
(9.30), the equations without time derivatives, resulting, first, in the intermedi
ate form
r£2 + £2
H = E ( ¿ (p* - ~ A(r<o)2 + + / (dr) 87T - p<t>
(9.36)
The first term here is the energy of the particles moving in the field [particle
energy— see (9.8)], so we might call the second term the field energy. The
ambiguity of these terms (whether the potential energy of particles is attributed
to them or to the fields, or to both) is evident from the existence of a simpler
form of the Hamiltonian
E2+ B2
7 A (r a)) + j ( d*) (9.37)
8?r
where we have used the equivalence of the two terms involving given in (9.22).
This apparently startling result suggests that the scalar potential has disap
peared from the dynamical description. But, in fact, it has not. If we vary the
Lagrangian (9.35), where H is given by (9.37), with respect to E we find
ít=-¿/<iir»sE-GlA+E)=o- <9-38>
Do we conclude that + E = 0? That would be true if the <$E(r,¿) were
arbitrary. They are not; E is subject to the restriction— the constraint— (9.24),
which means that any change in E must obey
(9.39)
^0
>
tí
0
II
The proper conclusion is that the vector multiplying <SE in (9.38) is the gradient
of a scalar function, just as in (9.28),
~ A + E = -V ¿, (9.40)
c ot
as required.
The fact that the energy is conserved,
dH A
(9.42)
-3T = ° '
where
(9.43)
a J
102 CHAPTER 9. ACTION PRINCIPLE FOR ELECTRODYNAMICS
is a simple sum of particle kinetic energy and integrated field energy density,
can be verified directly by taking the time derivative of (9.36). The time rate
of change of the particle energy was computed in (6.14),
We can compute the time derivative of the field energy by using the equation
of energy conservation, (3.7),
J (d r ) U = - J (d r ) j •E, (9.45)
to be
= - /< * ) N í* - ? - s a]
a N x
(9.46)
= 0 (9.47)
dr
by charge conservation. Observe that (9.44) and (9.46) are equal in magnitude
and opposite in sign, so that their sum is zero. This proves the statement of
energy conservation (9.42).
one. (See Fig. 9.1.) The response of the particle term in (9.34) is simple:
Se • Y2a f° r the field Parf >we require the change, 6A , of the vector potential
induced by the rigid coordinate displacement. The value of a field T at a
physical point P is unchanged under such a displacement, so that if r and r + <$€
are the coordinates of P in the two frames, there are corresponding functions F
and F such that
F {P ) = F ( r) = F ( r + Se), (9.48)
that is, the new function F of the new coordinate equals the old function F
of the old coordinate. The change in the function F at the same coordinate is
given by
F (r) = F (r) + 6F(r), (9.49)
so that
SF(r) = F (r - Se) - F (r) = - 8 e •V F { r), (9.50)
for a rigid translation (not a rotation).
As an example, consider the charge density
If the positions of all the particles, the r a, are displaced by ¿c, the charge density
changes to
p(r) + 6p(r) = ^ 2 ea6(r - r a - 6e), (9.52)
where
5(r - ra —Se ) = 6(r — ra) —6 e •V r 6 ( r — r a), (9.53)
104 CHAPTER 9. ACTION PRINCIPLE FOR ELECTRODYNAMICS
and therefore
<5p(r) = -<5e.Vp(r), (9.54)
in agreement with (9.50).
So the field part o f G in (9.34) is
where the last rearrangement makes use of (9.24) and (9.30), and the vector
identity
íex(VxA) = V(¿e.A)-(¿€. V)A. (9.56)
Including the particle part from (9.34) we find the generator corresponding to
a rigid coordinate displacement can be written as
G = 6 e- P, (9.57)
where
P = E(
a
p« - 7 A (r a)) + ¿ / ( * ) E X B = 2 > 0v a +
J a J
f (dr) G , (9.58)
with G the momentum density, (3.13). Since the action is invariant under a
rigid displacement,
0 = 6W = Gi - G2 = (P i - P 2) •«r, (9.59)
we see that
P i = P 2, (9.60)
that is, the total momentum, P , is conserved. This, of course, can also be
verified by explicit calculation:
£ ’ 1. „
J(dr)±-EXB = - J ( d r ) p E + - J X B
dt c
= - E e« (E(rfl)+JvaXB(rjj, (9.61)
which restates (3.15), from which the constancy of P follows from (6.1).
Similar arguments can be carried out for a rigid rotation for which the change
in the coordinate vector is
6r = Su>Xr, (9.62)
since a vector transforms in the same way as r, so the new function at the initial
numerical values of the coordinates is
G= (9.66)
the electric and magnetic fields defined by (9.28) and (9.30) remain unaltered,
for an arbitrary function A. This is called gauge invariance; the corresponding
substitution (9.68) is a gauge transformation. [The term has its origin in a now
obsolete theory (1918) of Hermann Weyl (1885-1955).]
This invariance of the action must imply a corresponding conservation law.
To determine what is conserved, we compute the change in the Lagrangian,
(9.13), explicitly. Trivially, the field part of L remains unchanged. In considering
the change of the particle part, we recognize that (9.68) is incomplete; since v is
a physical quantity, p —(e /c )A must be invariant under a gauge transformation,
which will only be true if (9.68) is supplemented by
(± X + * ± .V > )
\dt dt J
a
d
(9.70)
dtW’
where
w = J 2 ~ x (Va’ t)- (9.71)
W 12 = W \2 + ( w i - w 2 ) y (9.72)
G = G -f 6w. (9.73)
This alteration reflects the fact that the Lagrangian itself is ambiguous up to a
total time derivative term.
To ascertain the implication of gauge invariance, we rewrite the change in
the Lagrangian given in the first line of (9.70) by use of (9.20),
In view of the arbitrary nature of 6\(r, t ), the stationary action principle now
demands that, at every point,
s / . + v . j = o, (9.77)
that is, gauge invariance implies local charge conservation. (O f course, this
same result follows from Maxwell’s equations.) Then, the special situation
6X = constant, where 6A = 8<j) = 0, and W\2 is certainly invariant, implies
a conservation law, that of
Gsx = -6 X Q , (9.78)
in which
= /(* )/ (9.79)
{(VA) = - « ¿ ( V A ) + ( - I | a) cV í í, (9.82)
(9.83)
or, equivalently,
8A. — cStlij “I- V(<^c<5t), (9.86)
(9.88)
Then we can apply the field variation (9.88) directly, and get
¿ ¿ fie ld = S t^ C ñ M ~ ^ E 2 § ¡ St ~ ¿ E X B *
= -— (« ¿ a d d ) - ¿ ( # 2 + - ¿ E x B •VSt. (9.90)
(9.91)
where
(9.92)
a
and
(9.96)
9.6. GAUGE INVARIANCE AND LOCAL CONSERVATION LAWS 109
the last step exhibits both the explicit and the implicit dependences of 8t(ra,t)
on t. In computing the variation of <j>(ra, t ) } for example, we combine the po
tential variation given in (9.85) with the effect of 8r a:
t) = - M ~ ~ Stv« * ~ ( 9 -97)
and, similarly,
A 1
<$A(ra(t), t) == —8t— A + <j)cV8t — 8tva •V aA = —6 t ~ A + <j>cV8t. (9.98)
Ea = ^ mav2
a. (9.101)
We have retained the particle symbol d/dt to the last, but now, being firmly back
in the field, space-time viewpoint, it should be written as d/dt, referring to all
t dependence, with r being held fixed. The union of these various contributions
to the variation of the total Lagrange function is
p\ r\
¿A ot = — U tot-Q ^fo — S tot • V f t , (9.102)
and
S tot= ¿ E x B + I ^ ¿ (r - r «(0 )^ a V ai (9.104)
a
are physically transparent forms for the total energy density and total energy
flux vector.
To focus on what is new in this development, we ignore boundary effects in
the stationary action principle, by setting the otherwise arbitrary 8t(r, t) equal
to zero at t\ and 12 . Then, through partial integration, we conclude that
¿ C /t o t + V. S to t = 0 , (9.106)
( d~v \ 1 g
L ( r ,p ,v ,t ) = p* ( "¿i - v ) + 2 mv2 ~ e,K M ) + - v * A (r ,i).
(a) Reduce to Lagrangian form and then derive the equations of motion.
(b) Reduce to Hamiltonian form and then derive the equations of motion.
(c) Show the equivalence between a) and b).
2. Verify the transformation laws under rigid rotations for scalars and vec
tors,
6S(r) = - ( ¿ w x r ) •V S (r),
SV(r) = - ( i w x r ) •V V (r ) + 6w X V (r),
by considering the transformations of the charge and current densities
(9.21).
5. Verify directly the local conservation law obeyed by i/tot and Stot, (9.106).
Einsteinian Relativity
In contrast, the masses of the charged particles are fixed quantities that have
no reference to the particle’s state of motion and its associated energy. Another
way of expressing this lack of mechanical unity between fields and particles
comes from the physically evident expression for the total momentum density
(Problem 9.6)
The relation
G t ° t = ^2 S tot, ( 1 0 .3 )
which is valid for the field terms, does not hold for the particle contribution
[see (9.104)]. Could it be that Newtonian mechanics is mistaken, and that the
correct expressions for particle inertia and energy do satisfy m = E/c2? We now
follow this unifying suggestion— that the relation between inertia and energy,
which the electromagnetic field has disclosed, is, in fact, universally valid.
Consider a single particle in the absence of applied electromagnetic fields.
What we are proposing is that the connection between momentum and velocity
is actually
p = ;§v- (10-4)
ill
112 CHAPTER 10. EINSTEINIAN RELATIVITY
dE
(10.5)
V _ 5p’
and deduce that
c2p •dp = E dE, ( 1 0 .6 )
which is integrated to
E 2 = c2p2 + constant. (10.7)
We already know (Section 3.4) that the added constant is zero for an electromag
netic pulse, moving at the speed c. What is its value for an ordinary particle?
The energy (10.7) is smallest for p = 0, when the particle is at rest. Then we
write
p = 0: E = ra0c2, (10.8)
where mo is the mass appropriate to zero velocity— the rest mass. Therefore,
we have in general,
The last momentum construction exhibits the relation to, and the limitation
of, the initial Newtonian formulation of particle mechanics. For speeds small
in comparison to that of light, |v| <C c, the momentum is p = mov and the
particle inertia is constant. This is the domain of Newtonian mechanics. But
even here something is different, as we see from the energy derived from the
approximation
v ( 10. 12)
-<1
c
namely,
E « moc2 + im o v 2. (10.13)
¿i
In addition to the Newtonian kinetic energy ^mo^2 there is a constant, the
rest energy moc2, displacing the Newtonian origin of energy. The same thing
appears in the momentum form of E, (10.9), as
p < m 0c 2 :
9
moc 2 +
P‘
E (10.14)
2 mo
10.1. RELATIVISTIC MODIFICATION 113
For speeds approaching the speed of light we enter a new physical domain,
one where the speed of light is an impassable barrier. This we can see from the
particle velocity, exhibited as
p (10.15)
v= c
(p2 + m 2c 2 )V 2 ’
it is such that
v<c, (10.16)
with the equality sign occurring only for rao = 0. As for the last conclusion, it
is not unreasonable that a system, such as an electromagnetic pulse, which can
never be at rest, has no rest mass.
Now we must reconstruct the Lagrangian-Hamiltonian dynamics of particles.
If we omit the potential from the Lagrangian of (8.27), we have
1
L= P + 2 mv
2
(10.17)
E = p •v — L(v), (10.19)
m0v
o -s r
moc*
L(v) ( l _ * ; 2 / c2 ) l/2 (1 - „2/ c2 ) 1 / 2
= —rriQC ( 10 .20)
In the Newtonian regime, the L(v) reproduces the original form to within a
constant,
- < 1 : L(v) = —m 0c 2 + - m 0t;2 . (1 0 .2 1 )
c ¿
<$r(f) = 6 v t , (10.24)
8t = ^ 6 v 'T , (10.27)
the additional contribution, —H d(6t)) will cancel (10.26), and the action is
invariant under the combined space and time transformations
6r = 6 v t, 6t = ^ 6 v r. (10.28)
In view of the invariance of the action, the implied conservation law should now
follow directly. Indeed
where ( H = E )
N = p (10.30)
c1
is conserved:
dN E dr
(10.31)
dt ^ c 2 dt
and we have recovered our starting point, m = E/c2.
But, from this initial dynamical modification of Newtonian dynamics has
now emerged a change in Newtonian kinematics: The absolute distinction be
tween time and space has been removed. That is emphasized by the fact that
10.2. LORENTZ TRANSFORMATIONS 115
The fact that an observer in uniform relative motion will assign different values
to the elapsed time, and to the distance traversed, but agree that (10.33) is still
valid, means that he also measures the speed of light as c. This is Einsteinian
relativity.
One might object that it could all be true only for infinitesimal transforma
tions. But, from infinitesimal transformations, finite transformations grow. We
make this explicit by letting Sv point along the 2 -axis, so that
(10.35)
d2z(0) d2ct($)
= z(0), = ct(e), (10.37)
dd2 de2
which are solved by the hyperbolic functions, cosh 6 and sinh 9. The explicit
solutions o f these equations that obey the initial conditions
z( 0 ) = z, ct( 0 ) = ct,
are
ta n h 0 = ^ , (10.40)
which reduces to (10.35) for infinitesimal values of these parameters. Then, the
constructions
(1 <* + ” <)’
along with
x' = x, y' = y. (10.44)
We see that the point with coordinates
(10.45)
o
-Si
II
II
II
1
(10.46)
II
II
II
O
It is the origin of the new reference frame which therefore moves with velocity
—v relative to the initial one. See Fig. 10.1.
To see that r 2 —(ct ) 2 is left invariant by these finite transformations, it helps
to present (10.43) as
1/2
_ ( 1 + v /c\
z* + ct' > (z + ct),
1/2
_ fl-v / c \
z' - ct' 1 (z — ct), (10.47)
\1 + v/ c)
for it is immediately apparent, on multiplication, that
# = 0, y —0, z= 0 (10.50)
implies
x 1= 0, y( = 0 , z' — vtf. (10.51)
Then, should not the transformation that produces the unprimed coordinates
from the primed ones be of precisely the same form as (10.43), (10.44) but with
the sign of v reversed? Indeed it is, as is evident on rewriting (10.47) as
1/2
Í 1 — v/c
zA d
\1 + v/c
z — ct ( * '- 0 , (10.52)
(the other set emerges by the systematic substitution of —c for c) which imme
diately yields
(10.55)
with
(10.56)
For any of the square root factors, the variation of the value of the appropriate
v/c from — 1 to - f l changes the square root from 0 to oo; it is a positive number,
and the product of two positive numbers is again a positive number. In other
words, no succession of Lorentz transformations can produce a net transforma
tion with |v| > c. The specific value of v in (10.56) is identified by writing this
relation as
1 + v/c _ 1 + v\v2/c2 + (vi -f v2)/c
(10.57)
1 — v/c 1 + v\v2/c2 — (vi + v2)/c'
or
V\ + v2
(10.58)
1 + v\v2/c2 ‘
Simple addition of the velocities occurs only in the Newtonian regime, \v\¡2\<C c.
6A = xB + i« v r E + V Svt •A + -<$v •
The use o f the latter simplifies the calculation of the field variations; they emerge
as
SE = -S coorE - i t fv x B ,
c
SB = -¿ c o o r B + it fv x E . (10.62)
c
Then further differentiation in accordance with the field equations,
1
', = 4 i V ' E'
(10.63)
yields
Sp = SqooyP “b ^ V#j>
Sj = -¿coorj + Svp. (10.64)
Notice that the Lorentz transformation properties of j / c , p are the same as those
for A , <j>. If we introduce a new symbol ‘S’ that involves both changes of fields
(5) and of coordinates (<$Coor)> then its meaning as applied to any field F (r ,t) is
*S’F (r, t) = (F + SF)(r + fir, 1 4 - St) - F (r, t) = SF(r, t) + ScoorF (r , t). (10.65)
Note that the use of Scoor in (10.59) is consistent with this, since A(r,t) is a
scalar, ‘<$’A(r, t) = 0.
We must now examine the response of the various parts of the total Lagrange
function. First consider
¿ fie ld = ¿ ( £ 2 ~ B 2 ). ( 10 . 66 )
120 CHAPTER 10. EINSTEINIAN RELATIVITY
It is immediately apparent that the contributions of the last terms in the trans
formation equations (10.62) just cancel and
In contrast with the procedure of (9.91), our introduction of the Lagrange func
tion now is dictated by the impossibility of maintaining a common time for
particles at different spatial points— that Newtonian concept has disappeared
in the world of Einsteinian relativistic kinematics. Accordingly, we give each
particle its individual time coordinate, and present its contribution to the action
as
j dta La - J ( - m 0ac)[{c<Uay - ( d r a)^ = J dt(dr)Ca, (10.71)
where
[Unlike (9.93), interaction terms are not included here.] The last integral is
extended over the trajectory of the particle, with r a varying as a function of
ta, or, better, with r a and t a given as functions of some parameter that is
not changed by Lorentz transformations. Apart from a sign change and the
considerations of infinitesimals, the space-time structure of the square root has
just the invariant form in (10.32). Then, in response to
with the last form following from the delta function property, and the resulting
change is just
^coor[^(r ra)^(^ ” ^a)]* (10.75)
10.3. TRANSFORMATION OF FIELDS 121
so that the total variation in the sense of (10.65) is then zero. More generally
this can be expressed as follows: On introducing the transformed particle and
field variables associated with the transformed coordinates,
the Lagrange function of the system of interacting charges and fields is invariant
under the Lorentz transformations of Einsteinian relativity.
The action
W i2 = ^ dt{dr)C (10.78)
We shall not repeat the discussion of the various conservation laws in the light
of these relativistic modifications. It should be sufficiently evident that such ex
pressions as the total energy density and energy flux vector, (9.103) and (9.104),
will be regained with Ea replaced by the relativistic energy ( 1 0 . 1 1 ). And, cer
tainly (10.3) is now satisfied!
Finally, we supply the finite version of the Lorentz transformations for the
various fields. Proceeding in analogy with (10.36), we write (10.60), for a
Lorentz transformation along the z axis, as
K ir 'X ) = H *(r,<)) .
dE,(0) . dBm
dO ’ dd
122 CHAPTER 10. EINSTEINIAN RELATIVITY
dEx(0) dB,
By(0),
dO - i r = «-W '
dEy{0) dBx {6)
= -E y {9 ), (10.82)
d$ dO
and therefore
W O = ( i _ ' ¿ 2) 1 / 2 Í ^ ¿) + ^ ( M ) ] ,
s y(r'> 0 = (i •
ti2/c2)i/2[g y(r ’ *) + (10-83)
and
in L (r,p , v, t) in Problem 9.1. Use the Lagrangian viewpoint and find the
equations of motion.
3. Find the explicit form of 9 in terms of v/c as a logarithm. How does the
composition law of velocities appear in terms of 01
4. Verify (10.61) and use the results to derive the statements o f (10.62).
what is
« (V ), « (£ )?
7. Use these variations of derivatives, and the total variations of fields defined
in (10.65), to confirm that Maxwell’s equations are maintained under the
combined variations.
8. Show that the finite Lorentz transformations have the vectorial form [7 =
( l - i ^ / c 2) - 1/ 2]
72 1
r' = r + 7—------r w * r + yvt,
1 + 7 cJ
t' = 7 ( t + I v . r ) .
10. As in Problem 1.6, let Greek indices like range over // = 0 ,1 ,2 ,3 and
let repeated indices be summed. We have identified (<^,A) as a four-
vector [see (10.81)], which we will denote by A**. Problem 10.6 identifies
another four-vector that we may call d^ defined by do = d/d(ct), d% = V¿,
i = 1, 2, 3. Also note that the invariant scalar product of two four-vectors
is
ApB** = A •B — A qB q.
Here, upper index quantities ( “contravariant vectors” ) are related to lower
index quantities ( “covariant vectors” ) by, for example,
A^ = ( - A o , A ) = ^ M „
¿0123 = -hi.
124 CHAPTER 10. EINSTEINIAN RELATIVITY
(a) Identify the components of FyU and * FyU with the fields E and B.
(b) Show that by performing a vector transformation on each vector in
dex, as in (10.81), FyV transforms in a manner consistent with (10.83)
and (10.84).
(c) Write E 2 — B 2 and E •B in terms of Lorentz invariant combinations
of Fpj, and *FyV,
(d) Write Maxwell’s equations (1.65) in terms of F ^ , *FyU) dy and j** =
11. In covariant notation, the Lagrange function for the electromagnetic field
interacting with a prescribed current = (/>,j) is (we set c = 1 for
convenience)
C= T - dyA „) + + 8 ^ “ = 0.
4ir
The action is obtained from the Lagrange function by integrating over all
space and time, W = f(d x )£ , where (dx) = dt(dr). In £ the vector poten
tial, A[¿, and the field strength tensor, F ^ , are regarded as independent
variables.
6W = [ (dx)dy8xptfiU,
(e) Use the action principle to show that t^v is conserved, dyt^v = 0.
(f) Verify that t00 is the energy density, t 0t is the energy flux vector (or
the momentum density), and that t^ is the three-dimensional stress
dyadic discussed in Chapter 3.
(g) What is the trace of What is the significance of that result?
Chapter 11
E — E e \e c “b -É^mag) (li.i)
with
Eelec = J (d r ) ¿ (e • + ± E 2) ( 11.2 )
and
= J (*) [-ij •A + ¿ (b •V x A - ÍB>) (11.3)
125
126 CHAPTER 11. STATIONARY PRINCIPLES FOR ELECTROSTATICS
Derived as it is, by a specialization of the action, this energy must also have
the property of being stationary for variations about the solutions of the field
equations. Indeed, independent variations of <^(r) and E (r) gives
J
to read
SE = (dr) 64> (/> - ^ ^ ) + ¿ Í E •(V<¿ + E) ( 11 .8)
where we have ignored the surface term since we assume the integral extends
over the entire region where the fields are nonzero. (In particular, we explic
itly assume that (j) retains a zero value at infinity: S# = 0.) The stationary
requirement on the energy Ü7, SE = 0, now implies the two basic equations of
electrostatics,
V •E = 4ttp, E = - V < ¿ . (11.9)
Of course, the connection between E and <t> implies
V x E = 0. (11.10)
Having seen this, we can immediately modify the energy expression, (11.5),
to incorporate the effects of a dielectric medium,
eE -V<t> eE2
E = ( 1 1 .1 1 )
Air 8tt
The validity of this form is indicated by noting that if it is required to be
stationary under variations in (j) and D = eE, we recover the equations of
electrostatics in a dielectric. That is, the variation in the energy,
SE = J(dr) [ i ^ - i - V . D ) + — ¿ D . ( V ¿ + E) = o, (11.12)
47T
implies
V . D = 4ttp, E = -V <¿. (11.13)
(As before, in writing the variation in the form of (11.12), we have integrated
by parts and omitted the surface term.)
The discussion of action forms has accustomed us to the idea of restricting
the choice of variables by adopting one or more of the (field) equations as defini
tions. Here, we can use two restrictive versions of the above energy functional,
(li.ii).
11.1. STATIONARY PRINCIPLES FOR THE ENERGY 127
E = -V < ¿, (11.14)
the curl of which is zero. The energy, (11.11), as a functional of the potential,
for which
6E[4] = 6 \ J (d r ) [ p < ¿ - ¿ e E 2 = 0 , (11.19)
supplies information about E , the value of E[</>] for the correct field. It is
E = J (d r )^ e E \ ( 1 1 .2 0 )
since the linear term in A<^ vanishes by the stationary principle. The correct
energy, given by the physical <^o, is a maximum of the functional (11.15). Eval
uating the energy functional for an arbitrary potential bounds the energy from
below,
E > E[<j>]. ( 1 1 .2 2 )
V ♦D = 47rp. (11.23)
128 CHAPTER 11. STATIONARY PRINCIPLES FOR ELECTROSTATICS
D2
E[ D] = j (dr) (11.24)
Sttc 5
which is a functional of D . How does the stationary principle work here? The
variation of (11.24) is
6E = /(d r )^ D .5 D , (11.25)
where the displacement vector is varied subject to the restriction that (11.23)
be satisfied:
V •(D + 5D) = 47rp, (11.26)
or
V •<$D = 0. (11.27)
ÍD e V xíA (11.28)
V x E - 0. (11.30)
A second way of getting this result is based on the replacement of the local
restriction (11.27) by the equivalent integral statement
0 = J (d r)l^ < t> (r)V -6 D (r) = -J (d v )l^ 'V < l> .6 D , (11.31)
with the restriction (11.31) yields the condition E = —V<^, which is equivalent
to (11.30).
The advantage to this functional form of the energy, (11.24), is seen by again
considering finite variations (called A D ). If Do is the correct physical solution,
then i£[Do] = E is an absolute minimum,
E [D 0 + A D ] (11.33)
87T6
11.2. FORCE ON DIELECTRICS 129
(since the linear term in A D is zero due to the stationary principle). Therefore,
the correct energy is the minimum value of (11.24) while an arbitrary D [com
patible with (11.23)] will give an upper bound to E . These bounds, (11.21)
and (11.33),
E[4>\ < E < £[D ], (11.34)
are useful for finding approximate solutions when exact solutions are difficult or
impossible to obtain.
= o. (11.35)
E = ^ E 0, (11.36)
where
E0 = J (d r ) |¿0/, — i - ( V ¿ 0) 2 (11.37)
will be recognized as the energy expression for e = 1, the vacuum. The same
energy relation follows immediately from (11.24), inasmuch as D retains its
meaning in order to satisfy (11.23). That the presence of a uniform dielectric
medium reduces the energy by a factor of e is familiar in elementary presenta
tions of electrostatics as the corresponding reduction in the strength of the force
between charges.
Now we turn to a position-dependent dielectric constant, e(r), and examine
the effect of an infinitesimal alteration,
This will change the fields, <j>(r) or D (r), but these changes have no first order
effect on the energy, which is stationary. Accordingly, the only contribution to
the energy change is produced by the explicit appearance of e(r) in the respective
energy expressions. Using (11.15), we get
SeE 2
6E{<t>] = - J ( d v ) 6
- ^ £ = -J (d v ) (11.39)
8 tt
Equivalently, from the second form, (11.24), the first order variation in the
energy is
8eE 2
6E[D] = - J ( d r ) ¿ ( y ) 6e = - j ( d r ) (11.40)
8 ir ’
130 CHAPTER 11. STATIONARY PRINCIPLES FOR ELECTROSTATICS
or
<Se(r) = —Sr - V e(r). (11.42)
Then, the change in energy is
f E2
SE = S r- / (dr) ( V c ) — = - F •Sr. (11.43)
This result, (11.44), also shows the continued relevance of the electric field part
of the stress tensor, (7.15) with H = 0,
£^_£E E
(11.45)
8 7T 4ir ’
for spatially inhomogeneous dielectrics. Since the stress tensor (dyadic) de
scribes the outward flow of momentum across a directed unit area, the total
force on a body bounded by a closed surface S, the net flow of momentum into
that body, is
F
Í dS •T = V •T, (11.46)
where the volume integral extends over the body. The divergence of the stress
tensor is
E2 1 V D e _(D .V )E
V •T = ( V e ) - ---- b — Di'VEi — (11.47)
8 tt 4it 4 ít 47T
DCVEi - ( D • V ) E = D x ( V x E ) = 0, (11.48)
V •D = 4wp = 0, (11.49)
F F
E 2 small E 2 large
Ve ^7
(11.51)
The same result emerges from the energy form (11.15) if we introduce the ana
logue of (11.42) for the charge density,
(11.52)
In using the vector field form (11.24), we must recognize that the displacement
of the charge distribution requires a corresponding displacement of the field D,
in order to preserve (11.23):
/<*)
Then, the identity
and use of both electrostatic field equations then gives the anticipated result
(11.51).
132 CHAPTER 11. STATIONARY PRINCIPLES FOR ELECTROSTATICS
e2 ) Vi ci, Vi
Figure 11.2: Boundary between two regions with different dielectric constants.
ÍE = / y_ ( * ) V . ( ^ ) + / Vj( * ) V . ( § « )
= Is dS (ni * + n 2 * l w ^ 2) ’ (11.56)
where S is the common surface and n i and 112 are the oppositely directed
outward unit normal vectors for V\ and V2 , respectively:
ni = - n 2. (11.57)
that is E (r) remains finite in that transition region. In the static regime now
under consideration, where E (r) = —V<^(r), this implies continuity of the scalar
potential across the contact surface, or
^ 1 = ^ 2, (11.58)
J dS ar<f>. (11.61)
J dScrS<f>. (11.62)
SE = J d 5 ^ [ n 2 . ( D 2 - D i ) + 47rir]^ = 0 ) (11.63)
V . D = 4ttp. (11.65)
(n 2X V)(<^i — f c ) = 0 ( 11 . 66)
134 CHAPTER 11. STATIONARY PRINCIPLES FOR ELECTROSTATICS
or
n2x ( E i - E 2) = 0. (11.67)
Equation (11.67) states that the tangential component of the electric field is
continuous. This result can also be derived directly from the vector field form
by considering the surface terms that follow from (11.29).
11.4 Conductors
The surface of a conducting body provides an important example of these bound
ary conditions. Within a conductor electric current flows in response to the
presence of an electric field. Since in the static situation there is no flow of
charge, E = 0 everywhere inside the conductor; the interior of a conductor is a
region of constant potential </>. The continuity of the tangential component of
E, (11.67), then implies
nxE = 0 (11.68)
on the surface of the conductor. Indeed, if it were otherwise, charge would
flow on the surface, which must also be a region of constant potential, of value
equal to that in the interior. Moreover, since E and D vanish inside, the other
boundary condition, (11.64), implies
n *D = 47rcr (11.69)
just outside the conductor, where n is the outward normal to the surface, and
again cr is the surface charge density.
We use these properties of the field to examine the electrical forces exerted
on the conducting surface, as described by the stress dyadic (11.45)
__ „ eE 2 eEE
T — 1 ----------------- (11.70)
8 tt 4tt ’
where e is the dielectric constant of the medium surrounding the conductor.
The force acting on a unit element of area, with its outwardly directed normal
vector n, is
eE 2 n •D
—n •T E, (11.71)
‘ ■ s r n + 47T
where
eE 2 = e(nxE)2 + i ( n - D ) 2 (11.72)
is a helpful decomposition into tangential and normal field components. We
now recognize that
—n •T = + <rE, (11.73)
which force has no tangential components; there is only a normal component of
force per unit area:
/ v ™ / x 2tt(t2 cr 27r
(—n) •T •( - n ) = --------------n •D = (11.74)
v ' e e
11.5. PROBLEMS FOR CHAPTER 11 135
The negative sign attached to this normal force per unit area tells us that it is
not a pressure but a traction or tension, drawing the surface element into the
region occupied by the electric field.
Here is a simple application of this result. Two identical conducting plates of
cross sectional area A and negligible thickness are placed in parallel proximity,
just short of total contact. Equal and opposite charges are bought up and
placed on the respective conductors, resulting in the final charges Q and —Q.
In view of the almost complete cancellation of the electric fields produced by the
opposite charges, essentially no work is performed during the act of charging up
the plates. Over most of the area of each plate, the surface charge density is
uniform, of magnitude
a — (11.75)
A'
Therefore, the total force on each plate, drawing it toward its partner, has the
strength
2tt
F = (11.76)
e
where, again, e is the dielectric constant of the medium in which the plates are
inserted (only the material between the plates is physically significant).
Now holding one plate fixed, move the other away to a distance a, which
distance is very small on the linear scale provided by A 1/2. The work required,
the energy of the resulting parallel plate configuration, is just F a , or
1 Q2
(11.77)
2 C ’
where
cA
C = (11.78)
4ira
is identified as the capacitance of the electric capacitor. We will discuss capac
itance in general in Sec. 24.6.
2. Consider the boundary conditions between two different media with differ
ent dielectric constants. Prove that the continuity of the potential implies
that the tangential derivative is also continuous, hence that the tangential
component of the electric field is continuous.
3. Derive (11.67) from the vector field form of the energy functional (11.24).
136 CHAPTER 11. STATIONARY PRINCIPLES FOR ELECTROSTATICS
+<r
e> 1 e= 1
F = J (d r ) [ - ( V •P )E ].
1 [J(dr)p<¡>}2
4f(dr)e(V<f>y/8n'
Verify that this is indeed stationary.
7. Recall that for homogeneous e, <¡> = <¡>q/c. Now use E[</)\ and E[T>] to
find bounds for the energy of a charge distribution p(r) in the dielectric
medium e(r) where e is slowly varying:
|Ve(r)| |V<ft(r)| |V¿„(r)|
e(r) <j>(r) ~ <^0(r)
Here <j>o is the solution to Poisson’s equation in the absence of the dielec
tric,
~ v 2d>o = 4irp,
and the bounds should be expressed in terms of integrals involving e,
V<j)o, and Ve.
Chapter 12
Introduction to Green’s
Functions
The differential equation governing the scalar potential of a given charge distri
bution in a dielectric medium, the combination of
V •D = 47T/?,
D = eE, E = -V <¿, (12.1)
is
- V . ( e V < ¿ ) = 47rp. (12.2)
V 2(j) = 0. (12.4)
Our task is to find the solution of (12.2) for <^, for a given charge distribution.
Since (12.2) is a linear differential equation relating p and <^, the potential at a
point r is the additive contribution of all individual charge elements (dr,)/?(r/):
It is evident that G { r, r') is the potential at r arising from a unit point charge
at r', so that it satisfies the differential equation
137
138 CHAPTER 12. INTRODUCTION TO GREEN’S FUNCTIONS
Equation (12.5) expresses the fact that the potential due to a charge distribution
is simply the sum of the contributions of each of the charges. G (r, r') is an
example of a class of functions introduced by George Green (1793-1841). Once
we have Green’s function, the solution for any charge distribution in a given
dielectric arrangement is a matter of integration.
E = J ( d r ) e ^ = \j(dr)p<f>. (12.9)
[See also (11.24).] If we use the second of these expressions together with (12.5),
E = \ J (d r ) p(r)(j>(r)
(where the last form has used r —» r'), we see that only the symmetrical part
of G (r ,r /) contributes to the energy. This proves (12.7) because the energy
completely determines Green’s function. In other words, the reciprocity relation
states that Green’s function is the interaction energy of a pair of unit charges at
specified points, and is a property of that pair of points, not of their individual
labels. [Note that (12.8) and (12.10) imply the construction (12.5).]
12 .2 . PROBLEMS FOR CHAPTER 12 139
in which we have also applied the following identity for three arbitrary scalar
functions,
Let us stop for a moment and appreciate that this same volume integration,
applied to the differential equation ( 1 2 .6 ), gives
—V 20 = 47rp
- V 2G(r,r,) = 47r5(r-r')>
<j> is expressed in terms of the charge density and the boundary values by
where S is the closed surface bounding the volume V , and dS' points in
the direction of the inward normal. If <t> assumes specified values on 5,
and G vanishes there, then inside V
We will now derive (13.3) from the differential equation (13.2). In order
to do this, we require an integral representation for the delta function, the
charge density of a unit point charge. The relevant properties are that it vanish
everywhere except at a single point, while possessing a unit integrated value. In
one spatial dimension, these properties appear as
/
oo
dz6(x-z') = 1. (13.5)
-CO
141
142 CHAPTER 13. ELECTROSTATICS IN FREE SPACE
while the value of the integral of the function over the whole range of x is unity,
independent of the choice of e:
/
oo
i
-oo
d(a *')z
ir (x — a?')2 + e2 = 1,
•\L dt■_
OO t 2 + 1
(13.8)
or
6<(x - x 1) = r 7T (13.10)
J —oo
Can we now set e equal to zero? No; the resulting integral would not exist—
it would not converge at infinity. But we can think of giving e an arbitrarily
small positive value, in order to make the integral mathematically meaningful,
without significantly altering thereby the physical properties being represented.
After all, when we speak of a point charge, we mean no more than one of size
that is very small on the scale set by all the other significant lengths in the given
physical situation. In the following we shall keep in mind the necessary presence
of the convergence factor, exp(—e|Ar|), but not write it explicitly. Accordingly,
the desired expression for the delta function, in one dimension, is
the possible outcomes for the right-hand side are just zero and unity, the latter
being realized only if the range of ^-integration includes x', that of ¿ /-in te g r a tio n
includes t/, and the range of ^-integration includes z'— which is just to say that
the point r' is within the volume V. The use of the integral representation
(13.11) for the three individual delta functions, with the integration variables
labeled kX) ky) and kz, respectively, then produces the three-dimensional repre
sentation
/ i (d
* k ) ¿fc . ( r - r ;)
6(r —r')
J (2 * :
)3 (13.14)
13.1. 2 + 1 DIMENSIONS 143
We now employ the above representation of the delta function to solve (13.2)
for Green’s function. This can be very simply accomplished if we note that
V e ,k .( r - r ') = i k e * k .( r - r ') ) ( 1 3 .15)
or, effectively,
V — *k, (13.16)
so that we can read off a solution to (13.2),
/
fdk'l e*k * (r~r‘)
(m7)
We can verify that this is in fact the known result, (13.3), by using spherical
coordinates for k, with r — r; pointing along the 2 -axis,
so that
k2dk 27rd(cos 0) e ikRcos0
G (r ,r ') = 47t J
8 tt3 k2
-7TJof
1
/
,
ikR
1
dk----- AkR — e —ikR^
2 1 f°° sin#
7TR Jo x dx = R'
(13.19)
The convergence of the final integral here shows that only values of k of
order 1 /R are significant, which also applies to the individual components of k.
Accordingly, the implicit convergence factors will indeed effectively equal unity
if the individual value of e are restricted by e <C R- This is entirely consistent
with the physical context in which a point charge has a linear extent small in
comparison to the distance R.
13.1 2 + 1 Dimensions
In the above, we have treated all three directions of space on an equal footing.
However, we need not do this. We can separate out one direction, say that of 2 ,
and treat it differently. The reason this is useful is because there are geometries
in which physically interesting quantities vary in only a single direction. Singling
out the 2 direction, we can write Green’s function, (13.17), as
fOO dkz
(13.22)
J-oo 2* k± kz 2k± +
g{z,z'; (13.24)
Of course, we could also have derived this equation directly from the left-hand
side of (13.22), which is only to say that the latter expression is an obvious so
lution of the one-dimensional differential equation. Can one also arrive directly
13.1. 2 + 1 DIMENSIONS 145
at the form o f the solution given in (13.24) by solving the differential equation
(13.26)? Yes.
This differential equation is solved by noting that for z ^ z ',
2
kX f - g (z ,z ']k x ) = 0 , (13.27)
dz2
while the derivative of g at z = zf is discontinuous. Indeed, if we merely accept
that g is bounded when 2 is in the infinitesimal neighborhood of integration
of (13.26) over the interval between z 1 — 0 and z' -f 0 yields
d z'+0
-^ -g (z ,z ';k ±) =1. (13.28)
OZ z 1- 0
+ 0 = Z (13.30)
z'-O z'-O
o'
KÍ
H
H
1
II
1
2. Show that
6e(x) - - x - ^ 6e(x)
is also a model of the delta function. Can you support otherwise the
inference that
- x - ^ - 6(x) = S(x)?
ax
3. Express
6(x 2 — a2), a> 0
in terms of 6(x ± a). What can you say about S (f(x )), where / is a real
function with a simple zero at x = a?
4. Perform the integral in (13.8) and the last integral in (13.19) by contour
integration.
Chapter 14
Semi-Infinite Dielectric
147
148 CHAPTER 14. SEMI-INFINITE DIELECTRIC
(14.4)
(14.8)
Z<0: C(“ ^2 ^ _
subject to the boundary conditions
= 9\m
z =.
= + O’
(14.9)
(14.10)
€Tz9 Z——Q = d~z9 z—-\-0
In the following we will first solve this problem by assuming that z' > 0
(that is, the unit charge lies in the vacuum and not in the dielectric). [For
the converse situation see Problem 14.1 and (14.55).] The solutions to (14.7),
(14.8), (14.9), and (14.10) can be expressed in terms of the solutions, ek±z and
e~k±z, of the corresponding homogeneous equation. The forms of the solution
in the three regions are as follows:
where the single exponentials in (14.11) and (14.12) are required by the bound
ary condition that g remain finite for \z\ —* oo. The boundary conditions at
z = 0, (14.9) and (14.10), require
a = c + d, (14.14)
ek± A = kx ( C - D ) , (14.15)
1 4 .1 . G R E E N ’S F U N C T I O N F O R C H A R G E O U T S I D E D I E L E C T R I C 149
Just as in the situation mentioned at the end of the last chapter [see (13.28)], the
singularity in the differential equation requires, at z — z*, that g be continuous,
while —dg/dz be discontinuous,
*=*'+°
z= z'-0 U> (14.17)
z—z1-j-0
= i, (14.18)
z —z' —O
C = — e —k ± z f (14.21)
2k±
1 -fcjy e — 1 1 0-k±z'
A = D = (14.22)
€ -J“ 1 2k± e + 1 2kj_
B = —Í _.________ (14.23)
e + l 2k x. 2 JbjL
Inserting these coefficients back into (14.11), (14.12), and (14.13), we find for
the reduced Green’s function, g, in the two media,
1 -k ,\ z -z e- 1
z >0 : g = >-k ± (z+ z‘ )
(14.24)
2 k± L’ e+l'
n:
z < 0 q — -------------
1 2 e - k ± ( z ' —z)
y 2k± e + l
-V <fo) ~ 1
P efí — Pbound — — V • P (14.28)
•( Air
or
<fo) ~ 1 e (r )~ l
P efi — —V • D •V { « * ) - n (14.29)
47re(r) e(r) P -
V <r) )'
which introduces the average free charge density p. The problem of finding the
potential that is produced in a dielectric medium by the free charge density is
therefore equivalent to determining the potential that is produced in the vacuum
by the total charge density
If e were completely uniform, the total charge density, and the potential it
produces, would be reduced below the vacuum values by the factor e, as already
given in (11.35).
In the situation before us, e(r) is discontinuous across the surface 2 = 0, so
that
dz7{r) = ( 1 _ l ) S^ ’ (14.31)
as one verifies by integrating over an interval that includes z = 0. Note that the
multiplier o f this derivative in (14.30), e(r)Ez(r), is continuous at z = 0 and can
be evaluated as Ez at 2 = -1-0, the vacuum side of the boundary. Accordingly,
the bound charge density has a surface charge component,
Abound = % M r x ) , (14.32)
Figure 14.2: The primary field, P, and secondary field, S', contributing to
Green’s function (14.26).
152 CHAPTER 14. SEMI-INFINITE DIELECTRIC
and therefore the multiplier of p in (14.30) is l/e (r '), which is unity for the
present arrangement (z 1 > 0). Accordingly, G for z* > 0 is the potential
produced in the vacuum by the combination of the point and surface charge
distributions:
g (r ix )
G(r,r') (14.37)
+ J s dSl tr -r ix l’
where the element of area on the boundary surface S, which is the interface at
z\ = 0 , is
dSi = (dr1±). (14.38)
This construction of G in terms of itself (through a) is an integral equa
tion. That we are able to solve it with ease is directly attributable to the
two-dimensional translational invariance embodied in the representation (14.6),
along with the one for l/|r — r7|, (13.35), which we now write as
Let us multiply the terms of (14.37) by exp(—¿kx t j l ) and integrate over rx-
We encounter the delta function
_d_
J (dr ix )e _ ik i *I‘1-Lcr(r ix ) = e_,kj- *r^- g(zi,z'-,kx )
dz\ z i= + 0
(14.42)
14.3. GREEN’S FUNCTION FOR CHARGE WITHIN DIELECTRIC 153
Thus, the one-dimensional reduction of the integral equation (14.37) is the al
gebraic relation
1 _ fc, zi 1 e —I f d
— e k±z + - , (14.44)
z' ,k ± ^z = + o 2 c z= + 0
to learn that
~ ^ g (z ,z ';k ±) ____ (14.45)
z= + 0 €+1
(14.50)
The above equation describes the ^-dependence of the potential produced in the
vacuum by the combination of two localized charge distributions: one at z = z',
the other at z = 0. Expressed in terms of the vacuum function go, which obeys
—^ 9 ( z ,z '- ,k x ) , (14.53)
z=+0
and therefore
(14.54)
The outcome is
zf > 0 and replace e by 1/e. That converts the region containing the point
charge into the one with the larger dielectric constant, in the ratio e : 1. On
restoring the value unity to the smaller dielectric constant— raising the overall
scale by the factor e— Green’s function becomes divided by e. And finally, to
have the region of higher dielectric constant in the half-space z < 0 , one replaces
all ^-coordinates by their negatives. Doing this in (14.26) yields (14.55).
F = - * '. (14.58)
Because (14.56) and (14.57) involve only the vacuum form of the one-dimensional
function, we can interpret (14.56) and (14.57) as follows. For 2 > 0, the Green’s
function appears to describe the potential due to two point charges, one, of
strength unity, at r' = (#', y\ ¿'), and another, the image charge, of strength
— at the image point r' = (x'^y^z1). For z < 0, only one point charge
appears, of strength located at r'. In either medium, the total effective
charge is the same. With this interpretation, the full Green’s function may be
written down immediately, {z! > 0 )
1 _ € —1 1
z > 0: G( r ,r ') = (14.59)
|r —r'| e+ 1 |r - F| ’
2 1
z < 0: G(r,r') = (14.60)
e + 1 |r —r'| ’
The related forms for z' < 0 , as produced, for example, by the substitution
£ —*•1 /e followed by the procedure in the preceding paragraph, are
1 e - 1 1
z < 0 : G(r,r')=i 1 ■+ -• (14.61)
e |r —r'| e e -f 1 |r — r'|
2 1
z > 0: G(r,r') = (14.62)
e + 1 I r - r 'l ’
One sees directly, particularly on comparing the z < 0, z* > 0 and 2 > 0, z* < 0
forms, that the symmetry of Green’s function is realized in these results.
Applications of these Green’s functions will be treated in the next chapter.
156 CHAPTER 14. SEMI-INFINITE DIELECTRIC
3. Now suppose a dielectric slab occupies the region 0 < z < a. Assume the
unit point charge is in the vacuum region z 1 > a. Compute the reduced
Green’s function in this case. Can you derive from this result additional
information about the arrangement of Problem 14.2?
Chapter 15
Application of Green’s
Function
Co(r-r') = ¡ ^ , (15-2)
157
158 CHAPTER 15. APPLICATION OF GREEN’S FUNCTION
fields, in agreement with the earlier discussion o f Section 11.2. The magnitude
of this force is
1 e-1
9Eint _ e - 1 e2
\p\
1 t + i e \e \
(15.6)
d ( - z 0) ~ € + 1 4zfi ( 2z 0)2
which can be interpreted as the force between the charge and the image charge.
The field point of view, as opposed to that of action-at-a-distance, provides
an alternate derivation of this result. To calculate the force on the dielectric,
we calculate the normal component of the flow of momentum into the dielectric.
In terms of the stress tensor, this force is
F —— (15.7)
2 (r - r0)x
z = +0 : Ex — e
C + l(p* + * 2 ) 3 / 2 ’
2e z0
Em —e (15.11)
6 + l ( p 2 +*2)3/2»
e2 / o \2 2
(ih )c
€ Zi
27tp dp
87r (,p2 + 4 ) 3
■- 1 - ( t - 1 ) _ c - i
(15.12)
(* + 1 ) 3 e + 1 4z% ’
V •T + p E = 0 . (15.13)
15.1. FORCE BETWEEN CHARGE AND DIELECTRIC 159
Now integrate the ^-component of this equation over the entire volume of the
semi-infinite region z > 0 to get (infinitely remote surfaces do not contribute):
Then, integration of the ¿-component of (15.13) over the region z < —0 relates
the last integral above to the force on the free charge within the dielectric.
There isn’t any! That effectively reduces the calculation to the one already
carried out in (15.12) or (15.14). Nevertheless, let us evaluate F by means of
the first part of (15.16), involving the discontinuity of Tzz across the interface
between dielectric and vacuum:
Tzz(z = - 0 ) = ± [ e E l - (15.18)
and then
Incidentally, the same expression for the force emerges from (11.44), with its
¿-component presented as
160 CHAPTER 15. APPLICATION OF GREEN’S FUNCTION
which involves the continuous transverse electric field and normal component
of D . With these field components given their z = +0 values, the resulting
^-integration over the discontinuity region yields
F = J dsL(e - l ) [ E l + -l E % (15.21)
F dp p 2tt €Zo + P
( P 2 + Zo
2) 3
€- l e2 f ° ° € + 1 + (t — 1 )
dt (15.22)
(e + 1 ) 2 2z l J 0 (t + 1 ) 3 ’
which is indeed equal to (15.12), in consequence of the null value of the integral
appearing in both (15.12) and (15.22),
i J í( í T ¡ j 3 = 0i <15'23>
as can be seen from the substitution t —* l/t. This property just expresses the
zero value possessed by the surface integral of Tzz(z = —0).
The version given in (15.17) directs attention to the physical interpretation
of F as the force on the bound charge that is localized on the interface between
the two regions. As such, the force should also equal
But what is the value of Ez(z = 0)? One way to answer that question has
us returning to (14.31), which describes how a surface distribution of charge
emerges in the limit where e(z) is discontinuous. If we regress to the continuous
description, through the replacement [the other factors in <r, (14.33), and the
reference to a particular point on the surface, are understood]
1 1
Dz (z = 0) = Dz{z = 0). (15.26)
< *) 2
That identifies Ez(z = 0) with
z0
E ,(z = 0) = (15.28)
V + * o 2)3/2’
which is just the field of the point charge. But we could have anticipated that;
the self-force of the surface charge distribution is zero. The above field is now
combined with (14.33), (15.11),
e c—1 zp
(15.29)
2 tt e + 1 ( p 2 + Z q ) 3/ 2
to produce
f e2 e - 1 f°° z2
zo 6 -1
F = j d S a E 2(z = 0) = — — J^ d pp 2*
(p2 + *o ) 3 6 + 1 ( 2 Z q) 2
(15.30)
Six derivations of the same quantity? Time to move on. Yes, but there is one
little question about the surface charge distribution that has not been answered.
What is the total amount of that charge? Obviously, one can compute it by
integration— and so we shall, for a particular example. But a more immediate
answer comes from Green’s function (z' > 0 ) as displayed in (14.59), when this is
regarded as the potential produced in the vacuum by the unit point charge and
the surface charge. At very large distances, on the scale set by z1, so that the
finite dimensions of the surface charge distribution are negligible, the behavior
of G in both half-spaces is simply
G( r ,r ') .
0 - 7TÍ)? (15.31)
z > 0,
G (r ,r ') = | ^ Q \*~*V (15.32)
2 < 0,
vacuum
+1
vacuum, and G vanishes on the surface, all of its points being at the same
distance from the equal and opposite charges.
For such a unit point charge, we now calculate the free surface charge density,
cr, induced on the conductor. We know, from (11.69), that
G(r,r') =
r -r r |r —r;
i
(dkx ) ,*k-L •( r - r 1
( 2 ’r) 2 '
(15.34)
This is the e —» oo limit of (14.26). The outcome for the form of a computed
from the first form can be read oif from (15.29) with e —» oo, e = 1, and Zq = zf:
1 2z*
(15.35)
4w (p2 + z /2)3/2 *
The second form gives
f (<lkx)
J (2*)2
ikj. . f r - r 'l | r - k j . z ' (15.36)
In using the second form, (15.36), one encounters the two-dimensional delta
function [dS = (drj.)], as already noted in (14.40),
/ *(r-r'U _ ( 15.38)
J y
That gives
Qind = ~ J (dk± )5(kx ) e - k^ ' = - 1 . (15.39)
In another exercise of this kind we give two derivations of the force pulling the
conductor toward the unit point charge, using the force per unit area presented
in ( 1 1 .7 4 ), but with e = 1 and omitting the minus sign:
F =
J dS 27TCT2. (15.40)
We need not trouble with the calculation that employs (15.35) for cr; it is just
what is exhibited in (15.30), with e —*•oo, e = 1, and z 0 replaced by z':
1
F = (15.41)
( 2z' ) 2
in which it is convenient to write —kj_ as the integration variable for the second
cr. Performing the rj_ integration with the aid of (15.38) we get the same result,
J? f (tflcx) (tflc'x) / o _ \ 2 _
C /1 . k/ V
F ~ 2 J (2ir)2 (2tt)2 ^ 5^kj-
- / =j f = (¿F- <15-43>
2. Prove the identity (15.23) and show that it expresses the fact that the last
integral in (15.16) vanishes.
Chapter 16
Bessel Functions
2z f k dk d<j> ikpcQB¿ kz
cP2 + z 2)3/2 J ( 2 tt) 2
(16.3)
The (j) integral in (16.3) is defined as the Bessel function of zeroth order, Jo,
it is the first member of the class of functions that are named for Friedrich
Wilhelm Bessel (1784-1846), although the infinite series that represents these
functions had appeared more than half a century earlier, in a work by Leonard
Euler (1707-1783). Now expressed in terms of this function, the equivalence
(16.3) appears as
z r°°
(p2 T z 2)3'/~2 = Jo k d k M k p ) e ~ k\ Z > 0. (16.5)
165
166 CHAPTER 16. BESSEL FUNCTIONS
Since this result was obtained by equating 2 derivatives [c/. (15.33)], it may be
immediately integrated to yield
1 r°°
■. / dk J0( k p )e -kM. (16.6)
Jo V ' '
(The integration constant vanishes since both sides go to zero as \z\ —> 00 .)
Note that (16.6) may be directly derived from the equality of the two forms of
the Coulomb potential in (13.35), which with r' = 0 reads
Here in (16.6) we have recognized two different forms for the potential of a
unit point charge at the origin, which satisfies Laplace’s equation,
V 2(j> = 0 , (16.8)
except at the origin. Actually, the integrand of the right hand side of (16.6) is
also a solution of Laplace’s equation, that is, for each positive value of k,
V 2 [J o (M e ~ * W] = 0, (16.9)
is as indicated by
167
1 3 d
(16.15)
duk
fc=i
\d_ f d_\ £_
(16.16)
p dp v dp) p2 d<j>2 dz 2 ’
'1 d_
+ k 2 J0(kp) = 0, (16.17)
.P dp
1 d
+ 1 Jo(t) = Jo(t) = 0 . (16.18)
t dt dt2 t dt~^
thereby again proving (16.9), and establishing the equation satisfied by the
Bessel function of order zero, (16.18).
To exploit the more general solution of Laplace’s equation exhibited in
(16.10), we take
introduce the symbol u = ie1^} and then encounter the generating function for
the Bessel functions of integer order,
oo
(16.22)
m = —oo
168 CHAPTER 16. BESSEL FUNCTIONS
which serves as a definition of all the Bessel functions of integer order m. Since
the generating function is invariant under the substitution
1
u (16.23)
u’
that is,
we learn that Bessel functions of positive and negative integer orders are related
by
= (-1 (16.25)
Now let us regard the generating function as the product of two exponentials,
with arguments proportional to u and —1/w, respectively. The insertion of the
appropriate infinite power series for these exponentials,
E
( i t ) m+n um+n
(m + n)! ■E n\
(16.26)
( i * ) m+2n
Jm(t) = E C " 1)” (m + n)! n! ’ (16.27)
n—0
J0
[
21T
<l> im<}> _ ^
-« m m ',
(16.29)
which contains the result for Jo, (16.4). Incidentally, changing the sign of m,
and replacing <f> by 27t — <j>, leaves the integral in (16.30) intact, or
= imJm(t), (16.31)
169
eim+Jm(kp)e~kW. (16.32)
The Laplacian, (16.16), acting on this solution yields the differential equation
satisfied by Jm,
'\d_ ( d_
J<m{kp) — 0 , (16.33)
p dp \ dp
or
d2 Id m2
Jm{t) — 0 . (16.34)
H? + t l t ~ 7 r +
One can, of course, verify that Jm(t)y as represented by the integral (16.30),
obeys this differential equation. See Problems 16.4 and 16.5.
Another way to present the differential equation uses the differential operator
relation (an operand is understood)
t- i l 2 ± tH 2 - 1 + 1 (16.35)
dt dt 2t ’
(16, 6)
~ + l) = 0. (16.37)
This version has the advantage of suggesting more clearly the behavior of the
Bessel functions at extreme values of t. For sufficiently small values of t y such
that unity is negligible in comparison with (m 2 — 1/4) f t 2y the differential equa
tion indicates that ~ ¿lml+ i/2? in accordance with the leading term of
(16.27),
( k ) m
m> 0 : Jm{t) = + (16.38)
(^ ) cos + 1/ 2 ) 0 . (16.39)
170 CHAPTER 16. BESSEL FUNCTIONS
by the construction
/ d2 o\
( - 5 ? + * ) - • * - = '<*>■ <1M6)
we arrive at
1 1 r°°
- V 2- = 4t dkkJo(kp). (16.47)
r 27T J0
What can we say about the latter integral? A glance at (16.5) shows that
TO if p ^ 0
-2ttÍJ0 d k k J o (k p )= lim
+o 27T (p2 -{- z2)3/2 \ oo if p = 0
(16.48)
16.1. DELTA FUNCTIONS AND COMPLETENESS 171
To these limiting results we append the value of the polar coordinate surface
integral [this is (15.37) again]
f°° 1 2 i r°° i
I dpp2*2^{p*+z*y/> = 2 Jo dt(t + 1)3/2 = 1 ’ (16-49)
independent of the particular choice oí z > 0 . Here, then, is the realization, in
polar coordinates, of S(x)S(y)) as required to complete the structure of <S(r) in
produces
f°° dk ikx
™mL y ° ° d x 'e -ikx' f ( x ') . (16.53)
This exhibits the completeness of the set of functions exp (ikx) in which k ranges
continuously from —oo to oo; any function of x can be constructed as a linear
combination of that set, with the coefficients as displayed in (16.53). Such
integral representations of functions were introduced by Jean Baptiste Joseph
Fourier (1768-1830), although the particular complex form appearing in (16.53)
can be attributed to the independent investigations of Augustin Louis Cauchy
(1789-1857). The analogous two- and three-dimensional completeness state
ments are similar consequences of the appropriate delta function constructions.
We now find the polar coordinate version of this completeness in two dimensions.
The starting point is the two-dimensional analog of (13.14) and (16.52),
where we have introduced polar coordinates r¿(/>, (f>) and r ¿ (¿ /, </>'). Corre
spondingly, if we use polar coordinates for kj.(Ar,a), (16.54) becomes
172 CHAPTER 16. BESSEL FUNCTIONS
We next expand the exponentials here by use of the generating function ex
pression, (16.28) [with <¡) —> <j) — a] together with its complex conjugate [with
<j) —►ft — a] and perform the a integration by means of (16.29):
27r da
Y ^ i mé m^ j m{kp)
1 27T
.m . m'
771,771'
OO
= Y eim^Jm{kp)e~im*' Jm{kp'). (16.56)
/ "o¿ir
—E
/»OO I JI oo -|
eim,t,Jm{kp)e-im*' Jm(kp') = U ( p - pf)6{4>- # ) . (16.57)
Jo P
Again this is a completeness statement, this time for the functions eim<^Jm(&p),
where k ranges continuously from 0 to oo, and m assumes all integer values:
0, ±1, ± 2 , --- That is made explicit on multiplying (16.57) by an arbitrary
function / ( / / , <^'), and by the area element pf dp' d<¡>\ followed by integration:
fM )= l
r°° b(jb
Jo 27r
Y
JE
eim*J™(kp) /
Jo
r°°
p'dp'
Jo
r 2lx .
Jm{k p ')f(p ',<!>')■
(16.58)
We may now easily isolate the individual p and <t> dependencies of (16.57).
If we multiply (16.57) by ) and integrate over <^, we select a particular
value of m, according to (16.29), so that we obtain
f°°
/ kdk Jm{kp)Jm{kpt) = 1-6(p - p'). (16.59)
Jo P
This states that Jm(kp) is a complete set of functions of p for any integer m:
f(p) = /
Jo
k dk Jmi^kp)
í
,Jo
p'dp' Jm{kp')f(p') (16.60)
Putting this information back into (16.57), we determine the completeness re
lation for the functions etmi
oo
Y — = 8{<j> - (16.61)
and correspondingly
16.1. DELTA FUNCTIONS AND COMPLETENESS 173
which is the Fourier series expansion for /(</>), that is, the statement of com
pleteness, over a 27T interval, of the set of functions exp(¿m<^), m integral.
Perhaps we should remark here that the validity of (16.61) depends on the
restricted 2ir range o f the variables (j) and <^'. But the restriction can be lifted
if the right-hand side of (16.61) is made periodic in its variable to match that
property o f the left-hand side. With this unlimited extension of <j) — ft now
called x y that is displayed by
1 oo oo
_ £ e<" » = X 8 {x -2 * v ), (16.63)
and, indeed, when x is restricted to the interval between —2ir and 27r, only v = 0
can contribute to the right-hand side. It’s interesting to check the correctness of
(16.63) by introducing the integral representation for the delta function, (16.52).
That produces, for the right-hand side of (16.63),
oo
/
dk ikx —i2iruk
e > (16.64)
-oo
which can equal the summation on the left-hand side of (16.63) only if
X e~i27tuk = 5^ S (k -m ). (16.65)
r2ir j 00
/ ^ e ik x ' (r_r'U = X Jm{kp'). (16.66)
On the other hand, we could also have specified on as the angle between the two
vectors kj_ and (r — r /)±. That would identify the left-hand side of (16.66) as
the Bessel function of zeroth order, (16.4),
Now, if we return to the completeness relation (16.57), and recall the integral
(16.50), the implication of (16.69) is that
U ( p - p l)8(<t>-t') = ± 6 ( P ) ± - , (16.70)
which is a quantitative version of the remark that coincident points have zero
spatial separation.
v'v=í ? ¿ ( ¿ 4
3. Derive the differential equation (16.34) satisfied by Jm, directly from the
generating function (16.22).
2 ~ (16.71)
TTl
2 = Jm -l(t) + Jm+1(t). (16.72)
d m —1 d m . _ ..
dt t ~Jt + T ) Jm(<)
= Jm(t)
6. Using the integral representation of Jm, (16.30), for m > 0, develop the
power series for Jm, (16.27).
16.2. PROBLEMS FOR CHAPTER 16 175
u" -f - u f + ( l - u = 0,
# \ x¿)
possesses two solutions. One of the them is Bessel’s function of the first
kind, Ju(x), which is regular at x = 0. Show that a series solution of
Bessel’s equation yields, up to a multiplicative constant
2m
( - 1 )”
*<■>-( D ' m
E= 0a i rv ! + ^ + l ) (I)
where T(^) is the gamma function, which generalizes the factorial,
where
Wi{x) ~ a o ^ i - i r c a n ^ + t o ^ C - i r c a n + i * " 2^ 1
n —0 n —0
oo oo
w2(x) ~ 60 ^ ( - l ) nC2„ * - 2n - a o X ; ( - l ) nc 2n+ia:- 2n- 1)
n=0 n=0
176 CHAPTER 16. BESSEL FUNCTIONS
Note that the series terminates when v = n + 1/2 and then these asymp
totic relations become exact. Jv(x ) is represented by the above with
a0 = ( 2 / 7r)1/ 2, 60 = 0 , and Nu by the above with a 0 = 0, 60 = ( 2 / 7r)1/ 2.
G= 0 at 2 = 0 , a. (17.2)
Since the geometry depends only on the 2 coordinate, Green’s function can be
written in the (2 + l)-dimensional form (13.23),
<¡>= 0
z = 0 z —a
177
178 CHAPTER 17. PARALLEL CONDUCTING PLATES
g (z ,z f]k±) = Q at 2 = 0, a. (17.5)
z = z '+ 0
g continuous, -S -g 1, (17.8)
oz z —z ' —Q
kC sinh ka = 1 . (17.12)
and k(a — z>). and, of course, ka, are considered to become very large. Then
every sinh» appearing in (17.13) is dominated by |e®, and
the vacuum form displayed in (13.24). Alternatively, we can withdraw just the
right-hand wall, for example, as accomplished by having ka become very large
without reference to z and z'. The outcome
Q in d (* = 0) = - ± 4 i r j ( d r ± ) j r'U ¿ * ( * . *)
z= 0
(17.18)
(17.19)
We have seen such integrals previously [see (15.39) and (15.38)]. The spatial
integrations over rx yields (27r)2<5(kx), while the subsequent k± integration sets
fc = 0 so that ^
g(z,z'; 0) = -z < (a - z > ) . (17.20)
with the total induced charge on both plates being —1, of course. The total
induced charge is divided between the two plates in inverse proportion to their
distances from the point charge:
17.3 Energy
The interaction energy between the conducting plates and a point charge e, at
the location ro, is given by [cf. (15.5)]
x — e —ka , %z 0
a = 11 -------- . (17.26)
2 f l , X a + x ~ a - 2x
a x --------------------- . (17.27)
l-x 2
1 e2 f 1 1 e2 ,
z0 — 2 ° ' ----- / dx--------- = -------In 2 . (17.28)
a Jo 1+ x a
To proceed with the general situation, we introduce the power series expan
sion for the denominator in (17.27), and get
Ei
■ 2 x ) J 2 x 2n
n=0
- I4 Vaf_ l_ 1 2
(17.29)
n
¿— i I J - 1±°L
+ n + lT L n+ 1J'
But before presenting this summation in terms o f a known transcendental func
tion, we use it as it stands to isolate the leading terms for the two situations,
1 —a <C 1 (^o < «) and l + <* <C 1 (a —zq •< a), which position the point charge
17.3. ENERGY 181
close to one of the conducting plates. To focus on the circumstance zo/a <C 1,
we first rewrite the summation (without the factor 1/4) as
1
n + (z0/a) n + 1 — (zo/a) n + l_
1 1
1 V + (17.30)
- izo/a) + 2 ^ n + (zo/a) n — (zo/a) nj ’
n= 1
(17.31)
O T + J( ? ) n—1
The summation encountered here is an example of the zeta function introduced
by Georg Friedrich Bernhard Riemann (1826-1866),
00 1
( ( ') = £ ? ! <17'32>
n= l
«,««: (17.34)
where the first term is recognized as that associated with the nearby conductor
[(15.5) with e = 00], thereby identifying the next term as the leading contri
bution of the second, distant conductor (more about this later). The analo
gous result for the other situation, a — zo <C a, requires only the substitution
z0 —> a - zq .
The gamma function, first studied by Euler, is conveniently defined in the
manner of Karl Wilhelm Theodor Weierstrass (1815-1897), by stating its recip
rocal as an infinite product:
1
(17.35)
r(<) = n
N 1 1
7 = lim ----- In iV = 0.5772. (17.36)
N -+ 00
Ln = l
1
iKO = ^ lnFW = - 7 - \ (17.37)
n+ 1 nJ
182 CHAPTER 17. PARALLEL CONDUCTING PLATES
(17.38)
t) =1-
we can present (17.37) as
(17.39)
W n=0n \xn + t n + l 'j
e>2 r
-^int — ^ ( — 77— ) + 2t (17.40)
4a
The comparison of this form with the particular example of (17.28), correspond
ing to a = 0 , supplies the value
T
^ \ 2 J = “ 7 — 2 In 2; (17.41)
this can be checked directly through the following evaluation of the summation:
i
lim x 2n ( ------------------- r ) = lim ( - In % ^ In - = 2 In 2 ,
V
\ n -f- 1 / 2 nn--f f 1i y/ x -+ i\\xx ^ 1l —
—xx xx2 1 — x 2J
n=0
(17.42)
which uses the series expansion of the logarithm,
oo n
in (i+ » ) = - ^ ( - i r ^ - . (17.43)
n=l
17.4 Force
The force exerted on the point charge is given by the negative derivative of the
interaction energy with respect to the coordinate Expressed in terms of the
Z q .
that force is
+ a —a
)]
1 1
■4>'
(17.45)
17.5. IMAGES 183
As one would expect, this force vanishes when the charge is equidistant between
the plates,Z =
q One also expects that a displacement from the equilibrium
^ a .
point should produce a larger force of attraction toward the nearer conducting
plate, resulting in a net force that acts to increase that displacement— the equi
librium at the midpoint is unstable. That is made explicit on expanding (17.45)
for small values of a,
1 00 1
- ^ " ( V 2 ) = £¿ -? ,i(T„ +T1 7/ 2™ (17.47)
)3 = 7« 3>'
although the qualitatively important point here is that the latter is a positive
number; the force is proportional to a small displacement from the equilibrium
point, and acts in the same sense.
The explicit expression for the force,
e2
OO o
17.5 Images
Let us look back at Green’s function in the physical region to the right of a
single conducting plate, as it is expressed by (15.34),
which makes explicit the unit charge at the point z' > 0. Now, we get a bounded
extrapolation of (17.49) into a fictitious vacuum region 2 < 0 by writing g as
the image charge appears here on the same footing as the physical charge.
The effect of the coordinate reflection 2 —» —2 is to reverse the sign of the
right-hand side of (17.52). Then the function </, being determined by the right-
hand side of its differential equation, will also reverse sign. Accordingly, the
continuous function g necessarily vanishes at z = 0. Here is the recognition
that symmetry considerations in extended regions can occasionally be applied
to satisfy boundary conditions. With this lesson in mind we turn back to the
pair of conducting plates.
In this circumstance the function g must vanish a,t z = a, as well as at
z = 0. Let us begin at the latter point and follow g to its zero at z — a. Now we
extrapolate beyond z = a, where the continuous function g will start out with
reversed sign. Suppose that the course of the function from z = a to 2 = 2a
duplicates that from z = a to z — 0, but with the opposite sign. Then the
function g will vanish at 2 = 2a. If we continue in the same way, with another
sign reversal at this zero, the behavior of g from z = 2 a to z = 3a duplicates
exactly that from z = 0 to z = a. In short, we are led to extrapolate the
function over the entire range of z %—oo to oo, by imposing the requirement of
periodicity, with period 2a. That periodicity of g must be matched by a like
periodicity of the point charges, so that the initial unit positive charge at zf is
supplemented by unit positive charges at z* ± 2a, z* ± 4 a ,___Doing the same
thing with the unit negative image charge at —z1 then replaces (17.50) with
Fine, but does it work? Will this charge structure reverse sign under reflec
tion, both at z = 0: ¿ —* —z , and at z = a: z —a^> —(z —a)? Without question
for 2 = 0: The replacement v —» —v leaves the charge structure intact. And
then we recognize that ¿ —* —2 + 2 a differs only in the additional displacement of
2a, which is just the periodicity interval of the charges. The solution of (17.53)
will indeed obey the boundary conditions for the pair of conducting plates. See
Fig. 17.2.
But before we construct that solution, let us look back at the force and
energy in light of the now known pattern of image charges. For the circum
stance of charge e stationed at , the image charges comprise charges e at
z q
v — 0, ±1, ± 2 , __ The image charges of the same sign as the physical charge
are disposed symmetrically about the latter, resulting in no net force. The dis
tances between the physical charge and the image charges of opposite sign that
stand on its right, and on its left, are 2(a — Zo), 2(a — Zq) -f 2a, . . . , and 2zo,
2zq -f- 2a, . . . , respectively. Inasmuch as forces of attraction from the right and
left are counted as positive and negative, respectively, the total force is precisely
as given in (17.48). Turning to the interaction energy given in (17.29), we first
rewrite it as
£int = JE ^ ena
1 00
n= l L
2
2 e2
2na — 2zo
e2
2 (n — l)a + 2z$m
(17.54)
Then we can recognize in the successive terms of the summation the electrostatic
energy of interaction between the physical charge and the image charges; first,
those of like sign, and then the oppositely signed charges, positioned to the
right and to the left, respectively. Very good, but why the additional factor of
1/2? [This question might well have been raised at our first encounter with the
interaction energy that was interpreted in terms of an image charge, (15.5).]
The electrostatic energy of the physical point charge can be identified with
the work performed in assembling successive infinitesimal multiples, ¿Ae, of
the final charge, with A increasing from zero to unity. At a stage where the
assembled charge is Ae, the image charges that it induces are also reduced by
the fraction A of their final values, and so also is the electrostatic potential of the
image charges. If <j) is the final value of that potential, the total work performed
in raising A from 0 to 1 is
d\e\<l>=le<f>; (17.55)
just one half of the electrostatic energy associated with the final charges, re
garded as independently assignable.
and a similar lattice composed of charges -f-e, which is displaced to the right
(along the 2 -axis), relative to the first lattice, by the distance 0 < 2zq < 2a.
In particular, for z0 = |a, the complete one-dimensional lattice is composed of
alternatingly signed charges with uniform spacing a. What is the electrostatic
energy of any individual charge? The charge pattern here is just that of a
physical charge e and its images in the parallel conducting plates, as shown
Fig. 17.2. Accordingly, the electrostatic energy per unit charge is twice that
given in (17.40); in the example with all charge spacings equal to a, the energy
per charge, twice that presented in (17.28), is —(e 2 /a)21n2.
which has already been used without comment in connection with (16.65). Thus
we have
00
= 1 £—00 K ^ - « - H
v~
+
_
1 00
\ s [ imw(z —z^/a ÍTmr(z-\-z, ) l a 1
(17.57)
~ 2a ^ i J’
which, in turn, can be rearranged as
lA r n 7r / n 7r / 2 ^ . nw . nir ,
- > c o s — (2 — 2 ) — c o s — (2 + 2 ) = - > sin — 2 sin — 2 . (17.58)
a n= 1 L a a J a {
n —1
a a
Should we now return to the physical range of the coordinates 2 and 2 ', so
that
—a < 2 — zf < a, 0 < 2 + 2 ' < 2a, (17.59)
only the physical charge contributes to (17.57) and this relation becomes
0 00
- ^ s i n — 2 sin — 2 ' = S(z — 2 '). (17.60)
n= 1
this interval,
. UTTZ
/(* ) = £ sin ----- (17.61)
a
n=l
g (z ,z ';k )= - £ (17.62)
And this function obviously obeys the boundary conditions, inasmuch as all
sin(mr/a)z vanish at z = 0 and z = a. Or is it really that cut and dried? Let’s
construct the Fourier series for the function 1 — (z/a):
oo
Z r -^ • n7r
1 - - = > sin — z (17.63)
a " a
n= 1
where the latter integral is evaluated as
a mr , / ' a . mr . a
— cos — z 1 ------ sin — 2 (17.64)
mr a \ a (rnr)2 a mr
The result is
2 rnr
— sin — z. (17.65)
> - : - E mr a
n=1
Certainly both sides of this equation vanish for z = a. But suppose we put
* = 0?
In order to refute the possible inference that the sine Fourier series applies
only to functions that vanish at 2 = 0 and a, let us evaluate explicitly the
right-hand side of (17.65):
£ — sin ^ - z = I m i £ I f e * - / * ) ” = Im ^ ln , 1 . , (17.66)
a T r^ n V / tt i _ eW<« ’ v >
n=1 n=l
which does not exist. We are being told that unlimitedly large values of n do
contribute, contradicting the assumption involved in replacing sm{nK/a)z by
(inir/a)z. But if the dominant contribution comes from very large values of n,
the summation can effectively be replaced by an integral,
dt 2 [°° dt
— sin(xT) = —€(x) I — sini, (17.70)
t n Jo t
where
/ x / 1, x > 0
(17.71)
Now we differentiate with respect to x and observe that the derivative of the
step function e(x) is 2 6(x), reflecting the value of the discontinuity at x = 0 :
9 i 00 9 f ° ° rli
—/ dt cos(xt) = 26(x)— I — sint, (17.72)
ft Jo TCJo t
where the left-hand side is recognized to be
1 f°°
- / dtelxt = 2 S(x); (17.73)
^ J —oo
2 [°°d t .
— / — smz = 1 . (17.74)
* Jo t
It is now clear that g(z, z'\ k) will indeed vanish at z = 0, say, if the infinite
series that results from the replacement of sin(n7r/a)^ by (nw/a)z is a convergent
one. This being a question referring to very large values of n, it suffices to make
the test for k = 0. Then, with zja < 1, we get just the sum given in (17.65),
UK z(a — z')
- < 1: g (z} 0) = * I T — sin — J = (17.75)
a n'K "a
n=1
z< ( a - z > )
g (z,z’ )0) = (17.76)
17.8. PROBLEMS FOR CHAPTER 17. 189
it is evident that (17.75) holds, not only for 2 <C a, but for all 2 < z*. More
generally, the equivalence of the two forms for g (z, z'\ Ar), (17.13) and (17.62),
can be checked by carrying out the integrations involved in representing the first
version as a Fourier series in 2 , or in z*. [See Problem 17.4]
z —a
2. Prove that
1
E (n + 1 / 2 )*
= (2 * - l)C (s ),
(p, •!>,*)
ka = inn, n = ±1 , ± 2, ± 3 , . . . .
Finally, by combining the contributions for n and —n, obtain the result
(17.62).
5. Consider an infinite conducting sheet, with a unit point charge above it.
Use circular cylindrical coordinates, so that the point charge is located
at ( p '.f t ,^ ) , as shown in Fig. 17.3. Find Green’s function by using ap
propriate representations of S(z — z') and 6(<f> — ft), Get the differential
equation for the radial functions g(p, //) . Now consider the potential of
an infinite line charge of unit density,
and the associated radial functions y(p,p/). Construct the latter from the
differential equation and boundary conditions. Show that the resulting
function can be expressed in terms of the complex quantities
17.8. PROBLEMS FOR CHAPTER 17. 191
C -C *
r = 2 Reln-^—
where
P = \ ( r - r')j. |= [p2 + pn - 2pp' cos(<j> - <j>')]lf2. (18.2)
A complete eigenfunction decomposition of G could be obtained by using the
addition theorem (16.69). In (18.1), we encounter a new type of Bessel func
tion, ATo, the modified Bessel function of zeroth order, defined by the integral
representation
Ko{Ap ) = J ~ dk k ^ Q . (18.3)
K ° W = Jo dss^ p - ( 18-4)
In terms of this new function, the Green’s function in the region between two
parallel plates is
193
194 CHAPTER 18. MODIFIED BESSEL FUNCTIONS
As always, that statement comprises two bits of information. The first one is
the homogeneous equation applicable for rx / , which is
or
|Jo(kP)
( - é ¿ Fé + * ) w p) = t ’ ( - p dp p f p + ^ j
pOO
= / d kkJ 0(kP), (18.10)
Jo
which is
p ( P ) = 0, P > o, (18.11)
according to (16.50).
The second bit of information, that carried by the inhomogeneous term in
(18.7), is extracted on integrating the latter equation [or (18.10), (18.11)] over
an arbitrarily small circle surrounding the point r'± , using the two-dimensional
form of the divergence theorem:
The value of the constant appearing here, along with other useful results,
can be obtained from alternative representations of K q. For these we return to
(18.6) and evaluate this rotationally invariant integral by using the direction of
(r — r ')x as the a?-axis:
eikxP
K »(AP) = i / (18.14)
dkmdk*k* + k¡ + W
195
We now have the options of integrating first with respect to kx, using the integral
of (13.22), which results in (ky = k)
K0 ^ , (18.15)
V F T I7
or, of applying a particular example of the cited integral to carry out the inte
gration with respect to ky (here k = kx)
1 Í 00 JkP
K 0(XP) = - j dk (18.16)
yjk2 4- A2 ‘
Equivalent versions of these two forms are
e-vs'+t* pc
d9 e —t co s h 6
(18.17)
K° V = l * 7y/s2
37 + ?
t2
(with the latter produced by s = tsinh#), and
[°° cos ss
COS f°°
Ko(t) = / ds-—======= = / <¿0cos(¿sinh0), (18.18)
Jo y s 2 + 12 Jo
respectively.
We apply the exponential version (18.17), for small t , by introducing a tran
sitional value of the integration variable s, S', such that < < S < 1 . Then the
integral can be divided into two parts, with appropriate approximations, as
In + V s2 + 1 2) . 25 (18.20)
nT ’
dss* i e (18.23)
196 CHAPTER 18. MODIFIED BESSEL FUNCTIONS
We now see that the integral appearing in (18.22) is r '(l), which is also ^(1),
inasmuch as T(l) ■= 1. Then on consulting (17.39), we learn that r '(l) is just
the negative of Euler’s constant 7 :
1
K 0(t) I d s-ex p (18.26)
Jo t
or
(18.27)
Of course, apart from the specific numerical factor, this asymptotic form is
immediately apparent in the version of the differential equation for Ko(t) that
is analogous (m = 0) to (16.37),
(18.28)
(18.29)
irP /a
P > a, G
The parallel conducting plates suppress most effectively the Coulomb field of
the point charge at large distances, on the scale set by the distance a. That
is qualitatively clear in the picture wherein the conductors are replaced by the
infinite set of image charges; the potentials of the two lattices composed of
positive and negative charges, respectively, will cancel almost completely at
large distances.
The image charge picture also makes evident that G(r, r') approaches Green’s
function for the vacuum— Coulomb’s potential— as |r — r'| becomes very small
on the scale set by a (except when the unit charge is quite close to one of the
18.1. MORE BESSEL FUNCTIONS 197
where Eint is the energy of interaction, with the conductors, of a unit charge
that is stationed at a point r « r'.
00 dkz
Ir - r 'l 2 tt k\+k2
poo pOO
1
Jo(kP)e,ikx Z
=- dkk (18.31)
* JO J -o c k2 + k2
1 _ 1
dkJ0( k P ) e - kW , (18.32)
|r-r'| “ ,/ W T Z 2
f
1 _ 1
dkzK 0{\kz \P)e i k x Z dkI<0(k P ) c o s k Z ) (18.33)
V P 2+ Z2 ~ *
where k replaces kz .
Observe that in one representation an exponentially decreasing function of
Z is combined with (for sufficiently large kP) an oscillating function of P, while
in the other representation it is an oscillating function of Z that combines with
(for sufficiently large k P ) an exponentially decreasing function of P. The neces
sity for this mixing in the functional character of the multiplicative constituents
stems from the significance of l/|r — r'|, r ^ r', as a solution of Laplace’s equa
tion:
/ d2 1 d d2 \ 1
+ p o p + dZ2) V F 2 + Z 2 ~ ° ' (18.34)
To the extent that the first derivative term is relatively negligible (P large), the
two second derivatives must be of opposite sign, requiring that a convex function
of one variable be combined with a concave function of the other variable.
Perhaps we should also point to the close relationship, indeed, equivalence of
(18.18) and (18.33). First notice that the statement of completeness in (16.53)
reduces, for an even function of x, to
■ pOO
2 f°°
/(# ) = — / dk cos kx / dx1cos kx1/ ( x 7) (18.35)
ft J o Jo
198 CHAPTER 18. MODIFIED BESSEL FUNCTIONS
Now regard (18.33), with Z playing the role of x, as such a Fourier integral
representation o f (P 2 + ¿J2) - 1/ 2 and conclude that
r°° i
K 0(kP) = J d Z 'z o s k Z '- — = = ; (18.36)
i— .r
-kZ
dk k Jo(kP)- (18.37)
+ Z 2 Jo k ’
that
- kZ
/ dt-
1
y/s2 + 12
= sinh
■(?)
2T
In — ,
s
T >^, (18.40)
f°° 1
/ dssln - Jo(s) = 1. (18.42)
Jo s
Or, we might multiply (18.39) by cos At before integrating with respect to t from
0 to oo. Then the following version of (18.18),
i At
Ko(Xs) = f dt (18.43)
Jo
yields
poo poo 1
h ( l ± % ) e'li =V k- ± . <«»•«»
from which follows
Although this is the preferred form for our present purposes, it is also possible to
remove the reference to <j> and present these just as relations among the Bessel
functions. [See Problem 18.1.] The two differential operators appearing here
behave as inverses when acting on the functions exp(im<¡>)Jm(kp):
I ( A ±4 1. (18.48)
ik \dx d y ) J [ ik \dx d y )_
It is in this sense that we begin with Jo(kp) and derive from it the following
constructions for both positive and negative values of m:
where the latter form uses a symbolic expression of a Taylor series expansion in
powers o f r^. It’s time to recognize that (18.48), or
(18,51)
when Jo(kp) is the operand. We can also assign to this unit vector a polar angle
4>, in the sense that
¿V i - 1, (18.57)
OO
lid d \ lm
+ h { k p )- (18'61)
Accordingly, this time we have
oo
e - r ^ * V _ L _ g fcrl • ( - l / f c ) V i ekp‘ co s($ -< ¿') __ e*m(^~^ ) / m (& //),
(18.62)
where
d .d \
e*'* (18.63)
dx *dy ) ’
and the outcome is
K 0(kp). (18.65)
This definition is analogous to (18.61), but differs in the additional sign factor
(—l ) m. That flaw is accepted in order to achieve the positiveness of both Im(t >
0 ) and K m(t > 0 ), which property of the Jm is evident in the power series
inferred from (16.27),
oo /J ,\ m + 2 n
Jm(t) = Y , ( [ M (18.66)
“ (m + n)! n!
1 ( d
+ RíkX ,r j- = R<ki *rj-. (18.68)
-]fc(*i + i* i)
k V dx '*)
202 CHAPTER 18. MODIFIED BESSEL FUNCTIONS
It is clear that integration over the polar angle of k'± [k'x -|- iky = k'e101] will
produce the factor of exp( 2m^) that stands on the left-hand side of (18.65).
Accordingly, we are free to specialize to <f> = 0, which is to say, x = /?, y = 0,
and get
JKp
+ k12 + k2
0 dkl eK p
2 tt k'2 + k'2 + k2
1 e - V k’y + k2P
(18.69)
J k '2 + k2
1 r°°
K m(kp) = - dO (cosh 0 + sinh $ y ne ~ kPcoshe. (18.70)
= K m(t). (18.72)
i i i
Jo 2 tt p2 p'2- pp>cos <j>~ [(p2 p'2)2 pp1)2]1/2~ p2- p'2'
+ 2 + - (2 ( 1 8 -7 6 )
18.2. PROBLEMS FOR CHAPTER 18 203
which is just what emerges from the second form of (18.44) after A —►iX.
Finally, we comment on the asymptotic form of K m(t) for any given m and
sufficiently large t. Approximations analogous to those in (18.26) (9 ~ s/t)
gives the leading term from (18.71),
2 . Show that, for m > 0, the differential operator in (18.49) acting on Jo(kp)
satisfies
JL
m
(kp'£t)-
-A Y Jm<f> jp
ik dx ^ %d y )
where F is a differential operator constructed from kp and d/dkp.
u" + - u ' - ( l + u = 0.
X y X J
/j/(<c) — 1 <7|/(23?),
K u(x) = ^ íi/+ 1 (J í/ (¿x ) -f iNv{ix)),
show that this asymptotic behavior is consistent with that found in Prob
lem 16.10, which holds for complex x ) \arg#| < 7r.
Cylindrical Conductors
19.1 Rectangle
In a step beyond the parallel conducting plates of Chapter 17, we consider
another set of plates that intersect the first set at right angles, thereby producing
a cylindrical region with a rectangular cross section. The latter is displayed in
Fig. 19.1 along with a convenient coordinate system. Again, we seek Green’s
function for a vacuum region,
(19.2)
205
206 CHAPTER 19. CYLINDRICAL CONDUCTORS
and
c 2 tti'k . mir ,
- y )= £ X u sin~ y sin~ y >
v — ' .
°{y (19.3)
m=1
combining them into the two-dimensional delta function
rmr .
h m{x ,y ) = ^ j\ s m l^ x y j \ s m — y, /,rn = 1 ,2 ,3 ,..., (19.5)
52
+ 7lm) 9lm{z, Z') = S(Z - z'), (19.8)
dz 2
in which
/7T
Tim = (19.9)
We know the solution of (19.8) for a cylinder of unlimited length [see (13.24)],
9 im (z,z') — 1 ep-Tlml*-*'!
= ------ (19.10)
2tí„
When |z — is large, on the scale set by the greater of the rectangular dimen
sions, it is the / = m — 1 term that dominates (19.7):
1 -yulz-z
G (r ,r ') ~ ^ u i x y y W n i x 1,!/) (19.11)
2th
Suppose, now, that conducting walls at z = 0 and c truncate the infinite
cylinder, thereby encasing the unit point charge in a rectangular conducting
enclosure or cavity. The solution of (19.8) under such circumstances is given in
Chapter 17. But there is a more symmetrical procedure here; begin with the
three-dimensional delta function construction
y
T
obeys
J (^r ) ) — b\\i8mmi8nni. (19.14)
in which
Yet this is at the heart of the possibility of constructing a new cylindrical shape
by applying the sort of symmetry consideration developed in Chapter 17. As a
glance at Fig. 19.2 indicates, Green’s function for a cylindrical conductor with
the cross section of an isosceles right angle triangle will be produced if we satisfy
208 CHAPTER 19. CYLINDRICAL CONDUCTORS
the boundary condition of vanishing on the diagonal line y = x. And that will
be accomplished by introducing an image charge through reflection in this line.
If the positive charge is stationed at the point with coordinates x ' , y' , the
negative image charge produced by the interchange of the x- and y-axes is found
at yf, x * . Thus the two-dimensional charge structure is
E
. hr . mir . Í7T . . m7T , /7T , m7T .
sin — x sin -----y sm — x sin -----y — sm — y sin ----- x
a a a a a a
lm
- y')> (19.18)
/< m
where now
2 . / tt . mw . m7T . /7T
<t>lm(x,y) — sin — x sin -----y - - s m -----x sm — y (19.19)
a a a a a
id 2 d2 \
- 2 + ^ 2) = y?m<l>im(x,y), (19.20)
where (/ < m)
(19.21)
1 » 1 1 .
r = -a ta n — = — -=a = - a . (19.23)
2 6 2i/3 3
ei
directed from the center of the inscribed circle to the three apexes. These three
vectors are presented in terms of the unit orthogonal vectors i and j by
1 > = °- (19.25)
a—\
One notes that
3 ... ... 3
= 2 (U + U) = 2 l x ’ (19.26)
2 3
r x = l x t x = - J 3 e a/xa, (19.27)
<2=1
where
Ma = e a • r j . . ( 1 9 .2 8 )
\ /3 1 \ /3 1
/¿i = 2/, = — Y x " 2 V) ~ ~ Y X ~ 2y' ( 19-30)
The trilinear coordinates of a point are the perpendicular distances from the
origin to the three lines, drawn through the point, parallel to the sides of the
triangle. A coordinate is negative if the associated line lies between the origin
and the related side. Thus, the three sides of the triangle are represented in
trilinear coordinates as /iX = —r, ^2 = —r, and ^3 = —r, respectively.
210 CHAPTER 19. CYLINDRICAL CONDUCTORS
according to
3
r'± + 2 h Y , » ae a, va = 0, ± 1 ,. (19.31)
a=l
= (^ 3 - ^ 2 )e 3 + (^1 - ^ 2 )e i, ( 1 9 .3 2 )
which means that any pair of trilinear basis vectors will do (no surprise in a
two-dimensional space). In actuality, then, we extend <5(rj_ — r±)> the initial
point charge density (in the transverse plane) to
¿ a = ea *kx, = 0. (19.34)
The i>2 and 1/3 summations in (19.33) are evaluated as [(16.63) and (17.56)]
1 00
—i2hvzkz
^ E
-i2hV2^2
(2tt) v3zz—oo
1 w TT
= (2*)i E
V 7 / 2 = — OO
E
/ 3 = — OO
(19.36)
which is
J dk2dkz6(k2 — Trl2/h)8{kz — xh/h) = -^=, (19.40)
according to the value of the Jacobian determinant [Karl Gustav Jacob Jacobi
(1804-1851)] that relates the integration variables k2, k3 and kX) ky:
=—
3ha Y
exp
.27r ^
a
i
exp
a
(19.42)
in which the /-summation extends over all integral values of the la that satisfy
the restrictive condition (19.37). It is worth noting here that a displacement of
rj_, for example, by 2Ae&, 6 = 1,2,3, alters the trilinear coordinates j±a by
_ 3 1 / a= b: 1 ( .
ea . e j - 2 ¿a6 2 _ \a#6:-l/2' ( 1944)
Accordingly, any exponential function of the (ia in (19.42) changes by the factor
.27r
exp /»2fc+ £ ( - / « / » ) = exp(i27r/6) = 1. (19.45)
l Zh
a?b
The next step supplements the lattice o f positive charges with a negative
charge distribution that is designed to satisfy the specific boundary conditions on
the three sides of the triangle. For that purpose we first observe that the image
of r'± in the triangle side perpendicular to ei has the rectangular coordinates
K = K + Vh + r ~ 3M / 4 + 9 = | “ ^ I ; ^ ’ (19.49)
2tt Oír
exp + ^3/^3 ” hh) — exp * + h ¿4 + — hh)
(19.50)
which, from its construction, vanishes when y!x = —r, for that is the line in
which reflection takes place. (See also Problem 19.1.) This property, being
independent of the particular choice of the /a, continues to hold when the indices
of the la are cyclically permuted: 1 —^ 2, 2 —^ 3, 3 —^ 1, thereby producing
9 7T .27T
exp —*3A — exp
^3^2 + “ hh) *3¿ (h u í + ¿3/4 + ^1/4 ~ ^A)
(19.51)
and
27T Oír
exp 3^03^1 + il /^2 + *2/4 - *3*0 exp ^ ^2/4 feft)
(19.52)
Furthermore, the sum of the three pairs, which of course also vanishes for
fi\ = —r, responds in a simple way to cyclic permutations of the ¡Ja indices.
One can check that the substitution //2 —► ¿4 —* M reproduces
the set of six terms, multiplied by the common factor
tT
1
1
1
s.
HJ
1-----
H!
And the implication of this property is that as the complete structure of six
terms vanishes for = —r, so does it vanish for fi2 = —r, and for /4 = —r; it
obeys the boundary conditions.
In an extension of what could be done without comment in (19.18), we
remark that the six term structure is either unaltered or reverses sign under
the six operations of cyclic permutation of the la (three in number, including
214 CHAPTER 19. CYLINDRICAL CONDUCTORS
the identity), and anticyclic permutations combined with sign reversal (three in
number), as illustrated by /i —►—/ i , ¡2 —* —/3 > h —» —h • These operations,
applied to an exponential function of the fia in (19.42), produce an analogous
six term structure, thereby presenting us with the final generalization of the
initial transverse delta function:
(3fca)1/3*»(/0
27T 27r
= exp + hP2 + ht*3 — hh) — exp
~ 23fi(^l^ 1 ^ 2 ^ 3 ~~
” 27T 27r
+ exp — exp
23 ^ 2 ^ 3 ~~ ^2^ 3ft(^2^i ^ 2 ^ 3 ~ ^2
2 tt 27t
+ exp i7rr{hpi + I1P2 4- ht*3 — hh) — exp
oh
(19.56)
The /-summation is extended over all sets of the la that are distinct with respect
to cyclic and anticyclic-reflection permutations. But one must not overlook the
fact that vanishes if any la = 0 (this is verified in Problem 19.2); the
possible values of la are ±1, ± 2 , ___Such are complete, and orthonormal
over the equilateral triangle, in the sense [analogous to (16.29)] that
Inasmuch as the functions <^¡(rj_) have their origin [(19.33)] in the exponen
tials exp[?’k_L •rx], they obey the differential equation
where [(19.38)]
(19.59)
is indeed a quantity associated with the set /, being unaltered by cyclic and
anticyclic-reflection permutations of the la. The smallest value of 7 2 is realized
for the /-set indicated by { 1 , 1 , —2 },
7? > ( x ) 2- ( 19-6°)
2 . 2w,
<MaO (19.61)
(3 h 7 a =l
19.4. CIRCLE 215
Its complete symmetry in the three trilinear coordinates is consistent with the
general observation [(19.53)] that <f>](fj), with any two la equal, is unchanged by
cyclic permutations of the /ia. [Note that the two other possibilities, multipli
cation by exp(±¿27r/3), are not options for a real function.]
It has been remarked that the </>i(fi) are normalized functions in the sense
of (19.57). We rise (partly) to the challenge of verifying this for the simplest
function, <¡)q(/¿). That presents us with the integral
+ X ! sin + sin +
a^b '
_ . 27r/ x . 2tt .
+ 2 sin — (/ii + r ) sin — (/i2 + r ) j. (19.62)
The second version exploits the threefold rotational symmetry of the equilateral
triangle; the three integrals involving fxa are independent of the particular choice
of direction, and the six integrals containing fia and fib are the same for all
pairs of directions. Now the latter trigonometric functions within braces can be
presented as
1 1 47T/ x 27T 27 r ,
2 - 2 cos ■y (a41 + r) “ cos y (A41 + r ) + COS — { m - /12), (19.63)
which also uses the equivalence under integration of —fi\ —/i2 = fi% and fi\. It is
left to the reader [Problem 19.3] to prove the null value that the area integrals
of these cosine functions possess. Then the outcome is just
(19.64)
which finally involves the area of the equilateral triangle. (See Chapter 25 for a
simpler treatment.)
19.4 Circle
Parallel plate, rectangle, triangle— we come to the circle, which is to say, the
circular cylinder. Let it be of radius a, with its axis identified as the 2 -axis of
a circular cylindrical coordinate system. The potential of a unit point charge
in vacuum within a conducting cylinder at zero potential— Green’s function— is
the sum of the Coulomb potential of the point charge and the potential of the
surface charge distribution:
G(r,r') dSi
F Lí i + / s | r- rx|’
ron S : G(r,r') = 0. (19.65)
216 CHAPTER 19. CYLINDRICAL CONDUCTORS
We make use o f the Coulomb potential construction (18.33) and the addition
theorem o f (18.64) in writing
-I -| pOO 00
/ OO /*27T
The boundary condition on Green’s function requires that the surface charge
potential annul the point charge potential on the surface S (p ~ a, z : —oo —►oo,
<j) : 0 —►27r). In view of the completeness of the functions exp{ikzz) exp(¿m<^)
over this domain, that null statement must apply to every value of kz and m (a
null function is uniquely represented by zero coefficients). Thus, we have
or
<rm(kz) = (19.71)
Im(k(i)
The result is presented, in the notation
/ OO rib 1 °°
¿ />';*), (19.72)
■ °° Til— — OO
The introduction of the construction (19.72) then yields the radial differential
equation satisfied by gm,
Km(ka)
P > p' • 9m(p,p'\k) = B Rm(kp) Imikp^j (19.79)
Im(k(l)
A = C [ffm( V ) - Im(kpr) ^ ^
B = C Im(kp‘ ). (19.80)
(19.82)
£ (19.83)
dt
£ = 0, (19.84)
dt
218 C H A P T E R 19. C Y L IN D R IC A L C O N D U C T O R S
or
t [ K m (t)I'm ( t ) - I m ( t ) K ‘m ( t ) ] = 1; (19.85)
the value of the constant on the right hand side is supplied by an inspection of
the asymptotic forms of the functions, (18.78), (18.79). Therefore C is unity.
When the forms for p < pf and p > pf are united we regain (19.73).
Incidentally, the Wronskian appears explicitly in verifying that the surface
charge density described by (19.71) is what it should be in terms of the normal
component of the electric field:
(19.86)
,T(ri)=¿ ¿ G(t't') p - a
The introduction of the representation (19.72) for G, along with the definition
of crm{kz), (19.68), converts the above relation into
Gm{kz) — C (19.87)
P~a
Itn jk pf)
a-i-9m {p,p']k) = Im{kp') ka
dp P~a Irn{ka) ’
(19.88)
on applying the Wronskian relation (19.85). The result for am{kz) reproduces
(19.71).
This reference to the surface density of induced charge naturally raises the
question o f its total amount. That is
/
oo p2 tt
dzi a / d<j>i <r(ri) = <ro(0) = -1 , (19.89)
-oo J0
according to (19.68) and (19.71), along with the fact that I o ( t ) is unity at t = 0.
We ask another question: What is the potential generated outside the radius
p = a by the combination of the internal point charge and the surface charge
distributions? A glance back at the forms appropriate to p > a will show
that the significant combination of the two constituents is just what appears
in (19.70), with the common factor K m{ka) replaced by K m(kp). In short, the
potential for all p > a is zero; the world outside the cylinder is unaware of the
balanced distribution of charge within.
Now let us turn back to the parallel conducting plates in order to develop
a useful correspondence with the circular cylinder. We recall from (17.13) that
the reduced Green’s function for the former pase is
sinh kz< sinh k(a — z>)
9 (z,z'\k) (19.90)
k sinh ka
or, equivalently,
Notice the close analogy between the pairs sinht, e_< and 7m(A), K m(t)> re
spectively. There is also an analogous Wronskian relation, appropriate to the
simpler differential equation,
All this leads us to search for an analogue of the alternative form of g {z, z'\ fc)
that is presented in (17.77). We stress that the statement of equivalence in the
latter is an identity, holding for all values of k, real and complex. As shown in
Problem 17.3, both sides of this equation have the same singularity structure,
poles along the imaginary axis at k = ±i(n7r/a) (n a positive integer), with the
same residues, and vanish asymptotically as k —►oo. The proof o f equivalence
of the two sides of (17.77) follows from the observation that, as a function of the
complex variable Ar, the difference between the two sides is everywhere regular
and bounded, which identifies it as a constant, according to the theorem of
Cauchy, ascribed to Joseph Liouville (1809-1882), said constant being zero in
virtue of the asymptotic behavior.
With this analogy before us, we now look back at gm{p, 7/; Ar), (19.73), and
recognize that it must be given equivalently as the sum over all the pole terms
that are associated with the zeros of Im(ka). Again, as with sinh Ara, these occur
for imaginary values of k [recall (18.58)]:
1 d d o m 21
~ — ImiJ^p) — 9?
p idp
~ pidp k2
' 7 j
Id d l*2 m2
7m(Ar>) = 0. (19.94)
P d>P dp
Then we refer to the Wronskian relation (19.85), for Im(t) = 0, to learn that
1 1
-^-m(d:¿7mn) = ~7~' r/ / i • , \* (19.100)
Tran)
(19.102)
k2a2 + 7mn'
The sum over all such pairs of poles then yields the desired construction
/ / I\ P m n (p )P m n (p ')
*»(/>, p ;* ) = ! , mn/a2 -
k2 + J2 (19.103)
in which
y/% J m { y m n p / a )
(19.104)
p / \
mAP) ~ a J U 7mn) ’
(19.105)
1
-S (p - p ' ) = Y l Pmn(p)Pmn(p'), (19.107)
P n= l
the statement of completeness of the functions Pmn{p) over the interval from
p zz 0 to p = a. These functions are also orthonormal in the sense that
( ¿ + 1 - m-2 ~ — ) = 0, (19.111)
/ / X 1/ 2
i y j Jm(t) = A (t)c o s $ (t). (19.113)
Introducing this into the differential equation, and setting the coefficients of
cos 4> and sin $ equal to zero, yields
^ _ ( ^ + ( , _ 2 ! ! l l i Z i ) A = 0,
where the stated value of the integration constant is provided by the asymptotic
form,
t \m\ : A (t) ~ 1, <&(t) ~ t — ( m 4- (19.116)
^ i2 = 1 - - W ± + T ’ (19-117)
in a sequence of approximations that begins with (19.116) as the first approxi
mation. For the second approximation we use A « 1 and t \m\ to get
(19.118)
and
<!>(< ) « t - ( m + l ) ! + l . (19.119)
(19.120)
A" 3m2 - |
_____ ______________ÍÉ_
(19.121)
A t4
and then
1 m2 - i 1 (m2 - l ) 2 3 m2 - j
(19.122)
t2 8 t4 + 4‘ t4
19.4. CIRCLE 223
leading to
. l\ 7r 1 m2 — 4 1 (m 2 — 4 ) 2 1 m2 — 4
* ( ‘ ) M<- ( r o + 2 ) 2 + 2 — f ^ + u — t^ L - 4 - i r L+- - <1£U23>
The roots of the Bessel function Jm(¿), the zeros of cos<$(tf), are inferred
from
/ 1\
$ ( 7) = ( n _ I ) n = 1,2....... (19.124)
1 ^
1 : 7 = + 2n —
2) ^ 2 ’
n i ^ m2 - \
2 : 7 = + 2n — (19.125)
2; ^ 2 m + 2n — \ ’
3 : 7 = + 2n —
2]^ 2 ' m + 2n — \
m'
2_ 1
1 ___ 4 7 ( TXl — - 7 I — 6
37T3 (m -f 2n — 1/2)3
1: = 2.3562,
*47T 1
(19.126)
2: T + 6^ = 2-4092’
3tt J _ 31
= 2.4031.
4 67T 162tt3
Compared to the actual value
2: X + T - = 5.5205, (19.128)
4 147r
3: — + —---------- = 5.52004,
4 14x 2058tt3
which are much more accurate approximations as compared with
1 f°°
-6 (p -p ') = / d k k J m(kp)Jm(kp')
P Jo
2 Jm(ymnP/Q^JmiSfrnnP I a) (19.132)
= £ a2 [^ m (T m n )]2
The first of these is valid for all / ? , / / > 0; the second one applies with 0 < p, p' <
a. Does the latter yield the first version in the limit a —> oo? Indeed. We begin
by remarking that, for any appreciable value of 7mn/u> which becomes k in
the limit, ymn is arbitrarily large. That, in turn, implies correspondingly large
values of the integer n, which is to say that the range of values— the spectrum—
of 7mn/u —►k is effectively a continuum. In view of these very large values of n,
the first approximation of (19.125) suffices, and the asymptotic form (19.112)
can be applied in evaluating J^(7mn)> where
1\ 7r
7mn = n7T + m (19.133)
2 /2
Accordingly, we have
(19.134)
Now we have only to recognize [(19.133)] that w/a is the interval between suc
cessive n-values of 7 m„ / a , for then, with the replacements
J in iL ^ k, — —*•dk, (19.136)
a a
we realize the integral form of (19.132).
19.5. CIRCLE AND SEPTUM 225
1
8(<j> - <j>') - - y ' s i n ^ s i n ^ ' . (19.137)
7T z ' Z Z
v—l
Indeed, the function sinm<^, which vanish at (j> = 0, will also vanish at <¡> =
provided
2m = v = 1 , 2 , 3 , 4 , . . . , or m = i t / = 1, 1, 2,.... (19.138)
/
OO J jL 1
same appearance as for the unobstructed circle, only the implicit reference to
the spectrum of m indicating the change. That directs our attention to what is
new, the values m = 1/2, 3/2, . . . .
As we know, the differential equation for Jm(t) can be presented as
£_ m2 — 1/4
+ 1 - ) = 0. (19.141)
dt2 ¡5
sini=V^rfj
/ 2 \ 1//2 / 2\
< /i/2 (0 = (^ J cos(< — ít/2 ), (19.142)
where the numerical factor is chosen to conform with the here everywhere valid
asymptotic form (19.112). All the roots of this Bessel function are given exactly
by
yn = n7r, n = l , 2 , . . . , (19.143)
which is also displayed in the three identical statements of (19.125) for m = 1/2.
Beginning with this simple form for J 1 / 2 , we now proceed to construct J3/ 2 ,
J 5/ 2 , ___ In doing this, we apply (18.47), which continues to hold for non
integral values of m. As we have mentioned, exp(im(j>)Jrn(kp) is a solution
of the differential equation in (19.140), irrespective of the choice of m, and
any derivative of this function is also a solution of the differential equation.
Furthermore, the combination of derivatives employed in (18.47) is such as to
introduce the additional exp(±i<j>) dependence:
(dp
±+-± e*** A ± £ A j (19.144)
dx dy pd<j>)
-<■+ w i ( s + ^ + <19-147>
where, in view o f (19.144) and (19.145),
Jm+l CO 1 d \ Jm(t)
(19.150)
tm+l \ t dt J
A tm
1 d\ sin t
w o - d y v ^ - i i (19.151)
which also displays the asymptotic form. For arbitrary /, the function yj2ir/t x
cos(t — is multiplied by a polynomial in l/ t} of even degree, that begins
with unity, while (2'jrft)1^2 sin(t — (m -f ^ )~ ) is multiplied by a odd polynomial
in 1/t.
We can give the construction (19.151) another form by observing that
sint _ 1 f 1
dfi eZfit (19.153)
~ T ~ 2 j_ 1
Then we write
1 d sint
dfi n e
t dt t
(19.154)
(19.155)
2V 2 . i+ x/2 1
1 (- d2 \ ’ s m t
228 CHAPTER 19. CYLINDRICAL CONDUCTORS
( d2 2 d \ sint
(19.157)
\ w + t 7 t + 1 ) ~ r - 0-
More generally, the connection between J1+ 3/2 and Jj+ 1/2 that is inferred from
(19.156),
Ji+3/2(t) _ 1 A ^ + 1/2(0 ( .
ti+3/2 — 2/ + 2 \ dt2) f '+ !/2 ’ 119.15»;
Jl+3/2(t) _ 1 d J/+ 1 / 2 CO
<1+3/2 ~ t dt tl+l/2 ’ (19.159)
(d 2 21 + 2 d W , + i / 2(<)
(19.160)
\dt2 t d t' ) <1+ 1/2 - u -
This is, of course, a version of the Bessel differential equation for m = / + 1/2.
One can rewrite it by noting the symbolic relations
Id d l
dt^ dt t’
ti— t ~i - ( i - i y - H i . * ( * + 1) (19.161)
dt2 t) dt2 t dt t2
d2 2d /(/ + 1 )
f " 1/2 J/+ i/2(¿) — 0; (19.162)
dt2 t dt t2
3. Prove that the area integrals of the cosine functions in (19.63) vanish.
4. Show that
d_ d d
pJm(kp)-7£-Jm(k p) pJmfó1
dp
= (k2 - k,2)pJm(kp)Jm(k'p).
19.6. PROBLEMS FOR CHAPTER 19 229
n 7^ n : Í dp p Pmn(p)Pmnt(p) = 0 .
Jo
Í dpp[Prnn(p)]2 1.
Jo
6. Again employ the first relation of Problem 4 , with kf = k*, to prove that
the roots of Jm are real.
7. Check the Wronskian relation between / m(tf) and K m(t), (19.85), using
the known t —> 0 behavior for m = 0 . Then, for m > 0 , use the known
t 0 behavior of Im(t) to deduce that of K m(t).
l e 2 f° ° di
Eint~ n a J0 U oW )2’
10. Find Green’s function for a unit point charge in vacuum outside an infinite
conducting circular cylinder of radius a.
f ( p) = (?„„£ ) ,
n= 1
Chapter 20
Spherical Harmonics
V 2i = 0, r > 0. (20.1)
In terms of this solution, we can generate a large number of others. For example,
taking a to be a constant vector,
V 2(a •V ) —= 0, (20.2)
r
we find
-1 a •r
a •V - y»3
(20.3)
r
is also a solution for r ^ 0. Continuing this operation, we see that
231
232 CHAPTER 20. SPHERICAL HARMONICS
l fl number of independent
solutions
0 1 i
1 X\y X2) X3 3
2 3xmXn 5
Why are there only five independent solutions for / = 2? A symmetrical tensor
has six independent components but because of the constraint that the tensor
satisfies Laplace’s equation, it must be traceless, leaving but five independent
components.
The general polynomial of degree / can be constructed from the monomials
k\ + k2 + k3 = /. (20.6)
How many of these monomials are there? To answer this, we first ask the
analogous question in two dimensions: how many monomials of the form
(20.7)
kt + k2 = / - k3. (20.8)
/ - Jf es+l , (20.9)
* i
£ ( / - f c 3 + l) = - ( / + / ) ( / + 2 ) . (20.10)
fc3=0
From this set of polynomials, we wish to find those combinations which are
solutions to Laplace’s equation. Since V 2 acting on a homogeneous polynomial
of degree l produces a homogeneous polynomial of degree / — 2, of which there
i ( f - 2 + l ) ( Z - 2 + 2) = i / ( / - l ) (20.11)
For the cases / = 0,1,2, this agrees with what we found above. The solutions
we find in this way are called solid harmonics, Y¡(r). To emphasize the fact
that they are homogeneous polynomials of degree /, the solid harmonic may be
written in terms of a surface (or spherical) harmonic, Y /(r/r):
- ¡¿ r tf ( ; ) ■ (20-M)
where the latter form, also a solid harmonic, results from inversion and is the
solution constructed in (20.5).
(a •r)', (20.15)
(again, note that the components a¿ are complex), suggesting that the condition
that a2 be zero can be automatically satisfied if we write
ai +«*2 = É-,
ü\ id2 — 5
*3 = í + í - , (20.19)
r 1, . . x 4 iy 1/ . . x — iy z
* --= 2^ai ~ + 2 (ai + + a3r
(a*r)* = rl , (20.21)
( 20.22)
where, in the last line, we have employed a convenient form of a Taylor ex
pansion. In this way, we have constructed, from polynomials of degree /, 2/ -b 1
independent functions which are solutions to Laplace’s equation, the coefficients
of the powers of in the expansion
£¡—m
(l + m)!
(sin 6)
r l , y/(l + m)! (/ - m)! Y 0 ~ m)-
l—m
(cos2 0 — 1)*
(20.23)
dcosO 2Ul
á +mt -
= V’ln (20.24)
a/ ( / + m)\ (l - m)\
( a « r )'
(20.25)
/! r' E
Y im (M ) - *íf- m ( - 0 , - ¿ ) , (20.29)
or, using the explicit form (20.26), we obtain the alternate version
where
1= 0 : Too = - ? = , (20.33)
V4?r
3 x + iy
1= 1: Yn = sin
8it
3 x — iy 3
Y i,-i = sin 0e
f i 15
b I(x + iy)2 _ Q L sini 0e2i*
1= 2 : Y22
~ V 32 ir 3 2 tt
15 z(x + iy) _ 15
Y2i = - —\l — cos 0 sin 0 e*^,
236 CHAPTER 20. SPHERICAL HARMONICS
(20.35)
with
díl = sin 0 dO d<j). (20.37)
This integral can only contain rotationally invariant combinations, that is, it has
to be a function of scalars constructed from a and a*. Since a2 = a*2 = 0, the
only such scalar is a •a*. Therefore, there must be an equal number of factors
of a and a*, which means that the integral (20.36) is zero unless / = we have
ljrm 2
~ 4 (2 / + 1 ) ! ’
(20.41)
h = / d x ( l - x 2)1, I0 = 1 , (20.42)
Jo
20.3. ORTHONORMALITY CONDITION 237
the last form being valid for / > 0 . Alternatively, we could evaluate // in terms
of the beta function, B {m ) n),
1 —x
= 221 f d tt'il-t)1 t =
Jo
= 22' 5 ( / + l , / + l )
/!/!
= 2 2'- (20.45)
( 2 / + 1 )! ’
where we have noted that for integer m and n,
B ( m ,n ) = i * d t f n- 1( l - t ) n~ l = -^ )!(W 1. ) ) ! . (20.46)
Jo (m + n - 1 ) !
= ¿ É düYrm{ 0 , m m'(O A ),
m = -l J
(20.47)
where we have used the generating function (20.25). What we now must do is
extract the coefficient of fr°m (a* *a )i>which is achieved as follows:
1 1
21 j ^ (20.48)
m ——l
238 CHAPTER 20. SPHERICAL HARMONICS
where we have used (20.19), (20.24), and the binomial expansion. Compar
ing this with (20.47), we obtain the orthonormality condition for the spherical
J
harmonics:
dnY,*m(0, ct>)Y,,m,(0, <j>) = 6w 6mm,. ( 2 0 .4 9 )
, m _ t/(2 7 T Ó !(s in 0 y
©i. - \ j— ------- j Ü T ’ (20.51)
2/ -f* 1
P/( cos 6). (20.53)
1 (cos 2 6 — 1 )/
Pi (cos 0) = (20.54)
_d( cos 0)_ 2 */!
P /(l) = 1. (20.55)
/
! 9
ci(cos 0)Pj (cos 0)Pj/(cos 0) = —— -Su». (20.56)
-i 2/ + 1
We will begin to explore the physical significance of Legendre’s polynomials in
the next chapter.
is also a solution.
20.5. PROBLEMS FOR CHAPTER 20 239
2. Check that
0
£+ — V>lm = V (l ~ m)(l + m + IJV’i m + l ,
= \/(i + m )(/ - m +
<7?+
Note that these are related by £+ 'ip¡m —* Show that
3 .1 3 V/ vi . 3 ,
a ? - c o t # 7 8 ? J ( a " r)
e -^ ( - A - c o t 4 ^ ) Ylm = W + ” * )(/- m + l ) y , m_ i.
3. Rewrite the conclusion of the previous problem for 0 /m(0). Derive the
differential equations
4. Prove
y*m = ( - l ) mVi-m,
by noting that (a •r)* is obtained from (a •r) by the replacement £+ —*
, £ - —►—*£J.. Show the consistency of this result with the conclusion
of Problem 2 .
This is equivalent to
a = (—¿cos a, —¿sin a, 1 ),
that is,
a •r = —i(x cos a + y sin a) + z.
r
i'2w
£(*-r)v””=
and, in particular, that
P7T ^
Pi(cos 0) = / — (cos 0 — ¿ sin 6 cos a )/ .
Jo ^
and arrive at a known result. Would you expect that P/(cos 0) is also given
by
i*IT J
I — (cos# — ¿sinocos a)
JO *
In any event, again work out ]T^ 0 t lPi(cos0), using the latter formula.
0 « 1 , / » 1 : P i( cqbO ) ~ J 0(W ).
+%) (é)
d_
dx
Coulomb’s Potential
The motivation for constructing the solid harmonics was that they formed,
in terms of homogeneous functions, a particular set of solutions to Laplace’s
equation. Since these harmonics are functions of the spherical angles 9 and
<t>, Laplace’s equation should be expressed in spherical coordinates, where the
Laplacian has the form
1 r 1 a (. &\ i d2
V2
r 2 dr \ d r) r2 sin 9 dO \ m dO) sin2 0 d(j>2
(2 1 .1 )
(This may be immediately inferred from (16.15) by noting that the distances
corresponding to infinitesimal changes of the spherical coordinates r, 9, and
<j> are dr, r d9, and rsin9d(j), respectively.) Thus, since the solid harmonics,
(20.13) and (20.14),
are solutions to
V 2y,m(r) = 0, r ¿ 0, (21.3)
and
(21.4)
Tr { r~ ]~l } - i(/ + 1 } { r ~ l~ 1 } ’
i d2
+ + 1(1+ 1 ) Yim(0,<f>) = 0 . (21.5)
sin2 0 d<f>2
When the 9 and <j> dependence of Y¡,n is separated as in (20.31), the differential
equation for 0 ;m is
1 d
sin 9 89
©im(0) = O. (21.6)
sin2 9
243
244 CHAPTER 21. COULOMB’S POTENTIAL
___________ 1___________
(polynomial of degree l in cos 7 ),
V r l + r< — 2 r>r< cos 7
(21-9)
where r> (r<) is the greater (lesser) of r and r*. (The recognition that the square
root is also the absolute value r>|l — (r</r>)e*7 | makes it evident that this
expansion converges.) The polynomial of degree / appearing here is a solution
to (21.5), and so must be a linear combination of Y¡m(0, ft)'s, —/ < m < L On
the other hand, as we will show below, this is just Legendre’s polynomial in
cos 7 , (20.54), that is
1
(21 .10)
| r-r'|
(In fact, this is how Legendre first introduced his polynomial in 1784.) For
7 = 0 , this expansion is trivially
(2U1)
which supplies the normalization condition
Pi( 1) = 1, ( 2 1 .1 2 )
while the Laplacian is given by (21.1). For r < r', (21.10) shows that the
solution to (21.14) can be expanded in powers of r, [cf. (21.2)]
and
r'+0 i
—r = 4w~— 6(0 - e')6U - (21.19)
2d~rG r'-0 Sm$
If we write
Aim — 1 Cr/m» Elm = T Clm) ( 2 1 .2 2 )
( 2 1 .20 ) is satisfied automatically, while ( 2 1 .2 1 ) reads
47T
Cim = ¿ T T T ^ O ^ ' ) . (21.24)
246 CHAPTER 21. COULOMB’S POTENTIAL
By substituting this into (21.23) we obtain the completeness statement for the
spherical harmonics,
which allows us to expand any function of 9 and <j) in terms of spherical har
monics. We therefore have obtained such an expansion for Green’s function
1 - ■■ Affr
G (r’ r '} = = £ T + T ^ T V fn fr m m W ,* ')- (21-26)
Comparing this with the alternative representation, (21.10), we obtain the re
lation
4 tt l l
P,(C0S7) = ^ E ^ M ) ^ ', ^ ) - (21-27)
m
The relation (21.27) is called the addition theorem for spherical harmonics.
Let us finally show explicitly that this function of cos 7 actually is Legen-
dre’s polynomial, (20.54). As noted above, this is easily done by considering a
particular coordinate system, where
y|m(0 , ^ ) = (21.30)
so that only the m = 0 term contributes to the right-hand side of (21.27), which
is therefore, by (20.53),
¡1 — cos 9 (21.32)
21.2. INFINITESIMAL ROTATIONS 247
provides the following form for the differential equation obeyed by Legendre’s
polynomial:
P M = o. (21.33)
One easily checks this equation for the first few polynomials,
We shall express the induced changes in Yim(9,(f)) and 'i/jim by a vector ro
tational operator R , which is defined initially by its action on r:
ir = (S w .R )r ; R = rxV . (21.37)
Correspondingly, we write
Thus, for the relatively simple situation of a rotation about the 2 -axis, given by
the coefficient of 6u>Zi we have, respectively, from (21.36),
<2L39>
For the rotations about the other axes, we see that
and
d
Rx ± iRy = e*’ * ( ± i “ cot # J ^ ) > R * ± iR y = - t £ (21.42)
T dt±
4>lm = (21.43)
■\/{l + m )!(i — m)!
£± - (f ^ (21.44)
(R x i i R y )ipi m — ®v u ± m ) ( / m + • (21.46)
If we take half the sum of these, from which RxRy and R yR x cancel, we get
R 2V»;m = - / ( / + l ) V w (21.49)
d
Oj¡Yi™(0> <t>) = imYim(0, <t>), (21.50)
** ( ± ¿ + ¿ c o t * A ) Ylm(e,<f>) = V ( l T m ) ( l ± m + l)Yl¡m±1(0,4>).
(21.51)
21.3. SPHERICAL BESSEL FUNCTIONS 249
And it will be clear that the analogue o f (21.49), still written compactly as
(/ — m)\ (/ -f ra -f Ar)!
(sin 0 )-m- * 0 ,,m+*(0). (21.56)
(/ —m —k)\ (/ 4- m)\
If we here set m = where [(20.51)]
in agreement with the lower sign choice in (20.32). Working down from m = /
produces the equivalent version.
the latter form appearing when the direction of k is adopted as the 2 -axis. Now,
for functions that depend only on the angle 0 , the completeness statement of
(21.25) reduces to [recall (20.31) and (20.53)]
oo fi pir
f(0 ) = V ) ( 2 / + l ) f i ( c o s 0 ) - / de( sin 0' Pi (cos 0f)/(O') (21.61)
where
i'M k r) = \ J x dp P,(p)eikr* . (21.63)
which includes
P /(—cos#) = (—l /P /( c o s 0). ( 2 1 .6 6 )
This tells us what happens when the complex conjugation of the right-hand
side of (21.63) is combined with the transformation ¡i —> —fi. The appearance
of a factor of (—1)* is compensated by the response of il. In short, the function
ji(k r) defined by (21.63) is a real function.
The introduction of the Legendre polynomial construction (20.54) into this
definition yields
d ^V - i y
M *) = r ' l j ' d\i d¡S l
dp J 2 ' /! 2 ' /!
(21.67)
the latter being produced by /-fold partial integration. [This can be turned into
a differential formula. See Problem 21.3.] But we recognize this! According to
(19.156), we have
which accounts for the name “spherical Bessel function” that is applied to j¡.
We note the translation into this notation of (19.158), (19.159):
<P_ 2 d_ *(* + 1 )
ji(t) = 0 , (21.70)
dt2 + td t + t2
(V 2 + k2)eik •r = 0; (21.71)
the spherical coordinate version of this equation, valid for each / in the expansion
of (21.62) is [see (21.1) and (21.5)]
L A . f. ji(kr) = 0 , (21.72)
r2 dr \ dr J
where a and (3 are the spherical coordinate angles for the vector k. The complex
conjugate version is, with r —» r',
oo /
g—*k = £ Y2 (21.74)
/=0 m——l
Q oo /»oo Í
= - £ / dkk2 j ^ n i k r ' ) Y , Ylm{eA)YL(0',<t>')- ( 2 1 .7 7 )
m tl /
252 CHAPTER 21. COULOMB’S POTENTIAL
1 2 f°°
-=*£(r — r') = — / dkk2 jj(k r)ji(k rf) ) (21.78)
r n Jo
the completeness property of the radial functions jj(k r), 0 < k < oo. It is not
a new statement, for on introducing ( 2 1 .68 ) we get
1 I°°
-8 (r -r ')- d k k J l+1/2(kr)Ji+1i 2(kr'), (21.79)
r Jo
the completeness statement for the Bessel functions Jm, with m = / 4 - 1/2 [cf.
(16.59)]. Incidentally, putting (21.78) back into (21.77), yields the completeness
relation for the spherical harmonics, (21.25).
The reverse side of the coin of completeness carries the image of orthonor
mality. We look to (21.77) for the basic functions,
(21.80)
which combines summation over l and m with integration over k. The corre
sponding orthonormality statement is contained in the values of
♦
r / 2 \ 1/2 i Y ‘A 1/2 1
(-) ii(* r )y /TO(tf,^) -J ji'(k'r)Yiimi(0, <j>)
[0
2 f°°
= 8>w8mm' ~ d r r 2 ji(k r)j¡(k 'r)
n Jo
where the radial integral is just (21.78), with the substitutions k -»• r r -* k
r' k'.
Let us not hurry on without noting the form taken by Coulomb’s potential
(13.17):
1
w i * (r-r')
|r^r'| ( 2 tt) 3 P e
v° ° s 9 ,oo
2 ^ (2 / + l)P /(co s 7 )--- / d kj](kr) ji(kr'), (21.82)
l-Q * JO
or, equivalently,
2 r°° 1
A< 1: - jf dtj,(X t) j,{t) = — A'. (21.84)
= = <2 1 8 5 >
/
Pi( cos7) = £ aimYlm{ e j ) Y ; m{e'
m——l
(± 0 1 _ V «2
2/+ 1 - 2 ^
m——l
2. Verify the rotations of the coordinates 0, <f>, £+, and in (21.36), and
thereby verify the rotation operators in (21.39)-(21.41).
3. Starting from the integral representation for the spherical Bessel function
ji(t) given in (21.67), show, by repeatedly integrating by parts so as to
eliminate the (1 — p 2)1 factor, that
[this is (19.155)], and from this give explicit forms for ji for / = 0, .. .3.
4. The expansion
5. As with the spherical Bessel function of the first kind, the spherical Neu
mann function may be defined in terms of cylinder functions of half-integer
order by
8. Prove that
kJ m {k p )J m {k p ) k J m+ l (k p )J m + l (k p )
A k ( J m + l { k p ) J m { k p ) -f- Jm ( k p ) J m + \ ( k p ) ) •
d k [ p 4* p f
Use this result to show that if the Jm{kp)1 k : 0 —» oo are complete, so also
are the Jm+i(&p)> i.e., the completeness of all the J/+ i/ 2 (&p), / = 1 , 2 , .. . ,
follows from that of the J\p¿{kp). [You might also want to think about a
convergence factor or something equivalent.]
Chapter 22
Multipoles
where, for convenience, we will choose the origin to lie within the charge dis
tribution. If r is large compared to the characteristic dimensions of the charge
distribution, we may expand Coulomb’s potential as follows:
1 1
—r' •v ! + l(r' •V )2- + ...
| r-r'| r r 2 r
1
+ + 1-lr »(3rV - lr '2) *r + ( 22 . 2)
r r 2 r
so that the potential, in its leading behavior for large distances, has the form
#/ \ e r •d 1 1
(22.3)
^(r ) = ¡: + — + 2 ^ r ' q< r + ....
•r
•r
Figure 22.1: Geometry of field point and source point for a bounded charge
distribution.
257
258 CHAPTER 22. MULTIPOLES
Occurring here are the first three moments of the charge distribution,
d = y W ) r V ( r'), (22.5)
q= / ( d r O ( 3 r V - l r ' 2M r '), ( 2 2 .6 )
which are the total charge, the dipole moment vector, and the quadrupole mo
ment dyadic, respectively.
Using this potential, we can now calculate the interaction energy of the
charge distribution with an additional point charge e\ located at a point r lying
far outside the charge distribution:
(22.9)
E = (22.10)
(We have seen this form of the dipole interaction energy before, in Section
4.1.) Note that the trace of q is zero, YLila = 0* This is a starting point
for considering the interaction of one charge distribution with another charge
distribution. For example, if one had a dipole di rather than a charge, e\)
interacting with a charge distribution which had only a dipole moment, d 2 , the
interaction energy deduced from ( 2 2 .8 ) would be
3r •d x r •d 2 — di •d 2r 2
E — —d 2 • ( 22. 11)
ip5
4?r
( 22. 12)
ir 1 lm
27+TYrm( e ' , n
259
implies
The connection with the previous definition, (22.4) and (22.5), is, for example,
given by [see (20.34)]
(22.15)
05
II
o
o
(22.16)
<51
II
1
1
p\o = d Z} (22.17)
E = J(dr)p(r)<f>(r). (22.19)
(j)]m being the expansion coefficients. Inserting this multipole expansion for the
potential back into (22.19) and using the definition (22.14) for the multipole
moments, we obtain the simple expression for the energy of interaction
E = '£pL<f>im, ( 2 2 .2 1 )
lm
generalizing ( 2 2 .8 ).
Rather than expressing the interaction energy in the unsymmetrical form
( 2 2 .2 1 ), let us formulate the energy in terms of the interaction of the charge
multipole moments of each distribution; that is, we seek a generalization of the
dipole-dipole interaction, (22.11). If we let ri and r 2 be measured from points
within pi and p2, respectively, while r measures the distance between these two
origins, as illustrated in Fig. 22.2, the interaction energy can be written as
E= /( d r i ) ( d r 2) f i ^ ^ 4 . (22.22)
J |r + r i —r 2|
260 CHAPTER 22. MULTIPOLES
Since the two charge distributions are non-overlapping, we can expand the de
nominator occurring here in a double Taylor series:
(22.24)
p h i = = E
(22.25)
m
Further, recall the generating function for the spherical harmonics, (20.25),
a —►V , (22.27)
aa _ * V 3i = 0 , r > 0. (22.28)
r
In this way, a comparison of (22.25) and (22.26) gives the identity
= (-l)V fc n (22.29)
the same way under rotations.) Using (22.26) twice with the above replacement,
we obtain
(r i . v y x (- r 2 . v ) M , ,9
h\ h\ r ~ (~ 1} ^ ____
x ^ 2 \ j2¡x ^2)
According to the definition of -0/m, (20.24), the product of two of these functions
is
V^imi = ^Iii2mim2 ^i+^,w i+m 2 ) (22.31)
where
1/2
(U + h + mi -b m2)! (h + h - rni - m2)!'
C\i l^rri\m>2 — (22.32)
.(¿i + mi)! (/i - mi)! (/2 + m2)! (¿2 - m2)!.
h+h ,
^/í+íajmi+ma —( 1)
2(/1 + /a) + r 11 h
(22.33)
Combining (22.22), (22.23), (22.30), (22.31), and (22.33), taking the complex
conjugate, and identifying p/m, (22.14), we find for the energy of interaction
47T l 1/2 l
E = Cl1l2mim2 /l + /2+1
y ( - 1 ) '1 .2(Zi + / 2) + 1J — r«
E = Y , = e<£ — d •E + - V •q •E + ----
262 CHAPTER 22. MULTIPOLES
[Answer:
This result holds whether the dipoles are permanent or induced. This is
the high-temperature, or classical, limit of the van der Waals, or Casimir-
Polder interaction between molecules.]
Yim(r) =
where the latter notation means the remaining terms have increasing pow-
ers of r 2r/2. Now replace r by V (with an operand 1 /r understood) and
arrive at
7. Consider
22.1. PROBLEMS FOR CHAPTER 22 263
(2Q! 1
tfm (V )- = (-1 )' V,m(r),
/! 2' r2i+1
from which the result (22.29) follows. Compare with the result of Problems
20.10 and 20.13.
Chapter 23
G — finite at r = 0 , (23.4)
and
G = 0 at r = a. (23.5)
To determine the expansion coefficients, Ajm and Bim, we use the equations for
the continuity of G, (21.18),
265
266 CHAPTER 23. CONDUCTING AND DIELECTRIC SPHERES
y ' | 7 /! +± !1 + !r
lr'l+1 \ „ , .,., ,
1 1 r" L ~
a 2 f+r ) Blm + lr A lm Ylm{6 A ) = 4 » £ y hl, ( M ) y ; ( i ' 1A
lm 1 V /m
(23.7)
at r = r; . Solving (23.6) by introducing C\m defined by
o - E ( ^ r - ^ ) « ~ T ) . P 3 11)
where we have used the addition theorem (21.27), y being the angle between r
and r', (21.8). Noticing that
rlr/l _ a rl
— > a > r, (23.12)
a^+T “ 7 (a 2 / r ' ) /+1 ’ r
a 1
G= (23.13)
|r-r'| r' |r — r'|
r (23.14)
which, of course, lies outside the sphere. Thus we have achieved for the sphere
the analog of the image solution given for the conducting plane in (15.32).
What is the induced charge density on the inside surface of the sphere? This
charge density is proportional to the radial electric field, according to (11.69),
since the normal is inward, and so in the negative radial direction. Differenti
ating (23.11) with respect to r = r>, we obtain
4™ = - £ ( 2/ + l ) ¿ Pi (cos 7 ). (23.16)
i a
23.1. INTERIOR OF CONDUCTING SPHERICAL SHELL 267
Alternatively, we could use the image charge form of Green’s function, (23.13),
to derive
_________ 1 - (r '/ a f_________
(23.17)
[1 — 2 (r '/a ) cos 7 -f (r'/a)2]3/2
which indicates that as r' —* a, the only significant charge buildup is near 7 = 0 .
The total charge can be computed from (23.16) by use of the orthonormality
condition, (20.56), which, for /' = 0 , implies
J a2 sin 7 dy d<j>
47ra2
= -l, (23.19)
jl(7ln) = 0. (23.21)
Exhibited in (23.20) are the radial functions that are orthonormal in the sense
of the integral
/ d rr2R i„(r)R i„i(r) = Snn>, (23.22)
Jo
namely
11/2 ji(7tnr/a) (23.23)
JÍ (7>n)
This orthonormality property can be verified easily for / = 0, jo(t-) =
1 sini,
where (23.21) reads
or
7 on = m r, n = l , 2,.... (23.25)
Then we have
fa i 9 2 / • W7rr . n'7rr
/ dr r2“ ^ (—1 ) ^ s i n ----- sin -------
Jo a3 r2 a a
orthonormal,
I (dr) mn (r)<£i7m 'n ' (r) — fim m ' fifin' j (23.30)
(r)<f>*mn(r')
G (r ,r ') = 47
Imn tL / « 2
= £ ( 2 '+ i w = o s r ) 5 i # M 2 . (23.32)
Jn 7' " /a
v2. Rm(r)R*n(r') 1 , ( 1 ^ \
(23.33)
¿ í t,2
„ /« 2 ~ 2/+r< U 1+1 o2,+i y
fdr.2 +
+ H
r dr! _ E ±
r 2 i r ,U
1 ^ (r, r')
r; =-
V - r Or j, (23.34)
23.1. INTERIOR OF CONDUCTING SPHERICAL SHELL 269
with
gi(a.,r) = 0 . (23.35)
Approximations for the roots of the spherical Bessel functions are obtained
from (19.125) by writing m a s i + | ,
7,„ * (I + 2 „ ) f - £ $ ± 2 - 1) * f t (23.36)
as illustrated by
(23.37)
3tt3 (n + 1/2 ) 3
A a2 _ 1 2/l r21 \ a? (I 1 \
39)
^ 7 ,2„ 2/ -h 1 70 dVr ( r a2, +1) 2/+1V2 2/ + 3 J ’ ( '
or
V 'J_-- I01
-v?
1
(23.40)
^ 1 1
(23.41)
2—1 (mr)2 6 ’
00 1 1 00 1
cot (23.42)
^2 i — n7r f ^ ^2 ( n7r) 2 __ f2 ’
n = —00 n= 1 v 7
the comparison o f the t terms in the small t expansion of the two sides yields
(23.41).
The second in the unlimited sequence o f such statements considers the square
of (23.33), integrated over both r and r'. Here, the orthonormality of the R¡n
270 CHAPTER 23. CONDUCTING AND DIELECTRIC SPHERES
yields
00 4 1 Ca Cr
r' 2r/2/
^ = 2w ^ l drr,l dr' r'+i a2* * 1
1 1 ra 0 / / r \ 2/ +
2
i\ 2
= 2W Í W ü T > l i r r { 1 - U )• (23A3)
where the symmetry in r and r' permits one to simply double the contribution
of the integrals for r* < r. The outcome of this integration is given by
1
V 1 = 1 1 (23.44)
¿^-s7 (4„ 2 (Í 2 /+ 3)2 2 /+ 5 ’
2
^ 1 _ 1
(23.45)
~r[ ( n7r) 4 90 ’
sin t
M *) = ~T
^2/+ 2fc
( 2 / + 1 + 2k)\
x 7 Jk=0
00
(l+ k )\ 2k (23.47)
= (2<)' ¿ ( - 1)kfc! ( 2 / + 1 + 2 *)!
0
/! 1 1
M<) = ( 2 Í)' 1 - 1 V t4 + . . . (23.48)
( 2 Í + 1 )! 2 21+ 3 + S (21 + 3)(21+ 5)
from which we regain (23.40) and (23.44) by examining the t and t3 terms in
the power series expansion of (23.46). We leave it to the reader to carry out the
next step and derive the sum of the inverse sixth powers of the y¡n [Problem
23.3].
One use of these summations,
oo 1
(23.50)
n= 1 l¡n
comes from the evident inequality
rW (23.51)
’ N r
or
1
7/i > (23.52)
(<r,(r)) l / 2r ’
these are lower bounds to the smallest of the 7 ’s for a given /.
As an example, consider (23.41), (23.45), and the next member of the se
quence,
00 1 1
S ( ^ = 945' <23'53>
n= 1 v 7
which yield, successively,
rapidly converging from below to the limiting value. One can also improve the
last member of such a sequence by an approximate extrapolation to the limit
that employs it and one or more of the preceding numbers. [See Problem 23.4.]
We defer the discussion of another application of such sums, one that focuses
on the second smallest 7 value. [See Chapter 25.]
* (n )
dSi
G(r'r,)- l v h i + l |r-rj|’
G {r, r') = 0 , r on S. (23.55)
»<
00 7
Alt
a
1
(23.56)
|r —r'| ¿ ¿ « 2 7 T T m= —l
272 CHAPTER 23. CONDUCTING AND DIELECTRIC SPHERES
and (r > a)
Js dSl\ r - r i| = 1 (23.57)
1 A| 1=0 m
where
«■im = J dS\Y*m(6\, ^i)<r(n). (23.58)
The requirement that G, the sum of these components, vanishes for all 6 and </>
at r = a leads immediately to
(23.59)
where
-/ u
r = — < a. (23.62)
Then, with the introduction of the vector r', with magnitude r' and the direction
of r', presents Green’s function in the image form
1 a 1
G (r,r') = (23.63)
|r —r'| r' |r —r'| ’
the same as (23.13).
With the information contained in (23.59), that
we can construct the surface charge density. According to the definition (23.58)
and the completeness of the spherical harmonics, it is [dSi = a2dQ]
One should notice, however, that an aspect of the surface charge distribution,
the total induced charge, is given directly by croo:
(23.67)
= ¿ ( - | G(ry))
The introduction of the Green’s function form (23.60) immediately reproduces
(23.65). Now let’s use the image version, (23.63), which we write as
[r/2 — 2 rr' cos 7 -f r 2] 1/2 [(r r '/a ) 2 — 2 rr'COS7 -f a2] 1/2 (^3-68)
l a 1 — (a /r ' ) 2
(23.69)
47ra2 r' [1 — 2 (a /r ') cos 7 + (a /r ' ) 2]3/2 ’
and we note particularly the ratios of the surface charge density to its average
value, ( —a / r ') / 47ra2, at the point nearest the unit charge:
14 .1
7= 0: (23.70)
(i-^ r
7 = tt : w < 1. (23.71)
(i + f ) 5
1
— id S a -(r) = ^ f ¿¿(cost ) rj— = i. (23.72)
ind J
Qind 2 J_x [1 - 2 (a /r /) cos 7 + (a /r ' ) 2]3/2
1 _ f2 00
is especially interesting in the limit t 1. For fi < 1, the limiting value of the
left-hand side is zero [as illustrated by (23.71)], whereas with /jl = 1 [see (23.70)]
the limit is infinite. And it is the content of (23.72) that, independently of the
value of t , half of the /i-integral of the left-hand side, from — 1 to 1 , is equal to
one. In short, we have learned that
(23.74)
274 CHAPTER 23. CONDUCTING AND DIELECTRIC SPHERES
= (23.76)
S in 7 Z7T Z7T
Eiint = Le2[
2 G ( r , r ' ) - G 0(r,r')]
1
2 r0 r0 - (a 2/ r 0)
(23.77)
2 ” r l - a 2‘
This is just the Coulomb energy between the charge e and the image charge
—(a/ro)e, multiplied by the characteristic factor of 1/2. [Recall Section 17.6.]
The magnitude of the force of attraction between charge and conductor,
Q ar0
F = ^int = e2
Eint
dr0 ( r 2 - a 2)2
2 00
6(<j) — (j)1) —►— sin m<j) sin m<¡)1
7T
m= 1
^
00 i
(23.79)
= E s
23.3. CONDUCTING PLATE AND HEMISPHERICAL BOSS 275
Figure 23.1: Sphere bisected by plane. Shown are the locations of the physical
charge at r' and the image charges at r', , and rfy .
which exhibits the equivalent image charge structure at reversed values of <j>') or
of y'. Accordingly, if images in the plane y = 0 are indicated by the subscript
y, Green’s function for this external situation is (see Fig. 23.1)
1 1 a / 1
G( r ,r ') = (23.80)
k - r '| lr ~ ryl r>
a/r0 a/r0
F = e2 (23.81)
_(2 r 0) 2 [ro - {a2/r0)]2 [r0 + (a 2 / r 0)]2.
One notes that as ro approaches a the first and third terms of F tend to cancel,
leaving just the force associated with the sphere, whereas, in the limit of large
ro/a, it is the first term, associated with the conducting plane, that dominates,
It should also be observed that, while the sum of the second and third
terms in (23.81) is positive, so that the force always exceeds that produced by
the conducting plane alone, a similar remark about the first and third terms,
and the force produced by the conducting sphere alone, requires that ro be
sufficiently large. Indeed, the contrary circumstance,
a/r0 ^ 1
(23.83)
[r0 + (a 2/ r 0)]2 > ( 2 r 0) 2 ’
or (x = ro/a )
4x3
> 1 (23.84)
(x 2 + 1 ) 2
276 CHAPTER 23. CONDUCTING AND DIELECTRIC SPHERES
here the effect of the conducting plate is to reduce the attractive force that the
conducting sphere alone produces. The largest reduction occurs for ro/a = 2.297
where it is 8.7%.
where we will take e to be a constant in the interior of the sphere, r < a. The
boundary conditions at r = a are, from (11.58) and (11.60),
G is continuous, (23.88)
and
(23.89)
~€T r G
G is continuous, (23.90)
and
r = r ; +0
= 47r-T^7:<5(0 - 0 ' W ~ *')• (23.91)
r = r '-0 SlIid
23.4. DIELECTRIC SPHERE 277
As is familiar by now, the solution in the three regions has the form
4 ^ ( 0 0 (23.95)
Aim
l(e + 1 ) + 1
,-,-r
Blm (23.96)
2/+1
( e - 1 )/ 4-KY*m(6', <f>') a 2,+1
C]m (23.97)
l(e + 1 ) + 1 2 1+ 1 r ' ,+ 1 ’
4
Dim (23.98)
]m+ 21 + 1 r •
Green’s function, outside the sphere, is therefore found to be
CO
(e - 1 )/ a21* 1
- E l(e + 1 ) + 1 rl+l r,l+l Pl{ cos 7 ).
1
r,r' > a : G( r, r') (23.99)
|r - r'|
We now ask what is the leading behavior of this potential when the separa
tion between the point charge and the sphere is large compared to the radius
of the sphere, r' a. Since the Ith term in the sum behaves as ( a /r ') i+1, only
small values of / contribute. The leading contribution arises from / = 1,
1 /N 4 e — 1 a3
r a: G (r ,r ) ~ ------- - 7 ---------- cos 7 . (23.100)
v ’ ' | r-r'| e+ 2 rV 2 1
r *r'
cos 7 = , (23.101)
rr
this asymptotic form o f Green’s function can be rewritten as
G (r ,r O = Í - T - 7 7 + 4 *d, (23.102)
|r —r'| ró
the two terms of which have simple physical interpretations. The first term
is due to the point charge while the second is the potential arising from the
induced electric dipole moment of the sphere [cf. (22.3)]. The latter is identified
from (23.100) to be
(23.103)
278 CHAPTER 23. CONDUCTING AND DIELECTRIC SPHERES
where —r '/ r /3 is interpreted as the electric field, Eo, at the center of the sphere
(in the absence of the dielectric) produced by the unit point charge. Since this
electric field is essentially constant over the sphere, we recognize that the electric
dipole moment induced in a dielectric sphere of radius a by a constant electric
field Eo is
d = ^a 3 E0. (23.104)
e+ 2
Finally, we write the expression for Green’s function inside the sphere:
r < a } 00 rl 2/ + 1
: G( r ,r ') = £ Pi (cos 7 ). (23.105)
r' > a r ,,+ 1 l(e + 1 ) + 1
' ;=o
Again, in the situation in which the point charge is located far from the sphere,
r' a, low values of / predominate:
I 3 r«r'
G ~ ^ 2 r E 0 = i —r •E, (23.106)
r' e + 2 r '3
where we identify the electric field in the dielectric as the negative gradient of
G, so that the field E in the dielectric,
E = 7 T 2 Eo’ (23-107)
is less than the applied field Eo if e > 1. This is equivalent to (5.87), (5.93)— See
Problem 23.7.
we recognize that surface charge density on the sphere induced by that Coulomb
potential at r> = a is proportional to (r< = r')*, and so we have the forms for
the reduced Green’s functions in the two regions:
1 r Ufi
V
rf < a, r < a : 9i(r,r') = 7 - f i j + Pi
e r! e a /+1
2 1
1 r"
rf < a, r > a : gi(r, r') = rj (23.109)
e r H-l •
23A. DIELECTRIC SPHERE 279
e ( /+ 1 - I p i ) = (/ + l)n - (23.111)
For comparison, let us write down the reduced Green’s function for the
previously considered situation in which the point charge is exterior to the
sphere:
/ dS •D = 47T Pi (23.116)
assures us that the radially directed electric field at distance r', in the vacuum
surrounding a spherically symmetrical arrangement containing a unit charge, is
indeed just that derived from the potential 1 /r '.
280 CHAPTER 23. CONDUCTING AND DIELECTRIC SPHERES
4. Verify the lower bounds for 1r shown in (23.54) and carry out the improve
ment referred to at the end of Section 23.1.1.
5. Construct Green’s function, from its differential equation, for the region
exterior to a conducting sphere. [That is, use the method given in Section
23.1.] Give the image interpretation of the result. What is the physical
significance of the two leading terms when one point is very far from the
sphere?
6. Calculate Green’s function for a dielectric sphere when the point charge
is inside the sphere, r' < a, using the direct discontinuity method given
in Section 23.4.
8. Consider the limit in which e —» oo, so that the dielectric sphere discussed
in Section 23.4 may be regarded as a conductor. Give the form of the
Green’s function in that situation for r' > a by taking the e —* oo limit
of (23.99) and (23.105). Show that indeed the interior of the sphere is an
equipotential region, but not one of zero potential. Also, show that in this
case there is zero charge on the conductor. Show that by adding a suitable
charge distribution to the sphere one can recover the situation considered
in Section 23.2.
e = 5 T T e <°>'
where E(0) is the field that the unit charge would produce at the origin
if there were no cavity in the dielectric medium. Also, discuss the limit
e —* oo.
23.5. PROBLEMS FOR CHAPTER 23 281
10. What is the statement of completeness for the functions 0 /m(0), as inferred
from that of the Y¡m(0, <¡>)1 Consider a conducting sphere that is bisected
by an infinitely thin conducting partition, as discussed in Section 23.3.
Find Green’s function for the interior of the conducting hemisphere.
Chapter 24
6E = 0 , (24.1)
where [see ( 1 1 . 1 1 )]
E = / ( * ) [* ¿ + ¿ ( e . v * + ! e 2) (24.2)
We now wish to generalize this situation to include conductors as well. The new
feature here is the existence of surface charges on various conductors implying
an additional contribution to the energy:
E = + ¿ ( e .v * + Ie’) ] + ¿ / ( ís ^ , (24.3)
where the volume integral extends over all space exterior to the conductors and
the surface integral is over all of the conductors, cr being the surface charge
density. (See Fig. 24.1.) This energy functional is to be supplemented by the
condition that the total charge on each conductor,
Qi - j dSior, ¿ = 1 , 2 , . . . , n, (24.4)
283
284 CHAPTER 24. DIELECTRICS AND CONDUCTORS
conductor conductor
Si S2
dielectric
E, and a is
(24.5)
which is subject to the condition that Qi be constant, that is
dSi 8a = 0. (24.6)
The implied surface integral here cannot be discarded since now there are contri
butions arising from the surfaces of the conductors. If we let n¿ be the outward
normal on the ¿th conductor, this surface term is
J (d r )V - ( £ « * ) = - ■ £ (24.8)
t
+V f dSi +8 a<j>+<j8<
j>. (24.9)
6E = ? / dSi6a<t> = 0 (24.13)
E = V(f>, (24.15)
The independent variables are <j> and ít, the latter of which is subject to the
condition (24.4). For the second form, D is regarded as an independent variable,
subject to the condition
while cr is determined by
E[D] = / ( * ) ¿ ^ . (24-19)
(24.20)
=/ dS,si r
286 CHAPTER 24. DIELECTRICS AND CONDUCTORS
V •A D = 0, (24.22)
J dSin- A H = o. (24.23)
r—^ f n •A D
0 = ^ 2 <f>i I dSi —— — (24.25)
i J
(a) (b)
of energy in going from the initial configuration (a) to the final configuration
(b). (In the following, the subscript 0 refers to the introduced conductor.) See
Fig. 24.2. The energy for (a) is
Ea (24.28)
where V is the volume exterior to the conductors and the charge on the ith
conductor is
J dSi = Qi • (24.29)
E t= [ (dr) £ * -, (24.30)
Jv-Vo bTre
where now the volume occupied by conductor 0 (Vo) is also excluded, and the
charges on the conductors are
Jds“iir”0" (24.31)
Ea (24.33)
Although D a is not the correct field for (b), it is an allowable trial function
to use in the energy functional, (24.19), because it satisfies all the necessary
conditions:
V -D „ = 47rp, (24.34)
/
n «D a
dSi = Q i, (24.35)
47T
/ dSo
n *D a V -D a
= 0, (24.36)
47T 4w
288 CHAPTER 24. DIELECTRICS AND CONDUCTORS
Ea > Eb (24.38)
(24.41)
which becomes, for the actual field values on the surfaces, <j>=
(24.42)
E = Qi<i>i - J 2 j dS<v * + j ( d l) ( 2 4 -4 3 )
% l
Here we regard (j> and a to be the variables while fa is specified [note that here
we impose no subsidiary restriction on a]. Under variations in (j> and a the
energy changes by
which becomes
8<t>: V . D = 0, (24.47)
These are the correct equations of electrostatics when there is no volume charge
density.
G (r ,r /) = 0 fo rro n S ¿. (24.51)
We will show that this Green’s function can be used to solve the electrostatics
problem in which the potentials on the conductors are specified. We wish to
consider a situation for which the free charge density is zero,
= fa on Si. (24.53)
- f a r )'V . [e(r)V G (r, r')] + G(r, r ')V . (c(r )V fa r )) = 4ir8(r - r')far). (24.54)
290 CHAPTER 24. DIELECTRICS AND CONDUCTORS
when we integrate over the entire volume, V, exterior to all of the conductors,
we obtain an integral over a surface S which is made up of all the surfaces of
the individual conductors, S¿,
(24.56)
The negative sign occurs since dS is directed out of the volume V, and so into
the conductors, while n* is the outward normal for the ¿th conductor. Deleted
here is the surface at infinity for which
dS ~ R2,
G < I | V G | < -^ ,
\V<j>\<±, (24.57)
so that the corresponding contribution goes to zero as the volume gets arbitrarily
large. Now imposing the boundary conditions, (24.51) and (24.53), we obtain
the desired expression for the potential,
24.6 Capacitance
Once we know the potential, we can compute the surface charge density on the
zth conductor by using
= ¿ n*•
Qi = Y 2 Ci^ i - (24.63)
j
E = \ n Q i*i = \ E • (24-64)
i ij
There is a consistency check between this expression and the variational prin
ciple which employs (24.44). Suppose we vary the potential on conductor i by
an amount 6(pi. Such a change induces variations in a and <p but the resulting
change in the energy from these induced variations is of second order due to the
stationary principle. So the first order variation in the energy arises only from
the explicit variation of
or,
i s = « ‘. (24.66)
BE
Q fi = ' E ' c n *i = Qi > (24-67)
j
(24.68)
where we have used the first line of (24.59), with <j) replaced by G, as well as the
differential equation satisfied by Green’s function, (24.50). This implies that
the coefficients of capacitance, C¿¿, (24.61), satisfy
= - J •V ') ( l ) = 0 , (24.69)
that is, the sum of all the coefficients of capacitance referring to a given con
ductor vanishes,
(24.70)
* 3
Consequently, the total charge on the conductors is zero when there is no volume
charge present:
J 2 Q í = J 2 C í^ j = 0. (24.71)
i ij
Furthermore, for this system, only relative values of the potential are significant.
If we were to add a common constant to all potentials, all charges would remain
the same:
Qi = + constant) = (24.72)
3 3
C n — —C 21 = —C 12 — C 22 = C ) (24.73)
24.7. PROBLEMS FOR CHAPTER 24 293
where C is called the capacitance of the system. The charges on the two con
ductors are
Qi = -Q 2 = C (4 1 - fa ) = C V (24.74)
where V is the potential difference between the two conductors, while the energy
is
= \ Y2 w a + i = \ c y 2 -
e (24-75)
n
As an application of these ideas, consider a capacitor constructed from two
parallel conducting plates of area A. The separation of the plates, a, is assumed
to be small compared to the transverse extent of the plates, a <C V A , the
approximate Green’s function therefore being that of two infinite plates as shown
in Fig. 24.4 [cf. (17.13)]. The material between the plates is characterized by
a dielectric constant, e . The above discussion applies to this situation so that
the system has a capacitance (7,
C = C\\
/ dSdS' (47r)2 (24.76)
The first surface integral here was previously evaluated in (17.18) and (17.21):
(24.77)
z= 0
where we have now included the presence of c in (24.50). The remaining surface
integral is trivial,
C = Í i í Íg = t a (2478)
yielding the well-known result for a parallel plate capacitor, derived in Section
11.4.
<fa{6'A Q
*,(r'#''6)=¿ ( ' - © O Í " ' l “ " '[ i2 4 r 2 - 2 ar cos 7 ]3/2 ’
where 7 is the angle between the direction specified by 0 , <^, and that
specified by 0', <^',
G ( r y ) = 4 ^ ^ (r )f ( r /), (25.2)
a
where the 4^ (r ) obey
set of functions.
The functions <f>a(r) describe the various possible field configurations, or
modes, such that the charge density is proportional to the potential that it
produces,
4npa(r) = 7 l<j>a(r). (25.6)
The coefficients of proportionality,
= 71, (25.7)
295
296 CHAPTER 25. MODES AND VARIATIONS
r on S : <f>(r) = 0 , (25.8)
namely, [(11.15)]
where n is the outwardly directed normal (dS = n dS). The additional terms
thereby produced in SE[</)] are
the first of which cancels the surface integral term of (25.10). Accordingly, this
E[<j>] is stationary for infinitesimal variations about a solution of the differential
equation for </> that obeys the boundary condition, without resort to surface
restrictions on the variations. One does, however, lose the general inequality of
(25.11), for finite variations.
For our immediate purposes, however, we want to retain that inequality, into
which we introduce two modifications. First, let us supply <j>with a scale factor,
AC
^(r) —+ K^(r), (25.14)
which is to be chosen to maximize E[<¡>\. Now,
vanishes at /c = 0 , and initially rises linearly with increasing k until the negative
quadratic term dominates, thus producing a maximum at the value of k given
by
J(dr)p<j) - 2nJ{dr) ^ ( V <£)2 = 0 . (25.16)
f(d r )p 2
^1 < J (dr)e(V<f>)2 (25.22)
[J(dr)p<t>r
which supplies an upper limit to A i, one that involves two arbitrary functions [</>,
of course, must obey the boundary condition]. For a given <f>, we can choose p to
minimize the right-hand side of (25.22). There is a general inequality, analogous
to that for a pair of vectors,
the equality sign applying only when p is a constant multiple of <j)—just the
situation described in (25.6). This gives
f (dr)e('V<j>)2
(25.24)
lS f(d r W
298 CHAPTER 25. MODES AND VARIATIONS
f(dr)e(V<j>)2
m = (25.26)
f(dr)<f>2
is required to be stationary, [6A = 0], so that
one infers that (j) is a mode function having the stationary value of A as its
eigenvalue. It is only the lowest eigenvalue, Ai, for which the stationary value
is a minimum.
As a prelude to another application of the inequality (25.18), let us, for a
given choose p according to
Then we have satisfied the differential equation connecting (j) with p, and the
equality sign in (25.18) applies. Indeed, one can verify directly that both sides
of this relation now coincide with
1
J (dr)e(V<j>)2 (25.29)
(4 tt) 2
In the next step we add some infinitesimal Sp to the p of (25.28). That would
require an infinitesimal change of (j) to maintain (25.28). However, the stationary
property tells us that ignoring the need for such a 8<j) only introduces an error
of second order. In short, (25.18), with the equality sign, continues to hold on
introducing the infinitesimal change of p) or
[The use of the latter is unnecessary if one returns to the E[<j>] of (25.9), as
already noted in Chapter 12.]
Now we choose 6p to be an infinitesimal multiple of
A'k{D(¡> — (25.32)
and
where
2 f (d r)(D </))2 /( d r ) e ( V ^ ) 212
>0, (25.35)
S(dr)4>2 /(dr)¿2 .
according to the inequality (25.25). So within a factor of f (dr) <j>2, the left-hand
side of (25.30), multiplied by A, is thus given by
= J(dr)4(D-M)*a
and
4ttJ ( d r ) f ap = J ( d r ) f aD<f>
= j{dr)(D<j>a)*cl>
= \a J ( d r ) r a<i>, (25.38)
that (j) and the <t>a obey. The result for the right-hand side of (25.30), also
multiplied by A, is
I r 2
Y (Aa - Ai)(A - Aa), (25.39)
A<*>A2 ^
The useful aspect of this inequality emerges when A[0], an upper limit to
Ai, and A2, a lower limit to A2, are both sufficiently close to their respective
referents that
A2 - \[</>] > 0. (25.42)
Then the inequality supplies a positive upper bound for A[<^] — Ai, which is a
lower bound to Ai. Indeed, Ai is now bracketed as
The complex conjugates of the expansions of these real functions are required
as well. First, let A = Ai and then use the orthonormality of the <j>a to evaluate
which is
or
1 1 1 1
r = 1 , 2 ,. > — + > =F + (25.51)
Ai Ai Ar2
(25.52)
w = ----------------- ü m w -----------------
Now we want to examine what happens to A«, the eigenvalue associated with
the (unnormalized) mode function <^a , when the boundary surface S is altered
infinitesimally by a displacement 6n(r) along the normal to the original surface.
According to the stationary property in which the boundary condition need not
be maintained, it suffices to extrapolate the original mode function into the new
region, or terminate it within the initial region. Thus we have
where
d . _
(25.54)
whereas
6 Jv (dr ) = °> (25.56)
cA <fs dS 6n e{d<j)aldn)2
(25.57)
l v (dr)<f>l
as the net variation: Any outward displacement generally lowers the value of
Aa ; any inward displacement generally increases it.
One application of (25.57) refers to situations where Xa has a known de
pendence on geometry. Then the normalization of the mode function (¡>a— the
requirement that the denominator in (25.57) be unity— can be achieved in terms
of the surface behavior of the mode function. This is illustrated by a sphere, of
radius a, where [(23.28),which refers to e = 1]
and
Rin(r) = Cinji(yinr/a), ji(yin) = 0; (25.59)
here Cin is the normalization constant left unspecified by the radial differential
equation. In this example we know that A/n = 7 ?n/a2 varies inversely with a2.
Hence, on choosing Sn = 6a, (25.57) reads
/ O X 1/2 1
(25.61)
C,n Va3/ j[ (yin) '
in agreement with (23.23).
Before pressing on let’s use (25.57) to fill a small gap in Section 19.3, where
we promised a simpler verification of the normalization constant in <¡>o{n)) the
mode function of lowest eigenvalue for an equilateral triangle. That function is
(e = 1)
OjJT . 2ir .
4>o{fj) sin — {y + r)
h
= C ^ s in — (//<, + r ) = C
a—1
. 2* f 3 1/ 2 1 \ . 2 tt f 3 1/ 2 1 V
according to (19.61) and (19.30). The dimensions of the triangle are specified
alternatively by the height /¿, the radius of the inscribed circle r = |/i, or the
base a = (2 /3 1f 2)h. Thus, the lowest eigenvalue is given equivalently as [(19.60)]
A (25.63)
dx 8n (25.64)
where the latter form exploits the threefold symmetry of the triangular sides. It
is clear from the geometry that the uniform displacement Sn needed to maintain
the triangular shape is
8n — 8r — ~ 8h. (25.65)
o
Now, the derivative appearing in (25.64) is
= A1 |«C'2. (25.67)
y= -r ¿
as stated in (19.61).
We turn from these quantitative uses of (25.57) to its qualitative side. An
outward displacement of the boundary of a cavity decreases the eigenvalue of any
mode. If the given region can be transformed in this way into another one, with
known properties, and a desired mode identified among those of the modified
region, the latter eigenvalue provides a lower bound to the eigenvalue of interest.
As a first example, consider a sphere of radius [to avoid confusion] 77, and the
enclosing cube of side 277. The mode functions of the sphere that have / = 0 are
spherically symmetrical [Ybo = ( 47T)""1/ 2], and we look for their counterparts in
304 CHAPTER 25. MODES AND VARIATIONS
the modes of the cube that are symmetrical in the three orthogonal directions
defined by the cube. Accordingly, we should have [(19.16)]
or
31/ 2 n7T
TOn > 0 . 866 ^ , (25.71)
~ ~ R
which is true, for [(23.25)] Ton = mr/R.
With cylindrical conductors, attention focuses on the two-dimensional cross
section. Let’s compare a circle of radius R with an enclosing equilateral triangle,
for which R is the radius of the inscribed circle, and with an enclosing square, of
side equal to 2R. The modes of the circle with m = 0 are axially symmetrical;
we look for the first few counterparts among the real mode functions of the
equilateral triangle, specifically those with l\ = h, say, where [(19.59)]
(25.72)
7? = 5 © 2 ( ' ? + ' ? + 4 ' ?>= i( S ' 0 !
9 /717T \ 2
(25.74)
7on > 2 ( 2 r ) ’
and
7o i# > 2 " 1 / 27t = 2.221, jo iR > 21/2?r = 4.443. (25.75)
A lower limit to 711 is provided by the first example of an unsymmetrical mode
of the square, [/ = 1 , m = 2, with a = b in (19.9)]
2 / 7T\ 2 5 / 7T\ 2
(25.76)
* ■ > ( £ ) + © - ? (5 )'
and
51/2
JuR > j ^ — 3.512, (25.77)
25.2 Iteration
A (real) mode function, say one with the smallest eigenvalue, Ai, obeys the
boundary condition, and the differential equation
where, for simplicity, we restrict this discussion to e = 1 . Suppose one has picked
some initial approximation, <£i°^(r), which satisfies the boundary condition but
not the differential equation. Is there a procedure for systematically improving
that initial choice? We begin our affirmative response by defining <^^(r) as the
solution of the differential equation
- V ^ / V ) = <^,0)(r) (25.79)
that satisfies the boundary condition. It is the first step in a process of iteration
wherein an approximation is introduced on the right-hand side of (25.78) and the
equation solved (with a change of scale) to produce an improved approximation.
The general statement of the iteration process is
[m -f n] = J (d r )^ m^ n\ (25.83)
The stated dependence on only the sum of the indices follows from the relations
Now we demonstrate that the iteration process does converge to the correct
eigenvalue and mode function, provided that the initial choice <j>^ is not or
thogonal to the desired mode function. We begin by applying the inequality
- ^(n+ 1)
(25.25) with ,<j>
[2 n] [2 n -f 1 ]
(25.85)
[ 2 n + l ] “ [2n + 2]’
Then we use the inequality, analogous to (25.23), that refers to vector fields,
or
[2n — 1] ^ __[2n]
(25.88)
[2 n] [2 n + 1 ] ‘
where
A(n+1 / 2) _ i l nJ (25.90)
1 [2 n + l] v '
is indeed obtained from (25.81) by the formal substitution n —» n + Thus,
the sequence of approximations to Ai is monotonically decreasing, but cannot
be less than Ai— it approaches a limit, fi > Ai. Preparatory to a discussion
of that limit, let us remark that one solution of the iteration equation given in
(25.80) can be displayed with the aid of Green’s function:
The only question about the integral involving the square of Green’s function is
its existence for a small region that includes the point r' = r. In such a region,
however, Green’s function is dominated by Coulomb’s potential, and the integral
of |r' — r |“ 2 does exist. Accordingly, (25.92) presents us with a bound of the
form
|<¿i"+ 1 )(r)| < C (r)[2n]1/2. (25.93)
Now consider the following series, for positive /?,
and so the function defined by the infinite summation will exist if the ratio of
successive terms in this upper bound approaches a limit less than unity,
J
1
1¡m P lim [A (n + l/ 2 )A( n + l) ] 1/2 < 1 , (25.96)
0
n- .00 y [2n]
25.2. ITERATION 307
> j { d v ) F l { - V 2 -P)4>i
Here is where we prove that /i = Ai, provided the initial choice is not
orthogonal to the actual mode function <t>\— the left-hand side of (25.99) does
not vanish. Assume the converse, that /j, > X\. Then, according to (25.97), the
function F i(r) exists for /? = Ai. But with that value of /? the right-hand side of
(25.99) does vanish. The contradiction shows that /i = Ai. And we know that
A[</>] = Ai can only be achieved with = (j>\. Perhaps the following afterthought
will be helpful. One can verify, with the aid of the completeness of the mode
functions, that the solution of the differential equation (25.98) for F\ is
Here, clearly displayed, is the first singularity that appears with increasing /?,
at ¡3 = Ai, unless this term is missing because is orthogonal to <j>\. In the
latter circumstance the first singularity— and the number to which the sequence
of approximants converges— appears at A2 unless___
How rapidly does the approximation sequence approach Ai? To answer that
let’s look at the inequality (25.49), with (j) = <^n+1^:
A(xn ) - A x
(25.104)
A(n+X/2) _ Ai - a (»+1/2)-
308 CHAPTER 25. MODES AND VARIATIONS
A(jn) - At A2
^(n +1) _ ^ — ^ ( n + l / 2 )^ (n + l) * (25.105)
The last result shows that the error of the (n + l)th iteration is smaller than
that of the nth iteration by a number that (for A2 = A2) approaches (A1 /A 2 )2 =
( 7 1 / 7 2 )4 as the iteration proceeds. Thus, the larger the ratio A2 /A 1 , the more
rapid is the convergence. We must point out here that the second eigenvalue re
ferred to in this convergence criterion can well exceed the second smallest eigen
value of the cavity. This occurs in the presence of spatial symmetry properties
that permit the decomposition of the modes into different symmetry classes. If
the initial function <j>^ possesses the particular symmetry that is characteristic
of <¡>\ so also will the successive approximations <j>^\ every member of this se
quence is automatically orthogonal to modes of other symmetry classes, and the
relevant second eigenvalue is that of a mode having the same symmetry as <j>\.
Furthermore, in consequence of this orthonormality, the method under develop
ment applies to the lowest eigenvalue of each symmetry class, independently.
These remarks are illustrated by the circular cylinder. A real mode function
has the angular dependence cosm<¡> or sinm<^, and modes associated with dif
ferent m values are orthogonal; each value of m is independent. Accordingly,
in discussing the lowest eigenvalue of the m = 0 class, A01 , the relevant sec
ond eigenvalue is not the actual next larger one, An, such that (An/ Aoi )2 =
(3.832/2.405)4 = 6.445, but rather it is A02 , with (A02/A 01)2 = (5.520/2.405)4 =
27.75.
As we have already noted, the inequality (25.41) provides a lower limit to
Ai if a reasonably accurate A2 is available. The evaluation of A 2 [(25.35)] for
<j) = ( f is
and we learn from (25.43) how Ai is bounded at the nth iteration stage:
—1/2) \(n)
A(!n) > Ai > A<"> - ^ (25.107)
a 2/ a ^ - i k
the lower limit also follows from (25.102), with n —> n —1. As always, the corre
sponding statement at the (n -f l/2 )th stage is produced by formal substitution.
We now want to recognize how the iteration process can be used to optimize
the bounds within which Ai is confined. Suppose that iteration has advanced
so far that (25.105) can be presented as
A ^ -A x
(25.108)
A(” +1) - Ai
25.2. ITERATION 309
which employs the optimum choice, A2 = A2, for that is what the process selects
(we return to this point shortly). This asymptotic ratio of iteration errors shows
that the approach to the limit is as
(25.109)
Now we use
(25.110)
A A 2n A2 - A i
A<” > * * , + « , ( £ ) - « ( £
A2 —Ai
2n
A A2 — A2 ( Ai
= A\ — -------- ; — CL (25.111)
A2 — Ai \A2
the error of the lower limit at any (sufficiently advanced) stage is less than that
of the upper limit. And one can improve matters, for the asymptotic ratio
A(n - l / 2 ) _ A(n) ^ A2
provides an internally generated estimate of A2 that can be used for A2. The
closer the latter is to A2, the more rapidly will the successive lower limits con
verge, according to (25.111). [As we shall see in a moment, with A2 = A2, the
error of A ^ is dominated by a multiple of (Ai/As)2n.] And the latter equation
also shows that too large a choice for A2 will betray itself—then the “lower
limits” will converge from above!
Before illustrating all this in a reasonably favorable situation, we shall supply
additional assurances concerning the convergence of the iteration process. We
begin by representing 4 in terms of the complete set of <j>ai
(25.114)
Ca
\ r) = T ^ « ( r ). (25.115)
310 CHAPTER 25. MODES AND VARIATIONS
and, in general,
Ca
A(n)W = £ -M * )- (25.116)
(Aq.)’
Notice that the evaluation of the integral defining [m + n], (25.83), confirms its
dependence on the sum of the indices,
Cl
\m + n\ = 'y 77—. a , (25.117)
v ( A« ) m+n
Now we have
\ («) _ [ 2 n — 1] _ ^ a { C a f A” ) 2
(25.118)
1 " [2 n ] " E „ ( C « / A » ) 2
or
\(n) _ \ I Y 2a (^a ~ A l ) ( C a / A £ ) 2
(25.119)
1 " 1+ Y , « { c a/ K Y •
As the iteration proceeds, the summations in the latter version will be dominated
increasingly by the leading terms:
in which
&= (A3 — M){C^/C\)2. (25.124)
We find that
. A3 — A2
(Sf-
A(in) Al — ------ ;— 0 (25.125)
A2 — Ai
and thus the correct localization of A2 is signaled by a marked increase (to the
extent that A3 > A2) in the convergence rate o f the lower limit.
25.3. EXAMPLE 311
25.3 Example
Consider the parallel plate geometry, 0 < 2 < a, for which the lowest mode
function is sin with corresponding eigenvalue 7 2 = 7r2/a 2. First consider the
variational bound (25.24). We must choose a trial function which vanishes at
z = 0, a. The simplest example is
SO
AͰ> =
XTdz [&(*(<» - z))]2
Jo* d zz2(a — z )2
_ 1 fp dt(l — 4< + At2)
~ a2 fg dt(t 2 - 213 + t4)
_ l i _ l £ . (25.127)
Q?“ -h.
U30 “
5
indeed, \/T0 = 3.162 is only 0.7% bigger than it. To get the next iterated bound,
we have to solve (25.79), or
d2 ,(i) , .
(25.128)
(25.129)
^ = T 2 z 4 - r 3 + T2z -
So, the second iterate is
A(1) - M (25.130)
1 - w
where
and
which gives
(i) _ 306 1 (i)
3.14181
(25.133)
aí
31 a2 ’ 7l
only 0.007% too high. As for the lower bound, let us simply take A2 = ^2 =
(27r/a)2, and then set n = 1 on the right-hand side of (25.107):
AC1/ 2) A(l)
\(i) _ \W A1 “~ Ai (25.134)
Ai - Ai - 7 rT T 7 T 7 (I) 7’
(27r)2/ a 2A\
312 CHAPTER 25. MODES AND VARIATIONS
3. Carry out similar estimates for the lowest eigenvalue of a spherical cavity.
Chapter 26
Magnetostatics
1, ' 1
£[A,B] = - J(dr) —J •A + ~—
c 4tt/jl )]■
-B. V x A + - B 2 (26.2)
where is the permeability of the medium. We now have to check that the sta
tionary principle applied to this form of the energy yields the correct equations
of magnetostatics. We are to regard A and B as the independent variables, so
the variation of the energy is
B = VxA, (26.4)
313
314 CHAPTER 26. MAGNETOSTATICS
which is equivalent to
V •B = 0. (26.5)
By making use o f the identity
® - V x 5 A = V . (<5AX®) + 5 A ‘ (V X ^ ) ’ (26.6)
and discarding the implied surface integral, we find from the vanishing of the
coefficient of <5A ,
4w
V x H — — J, (26.7)
c
a consequence of which is that only steady currents occur here:
V. J = 0. (26.8)
As appropriate to macroscopic media, we have introduced the magnetic field,
(26.9)
Thus we have recovered Maxwell’s equations in the static limit, (26.5) and
(26.7).
As in electrostatics, there is a restricted version of the stationary principle
for the energy. If we take
B = Vx A (26.10)
as the definition of B, the expression for the energy becomes
E [ A] = - J (dr) -J •A — — ( V x A ) 2 (26.11)
c 8tiV '
Regarding this as stationary under variations in A, we derive the equation
satisfied by the vector potential,
Vx Q v x a ) = (26.12)
6E = - / ( d r ) ^ i - ( V x A ) 2 = - / (dr) ^ H 2, (26.13)
F = - / ( d r ) ^ V M; (26.14)
or,
m x ( A i - A 2) = 0, (26.16)
where ni ( 112) is the outward normal to Vi (V2 ) so that 112 = —n i. The relation,
(26.15), is true for all points on the surface. Thus, when we take the divergence
of this expression, in which only tangential components of V occur, we find
ni •B i -f n 2 •B 2 = 0, (26.17)
or
n i . ( B ! - B 2 ) = 0, (26 .1 8 )
that is, the normal component of B is continuous across the boundary. [We
may regard this as a surface version of V •B = 0.]
If we include the possibility that there is a surface current, K, on the bound
ary between V\ and V2 , we must amend the energy expression, (26.2), to read
- B - V X A + Í B 2) j - J d S - K - A . (26.19)
E
= - /< * > [?J •A -f-
47Xfi
In our previous discussion of the variation in the energy, we discarded the surface
integral [see (26.6)]; this is no longer permissible because of the presence of the
boundary. Consequently, there is a new contribution to the variation of the
energy arising from the occurrence of the interface,
6E = ¿ A 2X ^ ) - t d S ^ K 'S A
<a '*Í7
316 CHAPTER 26. MAGNETOSTATICS
=
74/ JC| ÓAiXHi
/ a6 ni •----- ---------- b n2 •
47T
ihX
6A2XH2
H j f n2x H 2 , 1 ^
1 T^
;------------- K •0A
47T C
£A
¿ A i t = S A 2t = ¿ A t . (26.21)
4-tt
m x H i + n 2x H 2 + — K = 0, (26.22)
c
or
4-tt
ni X (H i - H 2) + — K = 0. (26.23)
H 2 = — B 2 -H. 0, (26.24)
P2
ni x H i = 0. (26.25)
This is the same condition satisfied by the electric field at the surface of a
conductor.
V x Q v x a ) = Í p í, (26.26)
This equation can be simplified by using the fact that it is invariant under a
gauge transformation [see (9.68)],
Because of this gauge freedom, we can usually choose some particular gauge to
simplify the problem at hand. In the present situation, the convenient choice of
gauge is one for which
V •A = 0, (26.29)
V •A 0 0. (26.30)
V . ( A o + VA) = 0, (26.32)
—V 2A = V - A 0. (26.33)
—v 2a = i l l , (26.34)
c
(26.35)
* ' > - ; / < * ' ) I,1- rV
where we have used (26.8), and the fact that the current distribution is localized.
318 CHAPTER 26. MAGNETOSTATICS
Once we have the vector potential, we can compute the magnetic induction'
B:
B = V X A = i v x [ (dr1) , J ^ /,
c J |r — r|
= iy<<ít')J(r')Xj¡í f I F . (26.38)
where R is the position of the particle, which produces the magnetic field
v e(r-R )
(26.40)
c |r — R|3
j+ j ¿ í A=v*
Show that this leads to (/i = 1)
V 2B = ^ B ,
which implies the Meissner effect, that a uniform magnetic field cannot
exist inside a superconductor. is called the London penetration depth,
and has the typical value of 10""5 cm.
Chapter 27
Macroscopic Current
Distributions
The expression for the magnetic induction, (26.38), a distance p from the wire,
is then reduced to
(27.2)
t I
319
320 CHAPTER 27. MACROSCOPIC CURRENT DISTRIBUTIONS
For a long wire of length 2L, L^> p, the integral occurring here is
fL t 1 [L dz' ( L \
(27.3)
J-L \/p
\fpi2 + zf2 Jo \Zz'2 + p2 \ P )
the gradient of which is
, L \ Vp 2
V 2 In — b constant = —2-----— — p. (27.4)
P ) P P
where p is a unit vector in the radial direction. The magnetic field produced by
this wire is therefore
B = — (n x p ), (27.5)
cp
which is concentric with the wire, and in the sense given by the right hand rule,
that is, if the thumb of the right hand points in the direction of current flow n,
the fingers curl in the sense of B.
The force exerted on a current distribution by a magnetic field is [see (7.7)]
F = ^J(dr)3xB, (27.6)
that is, the force is attractive. If the currents flow in opposite senses, the force
is repulsive.
(27.11)
A -> A + VA (27.12)
(27.13)
since we can integrate by parts and use (26.8).] Introducing the explicit form
for A , (26.35), we can write the energy in terms of the current density alone,
(27.14)
which is analogous to the electrostatic result, (12.10), or (1.1), except that its
sign is opposite. For the case of two current distributions,
the energy expression contains self energies as well as the mutual interaction
energy. We are here interested only in the latter, which is
(27.16)
as it is the sole term that contributes to the force, (27.10). For straight wires,
this becomes
(27.17)
where we have used the restriction L/p 1. The force can now be calculated
from (27.10), or, since E depends only on /?,
f = - | A £|a (27.20)
F_ _ 2h h p
(27.21)
2L ~ ' c2 p’
= (27.22)
1
(27.26)
2c
A —►AA , A = constant,
1For example, Faraday’s law, (1.70), gives the emf in the ¿th circuit as S% = j> d ( i • E =
However, such circuit analysis goes beyond statics, dealing as
it does with alternating currents. In this case, as noted above, the energy associated with an
inductor L is positive, E¿, = ^ L I 2, and thus an L C oscillator, for example, can be understood
in terms of the energy interchange between that in the inductor, and that in the capacitor,
E c = | Q 2 /C.
324 CHAPTER 27. MACROSCOPIC CURRENT DISTRIBUTIONS
where (¿**±1 ,2) are the cross-sectional area elements of the two filaments,
and pi 2 is the distance between the two filaments. The integral over In p
may be carried out by noticing that —2 In p is the potential of a unit point
charge at the origin in two dimensions (two-dimensional Green’s function),
so that
is the potential due to a uniform charge distribution. But the latter must
satisfy the two-dimensional Poisson’s equation [cf. (16.16)], so for p < b,
<t>= A + B In p — 1rp2,
By doing the integral at the center of the wire, determine the constants,
and thereby find the self-inductance
Magnetic Multipoles
1 __ 1 r*r'
(28.1)
|r —rx| r ^ r3
in the expression for A , (26.35). The resulting expansion for the vector potential
is then
which is analogous to the expansion for <£, (22.3). From current conservation
for steady currents,
V •J (r) = 0, (28.3)
325
326 CHAPTER 28. MAGNETIC MULTIPOLES
To evaluate this integral, we again use (28.3) and consider the integral
3^ /»
^ ^ 1 (dr) (#? Ji "h Jj) j (28.6)
Ar= l J
K( \ ^Xr (28.9)
AW = — >
(28.10)
**= ¿ / ( dr)r><J(r)'
[For a point charge, the current density is given by (26.39), so the magnetic
dipole moment is
(28.11)
B = Vx
since
- vV i ---
= - Q> (28.13)
r ró
and consequently
0 = MX ( v x ¿ ) = v ( ^ ) - O . . V ) i . (28.14)
28.2. ROTATING CHARGED SPHERICAL SHELL 327
In general, this is only the leading contribution, since there are higher multipoles.
We will not, however, explore these further here.
V - (j o X r;. (28.17)
<28-2°)
The expression for the vector potential, (26.35), becomes
e wXr '
A(r ) = \ f d S ? (28.21)
Aira2 |r — r'
Therefore, only the / = 1 term in (28.22) will contribute to the integral (28.21),
1 j=i r r « r ' 1
- — r -r (28.24)
| r-r'|
= < 2 8 -2 5 >
Using spherical symmetry, we easily evaluate the integral over the dyadic to be
(28.27)
A = ¿ (“ xr)s5B*r’
where, using the result of (6.22), we identify the magnetic field B as
2e
(28.28)
ea 1 w X r _ /xXr
(28.30)
3c r 3 r3 ’
which, upon comparison with (28.9), allows us to identify the magnetic dipole
moment as
_2
ea*
LL — ---- U?. (28.31)
^ 3c
The magnetic field B is then given by (28.16).
Notice that B is discontinuous across the spherical shell because there is a
surface current density. The values of B just outside and just inside the surface
(n = outward normal = r /a) are
so the discontinuity in B is
n •(B + — B _ ) = 0, (28.35)
Anr
n x ( B + - B _ ) = — K. (28.36)
c
Obviously, (28.35) is satisfied. From (28.36) and (28.34), we calculate the surface
current density,
K = (t v = (u;Xn)a = - ^ w X n . (28.38)
47ra2 47ra
where dv' is a directed line element tangential to the wire, in the direction of
the current flow. Now the vector potential, (29.1), becomes
<»*>
which implies for the magnetic induction,
B= civ * /F = 7 f <2£U>
dr'
331
332 CHAPTER 29. MAGNETIC SCALAR POTENTIAL
V -^ V x a (29.7)
The identity
a x ( V 'X V ) = V '( a ♦V ) - (a •V ') V , (29.9)
allows us to rewrite (29.8) as
(29.10) reduces to
/ d r 'x V = / ( ¿ S ' . V ') V . (29.12)
Jc Js
We will apply this result to rewrite (29.5) for which
1
V = V' -V -, (29.13)
|r - r'| r —r n )
which satisfies the conditions (29.11) as long as r / r', that is, at points outside
the wire. We therefore find
H = —V ^ m, (29.14)
where the magnetic scalar potential, <j)m) is
(29.15)
c Js , |r - r'l
333
where S' is a surface bounded by the current loop. According to (1.17), the
surface integral,
V x H = 0. (29.18)
± ^ J , YES,
= { 0, NO,
(29.20)
where YES means the wire passes once through the surface S, bounded by
the path C, (the ± sign refers to the relative orientations of dS and J) while
NO means the wire does not pass through the surface. Some examples of this
are supplied by the illustrations in Fig. 29.2. Therefore, contrary to our naive
expectation,
YES,
(29.21)
0, NO,
334 CHAPTER 29. MAGNETIC SCALAR POTENTIAL
or
f ± 4 tr, YES,
(29.22)
I 0, NO,
which means that <¡>m is a multivalued function, as required by the fact that
V x H is not zero everywhere. The discontinuity found in (29.22) corresponds
to the fact that when one crosses the surface S' defined by the current loop and
used in the evaluation of the solid angle in Q, there is a change of 4ir.
Very far away from the current loop, the solid angle subtended by it is, if
we assume that the points of S' are localized in the vicinity of the loop
ÍÍ = S' = J d S 1, (29.23)
H = -V (29.24)
which upon comparison with (28.12) identifies the magnetic moment of the
current loop to be
A* = I s . (29.25)
We obtain the same result if we use the definition of the magnetic moment,
(28.10),
where we have used (29.10) to evaluate the line integral, which is also obvious
geometrically.
I 2z 1
H0 = —
C p \/{a + p)2 + z 2
I 2 a2 _ ,,2 _ 2
H; ) + 7T
m (a - pL) 2 +, z 2
29.1. PROBLEMS FOR CHAPTER 29 335
where the complete elliptic integrals of the first and second kind are, re
spectively,
pn/2 i
m = / w h L2, 2 ,
Jo y\ — k2sm¿ ^
pn/2 / ... ..
E{k) = / difty 1 — k 2 sin2 -0,
Jo
and
p = 4a?
(a + p)2 + z2 '
3. By taking the large distance limit of the result found in Problem 2, de
termine the magnetic dipole moment of a circular current loop. Compare
the result with (29.26).
Chapter 30
Magnetic Charge II
V . B = 0, (30.3)
B = V x A. (30.4)
j ) d S ' V x A = J (d r ) V « V x A = 0, (30.5)
while (30.1) implies for a closed surface surrounding the magnetic charge
We now want to find a vector potential that satisfies (30.4) almost everywhere.
The simplest possibility is that this equation fails to hold on a line, which we
may take to be the + z axis. We apply Stokes’ theorem in the form
337
338 CHAPTER 30. MAGNETIC CHARGE II
(30.9)
►
II
•e-
where
g 1 + cos 6
A <t>~ • UQ ♦ (30.10)
r sin
The structure of the singularity on the 0 axis is now isolated by taking the limit
0 - + 0,
6 dr * A = f d S '. V x A = -4 * 0 , (30.11)
Je Js 1
where C' is an infinitesimal circle about the z axis and S' is the enclosed area,
as shown in Fig. 30.2. Since (30.11) shows that ( V x A ) ¿ has the singular
ity —4irg6(x)S(y) on the -fz axis, we conclude that the magnetic field can be
expressed everywhere by
B = V x A -f 4irg6(x)6(y)r)(z)z, (30.12)
(3013)
This result can be confirmed by noting that B has the correct divergence,
C'
z = 0
or, alternatively, that (30.6) is consistent with (30.5). The vector potential
(30.9) is an example of a class of potentials that yield the correct magnetic field
except for a one-dimensional set of points, a curve. On this curve, called a
string, A is singular, whereas the magnetic field is regular, being the curl of A
plus a compensating singularity on the string.
g nXr
A = ----------- -— .
r v — n •r
j ) dr •A ,
where C is a circle about the z axis as shown in Fig. 30.3. What is the
limit of this line integral as 0 —►0, 0 7r?
CHAPTER 30. MAGNETIC CHARGE II
^ „ 1 <9 4 tt.
V x B = 7c at
*7 ® + — (31.1)
c j,
V •E = 4717?, (31.2)
-V x E = ~ B , (31.3)
c ot
V •B = 0, (31.4)
where p is the charge, and j is the current density, and we have assumed that
no magnetic charge is present. Notice that the local conservation law
c\
v-j + d tp = 0’ (3L5)
is not an independent statement, but is derivable from (31.1) and (31.2).
To solve Maxwell's equations, we first recognize that the last two equations,
(31.3) and (31.4), make no reference to charge or current, and they can be
identically satisfied by introducing potentials through the definitions
B = V x A, (31.6)
E = _1JLA - V ¿ . (31.T)
As we have observed previously, in Section 9.5, the potentials A and (j) are not
uniquely defined. Since the magnetic field is the curl of A , it is unchanged when
a gradient is added to A ,
A A -f VA (31.8)
342 CHAPTER 31. RETARDED GREEN’S FUNCTION
where A is an arbitrary function. In order that this new choice of vector potential
not alter the electric field, (31.7), it is necessary to simultaneously replace the
scalar potential by
+ <31-9>
This new set of potentials, (31.8) and (31.9), is as acceptable as the original
one since only the fields B and E are physically measurable quantities. This
arbitrariness in the choice of potentials is called the gauge freedom of the theory,
while the corresponding transformations are called gauge transformations. In
the following, we will exploit this freedom in the process of solving the differential
equations for the potentials.
Upon substituting the constructions of B and E in terms of potentials, (31.6)
and (31.7), into the first set of Maxwell’s equations, we find, from (31.1),
_ /_ AN 1 d ( _ , 1 d \ 4 tt.
V x (V x A ) = - - - - ^ A j + T J, (31.10)
1 d2 \ . _ /_ . 1 d A 4tt.
- (v 2- ? » J a = - v (v -a + ; » ^ + t j’ <3111>
This is a pair of coupled second order differential equations for A and <£, which
may be simplified by utilizing the gauge freedom in defining the potentials. The
two most convenient and common choices of gauge are discussed below.
V •A = 0. (31.13)
That we can always make this choice was shown in Section 26.3. In this
gauge, (31.11) and (31.12) reduce to
where
m
a 2¿ = ------------
1 92 (31.16)
c2 dt2
is the d’Alembertian. (Jean d’Alembert’s Traite de Dynamique was pub
lished in 1758.) The equation for </>, (31.14), is just the same as that in
electrostatics (hence the origin of the term “Coulomb gauge” ) so that </>
is, in principle, known. The structure on the right hand side of (31.15) is
31.2. GREEN’S FUNCTION IN THE LORENTZ GAUGE 343
V
47t at
where the last equality follows from charge conservation, (31.5). This
relation also entails the consistency of the choice of the radiation gauge in
that if we set V •A equal to zero at one time, it remains zero for all time,
since
J_d_
- D 2( V - A ) j - V (31.18)
47t dt
2. The Lorentz gauge condition is a relation between vector and scalar po
tentials,
1
V * A + - J ^ = °- (31.19)
In this gauge, the equations for A and t}> have the symmetrical form,
- □ 2A = — j, (31.20)
The consistency of this gauge choice again follows from the fact that charge
is conserved,
- D2(v -A + ^ ) = T ( v -j + t ) = ° - <3L22>
This Green’s function, G (r —r', t —t'), is a function only of relative positions and
times because of translational invariance in unbounded space. Since <¡) satisfies
(31.21), this Green’s function obeys the differential equation
To solve (31.25), we will analyze its time dependence by making use of the
exponential representations (Fourier transforms in time)
= r (31.27)
J—OO ^
/
°0 J
r _ r'), (31.28)
-oo ^
In the static limit, u —> 0, (31.29) reduces to (31.26), the solution of which is
Coulomb’s potential, (13.3):
G a ,= o (r -r ,) = 1— [— r (31.30)
|r — r'|
Since Geo depends only o n r - r ' , we may set r' = 0 , without loss of generality
in the following discussion. Also, since we are now looking for a spherically
symmetrical solution for Gw, it is natural to use a spherical coordinate system
in which the Laplacian here reduces to
V2 (31.31)
1 . — ( r 2± U)
+ GU r) = 0, (31.32)
r 2 dr \ dr
subject to the boundary condition that there is a point charge at the origin. The
consequence of this requirement is most conveniently extracted by integrating
(31.29) over a sphere S of vanishing radius r 0 about the origin,
or
= 1. (31.34)
~ r 2T rG“ V
[We have noted that the u>2/c2 term in the differential equation does not con
tribute to the integral since
(31.35)
c¿r G w ~ -r,
— as r 0,
31.2. GREEN’S FUNCTION IN THE LORENTZ GAUGE 345
dn -
2 d (31.38)
ri G uj — ^~T~9u ~~ Qun
dr dr
1 d ( 2 d \ _ 1 d2
(31.39)
r 2 dr \ dr / r dr2^u
gu ~ e±lW/ c, (31.40)
while from (31.28) we now obtain the space-time form of the Green’s functions,
/
oo
-oo 2 tt | r-r'|
What is implied by the use of the + or — sign in (31.44)? The choice of the +
sign leads to the retarded Green’s function,
This means that the signal propagates with the speed of light c from the source
(at time tf) to the observer (at time t); the effect occurs later than the cause.
If we pick the — sign, we obtain the advanced Green’s function
<f>(r,t) = (31.49)
A (r ,t) = (31.50)
<l>(r,t) (31.51)
= / ( * ')
A (r ,t) (31.52)
= /<*'>
These results are elementary generalizations of the potentials for electrostatics
and magnetostatics, but now reflecting the finite propagation speed of light.
y ■■ 6 ......... , A =
(z - vt )2 + (l - ^ ) ( x 2 + J/2) C
- □ 2B = — V x j .
C
^ ^ k - r ( i ,) | - [ r - r ( </) ] ‘ 2^ i ’
A (M ) =
t* = t -
KOI
- ( ¿ ¿ k /)G(n)(X i>•••.*«) =
\k= 1
•••«(*»)•
G^n \ x u . . . , x n ) = / f [ ^ -e x p n 1 ,2 •
J m=l ¿1T \ m=l /
348 CHAPTER 31. RETARDED GREEN’S FUNCTION
/
oo
•OO
Check this explicitly for n = 4, and then in general, from the explicit
answer in the previous problem.
4 ( ± y n=
and is explicitly
1 1
G =
- ( v 2 - ¿ ¿ i ) D + (r » o = ¿(r )^(c<).
where r = (# 1 , a?2 >« 3 ). Starting from the solution in Problem 7, show that
(a)
1
D+ = 1
47T2 r 2 — (ct ) 2 + «e ++o
and
(b)
where the two terms here are the retarded and advanced Green’s
functions, respectively, apart from an overall factor of 1 / 47T.
31.3. PROBLEMS FOR CHAPTER 31 349
(dk)d 4 e’'(k-r+*4®4)
f (dk)dk
G
~J ( 2 »r)) 4 k 2 + k2 ’
jb4 -H. e— ( - ^)
to obtain
(dk)(dw /c) e*(k *r-a,<)
f (dk)(d
J ( 2V)4 k2 - 7T - e— + 0
s-«|k|c|<|
r °°
° ^ 1 _____ 1
7-0 0 c 2 ir k 2 - K - ie
and consequently
Cr t\ - i f W i ¿(k»,-cikiitn
^ 7 ( 2 7T)3 2 |k|e
and
V2 - ^ | O + (r,«) = 0
f°
1 duj t°
e i\u>\r/ c
D.
(t' , ) = c L 2w( 4ixr
to derive the asymptotic form of the potentials in a Lorentz gauge [cf. (31.51)
and (31.52)],
where
n = - (32.4)
351
352 CHAPTER 32. RADIATION— FIELD POINT OF VIEW
is the unit vector in the direction toward the observation point. In the above
equations, the n * r ' term in the expansion of l/|r — r'| has been deleted since
it gives rise to a 1 / r 2 term in the potential, while it has been retained in the
expression for the time of emission, t\
P « t — - -f ~n •r t = tr . (32.5)
c c
The last term in tr reflects the finite amount of time it takes radiation to propa
gate across the source, which can be significant if the source distribution changes
rapidly, or, more precisely, when a typical frequency of oscillation of the source
distribution is of order c/a.
The fields at large distances can now be calculated by substituting (32.2)
and (32.3) into (31.6) and (31.7), and by using the evaluation
r 1 n i l . /x d_
V - + -n * - + — n x (n x r ) l/(* r )
c c c cr dt r
n d
(32.6)
c dt
We see that because of the appearance of r in the time dependences, the fields
behave as 1 / r rather than the behavior 1 / r 2 characteristic of statics, and in
particular, for r a, the field strengths are
These two terms in (32.8) can be further combined by using the local charge
conservation condition (31.5),
where we have used the fact that the divergence operator in (31.5) acts only on
the spatial arguments of j, while V 7 in (32.9) also differentiates the r' depen
dence of tr , where from (32.5) V 't r = n /c . The first integral in (32.8) may then
be simplified through the use of
since the charge distribution is bounded, and the remaining terms involving the
time derivative of j can be combined by means of the identity
n (n •V ) - V = n x ( n x V ) (32.11)
to read
■ \n 1 f , , ,v 1 d , ‘
E(r,<) = nX l^ -X - J ( d r ) - —3(r ,tr) = -n x B (r ,i). (32.12)
32.2. ANGULAR DISTRIBUTION OF RADIATED POW ER 353
Figure 32.1: Electric and magnetic fields for a wave propagating in the direction
n.
We observe that, far from the source distribution, E and B are perpendicular
to each other,
E = — nxB, B = nxE, (32.13)
are perpendicular to the direction of propagation n, and have equal magnitude,
E 2 = B 2. (32.14)
These are the same characteristics seen in Section 3.4, where we considered the
propagation of electromagnetic waves along a single direction, in terms of the
flow of energy and momentum. (See Fig. 32.1.)
S= ExB, (32.15)
47T
n •S = — (nxE)*B = — B 2 (32.16)
47r 47T
Rather than the energy crossing an element of area dS, we would instead like
the energy radiated into a solid angle dfi,
dñ = (32.17)
since the latter measure is independent of how far away the observer is from the
source. Therefore, the amount of energy radiated per unit time (the power) per
354 CHAPTER 32. RADIATION— FIELD POINT OF VIEW
dP
n X J (dr') j(r',U) (32.18)
dll 47TC3 dt
while the total power radiated is obtained by integrating this over all solid
angles,
p=Jda(§)- <3219>
This finite energy flow at large distances is a consequence of the 1f r behavior of
the fields, which, in turn, arises from the time variation of the current density.
For this situation, the time of emission, (32.5), of the radiation may be approx
imated by
/ w ( * 0 jV , M = e^ ( ^ ) > (32.22)
dP_
díl
(32.23)
where 0 is the angle between the direction of observation n and the direction
of the acceleration at the emission time te. (See Fig. 32.2.) Evidently there is
32A. DIPOLE RADIATION 355
<32 24)
we obtain the total radiated power:
2e 2 f dv\ 2
P for - < 1 , (32.25)
3 ^ [~dt J c
d (i) = J ( d r ) r p ( r ,t ), (32.28)
356 CHAPTER 32. RADIATION— FIELD POINT OF VIEW
we recognize (32.26) as the second time derivative of d (/). Therefore, the angular
distribution is given by
and (32.23) and (32.25) are recovered, as expected. However, for a system of n
charged particles, the electric dipole moment is
n
d = ^ e ar a, (32.32)
a= 1
so the power radiated is not additive, but exhibits interference effects:
= d ( t e) - nX//(ie), (32.34)
where /x is the magnetic dipole moment, (28.10). [Here we have neglected the
contribution due to
¿ ( r 'j + j r ') (32.35)
from ( 2 2 .6 ), since the unit dyadic evidently does not contribute to the radiated
power because n X n = 0.] The angular distribution of the radiated power is
therefore more accurately given by
dP 1 , ...,0
d ñ = 4 í ? [“ x < d ~ n x , ‘ )1
= [(nxd)3 + (nxjj)2 + 2n >(dx/i)], (32.37)
where the last term represents interference between d and fx, which, since it
depends linearly on n, does not contribute to the total radiated power:
P = ¿ [ ( d )2 + ( ü n (32.38)
B ~ -4r-nx(d-nx/i), (32.39)
c1 r
E ------n x B ~ - " n x ( / ¿ + n x d ) . (32.40)
Note that these results are invariant under the replacements E —> B and B —►
—E together with d /x and /x —►—d, which is a manifestation of the symmetry
discussed in Chapter 2. We here have an indication of the connection between
the directions of the electric field for electric and magnetic dipole radiation.
V •A = 0. (32.43)
The electric and magnetic fields are obtained from these potentials by the rela
tions (31.6) and (31.7). In order to make contact with what has gone before, we
consider the fields at large distances, which are those of interest for radiation.
The solution of (32.41) is the Coulomb potential,
<j>~ - , (32.45)
r
with e the total charge. Since the gradient of this is inversely proportional to
the square of the distance,
-V<j> ~ ^ r , (32.46)
we can neglect the scalar potential in computing the radiation fields, which
decrease only as 1/r. Therefore, the vector potential alone determines the ra
diation fields:
1 <9A
E (32.47)
c dt ’
B = V x A. (32.48)
We also note that the gauge condition (32.43) enforces the transversality of these
fields. That is, as a consequence of V •A = 0, we recover the scalar Maxwell
equations outside the sources,
V •B = 0 and V - E = 0, (32.49)
¿ = _ V 2 4^ > (32.51)
9 4 tt / VV\ .
(32.53)
- DA= c l1- V » ) - j’
which makes the radiation gauge condition (32.43) transparent. The solution
to (32.53) may be obtained from that of (31.20) by applying the operator
VV
1 - (32.54)
V2
to (31.52):
At large distances, by making use o f (32.1) and (32.6), we have eifectively the
replacement
v - - ” s <32-56>
so that the operator 1 — W / V 2 can be replaced by
VV ( _ s JL) ( _üJL)
_ V _
1 - c dt) \ c dt)
(32.57)
V2 (_ “ A )2
V c dt)
The resulting electric and magnetic fields arq precisely the same as those found
in the Lorentz gauge, (32.7) and (32.12).
t = v = \e .
2
Show then that the power radiated, averaged over one cycle is
^initial — 10 cm.
(This instability was one of the reasons for the discovery of quantum me
chanics.)
4. Derive the alternative form for the angular distribution of radiated power,
(32.18)
dP _ 1
dD 4irc
Chapter 33
d / e2+ b 2
—j . E = + V (33.1)
at 8w ( ¿ exb)'
which is the local statement of energy conservation, (3.6). When (33.1) is inte
grated over a large volume enclosing the charge and current distributions, the
conservation of total energy follows:
and discarding total time derivative terms, which are not associated with ra
diation. From this point of view, we need to know the electric field inside the
current distribution, in contrast to the previous discussion, in which we com
puted the radiated power by evaluating the fields far from the source.
361
362 CHAPTER 33. RADIATION—SOURCE POINT OF VIEW
where we have omitted the second term on the right-hand side of the first line
because the total charge e is conserved,
where we have used (32.27) and (32.28) for the electric dipole moment d. The
contribution of the 1 /c 3 term in (j) to E can be simplified as follows:
-V ¿ /( d r 'X r -r y J L r t t ',» )
(1 ) (33.7)
The expression for the energy transfer, (33.3), becomes, in this approximation,
3 3 .3 . H A M IL T O N IA N 363
(33.8)
(33.9)
(33.10)
where we have again used (32.27) and have set aside total time derivative terms,
which do not contribute to the radiation. The power radiated, (33.10), is the
same as that found in the preceding chapter, given by (32.30).
33.3 Hamiltonian
As a byproduct of this source approach we may identify the total electromagnetic
energy of the system, to order 1/c2, by comparing (33.2) with the total time
derivative terms in (33.8),
where the first two terms have the form of the electrostatic and magnetostatic
field energies. However, the sign of the current-current interaction is opposite to
that of (27.14). The resolution of this apparent discrepancy requires the third
term in (33.11), which we rewrite by means of the local charge conservation
condition, (31.5):
364 C H A P T E R 33. R A D IA T IO N — S O U R C E P O IN T O F V I E W
which, when combined with the first two terms in (33.11) yields
(33.13)
For point charges, the charge and current densities are
so the terms in (33.13) referring to the mutual interaction of the particles are
Va'Vt, t ( r ai, . V a ) ( r a b ' V b )
Afield = I2J^2 tr ab
t + ¿4 c 2 Y^ , e « e i I 3 (33.16)
aj£b a^b
. ah ' ab
where
= ra - r6. (33.17)
To this we must add the particle kinetic energy, from (10.11),
¿particle = ^ m a C2 ^ ( l - ) - 1^
V—s f\ 2 ^ va \ va -t (33.18)
« E 2 m“,'« + 8 mV + - ’ ^2 < 1 ’
a ' '
Pa = (33.21)
dva
To determine the Lagrangian, we substitute (33.21) into (33.20),
Note that the particle term in (33.23) agrees with the expansion of (10.20). The
canonical momenta are given by (33.21),
Pa = m ax a +
, ^1r n a —vl va
2 ~ + ^ „2 ^ 2 e ae i
V» raj(rat •vt)
(33.24)
bjia f*ab ' ab
where we have consistently kept terms up to order 1/c2. This form of the
Hamiltonian is appropriate for small systems and has application in both atomic
and nuclear physics. It is called the Darwin Hamiltonian when it is applied
classically, and the Breit Hamiltonian when it is applied quantum mechanically
(with an accompanying re-expression in terms of Dirac matrices).
From the general discussion given in Chapter 9, the canonical momentum
can be expressed in terms of the vector potential
p0 = x a- l / a + ^ A ( r . , < ) , (33.26)
V 1 - v l /c 2 C
We notice that this vector potential satisfies the radiation gauge condition
(31.13):
vj •V
|r-n|
iv 6 .V (V 2| r -r 6|) = 0, (33.28)
366 C H A P T E R 33. R A D IA T IO N — S O U R C E P O IN T O F V IE W
since
V 2|r —rj| = (33.29)
|r-rt|’
and, as required, it also satisfies the differential equation (31.15) to order 1 /c 2:
- □ 2A k - V 2A
4»v,«(r - r.) + j V ( v , . V jT T ^ j) ]
- Eb ?c
4 ir .. 1 9 #/ x
(33.30)
which has the form of the magnetostatic field energy, (27.11), which is therefore
correctly given by (33.11). As we commented in Chapter 27, because of the
negative sign in (33.31), “like” currents attract each other.
Models of Antennas
V .j = — J = i ° for (34.2)
dz z \ # 0 for z = ± t.
<343)
we see that (3 4 .2 ) implies that there is an oscillating charge density at both
ends of the antenna. In any realistic model, Jz will depend on 2 . We lack this
dependence since our model assumes that the antenna is fed at every point along
its length. Even though this model is oversimplified, it possesses many of the
significant characteristics of a real antenna. To compute the power radiated,
(32.18), we evaluate the integral
' , 1 T .> A
/
d ( r
+ - n .r j
1
( * ') - d i Jl (r c
LÜ ,J/2 (' r 1 , A
= - I / d z' COSU) 1t ------ h - z cos# 1
c ,J—1/2 \, c c J
W/2 / r\ /'u>zf ü
= -/ / d z' COSCJ t ---- COS ---- cos 6
C ,J - l /2 \ c) \< c
367
368 C H A P T E R 34. M ODELS OF ANTENN AS
where, as indicated in Fig. 34.1, 6 denotes the angle between the direction of
observation and the antenna. The angular distribution of the radiated power,
at the observation time t, is then
(34.7)
Figure 34.2: Radiation pattern produced by a short antenna, / < A. In this and
the following figures axial symmetry about the 2 -axis is to be understood.
dP_ 12 / irl\ 2
(34.12)
^ g=x/2 2'7TC \ \ J
sin2 (y cos 0)
(34.13)
cos2 0
holds true for all angles. The resulting radiation pattern may be alterna
tively derived from the dipole radiation formula, (32.29), which is appro
priate to a small system:
2
dP _ 12 nl
sin2 0. (34.14)
df2 ~ 27tc T
2. A > /.
When y < 7r, the argument of the factor sin2 (y-cos0) goes from 0 to
something less than w when the angle 9 varies from “ to 0. Therefore,
the only angles at which the power radiated vanishes are 0 and 1r, so the
radiation pattern has a single lobe. (See Fig. 34.2.)
3. A < / < 2 A.
When 7T< y < 27Tthere is an additional zero in the radiated power at the
angle 9 — cos""1 ( j ) . Consequently the radiation pattern exhibits both a
main lobe and two side lobes. See Fig. 34.3. Evidently, as //A increases,
more and more side lobes appear.
370 C H A P T E R 34. M ODELS OF AN TEN N AS
Figure 34.3: Radiation from an antenna of intermediate length, A < / < 2A.
This diagram is meant to be understood schematically only. Actually, the side
lobes are far smaller than indicated.
f
I2 sin2 0 sin2 ( y cos 6)
27rsin 9 dO
27TC cos2 0
p ,*/2 sin2 ( j - s in x )
= — / cos x dx cos X
C J —7T/ 2 sin2 x
212 sin2 z
1- (34.15)
c ( tt/ / A ) 2 J
l =i
_
(34.18)
z Jo t 2
Here we observe that the total radiated power, (34.17), increases linearly with /,
while the power radiated in the direction perpendicular to the antenna, (3 4 .1 2 ),
is proportional to /2; that is, as the length of the antenna increases, a larger and
larger fraction of the radiated power is concentrated near 0 = 7r/ 2 .
To see how much energy is radiated into a very small angular range near
X = 0 (or 6 = 7r/ 2 ), we consider the power radiated into the main lobe by a
long antenna
0<X< j <1. (34.19)
3 4 .2 . C E N T E R -F E D A N T E N N A 371
Following the same procedure used to obtain the total power radiated, (34.15),
we find for the total power radiated into the main lobe
(34.20)
The fraction of the energy radiated into the main lobe is obtained by taking the
ratio of (34.20) to (34.17):
(34.21)
x coso;
coso; (t — £)
(?cosd)
21 kl
cos cos ■ (34.23)
sin2 $ I V2 J 2
The power radiated into a given solid angle is then given by (32.18), or
dP 1 4/2 cos2 ui kl \ kl
12
sin2 6 cos ( — cos 9 J —cos — (34.24)
dQ, 4 it c sin4 6
dP 1 kl A
— cos 9
«I2
—cos —
COS (34.25)
dO, 27tc sin2 9 2 J 2 J
dP __ (J/2 ) 2 / tt/
y ) sin2 $, (34.26)
dQ 27TC
which may alternatively be derived from the dipole radiation formula, (32.29).
See Problem 34.1.
The angular distribution of the power radiated, d P /d Q ,) vanishes whenever
47rn\
n = 1,2, — (34.27)
IF J
Thus side lobes appear in the radiation pattern whenever kl passes through a
multiple of 27r. Plots of d P /d Q , are given in Figures 34.4 and 34.5. It will be
noted that now as //A —►oo, substantial energy is radiated into the extreme
side lobes, increasingly near the z axis, the antenna direction. See Problem 34.5.
This is very similar to the radiation pattern produced by impulsive scattering-
see Chapter 37. Therefore, such an antenna is most useful in the half- or full-
wave regime.
Figure 34.4: Radiation pattern produced by center-fed antenna for kl = 7r. This
is called a half-wave antenna because / = A/2.
Prove that this current does not radiate if the propagation constant k
exceeds k , the intrinsic wavenumber of the external medium, supposed to
be uniform and infinitely extended. Verify that radiation does occur if
k < & , and that the time-averaged power radiated per unit length of the
wire is given by
Spectral Distribution of
Radiation
A (M ) = J(dv') dt (35.2)
•r'| c
In deriving these results, we had used the spectral representation [cf. (31.44)]
If we now reinsert (35.3) into (35.1) and (35.2), and carry out the i' integration
by introducing the temporal Fourier transform
f dt e 7(0 = /(*),
/ A(jJt (35.4)
du -icot
l-oo ( 7(w) = /(<').
2 tt
(35.5)
/
piw\Y—r |¡c
^ ú id r—r 'l/ c
(35.6)
375
376 C H A P T E R 35. S P E C T R A L D IS T R IB U T IO N O F R A D IA T IO N
/
i w \ r - r'|/c i
(35.7)
We observe that if f(t') is a real function of t' , its Fourier transform, /(u>),
satisfies the condition
/(«)* = / ( - " ) ; (35.8)
consequently,
/ ( « ) * /( « ) = |/(^)|2 = /(-« )/(« ) (35.9)
is a real positive number, symmetric under the interchange w —►—u>. This
implies that the algebraic sign of u is not significant, since only its magnitude
enters into physical quantities.
Let us focus our attention on the radiation fields, far from the sources.
Following the procedure given in Section 32.1, in particular, using the expansion
(32.1) for |r —r'| in (35.6) and (35.7), we obtain the asymptotic expression for
the potentials, in terms of spatial Fourier transforms,
Ju>r/c p
<¿(r,w) ~ (dr')e-iWn' r /cp(r',u), (35.10)
iu r/ c p i
A(r,w) - - —
e / ( d r O e - ^ '- ' / ^ j ^ » , (35.11)
k = ~n (35.12)
the propagation vector, in terms of which the potentials are written as Fourier
transforms in space and time,
(35.13)
The corresponding field strengths can now be computed from the time Fourier
transforms of (31.6) and (31.7):
ÍÜ
E(r,u>) = i—A(r,w) —V^(r,u>), (35.15)
c
B(r,u>) = V xA (r,w ), (35.16)
3 5 .1 . S P E C T R A L A N D A N G U L A R D IS T R IB U T IO N 377
because
/
OO Q
(35.18)
[recall (32.6)]. Consequently, the asymptotic forms of the electric and magnetic
fields are
u) eiu>r! c 1 u eiiürí c
E(r,w) ~ - j ( k , w ) - i - —— np(k,u>), (35.20)
e iu>r/c i
B(r,u>) ^ i ---------- nX-j(k,u>). (35.21)
c r c
By using the Fourier transformed version of the local charge conservation con
dition, (31.5),
<jüp(k,u>) = k •j(k,u>), (35.22)
we may rewrite the second term of (35.20) as
¿ Ju r/c • Ju>r/c i
-----------nk •j(k,o;) = -------------- nn* -j(k,a;). (35.23)
c r c r c
/
OO r poo 1*00 1.
dtS(v,t) = — / dt — E(r,o))*eia,ixB(r,i)
-oo 47T J —oo J —oq ^
oo J
/
dt n •S du |B(r,a>) (35.30)
where we have used the symmetry property (35.9). As before, it is more useful
to consider the total energy radiated into the solid angle dFl [see (32.17)],
/
oo r f°°
dt(n*S)r2dO, = ffir2— / dw |B(r,a;)|
-oo 47T2 J o
dE d2E
= dCl EEdft / (35.31)
dQ Jo du)dQ'
d2E or
é M = 3 Í v | n * j ( k ' " )l
UT
- ip(k »w)r (35.32)
47T2 C
is the general expression for the spectral distribution, the energy radiated per
unit frequency per unit solid angle in the direction of observation n. This
equation is the analog of (32.18) for dP/dD.
3 5 .2 . S P E C T R A L D IS T R IB U T IO N F O R D IP O L E R A D IA T IO N 379
whence the Fourier transform of the current becomes that of the time derivative
of the electric dipole moment, according to (32.27) and (32.28):
j(k,w) = j dt
J u t
d_
dt
d(<) = —ia)d(w). (35.34)
d2E
|nxd(w)| = ^ Sin2 ^|d(W)| , (35.35)
~
3
dudD 4 tt2 c 3
where 0 is the angle between the observation direction and the direction of
d(o>). For the small system discussed here, the only reference to the direction
of observation, n, occurs as a multiplicative factor, implying the sin# behavior
exhibited above, characteristic of dipole radiation. For larger systems, n also
enters in the exponential so that the angular distribution could be completely
different. The total energy radiated per unit frequency range can be obtained
by integrating (35.35) over all angles,
él - /V d%E (35.36)
du> ~ J d u d il
dE 2 o, / \i2 ^ , x,o
(35.40)
380 C H A P T E R 35. S P E C T R A L D IS T R IB U T IO N O F R A D IA T IO N
(35.41)
„ I 00 d E ; e2 2| .
(35.42)
2 2
(35.43)
(35.44)
Thus we see that we can calculate the total energy radiated by a small system
through the use either of the spectral distribution, (35.36), or of the Larmor for
mula for the power, (32.30). The equivalence of these two descriptions is demon
strated generally in Problem 35.1, where the spectral distribution d 2E/du>dQ,,
(35.32), is derived directly from the angular distribution of radiated power,
d P ( t ) /d S l , (32.18).
where, due to the radiation produced by the accelerating charged particle, there
is a damping force represented by —y r , which we will assume to be small:
(35.46)
which exhibits the fact that many oscillations are necessary before significant
damping occurs. The velocity of the particle,
v(f) « —a w0 sin wote ytl2, for t > 0, (35.48)
has the Fourier transform
POO
Jo
( ____ L__________I____ A
OJpa
2 \u> —u>o + iy/2
i y /2 u>
u>+ u>>00 -1- /0 )J ’
+ iy/2 (35.49)
Without loss of generality, we may assume u > 0 here, in which case the two
terms are very different. Only the first denominator can be small, implying
that radiation is predominantly emitted with frequencies u> « u>0 . Therefore, we
approximate the square of the magnitude of (3 5 .4 9 ) by
wna
2„2
1
|v(u;)|2 (35.50)
4 (a; —u>o)2 + 72 / 4 ’
implying for the energy radiated per unit frequency range, (35.40),
dE_ 2 e2 ,2^02 „2 1
for ' w0. (35.51)
dio 3 7TC3 ^ 4 (u; —o>0) 2 + 7 2/ 4 ’
F/rad
Jo
to dw
2_e* 2( r duj
3 7TC3 (to - up)2 + 72/4
2 e 2 ^o « 2 2 tt
____
(35.52)
3 7rc3W° 4 y ’
/
/ OO 27r
/
OO
du) dx
(w - w o )2 + 7 2 / 4
-2a;o/7 (t2 /4)(1 + x2) -col + X 2 7 *
(35.53)
382 C H A P T E R 35. S P E C T R A L D IS T R IB U T IO N O F R A D IA T IO N
CO,,
Figure 35.1: Lorentzian line shape for the energy radiated per unit frequency
range.
We wish to compare this radiated energy to the original energy of the oscillator.
Since the latter is given by
^initial — ’ (35.54)
£ ra d = i n i t i a l ( |
jL d (35.55)
m e3 7
Assuming that the Lorentzian peak adequately accounts for the energy radiated
and that there are no other forms of energy dissipation, we learn, from the
conservation of energy,
2
7
3 0
2
* (35.56)
Since this is the same result found in Problem 32.2 where we calculated the
power radiated by the oscillator and identified 7 from
From this derive the formulas for B and E, (35.21) and (35.24).
4. Derive a formula for the collisional broadening of spectral lines by com
puting the spectrum of a weakly damped oscillator that vibrates freely for
only a finite time T. Assume that the probability of an oscillation time in
excess of a particular value decreases exponentially,
d2E
dudCl
= ~ 2c3 jnX J dt e,wij(k,<) • JnX J dt' el(vt‘j(k,i') , (36.1)
where we have introduced the average time and the time difference,
From the exponential structure of (36.2), we infer that the important range of
r that contributes to the integral is of order 1 /u;, thus setting the time scale
for the emission of radiation. This microscopic time scale may be much smaller
than macroscopic time intervals; for example, for visible light, r ~ 1 0 “ 15 sec.
The time T is then interpreted as the average (macroscopic) time of emission,
385
386 C H A P T E R 36. P O W E R S P E C T R U M A N D C E R E N K O V R A D IA T IO N
which can be specified only to within a time of order r. Substituting (36.2) into
(36.1), we write
d2 E _ [ d2P ( T )
(36.4)
dojdD J dcudO, ’
from which we infer the power spectrum at time T,
£
d2P ( T ) _
dr e“ nxlj(k,T + r/2)« n x £ j(k ,T —r / 2 )
dcodSl 47r2c
(36.5)
or, alternatively, using the second form of (35.32),
d2P ( T ) ÜJ*
r dr e " , w ( ij(k , T + r/2)* . lj(k , T - r/2)
dwdQ, 47T2 C J-o o l c c
- p ( k , r + r /2 ) V ( k ,T - r / 2 ) | . (36.6)
p ( r ,t ) = e 6 (r -v t),
f ( d r ) e ±iujn *r/ c / ^ ^ T^ \
e ± i u n • v ( T ± r /2 ) /c
(36.8)
J {dv) Uj(r,T±r/2)J
When these are substituted into (36.6), the T dependence disappears, as ex
pected, and we obtain
d2P e¿u;n • v r /c
d r e - iuTe 2
dujdCl
uP_
4ir2c
e2 ( ^ - l ) J°° dr e-*u,T(1-n ’ v/ c)
u>2e 2
(^2 _ 1 ) 2tt5 ( u> ( 1 - ^ cos0)) , (36.9)
4r2c
where 0 is the angle between the direction of observation, n, and the velocity of
the particle, v. The 6 function implies that there is no radiation, since
v
—cos 6 < 1. (36.10)
c
3 6 .2 . C E R E N K O V R A D IA T IO N 387
This is the familiar result that a charged particle moving with a constant velocity
in vacuum does not radiate.
However, if it were possible that
- > 1, (36.11)
c
the argument of the delta function could vanish, and radiation would be emitted
by the charged particle. Is there any way of effectively satisfying (36.11)? In a
medium, light can move with a speed, c', less than c, and correspondingly the
speed of a particle can be greater than c'. Now does the particle radiate? Recall
that the macroscopic Maxwell’s equations, (4.60), for a medium with dielectric
constant e and magnetic permeability //, can be put into vacuum form (1.65),
by the redefinitions (recall Section 7.2)
Therefore, the power radiated when a charged particle is moving with constant
velocity in a nonmagnetic medium {¡i = 1 ) of index of refraction n = y/e can be
obtained immediately from (36.9) by the substitutions e e/ n and c —* c/n:
V > - (36.16)
n
is satisfied. Such a medium can be easily found for fast particles. The radiation
is emitted on a cone described by (36.15) [see Fig. 36.1], and because of its unique
characteristics, is especially suited for determining the velocities of relativistic
charged particles. This phenomenon is called Cerenkov radiation. We emphasize
that the condition (36.15) can only be satisfied when n > 1. Because the index
of refraction depends on frequency, that is, media are dispersive, this means that
the condition (36.16) can only hold for a finite range of frequencies (typically, in
the optical region). Moreover, this dispersion implies that different frequencies
are emitted at different angles. Cerenkov radiation is commonly seen in water
moderated nuclear reactors as blue light surrounding the core.
388 C H A P T E R 36. P O W E R S P E C T R U M A N D C E R E N K O V R A D IA T IO N
Historically, this radiation was first observed by Marie Curie in 1910, studied
deliberately by L. Mallet in the 1920s, but only definitively explored by P. A.
Cerenkov from 1934 through 1938. The theoretical explanation was given by
I. M. Frank and I. Tamm in 1937. For a complete account of the history and
application of the Cerenkov effect through the 1950s see J. V. Jelley, C eren k ov
Radiation and its A p p lica tion s , Pergamon, New York, 1958.
The frequency spectrum of the radiated power can be obtained by integrating
(36.13) over all angles,
±
d = f da*£-
du J dtodCl
2 e 2 n 2v 2
nc c2 = H )
c2 \ C
"2~;”x~2 . if »(w) > v
n 2(u))v¿ J
(36.17)
dE
P
dt
(36.18)
dE_
(36.19)
dz
In (36.18) and (36.19) it is understood that the u> integration extends only over
the range where n(u>) > c / v . Finally we note that detection is a quantum
3 6 .3 . P R O B L E M S F O R C H A P T E R 36 389
(36.20)
where ¡3c is the velocity of the particle moving in a medium with index of
refraction n.
2. Show that the number of photons produced per unit path length of a
particle with charge Z e and per unit energy interval of the photons, d E }
is
d2N _ a Z 2 . 2 _ a 2Z 2 f _ 1 \
dEdz h e Sm c romec2 \ f32 n 2( E ) ) ’
du
=
*cj_„ [ (dr)(dr^
J'
p - p
A ' “ |r-r'|
I j ( t ,r + i ) . I j ( , , r _ : ) _ P (r , r + i ) , ( , , r - i ) -
r(i) = t 2, (37.1)
v(t) = at, (37.2)
391
392 C H A P T E R 37. C O N S T A N T A C C E L E R A T IO N A N D IM P U L S E
For u) > 0, the argument of the delta functions in (37.4) never vanishes,
The power, (37.4), is nonzero only for w = 0, which corresponds to static fields.
Such fields do not correspond to radiation, and a uniformly accelerated charged
particle does not radiate. On the other hand, the total radiated power can be
computed from the Larmor formula, (32.25):
-/( * ) <37'8>
where we identified E as the electromagnetic field energy while the remaining
term is the power radiated. To obtain the Larmor formula, we neglected a
further total time derivative. If we now use the above expression for the power
radiated,
d = er, (37.10)
3 7 .1 . R A D IA T IO N B Y A U N IF O R M L Y A C C E L E R A T E D P A R T IC L E 393
E = constant
v (0) v(T )
t = 0 t = T
v(w) dt eiwtv ( t )
394 C H A P T E R 37. C O N S T A N T A C C E L E R A T IO N A N D IM P U L S E
/
OO
dt e‘wt[a<5(í) —a<5(í - T )]
•CO
= a (1 - etVr)
|v(cj)p = 4u siu
1
iu T , (37.16)
¿4
u - 1 /T, (37.18)
37.1. RADIATION B Y A UNIFORMLY ACCELERATED PARTICLE 395
that is, as T becomes longer, dE/du> is more and more concentrated at low values
of a;. In the limit when T —►oo, we recover the situation of uniform acceleration
discussed above. Moreover, we see again that the emission spectrum is not time
analyzable, since the relation (37.18) implies that it takes a time of the order
of the whole process to determine the frequency. The total energy radiated is
obtained from (37.17) by integrating over all frequencies,
. dE 2cM a2 f du . 2 loT
Er&d —r sin
JQ U du) 3 C 3 7T Jo (jü¿ 2
2 e2 2 /2 f°° dx . 2
3c^a T \7T Jo ^ Smi
-2 e- a * T (37.19)
3e3 ’
where we have used the integral (34.18). From (37.19), we observe that the
total energy radiated per unit time is
E yeló. 2 2
(37.20)
3^ a
which is precisely that obtained from the Larmor formula, (37.7), which refers
only to uniform acceleration.
Finally let us calculate, from the source point of view, the radiated power
using (37.9),
2 e2 ..
P
3 ^ V *V
which incorrectly attributes the radiation entirely to the beginning and the end
o f the acceleration process. However, the total radiated energy,
f?rad = Í d tp
J —oo
2 e2
2 e2 2
(37.22)
= 3 c3 *1 T ’
t = 0
A
>
>
^1
U
II
The spectral distribution for the radiated energy can be computed by substi
tuting (37.23) and (37.24) into (35.32), where we encounter the integral
L d“ ,u - v (37.26)
L Jt‘M=- ly (37.27)
.e f jn .v , _ jn .v 2 \ = .c ( 1___________ 1 _ \
w \1 —n^Vj /c 1 —n * v 2/ c y ^l —n - V j / c 1 —n * v 2/ c / ‘
(37.30)
We now immediately obtain the spectral distribution from (35.32),
d2£ / v i____________ V2 \
(37.31)
d u ;d f2 47r2 c 3 n \1 —n -V j/ c 1 —n » v 2/ c /
Since (37.31) is independent of the frequency, the implied total radiated energy
is unbounded. This unphysical result is due to the idealization that the scat
tering occurs instantaneously, that is, our assumption that the particle changes
its velocity abruptly. Realistically, this change occurs over some finite period of
time, T, so that our result, (37.31) holds only when u <C 1/T . In the nonrela-
tivistic limit, where ¡vi^/c] <C 1, (37.31) reduces to
d2E ____
-|nx(vj —v2)| = ^ 2^ ( v i - v 2) sin 0, (37.32)
dudQ ~ 47T2 c :
where 9 is the angle between vi —V2 and n. Integrating this over all angles, we
obtain the energy radiated at the frequency u
dE_ 2e^
•(vi - V 2) 2, (37.33)
du> 3 7rc:
which can also be easily obtained from (35.40), since the Fourier transform of
the derivative here is simply
/
OO poo
dt eZUJt\(t) = / dt elwt(vi — V2 )^(t) = v i —V2 - (37.34)
-oo J—oo
Because either denominator in (37.31) has the structure
1 —- n . v¿ = 1 — — cos 9 i , (37.35)
c c
where is either Vi or V2 , and $i is the angle between n and v¿, we see that the
radiation in the ultrarelativistic limit (|v¿| ~ c) is preferentially emitted near
the direction of the velocity of the particle (the forward direction), either before
or after the scattering act. On the other hand, this behavior is softened by
the numerator factor |nXv¿| = t;¿sin which forbids radiation in the exactly
forward direction, = 0. In either region of significant radiation, characterized
398 CHAPTER 37. CONSTANT ACCELERATION AND IMPULSE
by Oí -C 1, only one term in (37.31) makes a major contribution (in the following
we drop the subscript ;):
/ n X v /c \2 J . ___________
V I - fe o s fe o s 6») 2
462
( i- f + f ) 2 ~ ( i- £ + * 2) 2
f 4/02, if y/1 - (v/c)2 <C 0 < 1,
(37.36)
X 0, if 0 = 0 .
* = V I - («/<02 = ^ (37-37)
(a) Write j(k,u>) and p(k,a;) as integrals over the time interval t\ < t <
t2-
(b) Show that charge conservation, (35.22), is satisfied.
(c) Derive the general formula for dE/dil as an integral over the time
interval t\ < t < ¿2 ,
d E _ _ _ f _ y*2 , 1 id . n X v (Q V
dCl 47re3 Jtl 1 — n •v (t )/c \di 1 — n •v (t)/ c)
(d) Integrate over Q in the above general formula for dE/dil and thereby
derive the general formula for the total energy radiated,
3. Apply the results of the previous problem to treat the radiation by a rel
ativistic, uniformly accelerated1 particle, undergoing so-called hyperbolic
motion, for a finite time interval,
_ cr . - cr ^
z — a cosh — , ct — asinh — , —tq < r < to.
a a
(a) By considering short times (or r « 0), make connection with the
nonrelativistic formulas and identify the parameter a.
(b) Show that
-Í® . = i l L
du> dil n #2_ 0 tt2c i c V c/J
in the Tq —►oo limit. Use the second integral representation in (18.18)
and note that
K t (x) = -K '0(z)
[cf- (18.71)].
(c) For 6 / 0, 7T (n *z = cos#) and tq —> oo, show that
dE_ _ 3e2 1
dO, 32a sin3 6
jp _ 4 g2 Pm
Synchrotron Radiation I
401
402 CHAPTER 38. SYNCHROTRON RADIATION I
or
(38.6)
This implies a constant deflection of the velocity vector, that is, v precesses
with angular velocity u?o>
6C
« o = —-g B , (38.7)
about the direction of ( —e x ) the magnetic field. (See Fig. 38.1.) The angular
speed of this precession is the Larmor frequency,
u>o = (38.8)
(38.9)
me ’
from a Cerenkov counter, which measures the speed of the particle according to
(36.14), we can determine the mass of the particle:
m — ------------ p (38.12)
/?(r, t) = e 6 ( r - r ( t ) ) y (38.13)
j ( r ,t ) = e v ( t ) 6 ( r (38.14)
where r(t) is the position vector of the particle at time t , and v(tf) is its velocity.
Substituting (38.13) and (38.14) into (36.6), we obtain the spectrum of the
power, emitted at time T, into the element of solid angle in the direction n,
/
d?P(T) _ w V
oo
dr e~ ’ lv (T + r /2 ).v (T -r /2 )-l
dudil 4tt2c
-OO
x exp (38.15)
{ ¿ ^ n . [r(T + r/2) - r(T - r/2)]} .
This description in terms of an average macroscopic time can be important, for
then it is possible to consider the effect of a slow alteration in the parameters
describing the motion.
It is convenient to choose the coordinate system such that the particle is
moving in a circle of radius R about the origin in the xy plane, and the magnetic
field B is directed along the z direction. Also, without loss of generality, we
choose the observation direction, n, to lie in the xz plane, making an angle 6
with the -j-z axis, as shown in Fig. 38.2. Then we have, for a convenient choice
of initial conditions,
where wo and Ü are given by (38.8) and (38.10), respectively, The corresponding
velocity is given by
In the last equation, we have used the generating function for the Bessel func
tions of integer order, (16.28),
oo
eizcos* = imeim*Jm(z). (38.27)
With these evaluations, we obtain the power spectrum, (38.15), in the form
x J2 S imeimUo(T+T/2)Jm ( - f l s i n f l )
m = —oo m ,— — oo
Now we recall that the emission time T represents an average over many periods
of the motion, since many oscillations are required to identify a frequency. Here,
it is sufficient to consider the average of (38.28) over one period, by using the
relation
{e¿(m -m > oT )oneperiod = 5mmi (38.29)
38.2. SPECTRUM OF SYNCHROTRON RADIATION 405
d2P 'u
' 2° 2 r°°
2e2 r
diüdíí = COSa' 0 T ' 11
OO
x £ eim“°r [/m 0 fls in 0 )] . (38.30)
Incidentally, we remark that this result, (38.30), can also be obtained by com
bining the two exponentials directly as follows:
i(u i/ c )n ♦[r (T + r /2 ) - r ( T - t/ 2)]^
(e one period
— / /c)Rsin 0 ( - 2 sin u>0T sin ^ o r ) \
\c /one period
f2 u R . . u 0r
= Jo ------ sin 6 sin — —
\ c 2 )
gimtd0r (38.31)
= £ J m ^ s in ^ l
m = —oo ' '
j°(a:)=C (38.32)
(38.33)
with p = / / = R, k = (u>/c) sin 0 , and </>— 0 ' = u/0t .
Exploiting the exponential representation for cosu >0 t ,
frequency, (38.8), are radiated. That is, the emitted frequencies are multiples
o f u>o, u> = mw0, where the integer m is positive since u > 0. Thus the power
spectrum can be written as a sum of the contributions of each harmonic,
d2P (T ) ^ £/ ,d P m(T)
(38.36)
dudO,
m= 1
406 CHAPTER 38. SYNCHROTRON RADIATION I
Therefore, the angular distribution for the power radiated into the mth harmonic
is
dPm Jm(m/3 sin 0) 1 2
(m/? sin 0)] + (38.42)
dQ ¡3 tan 9
The resulting frequency distribution of the radiated power is, from (38.15),
R J_c
= — — r dre~WTf(w0T). (38.46)
~ 2 tt
The challenge now is to carry out the r integration. We observe that /(u>or) is
a periodic function,
/m = (38.49)
Inserting the representation (38.48) into (38.46), and carrying out the then
trivial r integration, (38.35), we obtain the spectrum of the radiated power:
dP x°°A
- muJo)Pm, (38.50)
m= 1
where the total power radiated in the mth harmonic is expressed as a Fourier
coefficient,
Breaking this integral at n/2 and substituting <f>—*•ir —(j>for the range tt/2 to
we finally find
7T,
I*tt/2
Jm{z) = / — [cos(;zsin<£ — m<j>) + cos(^sin<^ -f m<f> — m 7r)]. (38.54)
Jo *
This supplies the following integral representation for the Bessel functions of
even order:
f n/2
J2m{z) = / — [cos(;zsin <¡) — 2 m<^) + cos(zsin <¡) -f 2 m<j))]
Jo ft
r d<j> ( . <f>\ ,
= / — cos ( zsm — 1cosrrnp, (38.55)
Jo ft \ */
where <j>—►^ in the second line. The integral in (38.51) can now be expressed
in terms of the derivative and the integral of J2m >which are represented by
t/ / \ r d * . * ( . <a
(38.56)
- -J ~ sm2 cosm(Psm l *sm 2 y ’
and
/■* rw ¿<t> sin ( * sin I)
/ dzJ2m{z) = / — cos m<¿-------r -r ------. (38.57)
7o Vo f sin §
Therefore, the total power radiated into the mth harmonic is
p2 m(3
Pm = 2^2mPm^) “ (1 — ft2)JOI dzJ2m(z) (38.58)
f°° , dP 1 f°° . dP
= / OW— = - / d u -r -
Jo du 2 du
-~ eiRL r
2 ir J _ 0
f3 2 COS U qT
sm 7 j U o t
— 1 1l í°°
- /
2 J-oo
1
du u e~lWT —
2*
p ^ fis in ^ _ e-i2 *R Sm
(38.59)
38.4. TOTAL RADIATED POW ER 409
/
OO J p CO
du> u e~iuX = i — / dw e~iuX = 2irtf'(A). (38.60)
•CO dX J — OO
where
y = <j>+ 2A/?sin^,
(38.62)
so that we may rewrite the total power as
(38.63)
lit y -00 1 + A pcosf
Here the y integration can be performed by integrating by parts twice and noting
that the support of the delta function occurs at </>= 0 :
£+¥&$■ <*«>
2
(1 + A/?)3
The remaining A integral can now be easily evaluated, yielding [see Problem
38.1]
/?2 —\
P = u>0/3
/ > J(T T W (- 2 + 8 1 + A/?
e2 2 /?3
= oTTw 0
3 i? ( I - / ? 2)2
2 e2 ~/ E
= (— 2 (38.65)
3 /£ Vme2
which is the exact result for the total radiated power.
In the nonrelativistic limit, /? <C 1, (38.65) reduces to the Larmor formula,
(32.25),
Pn .k . = z~jiUoP3
- i t (tt\ (38.66)
~ 3 c3 \dt ) ’
410 CHAPTER 38. SYNCHROTRON RADIATION I
For a relativistic particle, the power radiated is larger by the factor (E /m c 2)4,
signifying that high energy charged particles moving in circular orbits emit
substantial synchrotron radiation. In one period, T = 27r/u>o, the total energy
radiated is
where
1 GeV = 106 keV = 109 eV. (38.70)
Inserting typical numbers for an electron synchrotron, R = 10 meters and E =
10 GeV, we find for the energy loss per cycle
A E = 88.5 MeV, (38.71)
which is quite substantial. For this reason electron synchrotrons are impractical
for energies greater than ~ 10 GeV. [Roughly twice this energy is radiated
from LEP, where E = 55 GeV and R = 4.25 km.] On the other hand, the
radiation from any existing or projected proton synchrotron is quite negligible,
being smaller, for the same energy, than electron synchrotron radiation by the
factor (me/m p )4 ~ 10“ 13. [For the LHC, where E — 7 TeV and R = 4.3 km,
the energy loss per cycle is only about 4 keV.]
where
r = r(T + r / 2 ), r' = r ( T - r / 2 ) ,
thereby obtaining the frequency distribution
’2- 2 r°° ^ v (T H- t /2 ) •v (T — r /2 )
^ = ^ r dTe— (:
du 2 ir e J _ 0O \
xf d \ e - i(-uteW r~r'l,
38.5. PROBLEMS FOR CHAPTER 38 411
p= jrL ix -0 s" ( t + A
|r- r'|/c)-
Specialize this general result to circular motion, and let ujot = <^, to obtain
(38.63) and hence (38.65). (In this way, the second derivative of the 8
function emerges automatically.)
3. Verify the numerical result for the energy loss per cycle, (38.69). (1 esu
unit of potential = 300 volts— see Appendix A.)
6. Derive the total power radiated, (38.65), from the result of Problem 37.2(d).
Chapter 39
Synchrotron Radiation
II— Polarization
where the terms in the square brackets are proportional to the electric field
strength.
To isolate the effect of each polarization state, we look at the two components
of E, in the plane perpendicular to n, separately. We choose the coordinate
system (which is naturally set by the synchrotron) as before, with n given
by (38.16). We define two polarization orientations as follows: For “parallel
polarization” the electric field is in the direction of e||, which lies in the orbital
plane and is perpendicular to n, that is, ey points in the + y direction,
e || = ( 0 , 1 , 0 ). (39.2)
413
414 CHAPTER 39. SYNCHROTRON RADIATION II— POLARIZATION
and the resulting contribution to the power spectrum radiated in the parallel
polarization state is
x ^2vy ( T + r / 2 ) v y ( T - T / 2 ) . (39.6)
We now can evaluate (39.6) by following closely the procedure given in the
previous chapter. Therefore, instead of the factor
appearing in (38.15), we now encounter, because the position vector and velocity
are given by (38.17)—(38.19) and (38.20)-(38.22), respectively,
02
= — (cos 2u>oT + costa r). (39.8)
The time average of (39.6) involves not only (38.31) but also the new evaluation
[most easily done starting with (38.26)],
00 2
(/?2 cosw0r - 1 ) ^ eim" ir [<7m ^ ñ s i n ^ j , (39.10)
m ——00
1 00
-/? 2 eima,oT[ - J m+1 Jm _ 1 + cosw ot J £ ]. (3 9 .1 1 )
The r integration now leads to the power radiated into the mth harmonic being
proportional to
y ( ■'” « - 7 „ +1
(39.12)
where we have used the recurrence relation (38.39). Thus, the two terms in
(38.42) represent, respectively, the radiation in the states characterized by ey
and e± :
dPm
sin6))2, (39.13)
dQ,
2. Compute the power radiated into the mth harmonic for a given polar
ization, (Pm)|| and (Pm)j_, starting froin (39.1). Note carefully that ||
polarization means that only the <j> component of j enters, so that
r 2m>3 dz
2m
p 2m p j p2m p
(Pm)x = 0 2 m/? / - ? 2m W - / dzJ2m(z)
Jo z Jo
3. Compute the total radiated power for a given polarization, and P±.
The result is
6 + B2 2 - B2
= P±= P D
8 8
so nonrelativistically, and ultrarelativisitically, respectively, the ratio of
power in the two polarizations is
¿ = 3 (/?<!), = 7 (/?«!).
P±
Synchrotron Radiation
III— High Energies
e2
Pm « W0— 2mJ'2m(2m), (40.1)
q1/6 e2
Pm ---- r(2/3)u>0— m 1/3, (40.2)
7r rt
ol/6
^2m(2 m) ~ — r(2/3)m"2/3, m > 1, (40.3)
oo mc
P= £ Pm ~ £ ml/3> (40-4)
m=l m=l
where we have noted that the continuing increase in the power radiated into
higher and higher harmonics must break down for sufficiently large m, since the
total power radiated, (38.65), is finite. Consequently, we have cut off the sum
417
418 CHAPTER 40. SYNCHROTRON RADIATION III— HIGH ENERGIES
at mc, the critical harmonic number. In terms of this cutoff, the total power
radiated is
P ~ m*/3 , (40.5)
which, on the other hand, is, by (38.65), proportional to (E/mc2)4:
— (40.6)
me2 Jv
P
(40.7)
'm e21 V
E
-J R, (40.9)
R( m)
^c(A) (40.11)
£ 3 ( Ge V) '
J2m(2 ra) = — f
— sin<f)sm[2m(sm<f) — <j))].
Jo ft
(40.14)
d
— (sin - 0, (40.15)
<f>3
sm — —, (40.17)
sin<^ « <^, (40.18)
and noting the range of integration can be extended to infinity with negligible
error, we obtain
= -I m t 'f h - *
(40.20)
approximation (40.2) breaks down for sufficiently large m. This can be traced
to the fact that, in (38.58), /? is not exactly equal to 1. Consequently, both terms
there contribute but we will concentrate on the first one as it contains all the
essential characteristics of the radiation. [See Problem 40.2.] Thus we seek an
asymptotic expression for J 2m( 2 m/?), starting from the integral representation
(40.13):
sin <j) sin 2 m(/3 sin (j) — <¡>) « (j) sin 2 m -<t>
9 3!
1
= (j) sin ( 2 m - 4 ( 1 - f}) -
>sm m ( ( l - / 3 2 ) ¿ + i ¿ 3
(40.22)
For m fixed and /? approaching unity in such a way that m( 1 —/?2)3/ 2 <C 1, the
significant contribution to (40.24) comes from the region where x is large, and
(40.24) reduces to (40.19):
=T sin(?*’)■ (4°'25)
where all reference to the speed of the particle has disappeared. However, for
sufficiently large m, the parameter ra (l—/?2) 3/2 becomes large, and the integrand
undergoes rapid oscillations in x except near the stationary points, which satisfy
A 1 + x 2 = 0; (40.26)
dx
x = ± i. (40.27)
or
m« = ( l - ^ ) - ^ = ( - ¿ ) 3 , (40.33)
supporting our previous estimate, (40.7). The bulk of the radiation is emitted
with harmonic numbers near mc. The qualitative shape of the spectrum is
shown in Fig. 40.3.
becomes
^ <4°-34>
If we further let /? —> 1 and mc ^> m^> 1, we find from (40.3) and (40.2),
1o l / 6 T2
dPm CJo
m — r(2/3)(2/m )2/ 3
2w d9 27r R
.2/3 (40.35)
Here we compare Pm, which increases as ra1/ 3, with dPm/dOy which behaves as
m2/ 3, from which it is evident that the radiation, for large harmonic numbers,
is confined to a small angular range around 0 = 7r/2 of width
AO ~ m “ 1/3. (40.36)
Since most of the radiation is emitted with harmonic numbers in the neighbor
hood of
mc (40.37)
the radiation is concentrated in an angular range about the plane of the electron
orbit of the order
2
A0 ~ —— = \/\ —/32. (40.38)
E
A plot of the distribution in 0 is given in Fig. 40.4.
This radiation, concentrated around 0 = 7r / 2 , is predominantly polarized in
the plane of the orbit, according to (39.13) and (39.14). More precisely, it can be
shown (see Problem 39.3) that in the ultrarelativistic limit the ratio of the power
radiated with parallel polarization to that with perpendicular polarization is
P\
7, for / ? « 1, (40.39)
P±
P\
3, for ¡3 <C 1. (40.40)
P.L
That is, for any value of /?, the radiation is strongly polarized, the degree of po
larization increasing with the speed of the charged particle. This characteristic
distinguishes synchrotron radiation from thermal radiation, and was, for exam
ple, the clue to understanding the origin of the nonthermal radiation emitted
by the Crab Nebula.
424 CHAPTER 40. SYNCHROTRON RADIATION III— HIGH ENERGIES
where r (tf) is the position vector of the particle at the emission time t', while t
is the detection time, which are related by (31.46),
r - r(t')
n= (40.43)
| r - r ( t ') r
40.5. QUALITATIVE DESCRIPTION 425
me 2
<t>~ x/ 1 - /?2 = (40.46)
HT'
which is just the behavior found in the preceding section.
Next, we seek a qualitative understanding of the characteristic frequencies
radiated by a high energy charged particle moving in a circle with frequency
cj0. Because of the strong directionality of the emitted radiation as expressed
by (40.46), the radiation detected at a particular point only arises from a small
portion of the orbit, or equivalently, is only emitted during a time interval
small compared with the period of revolution, 2ir/u)o. (See Fig. 40.5.) This
effective emission time interval is of order (27r/u>o)<^, so the important frequencies
radiated are ~ (emission time) - 1 ~ (u>o/27r)(l/</>). That is, the smaller the time
interval involved in the emission, the higher the frequency emitted. Therefore
a typical frequency emitted is
ÍÜQ E
—7 7 = ----- 0 ^ 0- (40.47)
a/ 1 - ff2 me2
However, this is not what the observer sees since detection time intervals are not
equal to emission time intervals. To see the connection between these intervals,
the Doppler effect [Christian Doppler (1803-1853)], we recall that these two
times are related by (40.42), which implies for the respective time intervals
1 1 E
Ud = Wo ( (40.50)
( l - £ 2) ~ e (1 — /32)3/2 K m c <
Propagation in a Dielectric
Medium
V x H = lf>+ — J, (41.1)
c c
— V x E = -B, (41.2)
c
V •D = 47T/0, (41.3)
V •B = 0. (41.4)
Before proceeding, we note that here the scalar equations, (41.3), (41.4), are
not independent of the first set, (41.1), (41.2), since taking the divergence of
the latter, and making use of the current conservation condition, (31.5), we find
0 = J ^ (V •D — 47iy>), (41.5)
0=^(V.B). (41.6)
Thus, excluding statics, the divergence equations, (41.3), (41.4), are subsumed
within the curl equations, (41.1), (41.2).
To concentrate our attention on light of a definite frequency cj, we introduce
the Fourier transform
427
428 CHAPTER 41. PROPAGATION IN A DIELECTRIC MEDIUM
—iu), (41.8)
dt
yielding for the vector equations
tco,
V x H = ----- D +
IÜJ
—V X E = B. (41.9)
where //(r,u>) and e(r,u>) are complex functions. We will further assume
K r ,w ) « l, ' (41.11)
since a ferromagnet cannot follow the rapid oscillations of the electromagnetic
field in a light wave. It will be sufficient for our purposes to suppose that spatial
variation of e occurs in the z direction only,
, _ V tU /C A '1 T
—ik j..( n x B j_ ) = ------- Ez -\------ Jz ,
c c
2Ll)
.(nxEx) = B z, (41.15)
Bz = ——k x •( n x E x ),
U)
c A ir
Ez = — k± •( n x B ± ) - i — Jz . (41.17)
UJC WC
Inserting these expressions into the J_ set, (41.16), we have the following equa
tions for the J. components:
dn ÍU€ „ C, , „ . 47T , ,^
— B x -------- n x E x + « k x —kx * (n x E x ) = ------ n x J x , (41.18)
oz c w c
d ^ iu> ^ c , / ^ x ,47rT
^ -E x H-----n x B x - « k x — k x * ( n x B x ) = k x — Jz , (41.19)
OZ C (jJ€ we
d iwe c 47T
■^nxB i + Ex + i n x k x - k x - ( n x E x ) = — J x , (41.20)
oz c to c
d m zw , , c _ . _ v , A tt T / i< \
— n x E x ------ B x — m x k x — kx •( n x B x ) = n x k ± — Jz , (41.21)
oz c we we
^ - k x -E x + — ( l - ^ ~ ) kx •(n X B x ) = k \ — Jz , (41.24)
oz c \ to2e J toe
The vectors n and kx define a plane, called the plane of incidence, so that the
system (41.22) and (41.23) relates the component of B x in this plane to the
component Ex perpendicular to this plane, and vice versa for system (41.24),
(41.25). That is, if we take kx in the x direction, we have equations governing Bx
and Ey, and By and EX1 respectively. These perpendicular components belong
together physically [see Section 7.2 and also the relativistic transformations
in (10.83) and (10.84)]. By combining (41.22) with (41.23), and (41.24) with
430 CHAPTER 41. PROPAGATION IN A DIELECTRIC MEDIUM
(41.25), we convert these systems of first order differential equations into second
order ones:
d2 u>2c 2
(41.28)
+ ~c? ~ k ± '
We further assume that e is real. Then, depending on the sign of u>2e /c2— , the
solutions to the corresponding homogeneous differential equations are different:
U)2€
2---- &x > 0 : exp ±i (41.30)
k2 -k i, (41.31)
e*kjL • r ± e ikzz __ e ¿k • r
(41.32)
which represents a plane wave moving in the direction k, since the phase is
constant on a plane perpendicular to k. The wavelength, A, is defined as the
distance over which the phase advances by 27T, so
27T 1
|k|A = 2tt, or |k| = (41.33)
A5
41.2. REFLECTION AND REFRACTION: JL POLARIZATION 431
|k| is called the wavenumber, k the propagation vector, and A the reduced
wavelength. Including the time factor e“ *wt, we find the dependence of a plane
wave on space and time to be
«(k • r-w t)
(41.34)
. dr
(41.35)
k '3 i = " ;
the phase speed thus is
v — — = o>A = v\. (41.36)
|k|
\k\2 = k2
± +k2
0= ^ , (41.37)
as we saw in Section 7 .2 .
Since the relation between field and source is linear, the field produced by a
prescribed current can be expressed in terms of an appropriate Green*s function.
Green’s function corresponding to (41.26) satisfies
(41.42)
z’ > z > 0 : ti k t 2 z
g = B eik“ z + Ce~ik‘ *z (41.43)
z< 0: g = D e~ik‘ lZ, (41.44)
where the subscripts 1 and 2 refer to the value of kZ) as defined by (41.31), in
medium 1 and 2, respectively. Note that these equations are very similar to
those considered in electrostatics; in fact, when in both dielectrics
(41.45)
(41.46)
B + C=D (41.47)
B -C = -~ D . (41.48)
kZ2
(41.49)
g is continuous, and
* Q • Z — z'-\-Q
(41.50)
41.2. REFLECTION AND REFRACTION: ± POLARIZATION 433
1 Jkyz' (41.53)
C =
2k2 ’
k2 - h i
B = (41.54)
k2 + k\ 2k2
2k2 i cikzz'
D = (41.55)
k2 ■+■k\ 2k2
where we have simplified the notation by dropping all the z subscripts on the
k ’s. Combining these results, we may now write down the Green’s function in
the two regions,
&2 — k\
(41.59)
k2 + fa ’
2k2
t = (41.60)
¿2 + &1
The two terms in (41.57) have a simple physical interpretation. The first refers
to the direct wave from the source, since \z — z*\ is (the ¿-projection of) the
distance from the source, while the second represents the wave reflected from
the interface, since ¿ + ¿ ; is (the ¿-projection of) the distance from the source to
the interface and back to the observation point ¿. [See Fig. 41.1.] The following
algebraic identities relating t and r,
t = 1 + r, (41.61)
M = Jfe2' ( l - r ) , (41.62)
k±
The common factors in parentheses refer to how the wave is produced, which
is not of interest to us at the moment. The other remaining factors describe
the reflection and transmission of a plane wave: unit amplitude incident on the
interface, amplitude r reflected, amplitude t transmitted.
We recall that Green’s function, g (z) zf) ) describes only the ^-dependence of
the wave propagation; the x and y dependence is contained in the factor e*kj- *rJ-
[recall (41.32)]. Since the transverse momenta are unaltered, while the longitu
dinal momenta (kz) change from kz 2 to kz\ as the wave crosses the interface,
we have the geometrical picture shown in Fig. 41.2 for the incident, reflected,
and transmitted propagation vectors. Under reflection kz merely changes sign,
so the angle of incidence equals the angle of reflection. The relation between
41.2. REFLECTION AND REFRACTION: ± POLARIZATION 435
|k| = — , (41.66)
c
with n = yfl. The continuity of kx at the interface,
(41.69)
i k x r cot9 '
to obtain
_ sin(#i — 62)
(41.70)
sin(0i + 02) ’
and
2 sin 0 i cos 02
(41.71)
sin(0i + 02)
but the wavenumber form is usually preferable. This is particularly evident
for normal incidence, where 6\ = 02 = 0 and (41.70)-(41.71) is indeterminate.
On the other hand the wavenumber form, (41.59)—(41.60), is well defined, since
when = 0 , kz = which yields
ni - n2
(41.72)
ni + n2 ’
2n2
(41.73)
ni + n2'
Also from the wavenumber form it is easy to see that if n\/n2 —» 00 , the
coefficients become
r = —1, t = 0; (41.74)
this corresponds to the situation of total reflection by a perfect conductor. We
verify this last statement by recalling the plane wave part of the Green’s function
between the source and the interface, (41.64):
0 o c (l + r) = O, (41.76)
(41.77)
1
1
dz (\el
II
8*1
-
where, again,
" 2
k \ ^ e -k \ . (41.78)
With the plane interface between the two dielectric media located at 2 = 0, the
form of the solution in the three regions, when z‘ > 0, is
B + C = £>, (41.82)
B - C = - — h-£> (41.83)
£1 h
implying
c“Ki+Hs.,a (41.84)
1
B , . 1- ^ 10 . (41.85)
2 V £i h ,
z —z'+O
\_d_
= 1, (41.86)
€2 d z* z= z'~ 0
so
B eik*z' + C e~ik2Z' = Aeik2*', (41.87)
and
ik2 ( - A e ik2Z> + B eik2Z' - C e~ik' z') = e2- (41.88)
41.3. REFLECTION AND REFRACTION: || POLARIZATION 437
^ikzz1 (41.89)
C
2h6 ’
1 É2. kl •
_____ £i k2 ik2z'
B (41.90)
i + '
D (41.91)
l + f f f t 2*2
i _ e_2_ÍLL *
£i k2 c ¿fc2^; i ^2 —jk2z '
A (41.92)
1 + iajp. 2A;2 2*2
£l «2
. - i kiz gik^z1
z < 0 : g — t ——e (4 1 .9 4 )
2k2
Green’s functions for the two polarizations, J_ and ||, are given by (41.57)-
(41.58) and (41.93)—(41.94), respectively, with the corresponding reflection and
transmission coefficients given by
_L mode mode
ki - * 2 . * iA i ~ k2/e2
^1 + ^2 ' ki/ei + k2/c2 '
2Ar2 2 * 2/62
(41.95)
*1 + *2) * iA i + * 2/^2
t = 1 + r, (41.96)
-L mode || mode
where, again, the second relation is obtained from the first by the substitution
kz ►kzf 6 .
It is particularly interesting to ask when the reflection coefficient can vanish.
For the _L polarization, r = 0 only when kz\ = kZ2, or 61 = 62 , that is, in the
438 CHAPTER 41. PROPAGATION IN A DIELECTRIC MEDIUM
kzi _ kz 2
(41.98)
^1 ^2
or, since
*2 lo21 k\
(41.99)
e2 c2 e e2 ’
the condition becomes
("D i (41.100)
- = . / í i ± í i |kx|. (41.101)
c V eie2
Geometrically, since
|kx| = ~ n sin 0, n = y/l, (41.102)
c
(41.101) is satisfied when the angle of incidence, 02 equals 0#, where
£1__
sin Ob = (41.103)
+ ^2 *
or,
tan Ob = — . (41.104)
n2
At the incident angle 0b , the || mode is completely transmitted, so that the
reflected wave is completely polarized perpendicular to the plane of incidence.
The phenomenon was discovered by Sir David Brewster about the year 1800,
and in consequence 0b is called Brewster’s angle.
To express r and t again in terms of the angle of incidence O2 and the angle
of refraction 0 i, we use the geometrical relations
to write
kz
(41.106)
M 2
Consequently, for the || mode, the reflection and transmission coefficients can
be given as
B B
Incident:
k Reflected:
Figure 41.3: Relation between field orientations for incident and reflected waves.
Here we observe, from (41.107), that Brewster’s angle occurs when the sum of
the angle of incidence and the angle of refraction forms a right angle,
0i + 02 = */2. (41.109)
± mode mode
n\ - n2 n\ — n2
r = — , r = --------------------
n\ + n 2 ni + n 2
t= 2 n 2 . (41.111)
ni + n2
< = 2 n i .
ni 4- n 2
But for normal incidence there can be no physical difference between the two
polarizations, since the plane of incidence is not defined. To reconcile the two
forms in (41.111), we recognize that the reflection and transmission coefficients
for the ||mode refer to B, while those for the Jl mode refer to E [see (41.26) and
(41.27)]. Under reflection, E reverses its sense relative to B, since the direction
of propagation is reversed. [See Fig. 41.3.] So the two forms for the reflection
coefficient in (41.111) are physically equivalent. To reach the same conclusion
for the transmission coefficients, we recall from Section 7.2 that the magnitudes
of the electric and magnetic fields in the plane wave are connected by
(41.120)
-Sé
o
11
II
*4
k^ = ] / ^ ~ kl (41.121)
becomes imaginary,
kli < 0 or kz i = i/c, (41.122)
where k is real. The propagation of the wave in medium 1 is now described by
gjfc.ikl _ e- KU\t (41.123)
a decreasing exponential: the field penetrates only a short distance into region
1, and no energy is transmitted into that region, it being all reflected. This last
fact follows from the form of the reflection coefficients,
kZ2 —
1: r = (41.124)
kZ2 + i’
r — (41.125)
kz2U 2-iK i/ei
fc*2/£2 + * « l A l ’
both of which have unit magnitude,
|r|2 = 1. (41.126)
In the next section, we will prove that this implies that all the energy is reflected.
We have here the situation of total (internal) reflection.
41.5. ENERGY CONSERVATION 441
= -7 - 2 Re / ^ n .E (r ,« )* x B (r ,w ), (41.129)
4 7T J o 27r
We next integrate over the xy plane by introducing the Fourier transform in the
kx space:
J d xd yd tS 2 = ¿ R e j ~ ^ j l | ^ n . E ( * , k x , w ) * x B ( * ,k x ,« ) .
(41.130)
The integrand in (41.130) states that the flow of energy in the + z direction per
unit frequency interval and per unit kx volume is
Bx — i 7 T“ Ey , (41.133)
UJ oz
Ey = te~iklZ,
.d _ . _
l — Ey — k\Ey^ (41.135)
442 CHAPTER 41. PROPAGATION IN A DIELECTRIC MEDIUM
so that A
Ey = 1 + r>
i - E y = k2( l - r ) , (41.139)
M l - M a). (41.141)
The quantities (41.136) (at z = 0—) and (41.141) (at z — 0 + ) therefore must
be identical for two reasons:
0 = M l - M 2) (41.143)
or
|r|2 = 1, (41.144)
which is the situation of total reflection considered in Section 41.4.
For || polarization, the flow of energy is given in terms of
—n « E * x B = —E*By , (41.145)
41.6. PROBLEMS FOR CHAPTER 41 443
raC- i k 2( z + a / 2 )
e ik2( z + a / 2 )
2 = —a/2 z = 0
Ex = - ^ ^ - B y . (41.146)
lüJ€ OZ
Thus the flow of energy is proportional to
Re[—B*EX] oc (41.147)
with the same constant of proportionality as before. Since now the transmission
and reflection coefficients, and the Green’s function, refer to By) we may obtain
the result by simply replacing
(41.148)
in the previous form. And so, again, the ||form of (41.97) expresses the conser
vation of energy when k\ is real; generally, for this polarization, the conservation
of energy for a wave reflected and transmitted by an interface between plane
dielectrics is stated by
¿2 Cl ¿2
r e - i k 2( z + a / 2)
teik2(z-a/2)
eifc2(*+a/2)
z = —a/2 z = a/2
Find the reflection coefficient ra. What is the form of the statement of
energy conservation here?
3. Consider the same situation given in Problem 2 except that the perfect
electric conductor is replaced by a perfect magnetic conductor, for which
the tangential component of B vanishes on the surface. (In the coordinate
system given in Section 41.5, this means that dEy/dz = 0 on the magnetic
conductor for _L polarization.) Calculate the reflection coefficient for this
situation, r&. What is the form of energy conservation here?
r = l(r a+ n)
t = l(p j - r„).
Reflection by an Imperfect
Conductor
where J COnd is the conduction current and cr is the conductivity. The effect of
this current can be incorporated into the previous discussion by noting that
Maxwell’s equation for the curl of H now becomes
__ ___ 18 ^ 47t 4 tt
V x H — D H ------Jcond H------J> (42.2)
C C/t c c
where J is a source, external to the conductor, for E and H. For fields of
a definite frequency, that is, whose time dependence is given by e~lU)t, this
equation becomes
_ „ iu ( .4 tt \ _ 47t ,
V x H — ---(6 + %— cr ) E H--- J. (42.3)
c \ U) J c
This shows that the conductor can be described by an effective dielectric con
stant,
47T
e 4- i — <r, (42.4)
u>
which is complex, expressing the dissipative nature of the conductor. (This
was implicit in the considerations in Section 5.2.) According to (41.31), the
component of the propagation vector normal to the surface of a planar conductor
is then also complex:
n i/2
)
.47T 2
kz e+ i— cr - k± (42.5)
(jj
445
446 CHAPTER 42. REFLECTION B Y AN IMPERFECT CONDUCTOR
47TCT
(42.6)
u>
In the following we will assume that a is real, which can be approximately
valid only for sufficiently low frequencies, as (5.12) states. Since there is wave
propagation outside the conductor, k\ < u>2/c2, and (42.5) is, approximately,
where
6= 1. (42.9)
v27ro)(r
The z dependence of the propagation of a plane wave through the conductor is
then given by
e ik z \z\ _ e -\z\/6e i\z\/6 (42.10)
The first factor on the right side of (42.10) represents the exponential attenua
tion of the wave; the amplitude is reduced by a factor of e_1 in the distance 6.
This characteristic distance 6 is called the skin depth. As indicated by (42.1),
the current in a conductor is confined to a region within a distance ~ S of the
surface, a distance which becomes smaller as the frequency increases. For a good
conductor characterized by (42.6), the ratio of the skin depth to the reduced
wavelength of the impinging radiation, A « c/u>, is very small:
Thus, only radiation of long wavelengths (low frequencies) penetrates the con
ductor appreciably.
The discussion of transmission and reflection given in the previous chapter
remains essentially unchanged except for the replacement of kz by (42.8) inside
the conductor. Therefore, for _L polarization, the reflection coefficient for a plane
wave impinging from a dielectric (labeled 2) onto a flat conductor (labeled 1)
is, from (41.59) (again the z subscript is suppressed),
&2 — &i _ ki
r —[1 — 2^(1 *)L (42.12)
k\
^ “
&2 &2 + ” 1^
since
> f c 2. (42.13)
447
r = -1 (42.14)
d uj ic
— E _l + i —n X B i -------kj_[kx • (n X B j.)l = (external currents), (42.16)
oz c we
in which the effects of conduction currents are incorporated in the replacement
.47rcr
e —» e + i ----- , (42.17)
U)
which has a magnitude much greater than that of e ~ 1 inside a good conductor.
Under such circumstances, the third term in (42.16) is negligible, implying,
inside the conductor, r\
^ -E x + t-n x B j. « 0. (42.18)
OZ c
The derivative here is proportional to E x , since the electric field in the conductor
attenuates exponentially according to (42.10),
¿ E x = l^Ex, (42.19)
„ 1 — i ujS ^
E± ~ — ---- ~ n x B x , (42.20)
relating the tangential components of the electric and magnetic fields. This re
lation is valid for an arbitrarily shaped surface as long as the radius of curvature
of the surface is much larger than the skin depth 8. For a perfect conductor, this
supplies the boundary condition that the tangential component of the electric
field vanishes on the surface,
Ex = 0. (42.21)
The energy flow in the —2 direction just inside the conductor is now seen to
be proportional to [recall (41.131)]
- R e n - E ’ XB = R « i y ^ ( n x B i ) •( n x B j
- w5 ir I2 (42.22)
- 2 7 |B±I '
448 CHAPTER 42. REFLECTION B Y AN IMPERFECT CONDUCTOR
=i { i |B
'("=0)|1'S;|S!'(2=0)|i}' <42-25)
The tangential magnetic field is continuous since, for an imperfect conductor,
there is no surface current. Consequently, to determine the fractional amount
of power absorbed, we take the ratio of (42.22) to (42.25):
i¿8/2c
= 2k2S, (42.26)
U > /4 c &2
u>8/2c
(42.27)
ck2/4u)€2
2. Show that the rate of power loss per unit area by an electromagnetic wave
impinging on a good conductor is
^ oss _ 1 lzr\2
dS 4<r<5|K| ’
where K is the effective surface current,
K = J.
Chapter 43
Cylindrical Coordinates
eiu\r-r'\/c
Gu ( r , r ' ) = |r-lr /| ■ (43.2)
449
450 CHAPTER 43. CYLINDRICAL COORDINATES
We have seen this differential equation before; it is (41.41) with e = 1, and its
solution is given by (41.57) with r = 0:
1 ikz\z—z>\ (43.5)
g{*, *') 2 kz
if * > |kx|,
k, (43.6)
*V *i - if 7 < |kxl-
In statics (uj = 0) the second possibility in (43.6) occurs. [See (13.24).] Without
loss of generality, we set r' = 0, and define k = u>/c > 0. Then, comparing
(43.2) with (43.3) and (43.5), we obtain the identity
g i k x T j L e i X p COS (f>
Making use of the integral representation for the Bessel function Jo, (16.4),
p ik y / p 2+ Z 2 fO O eW k 2- \ 2\z\
= = t / A dX Jq(Xp ) —■ = = = = -. (43.11)
yjp 2 + *2 Jo Vfc2 - A2
/ dAJo(Ap)e-Al*l. (43.12)
Jo
43.2. THREE DIMENSIONAL FOURIER REPRESENTATION 451
The above expression is not well defined since the integrand has a singularity
wherever |k| = w/c, corresponding to real wave propagation. To assign meaning
to this expression, we must specify the boundary conditions. Recall that the
retarded and advanced Green’s functions have the explicit forms given in (31.43),
(dk) eik
Gret(v,r') = 4 * J g (43.19)
)3 j^2 _ ^v+iey
^+0
We can quickly check the validity of this result by explicitly evaluating the
integral using spherical coordinates [recall the spherical average of a plane wave
seen, for example, in (38.45)]:
47vk2dk sin kR
2
(2 tr)3 kR ~
452 CHAPTER 43. CYLINDRICAL COORDINATES
cos kR
IL
w RdR
jL
J[°°
0
dk
C
= = r <43-21>
which is the retarded Green’s function given in (43.14).
J
}_eiwr/c _ 47r (dk)
(43.22)
r W £->+0
Note that if we now perform the integral on kZ) and use the integral represen
tation for Jo, (43.10), we rederive the identity (43.11). [See Problem 1.] If,
instead, we rewrite (43.22) in the form of a 1 + 2 dimensional decomposition,
and first integrate over the angle associated with kj_, we find
1 ,iu>r/c _ 4jr f°° Xd\ Jo(X p)
(43.24)
r 2x Jo 2ir X2 _ Í(íü±i¿)2 - ife2
43.3. HANKEL FUNCTIONS 453
id 2 Id Jo(Xp)
dXX = 0, (43.25)
\dp2 + p d p + K A2 — K2
for p ^ 0. We may also obtain this result explicitly from the differential equation
satisfied by Jo, (16.17), since
( d2 i d 2\ r dXX Jo(*p)
W + pdp + K ) J 0 A2 — K2
1
l dAAA2 - « 2
(—A2 + K2)J0(Xp) = 0, for p # 0, (43.26)
Jo(Ap)
f XdX
A2 _ [ ( » ± i e ) 2
(43.28)
When oj/ c < \kz\, the square root in (43.28) becomes imaginary, — w2/ c 2,
and in this case, the Hankel function becomes the modified Bessel function,
(18.3),
in
Hn1^ ( \ i—r — k2 , Ko (43.29)
k* - ^ P
In terms of the Hankel function defined by (43.28), the equality (43.24) reads
\^iu xr/ c
(43.30)
poo eiky/p2+z2
Í7tH0 (kp) = / dz . . :♦ (43.32)
J-00 \/p2 + z2
We will now use this representation to find the asymptotic form of the Hankel
function for kp ^ 1. As we will justify a posteriori, the main contribution of
the integral comes from small values of z jp ) so that we may expand the square
root in the exponential as
p2 + z2 (43.33)
1 f°°
Í7rHQ1\kp) ~ - e ikp / d z e - kz2/2ip
P J-00
ir/4) (43.34)
The consistency of our approximation hinges on the fact that the dominant
contribution to the integral, (43.35), comes from small values of z/p, for which
< 1. (43.36)
H ^ \ k P ) = e~r ± * V j -H ^ (k p ), (43.38)
with the introduction of the mth order Hankel functions of the first kind,
-i zfcm
l(d _ .d_
eim4,H£ )(k p ) = imeim* H £ \ k p ) = H ^ ik p ), (43.41)
^ k \dx 1dy
for m > 0 (m < 0), respectively. Notice that since the whole construction is
identical to the Jm construction, Hm^ satisfies the same differential equation,
(16.33), and the same recurrence relations, (16.71) and (16.72), as does Jm.
sin y jt 2 + x 2
M t) = \ f dx- -------- -
T* J Oo y/t2 + x 2
2 [°°
— I dO sin(t cosh 0)
TCJo
_ 2 f° sin ty
dy
~ *Ji \Jy2 - 1 ’
456 CHAPTER 43. CYLINDRICAL COORDINATES
Reconcile the third form of J0(t) above with the fact that Jo(0) = 1.
= Jm(kp).
4. In general, Hankel functions of the first and second kind are defined by
= M x ) + iN „(x)
H l2\ x ) = M x ) - iNv(x).
Show from Problem 16.10 that these have the leading asymptotic behav
iors as x —> oo:
00 1
G *(r-r')= ^ eim(*~4'>Prn(p,p'),
m ——oo
and solve the resulting differential equation for pm, by using the
discontinuity method. Note that the boundary condition for large p
is to be an outgoing wave, so the appropriate behavior there is given
by Hm\kp).
6. Show that the relation between the modified Bessel function K m, and the
Hankel function H ^ } is
Use (43.41) and (18.65). This result is already given in Problem 18.3.
Chapter 44
Waveguides
The theory of microwave cavities and their application to radar technology was
largely developed at the MIT Radiation Laboratory during World War II. A key
player in that development was Julian Schwinger. In this chapter we consider
some o f the elementary aspects of electromagnetic waveguides, which we will
see is closely tied to the electrostatic mode theory discussed in Chapter 25.
Bz -V-nxE, (44.1)
ic ^ _ .4tt t
Ez --- V •n x B — i — Jz (44.2)
LUC UJC
VjJ A irm
Ex = ------ ( 1 _l H— r V x V x * n x B - i — V ± J Z, (44.3)
dz c \ u)¿e ) uc
zcoe / c2 \ „ 47T
---- (
I jl + i - V x V x •n x E -------- nxJ, (44.4)
c \ u ¿e J c
where we have used the gradient operator V j. in place of ¿kj_, that is, we
are undoing the transverse Fourier transform. Now in the transform space the
two independent transverse vectors are kj_ and n x k i , so correspondingly, the
trasverse fields Ex and B x can be written in terms of two scalar functions each,
Ex = - V ± V ‘ - V x n l / " , (44.5)
Bx = V x n / '- V x / " , (44.6)
459
460 CHAPTER 44. WAVEGUIDES
When we insert these constructions into (44.3) we obtain, for the Vj_ and the
V x n components,
4 - V = — ( l + - ¿ - V i ) / ' + «— X (44.10)
OZ C \ L )¿ 6 J LÚC
V x - n x J = - V i J ", (44.13)
- V . J j . = V\J'. (44.14)
A ;- = ¿ íü v _ 1l r , (44.15)
dzz e e
d_
-I" = i— ( 1 + 4 - V i ) v " - — J". (44.16)
;
dz' c \ u¡¿e J c
We thus have four coupled differential equations for the four functions I",
V ' , V " . Once those functions are found, the longitudinal fields can be found
from (44.1) and (44.2):
Et = - i — Jz , (44.17)
we ue
Bz = - i ^ 7 \ V " . (44.18)
Because there is no coupling between the sets V* and / 7/, V n we have two sets
of independent modes, characterized by the values of Ez and Bz , respectively.
E modes: Ex = - V x V ' ,
Sh
ii
$
nxB = V i / ' , Bz = 0 , (44.19)
H modes: BX = - V x / " ,
S3
II
o
( V i + ■y1)<j> = 0, (44.23)
<f)a = 0 on S, 73 0 (44.24)
<pa = constant on 5, 7^ = 0. (44.25)
— ^ = 0 on 5. (44.26)
± </>= 0.
V2 (44.32)
The boundary conditions for the T modes are that <j> is constant on 5, or
d<¡) dip
= 0, (44.33)
Os dv
where ds is a tangent vector to the surface. If we write the differential relation
between <j> and ip in Cartesian coordinates,
d , d . d , d
(44.34)
r J = d i* ' s i* =
we see that these are just the Riemann conditions satisfied by the real and the
imaginary parts of an analytic function; that is, because
44.2. BOUNDARY CONDITIONS 463
Figure 44.2: Cross section of waveguide consisting of inner and outer conductor.
we see that
<j>4- = F (x 4- iy) (44.36)
is an analytic function.
How general is the existence of T modes? Because they are characterized
by a harmonic function (that is, a solution of Laplace’s equation), if the region
enclosed by S is simply connected, 4> = constant, and hence there is no electric
field. The proof of this assertion follows from the divergence theorem:
/ d S ( V x <f>f
Js
<f)c f dSV2
x <f>= 0, (44.38)
1. For parallel plates, parallel to the x axis, the analytic function is merely
a linear one,
dll’ d
- = - _ v * = o, (44.40)
(44.43)
d v ~ dp ~
44.3 Modes
Let us summarize where we are. There are three types of modes: E, H (with
the eigenvalue 7 ^ 0 ) and T (with 7 = 0 ). The modes are characterized by the
equations
( V i + 7 2) { $ } = 0. (44.44)
For each mode, labeled by a, the ¿-dependence is given by the functions / ' ,
(E) or V¡¡ (H), which satisfy the equations
dz c a’ C ( * - u 2e ) a'
Note that if 7 ^ = 0, these two sets of equations are identical, that is, the E and
H modes reduce to the common T mode, where
(44.48)
the solutions of which are e±tkz with k = Uyfe/c, reflecting the expected speed
of light in the medium, c/y/e.
Let us henceforth confine ourselves to a situation where there are no T
modes. Consider a rectangular cavity, with sides a (along the x axis) and b
(along the y axis). For the E modes, where </> must vanish on the boundaries,
we have
m'KX mry
<f>mn oc sin ^ sin - y - , m ,n = 1 ,2 ,3 ,— (44.49)
44.3. MODES 465
o /m7r\ 2 /ri7r\ 2 9
(44.50)
Tmn - ( ~ ) + ( t ) - 7lu
a relation noted in Chapter 19. In contrast, for the H modes, the normal
derivative must vanish, so
. m'Kx niry ~ ,
Wmn cos cos , m, n = 0 , 1 , 2 , . . . (m, n not both = 0). (44.51)
a o
9 /m7r \ 2 /n7T\ 2 9
Tmn = J + ( - j - J > Tio, when a > b. (44.52)
( ¿ + k , - /' ) { r } = 0’ <44-53>
for which the solutions are
k > -, (44.55)
a
A < 2 a, (44.56)
Ex = VxXnV>, (44.57)
Of course, the above relation between E and B follows directly from Maxwell’s
equation
V xE = ~ B . (44.61)
H: = 0 when p — a. (44.63)
dp
The solutions to the mode equations are of the general form
= (44.64)
7 ^ 0 .= 2.405, (44.67)
7 o2a = 5.520, (44.68)
y'n a = 3.832. (44.69)
J i(x) = (44.70)
so 7o„ = Tin) so
7^ 0 = 3.832. (44.71)
A new situation emerges, however, for m = 1 . We could estimate the zeroes in
this case by using the method described in Section 19.4; a better approximation
is given in Problem 44.3. The result is
7 n a — 1-8412. (44.72)
Thus the lowest mode for a circular waveguide is the H mode with m = n = 1,
H u. The cutoff wavelength for such a waveguide is Ac = 3.4125a.
E = -y-rVx ( v x - / ' n ] ,
e(z) V « /
B = Vx(J'n),
44.4. PRO BLEM S FO R C H A P T E R 44 467
2. This problem refers to a coaxial waveguide. Let the inner radius be 6 , the
outer radius a. Assume a —b <C a, and find the cutoff wavenumbers 7 ',
7 " for the E and H modes. [Hint: You do not need Bessel functions. Look
at the differential equations for the mode functions. Which mode has the
lowest cutoff wavenumber (other than the T mode)?]
3. This problem illustrates an approximate method for finding the lowest
cutoff wavelength for a circular waveguide, that is, for finding the first
zero of J [.
(a) Derive the following formula for the eigenvalue from the differential
equation satisfied by the mode function ip from (25.26):
Notice that if this is varied with respect the ip we recover the differ
ential equation
d2 Id / X2 m2
ip = 0.
5 ? + r s + (T a ) - - F
V'(o) = o, v-/(i) = o.
(Where do these boundary conditions come from?) [Actually, the
estimate so obtained is an upper bound to 7 . See Chapter 25.] The
function ip(t) = t ( 1 —t 2/ 3) satisfies these boundary conditions; show
that if it is inserted into the formula for 7 it gives
7"a « 1.842265.
(c) Show that if, in the spirit of Section 25.2, we now write the differential
equation as
1£ 1_
'tp — —(7 a ) 2ip « —(7 a)2
d t2 t dt t2
7 "a = 1.841198,
4. Show generally that in source-free regions the electric and magnetic fields
can be represented by two Hertz vectors II and II* ,
e = vx(vxn) + —vxn*,
C
ÍüJ€
b = -----vxn
c
+ vx(vxn*).
5. Show that the fields in a cylindrical waveguide are represented by Hertz
vectors where
n - Ez
“ fcf - ew2/c2 ’
n* Bz
n &f — ew2/c 2
What is the connection between II, II* and the functions l7, V7, V nl
n X A a(r) = 0 on 5,
where n is the normal to S. If we use real mode functions, show that the
following orthonormality condition holds,
V •[15(r — r')]x = 0 .
44.4. PROBLEMS FOR CHAPTER 44 469
8. Now let the vector potential be given in terms of the current by a Green’s
dyadic G ,
A (r,w ) = / ( d r / G ^ r 'j w ) ♦-J (r ',w ).
Jv c
Show that, in vacuum,
9. Prove that
/ \ . v -' A a(r )A 0 (r')
G (r, r ;o>) = 4 tr > ,2 ’ ,\ 2 •
“ ¿ 2 _ ( w / c )2
Show that, with u —>u> + Oi, the resulting Green’s dyadic is the retarded
function,
0, t - t ' < 0,
G (r,t;
,t] r ' , 0 = | 4 x ^ 0, A a (r )A a (r/)^-sin[fcac ( f - t ') ] , t - t ' > 0.
10. A current J(r, i) is turned on and eventually off. Demonstrate that the
total amount of electromagnetic energy created in the cavity is
i p 2
A E = 27T^ / (dr) dt A a(r)etkoiet ♦J(r, t) .
IJ
2 _ /(d r ) ( V x E ) 2
or
~ /(d r ) E 2
2 _ /(d r ) ( V x H ) 2
" /(d r ) H 2
Verify that the first expression is stationary with respect to variations that
do not violate the boundary condition, while the second form is stationary
with respect to unrestricted variations.
1
H mode :
V cavity Q guide
1 1
E mode :
Q cavity Qguide
where 6 is the skin depth, and k is the guide propagation constant for
the mode in question. [For an oscillator, Q is defined as the ratio of the
resonant frequency to the dissipation constant, Q = ujq/j . Here, this is
27T times the average energy stored in the cavity divided by the power lost
per cycle.]
Scattering by Small
Obstacles
dv
(45.1)
mS = eE"
(as long as |v/c| <C 1). An accelerated charge radiates, which gives rise to
the “scattered” radiation; the power radiated is given by the Larmor formula,
(32.25),
<*•»
This scattered power is to be compared to the incoming power per unit area of
the plane wave, which is, in vacuum, given by the magnitude of the Poynting
vector,
|S| = ¿ | E X B | = ¿| E | 2. (45.3)
Fscatt
= (45.4)
|s| ’
(T
471
472 CHAPTER 45. SCATTERING B Y SMALL OBSTACLES
which is the effective area presented by the obstacle to the wave. For the
situation considered here, the cross section is
cr (45.5)
the so-called Thomson cross section. For an electron, this cross section is con
veniently expressed in terms of the length, ro,
g 2
a « - X 10 24cm 2 (45.7)
o
What is the angular distribution of the radiation scattered by the charge?
We recall the nonrelativistic expression (32.23),
d-Pscatt
(45.8)
d il
which gives the power radiated in the direction n per unit solid angle. Using
(45.1), we rewrite this as
dPsc 1
r(n X E )2, (45.9)
dfi 4tt m 2 c3
or, introducing e as a unit vector along the direction of E,
dPes catt 1
E 2[l — ( n « e ) 2]. (45.10)
dfi 47Tm 2 c3
Notice that there is no scattering at 0 = n/2 for the second choice, which just
re-expresses the fact that there is no radiation emitted in the direction of the
acceleration. If the incoming radiation is unpolarized, we have the average of
the two possibilities given in (45.11), so that
1 + cos2 9
1 — (n •e ) 2 (45.12)
2
45.2. SCATTERING B Y A BOUND CHARGE 473
The differential cross section, the effective area for scattering into a given ele
ment of solid angle, is defined by
d(T d P SCQ.ttl d il
(45.13)
dñ= fsj ’
For an unpolarized wave, the differential Thomson cross section is then
dcr / e2 \ 2 1 + cos2 6
(45.14)
dñ = \ ^ J 2 *
The total Thomson cross section, (45.5), is recovered when (45.14) is integrated
over all solid angles,
a= [ ^ d S l. (45.15)
J dS2
ip j» »
m — r + mu or 4- m y — = eE. (45.16)
at¿ at
Suppose the electric field is due to an incoming light wave of a definite frequency
w,
E (t) = Re Ee- *"* = \ (V e~iut + E V 'wt) , (45.17)
it
so the deviation of the particle from the center of the harmonic force is given
by (5.18):
474 CHAPTER 45. SCATTERING B Y SMALL OBSTACLES
2 e2 s2
-Pscatt —3 (45.19)
involves
e„ w2 Ee~iut
r = Re 9 9 (45.20)
m cuq —u)¿ — %yu)
We are interested, not in the instantaneous scattered power, but rather in its
time average over one cycle. For a typical time varying quantity,
Thus, the time averaged power scattered is, from (45.19) and (45.20),
2 e2 1 e2 |EP
P sca tt — ^ o ~ (45.23)
scatt 3 c3 2 m 2 U (u>2 — w2) 2 + ( t ^ ) 2 ’
(45.24)
both expressed in terms of the complex amplitude E. The resulting cross section
is
_ 87t Í e2 or
(45.25)
a 3 \ me 2 - w2) 2 + ( 7 U>)2 ‘
This reduces to the Thomson cross section for a free particle when both u>o and
7 are zero. More realistically, the above cross section approaches the Thomson
limit,
*-*<■> T h , (45.26)
when the frequency of light is large compared to the natural frequencies of the
bound system,
u > » a ; 0,7. (45.27)
Physically, in this limit, the time scale is set by the period of the incident light;
if this is very short, the electron does not have time to be influenced by internal
forces so it behaves as though it were free.
In the opposite limit of very low frequencies,
(45.29)
which is called the Rayleigh cross section. In this domain, the highest frequencies
are scattered the most strongly. This frequency behavior is presumably one of
the reasons the sky is blue.
We have the situation of resonant scattering when uj = u>o, at which point
the cross section becomes
(45.30)
For small damping, u>o 7 , (45.30) is much larger than the Thomson cross
section. Of course, 7 cannot be zero; energy is necessarily dissipated because, if
for no other reason, the particle radiates due to its acceleration (recall Section
35.3). We will return to this issue shortly.
(45.31)
(45.32)
A> a (45.33)
(45.34)
(45.35)
Again we see the uj4 dependence of the cross section characteristic of the scat
tering of long wavelength radiation.
476 CHAPTER 45. SCATTERING B Y SMALL OBSTACLES
F=|^v. (45.36)
2 e2 2
“ 3^ v = ~ m> v , (45.37)
where here yr is the radiative part of the dissipation, which must be present
in order that energy be conserved. It coincides with the previously determined
damping constant for radiation by a harmonic oscillator, (35.56), as long as the
characteristic radiated frequencies are near the resonant frequency u>q. Recog
nizing that there may be other forms of dissipation (for example, collisions), we
write the total dissipation constant as
2 e2
7 = x — 3 W2 + 7d = 7r + 7d . ( 4 5 .3 8 )
ó mc°
The energy removed from the incident electromagnetic field feeds both j r and
jd . The rate of energy transfer from this field is, according to (45.18),
e2 1 , —VjJ
( e E « v ) = — -| E | 2 R e ^2-------2----- •— * (45.41)
m2 uJq —u>¿ — «70 ;
Consequently, the total power removed from the incident field (not just scat
tered), is
yu>
(45.42)
(u,2 - W2) 2 + (Tu;)2 ’
which must exceed the scattered power, (45.23),
PtOt _ e2 yu>2
(45.44)
|S| atot * me {u l - w2) 2 + (y u )2
If we use (45.38),
(45.45)
_ A 3 2 7r7
(45.46)
a to t ^ 2 ° (u >1 — a ;2 ) 2 + (7 a ;)2 '
^ < 1, ( 4 5 .4 8 )
7 ~
We now break up Otot into two pieces corresponding to the two channels for
energy loss, given in (45.38),
0"scatt — 47T-C
2 (a)l - u>2) 2 -f- ( 7 a;)2
__ 87t / e2 V LO
(45.51)
3 \mc2) (w2 “ ^ 2) 2 + ( j v ) 2
[which is the same result found in (45.25)], while the dissipation cross section is
7r7d
0rdiss = 47T--C (45.52)
2 (u>q —oj2)2 + (Tw) 2
If we ignore 7 , and consider the high frequency limit u u>o, the scattering
cross section becomes
1 0rTh
0-scatt — 0”Th (45.53)
1 + ( 7 /w ) 2 l + [(2/3)(e 2/m c 2 ) ( l /A )]2 '
478 CHAPTER 45. SCATTERING B Y SMALL OBSTACLES
If X e2/me2, the correction to the cross section is very small. But if we were
to take literally the limit o f very short wavelengths,
which has the same form as the upper bound (45.49). This limit is completely
unbelievable, however, since quantum effects become significant already at the
considerably larger distance
h e*
= 137- 2 (45.56)
mc me *
Partial-Wave Analysis of
Scattering
r X ( V x B ) = V ± ( r . B ) - B j _ - r — Bj_. (46.3)
If we then divide by r, the transverse parts of the two Maxwell equations above
are
_ „ id 1\ _ vjj r „ 47r r ,
1V ±B r — l —— |— ) Bj_ — ------c- x E H-------- x J i , (46.4)
\or rJ c r c r
479
480 CHAPTER 46. PARTIAL-WAVE ANALYSIS OF SCATTERING
„ IU v
Ex = -----xB. (46.5)
c r
i r . V x B = i r x V * B = r x V « -B = — V x - «B = — V •- xB, (46.6)
r
to give
~ r ^ . 4tt t
—V i •- x B — ----- cEr -\----- J r, (46.7)
i
r c c
__ r „ iu ^
- V x •- x E = — B r . (46.8)
r c
We can combine these equations to obtain
Ex = - V x V " - V i X - r , (46.11)
^ i 9 ¿w A c2 o\ w i4w _
E modes: -F = _ (^1 + _ V i J / + — X, (46.15)
1 _ i_ J l ( . a d i d2
vl = sin 9 89 \Sm°d9
+
sin2 0 d<f)2m
(46.21)
(46.22)
and similarly for V , V ", I". Then the mode equations become
^2 ¡a _l i\\ ¿4 ir
E modes: = 7 (l , + (4624>
(46.25)
¿ U ) = 7 t W - 7 4 ( ') -
C2 / (/ + 1 )
H modes: ~ ~ ~ f l ---- " (46-26)
dr ‘ v ' c V w
(46.27)
Let us see what the differential equations are when we are outside the sources.
Let us set J (r) = 0 and take e to be a constant. Then we can turn the above
coupled system of first-order equations into the second-order one,
or / ( '+ ! )
+ T 6' (46.28)
dr2
E x = 0 for r = a. (46.29)
'S'
is
'’S'
o'
o
II
II
u
II
II
or
H modes: V L (r = a) = 0 , (46.31)
E modes: dp —0 (46.32)
d ¿I,m r = a
u ~
482 CHAPTER 46. PARTIAL-WAVE ANALYSIS OF SCATTERING
The solutions of the differential equation (46.28) are rji(k r), where j¡ is the
spherical Bessel function, and k = u 2e/c2. For the two modes, then, the fre
quencies are determined by
H mode: ji(ka) = 0, (46.33)
d
E mode: — (rjA k r)) = 0. (46.34)
dr
Let us make quick estimates of the characteristic frequencies. First, in the
spirit of Chapter 19, we recall the asymptotic behavior
cos(t —(/ + 1)7t/2)
3l{i) ~ oo, (46.35)
where we are never concerned with l = 0 because then V^Yoo = 0. Thus, the
lowest H mode has
k"a = 7'/=1}n=1* y = 4.71. (46.37)
And, for E modes, where it is the zero of ^ tj\(t) that is sought, we have ap
proximate zeroes at
k'a — 7 / - i >n=0
4^
II
II
Let us improve on these estimates using the methods of Chapter 25. The
eigenvalue functional corresponding to the differential equation
*(/ + !) = 0 (46.40)
r2
is
(46.41)
The small t behavior of the solution is u ~ t1+1. For the H modes, consider the
test function, for / = 1 ,
u = t2{ l - t 2). (46.42)
When this is inserted into the functional (46.40) we obtain the upper bound
7 " = 4.74, which is worse than our previous estimate. (The exact answer is
4.49341.) For the E mode we can try
(46.43)
for this has vanishing derivative at t = 1. This does give a considerably improved
bound, 7 ' = 2.7514, compared to the exact root 2.74371. In Problem 46.1 these
bounds are improved by iteration.
46.3. SPHERICAL HANKEL FUNCTIONS 483
(V 2 + k2)<j> = 0, k2 = ^ j , (46.44)
so, in the spirit of Chapter 18, we can regard zV'/Ar as a symbolic unit vector
when acting on etkr /r'. Thus, we rewrite the Taylor expansion (46.46) as
p«A;|r—r'| A k r'
e _ eikr*(iV /k)C___ (46.48)
r —r
Now recall the expansion of the plane wave in terms of spherical Bessel functions,
(21.62), or (21.73),
OO l / \
eik •:' = 4 ttJ 2 E Y *m 0 ) Ytm ( T i'ji(kr). (46.49)
J=0 m ——l ' '
e ¿fc|r-r'|
rpf = T , 4’ ‘'M tr)Y ^ (S) r lm (iv ') (46.60)
|r• - r 'l ~ Im
What is the latter symbolic operator? We recall the generating function for the
spherical harmonics, (20.25),
i = <46-51>
(a
lm
/ i .\ / r ^i
V/m
( H =G) r"n”(?) (G¿) ■ <46 53>
and so we have the expansion
e ik\r-r'\ J gikr1
lr '
(46.54)
Let us define a set of functions by
(46.55)
these functions, which are defined analogously to the spherical Bessel functions
ji (see Problem 21.3), are called the spherical Hankel functions. As with the
cylinder functions we have the decomposition into real and imaginary
parts,
h i ( i ) = ji ( t ) + irn(t), (46.56)
the imaginary part being the spherical Neumann function. Thus we have de
duced the addition theorem
the spherical analog to the addition theorem for the cylinder functions, (16.69).
We might also note the leading asymptotic behavior of the spherical Bessel
function:
Jt «(t —(/ + 1 ) tt/ 2 )
h,(t) ~ ( -* ) — = * » 1, (46.59)
it t
the real part of which coincides with (46.35).
46.4 Scattering
We set the stage for our discussion by recalling the result from the previous
chapter for the scattering by a dielectric sphere. In particular, if we take the
limit e —►oo of (45.35) we would expect to recover the situation of a perfectly
conducting sphere, which limit gives
87r a 6
(46.60)
46.4. SCATTERING 485
However, the well-known result for the scattering cross section by a perfectly
conducting sphere is that of Gustav Mie and Peter Debye (1908-1909),
IOtt a 6
cr = (46.61)
What is the reason for this discrepancy? It is simply that in this case there is also
magnetic dipole scattering, which gives the additional contribution 27ra6/3A4.
(See Problem 45.1). So this poses a nice problem for our general formulation.
Consider a circularly polarized electromagnetic wave propagating in the 2
direction, for which
Er = l . E = Í ± Í Ü , ‘ *«. (46.64)
(46.65)
r r
We express this in terms of spherical harmonics and spherical Bessel functions
as follows: First we recall the expansion of the plane wave, (21.62),
We note that
kx d f(k )
m 0, (46.68)
dkx kx=0 k dk kx= 0
so in taking the derivative in (46.67) we can neglect its action on ji(kr). In
working out
á T + ' Jdkv
-V -™ ® (46'69)
we are looking for stuctures like (kx + iky)/k = sin Oe1^ in Y/m(k), so that picks
out m = 1 ; from (20.30) we therefore have only
¡21+1 i /+i
k,-ik, ( d \ ‘" (r-iy^+iy
(46.70)
486 CHAPTER 46. PARTIAL-WAVE ANALYSIS OF SCATTERING
121 + 1
(
1
1(1 + 1). (46.71)
ky 4ir
Thus, the radial electric and magnetic fields are
T - ¿ V W T / i ( r ) ^ / + ^ r j¡(k r ), (46.74)
21+ 1 ji(ka)
V n = —i ^ V+7rilYn(r) rji(k r) — rhi(kr) (46.80)
Ki + 1 ) hi(ka)
i=i
hi represents the outgoing spherical wave. Let us express the coefficient of the
scattered wave,
jl(ka) _ ji(ka)
(46.81)
hi(ka) ji(ka) + ini(ka) ’
46.4. SCATTERING 487
ji(ka)
tan 8" (46.82)
ni(ka) ’
so that
ji(ka) //e i6i
—i sin 6f/ (46.83)
hi(ka)
The scattered wave, the second term in (46.80), is then expressed asymptotically
by the large t behavior given in (46.59):
ikr °° ___ / o/ i i
Kc ~ - i — E (i)y sin 6¡'eis" . (46.84)
Similarly, for the E modes, the mode function which satisfies the appropriate
boundary condition, and therefore contains both the incident and scattered
waves, is
2/ + 1 (rji(kr)Y\r=a 1
E V ^ i'Y n (r) rji(kr) — rhi(kr) (46.85)
i=i ¡(1 + 1 ) (rA,(A:r)),|r=<1J '
= *“ *». (46-86)
ikr 00 / 2 / 4- 1
4 ~ Kc ~ — E (f ) y j y f j j sin • (46.87)
which consists of an incident part and a scattered part. The radial component
of the Poynting vector is proportional to
and further
1 00
psc OC J 2 y ' / 4 ,r ( 2 J + l)(s in 2 ó'i + sin2 6 " ). (46 .95 )
9 _ 00
a = T 2 ' H i 21 + W 85» 12 s! + sin2 8" )- (46 .96 )
K 1=1
N ow we want to look at tw o special cases, where ka = a/\ is either m uch
sm aller, or m uch larger, than one. In the form er case, where ka <C l , the
dom inant term is / = 1, for which
. /iX d sin t t
(46 .97 )
■»<'> = n t < < 1 ’
/fv d cost 1
(46 .98 )
" , ( , ) = d« 1 * ( >’ , < i '
so
tan 8 " « 8 " « — i t 3 = —i ( & a ) 3 . (46 .99 )
o 0
Sim ilarly,
w _ (< V 3 )' _ 2 3
(46.100)
W l -(_ !/< )/“ Z ’
°r 2
«í = ¡ ( k a f . (46.101)
If we add together the contributions o f the E m odes and the H m odes, we find
for the cross section in the long wavelength lim it
2 tt„ [ / % , A 2 /1 n3\ 21 10 a6 ,
(46.102)
17 ~ W 3 (3 (U>) + ( f (fca) J = T F '
46.4. SCATTERING 489
OO
+ = (46 108)
Therefore,
This result is not unexpected. In the short wavelength regime we should see
just the cross sectional area presented to the beam by the sphere, 7Ta2. The
extra factor of 2 comes from the fact that we have a shadow behind the sphere.
The sphere must throw enough energy forward (with opposite sign) to cancel
the incident light beam. Thus we get
e •Esc ~
r
where e is the polarization vector for the scattered wave. Obtain an
expression for the scattering amplitude in terms of Legendre’s function,
P/(cos0), and the phase shifts 6¡ and 6'/. Show that the differential cross
section is given by
to = IWI2-
Show that if we sum over final polarizations and integrate over all angles
of scattering, we obtain the total cross section (46.96). Show that the
optical theorem is satisfied,
3. (a) From the discontinuity conditions (26.22) [for the surface current K]
and (11.64) [for the surface charge density a] show the conservation
of charge,
V_l •K — iujar = 0.
d — u Einc(r = 0),
Diffraction I
V xB = ÍÉ , V •E = 0,
c
—V x E = Í b . (47.1)
c
-V x ( V x E) = - V ( V . E ) + V2E = ^ E , (47.2)
or
(v ! - ? ^ ) e=o' <47-3)
Specifically, we consider a plane polarized wave normally incident on a perfectly
conducting screen, with an aperture. We choose the coordinate system (see Fig.
47.1) such that x is in the direction of propagation and z is along the direction
of polarization of the incident wave. From our choice of coordinate system,
far from the source, the incident electric field has only a ^-component. For
491
492 CHAPTER 47. DIFFRACTION I
z
X
Figure 47.1: Plane wave incident on an aperture. Here, the z direction is out of
the page.
x = 0
We are interested in finding the electric field to the right of the screen (x > 0),
subject to the boundary condition that the tangential electric field vanishes on
the screen (since it is a perfect conductor).
The solution to the differential equation (47.4) can be expressed in terms of
Green’s function, for x > 0, which satisfies (31.29),
The electric field to the right of the screen can now be obtained in terms of its
boundary values by integrating (47.8) over the volume to the right of the plane
{x* > 0) and using the divergence theorem:
+ [ dS' (47.9)
Js0
The first integral extends over the entire x' = 0 plane, the minus sign appearing
because the outward normal to the volume is in the —a?' direction. The second
integral is over a surface which can be taken to be a hemisphere with radius
tending toward infinity. Far from the aperture, which acts as the source of the
scattered wave, the leading behavior of the electric field, an outgoing spherical
wave, is the same as that of the Green’s function:
pikr‘ p ikr'
G (r', * ) - * — , m ~ — (47.10)
ik, (47.11)
drf
so that the integral in (47.9) over the hemisphere at infinity vanishes. Physically,
this follows from the fact that the source of the electric field is not at infinitely
remote points, but at the aperture.
Now we incorporate the boundary conditions that G = 0 on the entire x' = 0
surface, while E = 0 on the surface x f = 0 except for the aperture (since we are
talking about Ez here). Thus the electric field to the right of the screen is
The derivative o f Green’s function can be evaluated from (47.7) by noting that,
for the first term,
¿ =- h <4713>
since |r' — r| depends on x f — x y while in the second term,
A - A (47.14)
dx' dx ’
= cos 0, (47.16)
ox r
the normal derivative of the Green’s function is seen to be
ik pikr Í i
E (r )~ -f-c o s 0 — / dS' e~tkn' r E { t'). (47.18)
This expression is valid for any shape of the aperture. This form holds for
Ey as well as EZ) but for Ex one must integrate over the whole plane. In the
following sections, we will apply this result to discuss the diffraction by a circular
hole, a slit, and a straight edge. We will do this in the approximation that the
wavelength of the radiation is small compared to the dimensions of the aperture.
Consequently, the wave travels mostly forward, with small angular deviation.
We may further reasonably suppose that only Ez is present, and that it has
very little ^-dependence,
3E
Ex & 0, Ey « 0 , =► “a f » 0. (47.19)
In the same approximation, we may assume that the field in the hole is just
that of the incident wave,
-^hole ^ E/'mC) (47.20)
which is valid except near the edges. Since only relative amplitudes enter into
our discussion, we take the normally incident wave to have unit amplitude,
-S’inc := 1.
47.2. DIFFRACTION B Y A CIRCULAR APERTURE 495
n
ikr pa p 2 tt j ,
E ~ - i k ----- / pdp _ z e-ikpsine c o s t ^ (47.21)
r Jo Jo 27t
where # is the angle between n and the ar-axis, and <j> is the angle between r'
and the projection of n on the plane defined by the hole (see Fig. 47.3). Because
the integration over <j) is identified as Jo, according to (16.4), the electric field
is proportional to
pa pkasm.9
Jq p d p j a{k p s m 6 )= Jo zd zJ o(z). (47.22)
The remaining ^-integration can be performed with the aid of Bessel’s equation
of zeroth order, (16.18),
d_
Z -M z ) + zJ0(z) = 0, (47.23)
dz
that is,
fz d
j d z z J 0(z) = - z — J0(z) = zJi(z), (47.24)
(47.25)
~ d z Ja('Z^ = J l^ ’
which follows from (38.39) and (16.25). Thus the relative diffracted electric field
is
eikr a
E ~ —ik ----- — :—r J i (Arasin#), (47.26)
r Arsin# v
496 CHAPTER 47. DIFFRACTION I
(47.27)
r had
- (b n W ( Jl ( ka$) (47.29)
df2 \ kad
We anticipate that the total cross section, in the small wavelength limit,
k a ^ 1, or (47.30)
will be
ira (47.31)
since this is just the geometrical area of the circular hole, and we are in the
domain of “geometrical” optics. This can be shown explicitly by integrating
(47.29) over all solid angles. The element of solid angle is approximately
since physically the wave is mostly scattered near the forward direction. Math
ematically, this last fact is evident because the important values of 0 are those
for which the argument of the Bessel function is of order one, that is,
0~ = - < 1. (47.33)
ka a
f w/2
•r a 2 ’ l
(47.34)
(47.35)
is true.
47.2. DIFFRACTION B Y A CIRCULAR APERTURE 497
To prove (47.35), we first need an integral representation for J\. This can be
achieved by starting from the addition theorem for Bessel’s functions, (16.69),
OO
setting z = z',
(47.39)
2z (47.40)
1 1, 4
In • x In
sin ■ 2 (1 - e - ‘> ) ( l - é * )
(47.42)
1 I - - +
« ----- 6 d ó iz ------ ------- e 9
2m J 2
= 2’ for z < 1» (47.43)
which, of course, follows from the series representation (16.27). Therefore, very
near to the forward direction, the differential cross section is
da ^ 1
(47.44)
dñ * 4
Comparing this with the total cross section, (47.31), we see that, since a/A >> 1,
the diffraction is mostly forward, consistent with our assumption.
For scattering angles such that z = kaO ^ 1, we make use of the asymptotic
behavior of J\(z), which is given in (16.39) and used as a starting point for the
asymptotic analysis of the zeros of the Bessel functions in Sec. 19.4. A direct
derivation can be obtained by using the integral representation, (16.30), with
(j) = f t — 7r/2:
For z 1, the main contribution to (47.45) comes from the region near the
stationary phase point, ft « 7t/ 2 ; therefore, we reintroduce <j> = ft — 7t/ 2 to
obtain
/
7r/2
— c o s(z COS </) — <j) — 7t/ 2 )
■7T /2 ^
ss R,e / * /2 fí^eí ( * - '/2)e-< ^ a/3i (47.46)
J —n/2 ^
where we have expanded 2 cos —<j) about the stationary phase point, (j) « —l/z.
Evaluating the Gaussian integral, we find (for 2 > i )
7i(z) ~
ir y iz
= ^
¡2
_
,
cos(z _ _ _ _ ) .
7T 7TS ,
( 4 7 .4 7 )
Alternatively, we can obtain this same result by using the recurrence relation
(47.25) and the asymptotic form for J qj (43.37),
47.3. DIFFRACTION B Y A SLIT 499
Z=kaQ
Figure 47.4: Behavior of 2J\(z)/z and the differential cross section (normalized
to unity at $ = 0) given by (47.29) as a function of z = kaO.
Z\ (asymptotic) Z\ (exact)
¥ + ? = 3.927 3.832
= 7.069 7.016
¿2L + JUL = 10.210 10.172
Table 47.1: Table of the first three zeroes of Ji, compared with the estimate
from the asymptotic form (47.47).
infinitely long.]
(The justification for omitting the contribution of the surface at infinity is given
in Problem 47.1.) Green’s function is given by (47.7), or explicitly,
where
r± = (x,y), (47.51)
r± = (~ x ,y ). (47.52)
oo eiky/p2+ ( z ’- z ) 2
/ - oo
dz1
\Jp2 + (z1 - z )2
(47.53)
We will be concerned with the diffracted field far away from the slit, for which
we can use the asymptotic form (43.34),
ikix2+ (y -y 'f]1/2
(k^/x2 + (y - j/02) (kx > 1). (47.55)
ink [;x2 + (y — y')2]1/4 ’
characteristic of radiation fields. See, for example, (47.17).] When the wave
length is small compared to the slit, a A, we may again approximate the field
in the slit by the incident field,
eikp = - i k - e ikp
ox p
« —ikeikp. (47.61)
p af 2
E (x, y)
(47.63)
da _ 2 / sin (4 f sin#)
dO nk l sin#
„ ± { « n m
«5 y
(47.64)
irk \ 0 ) '
502 CHAPTER 47. DIFFRACTION I
0 n 2n 3n
kae/2
The latter holds for 0 <C 1, which is the only region of validity for our result.
The resulting diffraction pattern is shown in Fig. 47.6, where the zeroes occur
at equally spaced points:
(47.66)
In the short wavelength limit, we anticipate that the total cross section per
unit length is given by the width of the slit. In fact, for ka 1, most of the
contribution arises from values of 0 near zero, so that we indeed have
cr
(47.67)
1/2
E (* ,¡,)= Í /V W )(— ) f\iirk
— )1 (47.68)
[*3 + ( y - y 0 2]1/4
47.4. DIFFRACTION B Y A STRAIGHT EDGE 503
If we put E (y') = 1 for all y', we must recover the incident plane wave, E (x , y) =
etkx. We expect that yf ~ y gives the major contribution to the integral, since
the field should just advance with constant phase in y, thereby allowing us to
use
\Jx2 4- (y - y ')2 ~ % + - - ■— (47.69)
¿X
\ 1/2
( ^ } Jkx i k ( y - y ' ) 2/2x
\ srikx) 6 ~e
-j S * Ubj'
i k x e i k ( y - y ' ) 2/2x
/ Pi k ( y - y ’ ) 2/2x __ e ik x
dy'e (47.70)
=\ & -£
as is expected. Here we note, as a check of the approximation |y —y!\ <C that
the significant contributions to the Gaussian integral come from the values of y
satisfying
\ v-if\
(47.71)
x
E {y ) w Einc — 1, y >0,
E (yf) « 0 , y' < 0. (47.72)
= <47J5>
For sufficiently large $/,
y > V 5*, (47.76)
the second term in (47.75) may be neglected, andthe wave travels undisturbed:
E ( x , y ) ~ eik*. (47.77)
= >, (47.78)
< 1. (47.83)
x y x
A quite different limit of diffraction by a straight edge occurs when both x
and y are large, while the diffraction angle 6 is fixed:
— = tan 6. (47.84)
x
We anticipate that the dominant contribution to the scattered field comes from
the region near the edge, where yf is small, and in consequence [see (47.57)]
i eikp
(47.87)
27rkp 6
d* I W 1 1 (47.88)
<lf IE...A1
[In the next chapter we will provide an exact treatment of this diffraction prob
lem.]
The above discussion of diffraction by an edge provides a clarification of our
earlier consideration of diffraction by a slit. If we make observations by use of
a screen sufficiently close to the slit, so that
a \/X«, (47.89)
we see a geometrical image of the slit modulated by edge diffraction (see Fig.
47.8). However, if we move the screen so far back that
(k/nx)y
as
3. In the same approximation, give expressions for the electric field produced
by a slit, and verify the results shown in Figs. 47.8 and 47.9.
4. Show that when the slit disappears (a —►oo), (47.54) identically reduces
to eikx.
5. A perfectly conducting sheet has two infinitely long slits, of width a, that
are separated by a distance / a. An electromagnetic field of wavelength
A, polarized parallel to the slit edges, falls normally on the sheet. Using
a reasonable physical assumption that does not involve the magnitude
of a/A, and the formalism applied in the text to one slit, compare the
differential transmission cross section (per unit length) for the two slits
to that of one slit. What are the new features of the diffraction pattern,
particularly for //A >> 1? How does this pattern alter if there are a very
large number of equidistant slits?
Diffraction II
We now adopt a more physical approach to diffraction in which the currents that
give rise to the scattered wave are made explicit. We will reconsider diffraction
by a semi-infinite metal conductor, for which this method is capable of giving an
exact solution. The geometry of the situation is as given in Fig. 48.1. We con
sider E to possess only a ^-component (z subscript suppressed), and decompose
the electric field into incident and scattered parts,
We assume the incident field is a normalized plane wave with frequency u> = Arc,
The scattered field arises from the induced current that flows on the metal
plate, which by symmetry can have only a ^-component, J = J^, and has no
dependence on 2 ,
r\
a ;M * ,y ) = o, (48.3)
x' = 0
Figure 48.1: Geometry of diffraction by a perfectly conducting straight edge.
The z axis is perpendicular to the page.
509
510 CHAPTER 48. DIFFRACTION II
/
ik|r-r'| i
(dr7) - - p| - J ( r ;), (48.4)
which follows from (35.7) and (35.15). We consider an infinitesimally thin con
ductor for which we may reduce the three-dimensional integral in (48.4) to a
two-dimensional one by introducing the surface current,
J dx' Jz = K . (48.5)
For the semi-infinite perfect conductor being considered here, (48.6) is subject
to the boundary condition
which we will assume hold for all x < o, y < 0, and that the fields vanish for
x > 0, y < 0. From these forms for the electric and magnetic fields, we can find
the induced current through the use of Maxwell’s equation
V x B = - É + — J, (48.9)
c c
which becomes here
— Jz = dx By - - É z. (48.10)
c c
By integrating (48.10) across the conducting surface from just to the left to just
to the right,
47T
By (x = +0) — By(x — —0) = — K> (48.11)
48.1. APPROXIM ATE SOL UTION 511
K = (48.13)
2 -it'
If we further use the asymptotic form (47.55) for the Hankel function, together
with the approximation that when \x\ is large, only small values of \y — t/| are
significant,
V * 2 + (y - y ')2 « M + — ^ , (48.15)
’ rO
d y 'é Hy-y'YI 2k|
ik\x\
E ~ eikx - [ (48.16)
2iri\x\ J—oo
With the substitution y1 — y ►y', the integral in (48.16) has the form
f~y k y l2/2\x\
/ dy'e' (48.17)
J — OO
When y > 0 and far away from the edge, y the integral (48.17) is
negligible, and the wave propagates undisturbed,
E ~ é ikx (48.18)
On the other hand, sufficiently below the edge, y < 0, \y\ \/A]xJ, the integral
is ri\x\/k, and
E ^ e ikx — e ik\x\
_ ( 0, x > 0,
- y eikx _ e-ikx^ x < 0 i (48.19)
reproducing the boundary condition (48.8). On line with the edge, y = 0, the
electric field is
E ~ e ik x- (48.20)
These results are identical with those found earlier in Section 47.4, but now a
physical picture has been provided: The total electric field is the sum of the
incoming field plus the field produced by the currents induced in the metal.
512 CHAPTER 48. DIFFRACTION II
We next turn to the limit of large x and y where the angle of diffraction 0 is
fixed, as shown in Fig. 48.2. In this limit, the dominant contribution at (x, y) is
the sum of the incident wave plus the wave scattered by the edge. In terms of
the description employing currents, this is given approximately by (48.14), with
® 1 ,ikp_ 1
(48.22)
2irk y/p sin 6 ’
leading to a differential cross section per unit length (valid only for * < i )
da 1 1
(48.23)
dO 2tt& 92 ’
in agreement with (47.88).
/
plane,
OO 1
H ^o1}(*l» - y' I) - W ). (48-24)
-OO c
In order to solve this equation, we must recognize that an incident plane wave
is an over-idealization, one that can be removed by introducing an exponential
cutoff in y:
e —» 4-0. (48.25)
£W (* = 0 ) = l - e - ‘ l*l,
Note in (48.24) that we have two conditions:
48.2. E XAC T SOLUTION FOR CURRENT 513
d ye~tiyE {y),
II
-O O
('OO
(where the cuts are chosen not to cross the real k axis) we arrive at the following
representation for the Hankel function,
¿ky(y-y')
wiH^\k\y - j/|) = 2wi Í ^ - —7= (48.30)
J- ~ / ( k 4- ^ ) 2 — fcy
/ (48.31)
We now find for the Fourier transform of the integral equation (48.24)
where, if we make explicit the regions in which the integrands are nonzero,
/* o o
E (0 = / d y e -^ E (y ), (48.33)
Jo
If these integrals exist for real C> they will also exist for complex values of (.
Anticipating that E (y), K (y ) fall off like e~e\y\ as \y\ —* oo, we see that
E (Q exists for ImC < e,
K(C) exists for ImC > —e. (48.35)
We will call the half planes ImC < e, ImC > —e the lower half plane (LHP) and
upper half plane (UHP), respectively. It is essential to observe that the UHP
and the LHP overlap in a strip,
—e < ImC < (48.36)
Our physical requirements of boundedness ensure that
E (() is regular in the LHP,
Ar(C) is regular in the UHP. (48.37)
In order to examine clearly the analytic properties of (48.32), we multiply it by
y/k -f ie - C":
r.-----;----- 7n/>\ . y/k "p ÍC —C .y/k Í€ —C , km )
y j k -j- Í € — C E ( C ) ~ — í ------ -- ------ :-----------h l ------1—;— ;------------- 27Tk - . (48.38)
C — ie C+ y/k + ie -f C
The factors in (48.38) can be chosen to be regular in the following regions,
y/k + ie - C : LHP,
y/k + ie + C UHP,
y/k + ÍC — C
LHP,
y/k + ie - C
, . . - e...........*
< ImC < . . .e. (48.39)
C + te V '
The last combination can be written as the sum of terms regular in the LHP
and the UHP, respectively,
y/k + ie — C _ y/k -f ie —C — y/k + 2ie y/k -f 2ie
(48.40)
C+ ¿c C + ie C + ie
Thus we can reorganize (48.38) into parts that are regular in the LHP and in
the UHP:
n r - ----- 7nw\ , .y/k + i e - C .Vk + i e - { - V k + 2ie
y k + ie — QE(Q + i -----:---- :---------
(-ie C + ie
.y/k + 2ic km )
2irk (48.41)
C + l€ s/k + Í€ + C
48.2. E X AC T SOLUTION FOR CURRENT 515
The right hand side of (48.41) is regular in the UHP, the left hand side in the
LHP. Since the two functions are regular in a common region [the strip (48.36)],
they may be analytically continued into a function regular for all
We will now show that this function vanishes at infinity, so that it vanishes
everywhere. To this end, we examine (48.34) in the limit C —» oo, for which
only the behavior of K (y ') for y' —» 0 is significant. Because there is no intrinsic
length scale in this limit the current must behave as a power of yf,
~ y' - -0 , (48.42)
a < 1. (48.43)
The behavior of the corresponding Fourier transform of the current for large £
is therefore
f°° 1
E ( 0 = Jo e - ^ E ( y ) dy ~ ^ - 0, <- oo. (48.46)
— K<r\ - { V k + ie + ( (48.47)
C C + *e
(48.48)
~ ^ >
iV k 1
E (Q = i ( - r - L - - —1 __ (48.49)
VC + *e C -ie C 4" ic "\/k + ic — C
(48.50)
~ £3/2’ 1^1 ^ ^'
1
(48.51)
2’
516 CHAPTER 48. DIFFRACTION II
(48.52)
* (V ) ~ - M
E {y) ~ \ftj, (48.53)
1 i 1 ___ 1 y/k + C + if — V k
(48.54)
ch ^ 2n \ C + ie ' y/k C+
y < 0, (48.55)
2tr \ 0, y > 0,
which is the current, (48.13), we used in the first approximate solution to this
problem, corresponding to the neglect of edge effects. [Here, (48.55) includes
the exponential cutoff.] The second term in (48.54) thus gives the correction
that must be added in order to obtain the exact current. By considering the
behavior of (48.54) for £ small (which, as we will see below, corresponds to small
diffraction angles), we note that the first term is singular as £ —►0, while the
second is finite. Thus, the first approximation is valid for small angles, as we
have previously asserted.
The asymptotic scattered field at fixed angle 0 follows by use of (47.58) in
(48.6):
k cikp
27rip
(48.56)
where we have used the Fourier transform (48.34) with
C = fcsin0. (48.57)
Then from the solution (48.47) for the surface current, we find for the exact
asymptotic scattered field
da __ 1 1 + sin 0
(48.59)
dO 27rk sin2 9 ’
I L P ikp poo
where E (y) is the exact electric field in the aperture. The integral in (48.60) is
the Fourier transform (48.33),
according to (48.49). The resulting scattered electric field coincides with (48.58).
(v 2 + ¿ )E = °> (48-62)
near the edge. Since there the field is rapidly varying over a distance small
compared to the wavelength, we can omit the 1/A2 term. Thus our problem is
the electrostatic one of finding the field near the edge of a plane conductor. Using
the cylindrical coordinate system shown in Fig. 48.3, we write the Laplacian as
/ Id d 1 <92 \
(48.63)
\ p d ^ p dp + p¡ d 9 ^ ) ’
E Pm■ (48.64)
518 CHAPTER 48. DIFFRACTION II
The boundary conditions that the field must vanish on the conducting plane,
_ . 7T 37 T
E = 0 at 6 and — , (48.65)
with
sin27rm = 0. (48.67)
1
m=-, (48.68)
E = C^/psm - { 0 + » /2 ), (48.69)
where C is a constant. [Note that the solution (48.58) exhibits this same be
havior, since -y/l + sin6 = \/2sin |(0 + x/2 ).] For x = 0, y > 0 (6 = ir/2), the
electric field is
E(j>, ^ ) = C ^ p = Cy/y, (48.70)
= C \ m ^ i(x + iy )- (48.71)
( * L + * L ) e = o, (48.72)
\ dx2 dy2)
48.4. FIELD NEAR EDGE 519
a_
2
since there
E = C \m yji(x 4- iy) C = 0. (48.75)
Next, we apply the above ideas to the situation of a slit, as shown in Fig.
48.4. When the slit is small compared to the wavelength,
a < A, (48.76)
the solution to Laplace’s equation becomes relevant here. Since the appropriate
boundary conditions are
E = Clm-y/V2 — (a/2 ) 2
= C Im *\ /a2/4 — y2
\v\ > f. (48.79)
C \fa2/4 — y2, |y| < f .
In order to reproduce the static limit in the absence of the slit (a —►0), we
superimpose on this field another solution to Laplace’s equation, —C#, which
satisfies the same boundary conditions, (48.77), so that we have the following
asymptotic solution,
E ~ —C x + C\x\
_ r o, x > o,
(48.82)
~~ \ — 2C x ) x < 0 .
Far from the surface, we can no longer neglect k2 = 1/A2 in (48.62). Conse
quently, the approximate diffracted wave, the static limit of which is (48.82),
is
2c
E = — r-sinfc#, x < 0. (48.83)
K
This is exact for zero aperture, a = 0, since it represents a standing wave due
to total reflection by the conducting plane. The associated magnetic field,
d
ikBy = —— Ez = 2 C co sk x t (48.84)
for a narrow slit satisfying (48.76). The magnitude of the electric field in the
aperture is small compared to the magnetic field there,
and would be zero if (48.83) were exact. Equation (48.86) is the basis for treating
the diffraction by a narrow slit.
3. A slot of finite length / and very small width a is cut in a perfectly conduct
ing sheet. An electromagnetic field of wavelength A, polarized parallel to
the long axis of the slot ( 2 -axis), falls normally on the sheet. Find the an
gular (0, <j)) distribution of the transmitted radiation in an approximation
valid for //A 1. Compare the 0 dependence with that of the uniformly
excited antenna discussed in Section 34.1; explain the difference in the <j>
distribution.
Chapter 49
Babinet’s Principle
E (x) = £ ( - « ) . (49.2)
Ez: ___ , _ i
~ 2 + 3
(+ ) (+ ) (-) (+ ) (+ )
523
524 CHAPTER 49. BASIN ETS PRINCIPLE
I
I
Ez: _ l
~ 2
y < 0 : Ez = 0 at x = 0,
BE
y > 0: = 0 =* By = 0 at x = 0, (49.3)
where we have used the continuity of the derivative for y > 0. Thus the second
term corresponds to a perfect magnetic conductor (for which the tangential
magnetic field vanishes) for y > 0 and to a perfect electric conductor for y < 0 .
The situation for x > 0 is shown in Fig. 49.2, where a solid line represents a
perfect electric conductor, and a dashed line a perfect magnetic conductor.
We now consider another situation (the “dual” of the above), in which the
incoming wave has polarization Bz , and strikes a conducting plane x = 0 , y > 0 .
The decomposition into two terms may be made as before, as shown in Fig. 49.3,
where the + , — signs now refer to the sense of Bz . In the first term here, Bz is
even, so at x = 0 , y < 0 ,
r\
— Bz = 0 ^ E y = Q. (49.4)
y < 0 : Bz = 0 at x = 0,
49.1. PROBLEMS FOR CHAPTER 49 525
i
i
I
Ez Bz
y > 0 : Ey = 0 at x = 0, (49.5)
since Bz is continuous for y < 0. Thus the x = 0 plane acts as a perfect electric
conductor for y > 0, and as a perfect magnetic conductor for y < 0. So for this
polarization the situation for x > 0 appears as shown in Fig. 49.4. For both the
above circumstances, reflection from the infinite conducting planes corresponds
to exactly forward scattering, and can therefore be ignored when we calculate
the scattering cross section. The remaining contributions differ only in that
electric and magnetic quantities are interchanged,
E i- £,
B -E . (49.6)
Since Maxwell’s equations are invariant under this transformation (recall Chap
ter 2 ), the cross sections for these two problems are identical.
An immediate generalization of the above arguments shows that, for normal
incidence, the problems (E Zy aperture, screen) and (Bz , screen, aperture) cor
respond to the same cross sections, but with orthogonal polarizations. See Fig.
49.5. This is an example of what is called Babinet’s principle.
General Scattering
V x B = - i k E H------J, (50.1)
c
V x E = ikB, (50.2)
V x ( V x E ) = k2E + ik — J. (50.3)
c
ik V •E = J, (50.4)
(V 2 + fc2) E i „ c = 0 , (50.6)
527
528 CHAPTER 50. GENERAL SCATTERING
we can immediately write down the solution to (50.5) in terms of the retarded
Green’s function,
Etang = 0 on 5. (50.9)
e *n = 0, (50.13)
we have ^
e * Escatt ~ ~ ~ f > (50.14)
r
where / is called the scattering amplitude, and depends on the emission direction
and the polarization,
Note that the integral here is the three-dimensional Fourier transform of ~J(r/).
The differential cross section for scattering is immediately given by the absolute
square of / ,
^ = r 2 |Escatt- e |2 = |/|2. (50.16)
da
= k2 |e« J (dr')e~ik' r'~3(r) (50.17)
dQ,
the total cross section is obtained by summing over polarizations and scattering
angles,
y^ee + nn = 1 , (50.19)
y ee = 1 —nn —» i + p V V , (50.20)
e
where the latter form is assumed to act on e*k *(r" r/). The angular integral in
(50.18) is easily found to be [cf. (38.45)]
Air
* = T \mk2
'
J{dv){dv')-r{v). ( l + p V V j
f 1 / 1 \ e**|r-r#|
-j—
i
^ . - J ^ ') . (50.22)
We next use the boundary condition (50.9) together with the fact that J is
confined to the surface and therefore has only a tangential component to infer
that
-J * •E = 0. (50.23)
c
530 CHAPTER 50. GENERAL SCATTERING
Consequently, if we make use of the general expression for the field, (50.8), we
deduce the identity
0 = J (d r )h * * E
= J ( d r ) i j* •Einc
eik\r-r'\
+ ik J \ d r)(d r')-3 * (r)
1 +¿W | r-r'|
(50.24)
« - y Im J (dr)^J* •E inc
(50.25)
= T Im [ik / ( dr)E>*nc * \3 •
where we recall the expression (50.15) for the scattering amplitude. This result is
the “optical theorem:” the total cross section is proportional to the imaginary
part of the forward scattering amplitude. Note that this theorem is here a
consequence of the conservation of energy; the result follows from
j ( d r ) 3 •E = 0, (50.28)
which just says that the induced currents do no work in a perfect conductor.
As an indication that this theorem is actually much more general than the
derivation that we have supplied, let us return to the scattering of radiation by
an electron bound by a harmonic oscillator potential. The forward scattering
amplitude, (50.15), is, in this circumstance
e E(u>)
r(«) (50.30)
mu)Q — to2 — iu>y ’
(50.31)
m uIq — w2 — iuiy ’
50.3. BORN APPROXIMATION 531
implying
47r ________ u 27
(50.32)
k m ^m e ( wq — w2)2 + (wy)
- ik U - 1)E = — J. (50.34)
c
Inserting this in the formula (50.8) for the electric field, we obtain the following
integral equation determining the electric field in the dielectric,
/ i \ r eik\r-r'\ / r /\ __ i
E (r) = Einc(r) + ik (l + p V V j . J (dr’ ) ^ _ ^ ( - i k ) { E(rO-
(50.35)
Once E is known in the dielectric, the scattering amplitude, describing the field
far from the dielectric, can then be found from (50.15),
|e — 1| <C 1, (50.37)
we can employ a simple approximation based on the fact that inside the dielec
tric, E is then roughly equal to the incident field,
da
( e - e 0) 5 (dr)n(v)e i ( k 0 - k ) * r
(50.44)
díl
Note that for a single particle, for which n(r) = <S(r — R ), we obtain the Thom
son cross section, given by (45.14) [cf. (45.10)], when the sum over final polar
izations is performed.
K = iK{(¡>) = z Yj K meim*,
(a) Using (50.8), calculate the electric field everywhere (p > a) in terms
of the coefficients K m.
(b) Apply the boundary conditions (50.9) and determine K m.
50.4. PROBLEMS FOR CHAPTER 50 533
(c) Determine expressions for B for p > a and check the consistency of
K m by means of the boundary condition (26.22).
K = 4> K <neim*-
E = V x ( V X r ü i ) + — V x r n 2,
c
B = - y V x r l l i + V x ( V x r n 2),
1C
E(r) = £ / ,( r ) X im(Q) + - V x 3 ,(r)X ,m(Q)
u>
lm L
X/m(Q) = - - ^ = = L y /m(fi),
L = t X t V .
i
J da = sw &mm' j
£ |Xjm(íí)|2 =
Dl rd r2 r2 '
V xr(r,r') = f^-$(r,r'),
91 = 9h
fi =
r*
CD , + k 2)g, = i-c j dfl" X,*m(Q ") •[V "x li5 (r " — r)],
= ^ F ,( r ,r ') X ? m(fi),
Then show
r (r ,r O = £ { * 2 F K r y ) X ím(fi)Xrm(íí')
lm
(V 2 + k2)4>sc( r) = 0.
This expresses the scattered wave in V in terms of its values, and normal
gradients, on S. [It is necessary to show there is no contribution from
the hemisphere “at infinity” also bounding V .] Now make the “Fraun
hofer approximation” that the point of observation r is far away from the
portions of S that contribute significantly to this integral:
ptkv . .
r > r', so G (r, r') « ----- e~lk *r , k; = k~.
r r
Then writing
^sc(r) = ~— / ( M ) >
r
where / is the scattering amplitude, show that
/(0, <t>) = -L J dS '. {^ sc(r/)V 'e -‘k' *'' - e"ik' ’ r'V ^ sc(r0 } .
10. Apply the result of the previous problem to calculate the scattering am
plitude for a sphere in the “black” approximation, where
4 = eik ‘ , + 4tc = 0
50.4. PROBLEMS FOR CHAPTER 50 537
<^sc := 0
We obtained a dispersion relation for the dielectric constant, or for the electric
susceptibility, that is, an expression for e or x as a function of cj, in Chapter 5
[see (5.21), for example], by assuming a particular model for the motion of the
charges in the medium. Here we will, starting from general physical principles,
derive the form of the dispersion relation satisfied by e(u>) or %(o>), of which
(5.21) is a particular example. For simplicity, we will assume the medium to be
nonmagnetic, that is
¡ i n i or B « H , (51.1)
which is the usual situation. In a linear, isotropic, dielectric medium, the dis
placement, the polarization, and the electric field are related by (4.58) and
(5.31),
D = E + 47tP = eE, (51.2)
which is valid for most materials for sufficiently weak fields. The unit step
function, 7](t —¿'),
<m -5>
539
540CHAPTER 51. DISPERSION RELATIONS FOR THE SUSCEPTIBILITY
■+ 0 . (51.6)
v -f* ie
The function f ( t ) may be represented by a Fourier integral
m = /~ (51.7)
Substituting the representation (51.6) into (51.4), and using the Fourier trans
form (51.7), we identify
dw' 1
E(w). (51.9)
2w u> — o>' + ie / ( « ' )
♦E > 0. (51.14)
541
If the Fourier representation for the electric field is introduced, (51.14) becomes
-/—cj 2**
|E(o;)|2 > 0. (51.15)
On the other hand, since |E(u>)|2 is even in (51.15) can also be written
duj
f 2ir
( — jw ) x ( w ) | E ( w ) | 2 > 0. (51.16)
/
CO I POO J
= v4$ (w ) ~
> 0- (51.19)
I m — -— = 7t 6 ( ). (51.20)
x — it
x
ne
x (w ) OJ COatomic- (51.21)
mu2
(O f course, we assume that Tiw <C me2 so that the electron remains nonrela-
tivistic.) On the other hand, the dispersion relation (51.13) becomes, in this
limit,
, , 1 f ° ° dw’ /2 . , „
(51.22)
*<"> = “ l j 7 " *<" >■ U> W a to m ic ,
542CHAPTER 51. DISPERSION RELATIONS FOR THE SUSCEPTIBILITY
if the integral exists. In this way, we arrive at a sum rule for 4>:
f° ° du>' /2 _ ne2
/ — w'2 $(w ') = -----. (51.23)
Jo 2 tt m
(a/ 2 ne 2
(51.24)
and pOO
/ p(u/)cfu/ = 1 . (51.26)
Jo
The dispersion relation, (51.13), now appears as
, , ne 2 r , , P ("')
X{0J) = m y/0 du a;'2
,2 / -f, tc
• )\2j * (51.27)
— (a;
As a simple example, recall the situation when the system has only one
excitation frequency, wo, with a damping constant 7 . This is the Fermi model,
which we considered in Chapter 5. In particular, we found there the dielectric
constant to be [see (5.21)]
V2
e(u>) — 1 + 2 2 • ’ (51.28)
Uq — — iyw
9 47rne2
u 2 = ---------. (51.29)
F m
The spectral weight function p(u/) in (51.27) can be obtained from the imaginary
part of e:
a /27
(51.30)
,j'2 wg) 2 + (w'7 )2 '
For small 7 ( t < wo), this exhibits the structure of the Lorentzian line shape.
[Recall Section 35.3.]
It will be useful in the following to have a dispersion relation for the inverse
of e(u>) as well. Its form is suggested by the Fermi model. The reciprocal of
(51.28) is
1 O>0 — iljJJ — W 2
e(w) (u l + w2) - i'a>7 - u>2
1 _________ 1
(51.31)
p (w§ + w2) - JU>7 - w2 ’
51.1. PROBLEMS FOR CHAPTER 51 543
which differs from (51.28) by the sign of the second term, and the replacement
of cjq by Wq -f u>2. Therefore, we anticipate that l/e(u>) satisfies a dispersion
relation of the form
(51.32)
c(w) P Jo U)/2- (w + «02
where the spectral function <?(u/) satisfies the conditions
I q(u>f)du' = 1 . (51.34)
Jo
To show that (51.32) is the correct dispersion relation for l/e(u>), we may pro
ceed in the same manner as we did in deriving the dispersion relation for e(a>).
We apply the causality condition to write a general relationship between P and
D , leading to a dispersion relation for The positivity of the spectral
function, (51.33), follows from the physical requirement (51.14), expressed in
terms of |D|2, while the sum rule (51.34), results from the expression for l/e(u>)
at high frequencies, u u>atomic> which follows from (51.21)
1
(51.35)
e(w)
2. Show that the sum rule (5.36) follows from the dispersion relation (51.13)
and the sum rule (51.23).
Chapter 52
We will now apply the general dispersion relations given in the previous chapter
for the dielectric constant to discuss the energy loss by a charged particle when it
passes through a dielectric medium. The interaction between a charged particle
and a medium is described by the macroscopic Maxwell equations, (4.60). As
we remarked before, (4.60) can be transformed, formally, into the vacuum equa
tions, (1.65), by the redefinitions given in Section 7.2. It is simplest, therefore,
to start with a vacuum description and then transform to that of a medium.
J
dE_
(dr)(drf)dt/ ^je(r, t) •7¡¿G(r — r ',t — t') • ^ je(r', t')
dt
Occurring here is the retarded Green’s function given by (31.28) and (43.19):
545
546 CHAPTER 52. CHARGED PARTICLE ENERGY LOSS
The charge and current distributions for a single particle of charge Ze are
given by
? “ * ¿2 > (52‘6)
leading to the following formula for energy loss,
(dk)duj —iu
27r6(k •v —cj),
medium (27r)4 k 2 — a>2
(52.7)
where now e(u>) supplies the necessary imaginary part in the denominator. The
angular integration here is trivial:
We see here an intimation of Cerenkov radiation [recall Chapter 36]. This step
function supplies the lower limit for the |k| integration. As for the upper limit,
we note that momentum transfers must be limited in this description, since
we have ignored recoil. Large momentum transfer events must be dealt with
differently, as collisions with individual electrons (¿-rays). We therefore cut off
the |k| integration at some characteristic momentum transfer K , reflecting the
boundary between different ways of measuring energy loss. The energy loss rate
is therefore
dE
dt
(52.9)
where we have assumed that the important values of u which contribute to the
integral satisfy K 2 eu>2/ c 2. The energy loss per unit distance traveled, which
52.2. EVALUATION IN TERMS OF SPECTRAL FUNCTIONS 547
_ {Z e ? /■ °° / C2 \ / i 2 W2
r Im / duiu ( 1 ----- 5- 7 —7 I In ---- ;------- 3------ - , (52.10)
7TC
» 2
Jo V w2 ( l - £ c ( W)) v '
where we have used the reality condition, (51.8), for e(u>). Note that if e(u>) is
real, this implies the formula (36.19) for Cerenkov radiation.
_ ¿ E = (Z e £
(52.11)
dz 7rc2
where
I<2v2
pOO
I = Im 1 diOLO |
f i - 4 - ' I In (52.12)
- 2 (1 - £ ) ’
poo 1 _!¿
I I = Im 1 dujuj j 1 c2 (52.13)
1 In 1 y2 e'
1 ~
The imaginary part in (52.12) arises entirely from the imaginary part o f —l/e(o;),
which, from (51.32), is proportional to the spectral function g(u>),
(52.14)
V i(w )) '
ir 2 ( °2 \ i K'2y2 (52.16)
J ~ 2 ^ { ^ ) l n We2 ( i - $ y
u e = wo-\/<K0 ). (52.17)
548 CHAPTER 52. CHARGED PARTICLE ENERGY LOSS
v2 ... . v 1
(52.19)
? t(0)<1' or
there is no contribution to I I from the integration along the imaginary axis, C 2 .
This means that we may replace the integral I I by the contour integral along
the quarter circle, C 3 , where |w| —►00 . Since, in this circumstance, |w| is large
compared to atomic frequencies, we have there
e(u0 « 1 - (52.20)
1 V2 v2 1
In 1 ~ ^ (52.21)
1- ~ u2 c2 1 1
“
V2
P-
which results in
f ( c2 \ ( w? V2
= Im / dujio I 1
~ v 2) ( J c -
= (im
l . ) ■ rt-
(52.22)
52.2. EVALUATION IN TERMS OF SPECTRAL FUNCTIONS 549
Therefore, under the condition (52.19), we obtain the energy lost by the particle
per unit distance traveled,
dE _ 1 u>2(Z e )2 i I<2v2 v2 v 1
for (52.23)
d z~ 2 c2J
When the velocity of the particle is so large that (52.19) no longer holds,
v 1
(52.24)
c >
the integration along the imaginary axis C 2 , is no longer zero, since the argument
of the logarithm, 1 — p-e(u>) goes through zero. Now, there is an additional
contribution arising from C 2'.
dwu) (52.25)
where u = iv {u > 0). The value of is, vVl for which 1 — e(u>) vanishes can be
determined by the dispersion integral
w2 r dU' j ! t o _ = 1 _ i ! (52.26)
p Jo w' 2 + *'a C2
V
”2 1
(52.27)
L)„
4 - . 1 (52.28)
1
dE\ _ (Z e)2
-Im In------- ---------------- - .
) add _ ™ 2 Jo \c2 e(tv)J (-1 ) (2rc(tV) - 1)
(52.29)
To determine the branch of the logarithm involved here, we return to the general
spectral representation for e(u;), (51.27), and let u> approach the imaginary axis
as follows:
u -► vei(*l2- 6\ (52.30)
u/ 2 - w2 J 2 - ( v é ^ ! 2- ^ ) 2 = a/ 2 + v2 — i6, 6 ^ + 0. (52.31)
550 CHAPTER 52. CHARGED PARTICLE ENERGY LOSS
The imaginary part of e(a;), in this limit, is positive and tending toward zero,
which implies that
1
Im I In — Im I In
1
1—
vJc
< (l-^Ree)-z^Im e
( - £ ) = ( | ^ r '^ [ , - 4 _ u, ; r ^ _ « g L
V dz ) add 2v JO C2 PJ0 U i '2 + V 2
_{Ze)2 i ( 1 n2 \ 2 2 r J / , „ « '2 + ^ l
2v2
_ 1 h>¡{Ze)2
fii
r, u>2
(52.33)
u ve \ c2J up.
where u>2 is defined by (52.15), while w2e is given by the analogous equation,
f d u j'q ^ ) In
a/ 2 + V?,
= In ■ 2 •
w.
(52.34)
_ d E _ l a>2
p {Z e)2 I<2v2
L
In ■ . 2X
dz 4) c2 V c2) “ l
fo r (52.35)
c Ve(Ó)
A somewhat more convenient form of (52.35) can be obtained by expressing
ln l/u ;2e in terms of ln l/o ;2. We first observe that, from (52.26) and (52.34),
d , 2 f°° , / / /x 1 d 2
■ o T ~2\ = 1/
d( v2/c2)
l n u > v e
J/o d w
v 7j a
q ( u -T S —
J2 + v2o d(v2/c2)
2 'T i \ U v
I v2 dv2
_ 1 ~ p- (52.36)
>
U)2
p
d(ti2 / c 2)
d
v2A V2
i___ ___ ____ I
H 1^
___ ___ V
(52.37)
jS
d(v2/c2)
I
2
} c2) ( jj
P
ff vv2 /c
/c 2 /n/2\ u2 1 1 1 —”
(5238)
52.3. HIGH ENERGY LIMIT 551
Therefore, an alternative form for the energy loss per unit distance for v/c >
V y / f f l is
dE 1 u l{Z e )2 K 2v2
In _ _ ü ! _ (52.39)
’ dz ~ 2 v2 S) c2 JiK0) U K 2
where, in this form, we recognize that the last term is the added contribution
when v/c becomes larger than l/^ /e ( 0 ).
dE u 2 (Z e)2 K 2c2 v
In as — 1. (52.41)
dz c
Notice that this means that the energy loss saturates at a constant value not
much above the minimum ionization loss rate. This phenomenon, often called
the density effect, was first pointed out by Fermi in 1941.
Let us see, in the Fermi model, how this constant value (the Fermi plateau) is
approached. From the definition of i/Vy the value of v for which l/e(iv) = t>2 / c 2,
we obtain from (51.31),
(52.42)
Wq€(0) + yvv + v2 ’
because e(0 ) = 1 + w2¡w\, or, neglecting 7 in comparison to « 0 ,
7 (52.43)
From this explicit expression for v„ , we can then perform the v'2/c2 integration
in (52.39):
1/21
woe(0)
_ ^ 1/2
2 u€ u)p 1
(52.44)
552 CHAPTER 52. CHARGED PARTICLE ENERGY LOSS
where we have used the Fermi model expression for u>e, (52.17), which implies
1 — l/e (0 ) = cOp/u)2, and retained terms through order (1 —v2/c2). Substituting
(52.44) into (52.39), we obtain the energy loss per unit distance, in this model,
(52.45)
This approach to the Fermi plateau [deviation proportional to (1 —v2/c2) 112] is
not unrealistic.
We conclude this section by sketching the result for the energy loss by a
charged particle passing through matter in Fig. 52.2. This figure should be
taken as a qualitative indication only. For detailed discussion of the comparison
with experiment, and full references to the literature, the reader is referred to
The Review of Particle Physics, Particle Data Group [Physical Review D 54,
132 (1996)].
in 1975.] The formula can be immediately obtained from that given above
by exploiting the symmetry of Maxwell’s equations discussed in Chapter 2.
Specifically, we have the replacements, in vacuum,
dE K 2v2
(52.48)
dz 7TC2 J o V V ) W
f ’ 1
du p(ui) In — = In — .
1
(52.50)
Comparing (52.49) with (52.23), we see that the energy loss by a magnetically
charged particle is approximately obtained from that of an electrically charged
particle by the substitution
Ze g
(52.51)
V c )
provided that
2 2
ÜJe
w U) m • (52.52)
In the Fermi model, we recognize that
The general validity of this inequality can be established as follows. From (51.27)
and (51.32), we easily infer that
/ pK ) j
(52.54)
W>2 + Z/2 ^
and
= i - » l [ / ^ < 1 (52.55)
w) p Jo u>'2 + v 2 ’
554 CHAPTER 52. CHARGED PARTICLE ENERGY LOSS
(52.57)
in view o f (52.54). Therefore, the desired inequality is proved,
« e > Wm . (52.58)
Just as in the electrically charged particle situation, there is here an addi
tional contribution when ^ > —1—
2 „2
dE _ lWp</ , * 2t>2
In----- ;------- rr — 1 + (52.59)
dz 2 c2
. wm ( ! - v ^Add’
where, instead o f (52.25), we have
- = e(iuv) = 1 + wp2 j Q
C (52.61)
We can simplify (52.60) to obtain a form for the energy loss analogous to (52.39),
2 n2
dE _ 1 v*g- tfv i f « v Jf c 2\ v 2
v:
dz 2 c2
¿v.» W « i ] ’
V 1
for - > —= = , (52.62)
e.
C \/e(0)
[For the approach to the limiting high energy value, see Problem 52.4.]
4. Find the approach to the high energy limiting value of the energy loss
rate for a magnetically charged particle, in the Fermi model, in analogy
to (52.45).
Appendix A
Units
This book has been written entirely in the Gaussian system of units, which is
most convenient for theoretical purposes. However, for practical uses, and for
engineering applications, it is essential to make contact with the SI (Systeme
International) units, in which the meter, kilogram, second, and ampere are the
fundamental units. Fortunately, it is very easy to transform equations written
in the Gaussian system to the SI system. Here we will give that transformation,
as well as explain the Gaussian units. Thus, evaluations may be performed in
either system, or by using dimensionless quantities, such as the fine structure
constant.
The macroscopic Maxwell equations, (4.60), in Gaussian units, are
V •D = 4717?, V * B = 0,
(A .l)
where
D = E + 47rP,
H = B - 4ttM. (A.2)
(A.3)
V •D = p, V B = 0,
V X “ = ¿ D + J, - V X E = | B, (A.4)
where
D = e0E 4* P ,
H = —B -M . (A.5)
Ho
555
556 APPENDIX A. UNITS
eofio = c " 2,
Ho = 47T x 10~7, (A .6)
F = e(E + v x B ) . (A .8)
kj) kE = V 4 ttco ,
kfí \Z4npo,
1
kp k j — kp
\ /4 7 T 6 o ’
k]M (A.10)
So this means that all that is necessary to convert a formula in Gaussian units
into the corresponding formula in SI units is to make the scaling of the electro
magnetic quantities by the factors given in (A. 10).
To find the actual values of electromagnetic quantities in the two systems,
such as the charge on the electron, it is further necessary to convert the length
and mass units,
x —» Ai a?, m —* A2m, ( Al l )
where here, of course, Ai = 102 and A2 = 103. Thus, for example, an energy
scales by the factor A2A2, and hence the Coulomb potential of a electron at the
origin is, in the two systems, related by
p2 p2
£G _ \2 \ esi
(A.1'2)
ro 1 ^Treorsi’
557
CO
o
r*H
1
L H i x 1 0 -11
Table A .l: Relation between SI electromagnetic units and Gaussian units. Here
3 is an approximation for 2.99792458.
eo= es' V ^ \ ¡¥ ^ 2
= esiVlO-WIoVTo7
= 1.602 x 1(T 19 x 2.998 x 109 = 4.803 x 10- l o esu, (A.13)
• The formula may be transformed to SI units, using (A .10), and then eval
uated using familiar SI quantities.
seem less direct, has the virtue that one need not keep track of factors of
c, which is merely a conversion factor between time and space intervals,
and is naturally set equal to unity in relativistic calculations.
_ 3m3c5
(A.14)
T = 4e45 2 '
3ireom3c3
(A.15)
e4B 2
Now m = 9.11 x 10"31 kg, c = 3.00 x 108 m /s, and e = 1.60 x 10“ 19 C, so for
a magnetic field of IT, we obtain, again, r = 2.58 s.
The dimensionless method consists, first, in replacing the electric charge by
the fine structure constant,
e2 1
(A.16)
a ~ h e * 137’
and the magnetic field is to be expressed in terms of its ratio to the characteristic
(or critical) field strength,
m2c3
Bc = = 4.41 x 1013gauss. (A.17)
eh
Thus our formula becomes
3 ( B e\ 2 h
r
4 \B J am ec2 ’
(A.18)
so, because h = 6.58 x 10~22 MeV s, and mec2 = 0.511 MeV, we get the same
result, r = 2.58 s. Although this latter method may seem unnatural because it
introduces quantum quantities into a classical calculation, it has the virtue of
allowing the powers of h and c to be determined at the end of the calculation
by elementary considerations of dimensional analysis. Here because me2 has
dimensions of energy, exactly one factor of h must be present in the numerator
to obtain a quantity having the dimensions of time.
This advantage is lost if one uses a purely classical method of eliminating e
and J3, for example, by eliminating the former in favor of the electron mass in
terms of the classical electron radius, (45.6),
e2
ro = — - = 2.818 x 10” 13cm, (A.19)
mcz
559
(A.21)
This, of course, gives the same answer for B = 104 gauss, namely 2.58 s, but
now in terms of two dimensional quantities, ro and Wo, instead of the one, m,
in the atomic system. The latter is obviously preferable.
For a rather complete discussion of the vexing issue of electromagnetic units,
the reader is referred to Francis B. Silsbee, Systems of Electrical Unitsy National
Bureau of Standards Monograph 56 (U. S. Government Printing Office, Wash
ington, 1962).
Appendix B
Bibliography
561
Index
563
564 IN D E X
cavity, resonant 206, 295, 468, 469 straight edge 505, 517
Cerenkov, P.A. 388 Thomson scattering 472
Cerenkov radiation 386 two slits 507
charge density 6 Curie, Marie 388
bound 40 Curie, Pierre 68
effective 41 Curie temperature 70
induced 162, 179 Curie-Weiss law 72
magnetic 17, 337, 552 current density
surface 133 effective 41
charge quantization 19, 31 electric 8
classical radius of the electron 389, magnetic 18, 552
472 cutoff wavenumber 465
Clausius-Mossotti equation 58 cyclotron frequency 402
Clausius, Rudolf 58 cylindrical wave 501
collisonal broadening 383
completeness relation d’Alembert, Jean 342
Bessel functions 172, 221, 224 d’Alembertian 342
Fourier 171, 172 Darwin Hamiltonian 365
Fourier-Bessel 172, 230 Debye, Peter 485
Legendre’s polynomials 250 Debye potentials 533
sine functions 186 delta function 6, 141, 170, 186, 225
spherical Bessel functions 252, delta rays 546
267 density
spherical harmonics 246 angular momentum 25
conductivity 46 charge 6
conductor energy 22
force on 134, 163 force 22
imperfect 445 momentum 23, 76
conservation law density effect 551
and invariance 90 diamagnetism 65
angular momentum 25, 105 dielectric constant 50
charge 9, 105 differential cross section 473, 529
energy 22, 109 circular aperture 496
magnetic charge 18 circular disk 525
magnetic flux 61 slit 501
momentum 23, 104, 110 slot 521
Coulomb energy 1 spherical shell 490
Coulomb force law 1 straight edge 505, 517
Coulomb’s potential 141, 243, 257 Thomson scattering 473
critical harmonic number 418, 422 two slits 507
cross section 471 differential equations
circular aperture 496 Bessel functions 169
partial-wave expansion 488 Hankel functions 455
slit 501 Legendre’s polynomials 247
slot 521 modified Bessel functions 203
spherical shell 485 spherical harmonics 243
IN D E X 565
potentials 121
rotation 29, 104, 110
translation 102
velocity 118
transmission coefficient 433, 437
trilinear coordinates 208