Electromagnetic Fields and Waves, Lorraine & Colton
Electromagnetic Fields and Waves, Lorraine & Colton
ais
. oe
VECTOR DEFINITIONS,
IDENTITIES, AND THEOREMS
DEFINITIONS
Rectangular Coordinates
Erol ag OF ee, Oley.
I. a! oss
2 Vid = ey
Ox
S M
_ (94, _ 9A, 04, 94). (dA, OA \k
2 VAS Oy ar Oz ) Ox oy
vy? ot + Of3
: Vi Bi oe Oz?
5. V4 = VAG + aoe + W?A.k
Cylindrical Coordinates
of 1 of of|
iL: Vf= PAL ge de nn
1 0A,
2A yp Ae) + 2
p 0g
AVX
A= (OSS
10A, OA,
oe )ot oa
0A, Lyra
82a
Ten
4)
0A,
Spherical Coordinates
el ly
Mie ar"! 7 roo + rsin 8 dy i
pe I Peers
-A=——(r? 1 1 0A,
Meer so euerty iatSe an ede
0
ae E (A, sin 0) — oan
so He
rsin 6 (eX)
Vee igie. r
I /, of
(ao
Or
ree es, r?Ses
or r? sin 000 06 sin? § dg?
5); V?A ll= V(V-A) —V XV &X A (See Section 1.9.6.)
“TOM Cartson)
IDENTITIES
1. VUfg) =fVe+eVvf
Dy V(4-B) = (B-V)A+ (4:V)B + BX(V XA +A4AX(V X B),
2f V-(fA) = (Vf)-A4+f(V-A),
4. V-(A X B) =B-(V X A) —A-(V X B),
=F Vi xt) = (Vx 4/0 x A),
6. V X (A X B) = (B-V)A — (4-V)B + (V-B)A — (V- ADB.
7 VXVXA= ee — V?A (See Section 1.9.6.)
oe OB,).
oe (A-V)B = |eae
ae F 2: i
aB,
+A. ax te “6gore oi
aB,
OB, OB, OB,
+| Ag+ Ars Blk
Devs Gty =!—, where the gradient is calculated at the source point (x’, y’,z’) and
r, is the iia vector from the source point (x’, y’, 2’) to the field point (x, y, z).
LOD V. (;) = a where the gradient is calculated at the field point, with the
1. [ vrar = [faa
12. Je pee f AX das
where S is the surface which bounds the volume rT.
THEOREMS
S.
where C is the closed curve which bounds the surface
https://archive.org/details/electromagneticfOO0Olorr_x6d3
ELECTROMAGNETIC FIELDS AND WAVES
Second Edition
ELECTROMAGNETIC
FIELDS AND WAVES
Paul Lorrain
University of Montreal
Dale R. Corson
Cornell University
LOROSS7,
Contents
Preface XIX
List of Symbols XXII
VECTORS 1
eat! Vector Algebra 2
Examples 83)
EZ Invariance 5
1.3 The Time Derivative 8
Example 9
1.4 The Gradient 10
Example 1]
Flux and Divergence. The Divergence Theorem 13
Examples 13, 16
Line Integral and Curl 16
Examples 16, 19
1 Curvilinear Coordinates 23
1.9.1 Cylindrical Coordinates 24
1.9.2 Spherical Coordinates 26
1.9.3 The Gradient 27
1.9.4 The Divergence 28
1.9.5 The Curl 30
1.9.6 The Laplacian 31
1.10 Summary 3
Problems 36
2
PEEGLROSTATIG AFIELD Ss! 40
Electrostatic Fields in a Vacuum
3
ELECTROSTATIC FIELDS II wl
Dielectric Materials
4
ELECTROSTATIC FIELDS III 138
General Methods for Solving Laplace’s and Poisson’s Equations
4.1 Continuity of V, D,, E, at the Interface Between Two Different
Media 139
4.1.1 Potential 139
4.1.2 Normal Component of the Electric Displacement 139
4.1.3 Tangential Component of the Electric Field Intensity 140
4.1.4 Bending of Lines of Force 141
4.2 The Uniqueness Theorem 142
4.3 Images 144
Example: Point Charge Near an Infinite Grounded
Conducting Plane 144
Example: Point Charge Near a Grounded Conducting
Sphere 146
Example: Point Charge Near a Charged Conducting
Sphere 149
Example: Charged Sphere Near a Grounded Conducting
Plane 150
Example: Charge Near a Semi-infinite Dielectric 153
4.4 Solution of Laplace’s Equation in Rectangular Coordinates 156
Example: Field Between Two Grounded Semzi-infinite Parallel
Electrodes Terminated by a Plane Electrode
at Potential V, 158
Example: Field Between Two Grounded Parallel Electrodes
Terminated on Two Opposite Sides by Plates
at Potentials V; and Vy 161
4.5 Solution of Laplace’s Equation in Spherical Coordinates. Legendre’s
Equation. Legendre Polynomials 163
Example: Conducting Sphere in a Uniform Electric Field 168
Example; Dielectric Sphere in a Uniform Electric Field 173
Solution of Poisson’s Equation for V 176
Example: p-n Junction Diode in Silicon 176
Solution of Poisson’s Equation for E 180
Examples 180
Summary 181
Problems 183
5
RBEEATIVIT Yel LoS
The Basic Concepts
yall The Galilean Transformation 194
ee Breakdown of the Galilean Transformation and of Classical
Mechanics at High Velocities 195
CONTENTS
1x
6
RELATIVITY II 247
The Electric and Magnetic Fields of Moving Electric Charges
6.1 Force Exerted on a Moving Charge by Another Charge Moving
at the Same Constant Velocity 0 248
6.2 Field of a Charge Moving at a Constant Velocity 251
6.2.1 The Electric Field 256
6.2.2 The Magnetic Field 259
Example: The Field of a 10 GeV Electron
ati — 90. 260
6.3 Transformation of Electric and Magnetic Fields 261
Example: The Parallel-Plate Capacitor 263
6.4 The Vector Potential 4 264
Example: The Vector Potential for a 10-GeV Electron
at @ = 90° 266
6.5 The Scalar Potential V. The Electric Field Intensity E Expressed
in Terms of V and 4 266
Example: V,VV, and dA/dt for a 10-GeV Electron 268
6.6 Transformation of the Electromagnetic Potentials V and 4. The Four-
Potential A 269
Example: The Parallel-Plate Capacitor 270
6.7 The Lorentz Condition 271
6.8 Gauss’s Law 271
Example 272
6.9 The Divergence of B Pe
6.10 The Curl of E 273
6.11 The Curl of B 274
Onl Maxwell’s Equations 276
Example: The Magnetic Field Near a Straight Wire Carrying
a Steady Electric Current 277
Example: The Magnetic Field of a Short Element of Wire
Carrying an Electric Current 280
Example: Force on a Second Element J dl 282
Example: force on a Moving Charged Particle 283
6.3 Summary 283
Problems 286
i
MAGNETIC FIELDS I 292
Steady Currents and Nonmagnetic Materials
ae Magnetic Forces 293
7.2 The Magnetic Induction B. The Biot-Savart Law 295
CONTENTS
xi
8
MAGNETIC FIELDS II a2
Induced Electromotance and Magnetic Energy
8.1 The Faraday Induction Law B32
Example: The Expanding Loop 335
y
MAGNETIC FIELDS III 8 wo
(Fe)
Magnetic Materials
10
MAXWELL’S EQUATIONS 422
10.1 The Conservation of Electric Charge 422
Example: Charge Density in a Conductor 424
Ii
PROPAGATION OF
ELECTROMAGNETIC WAVES I 459
Plane Waves in Infinite Media
12
PROPAGATION OF
ELECTROMAGNETIC WAVES II 504
Reflection and Refraction
12.1 The Laws of Reflection and Snell’s Law of Refraction 505
12.2 Fresnel’s Equations 508
: 12.2.1 Incident Wave Polarized with its E Vector Normal
to the Plane of Incidence 508
12.2.2 Incident Wave Polarized with its E Vector Parallel
to the Plane of Incidence 510
12.3 Reflection and Refraction at the Interface Between Two Nonmagnetic
Nonconductors pill
12.3.1 The Brewster Angle 515
Example: Measuring the Relative Permittivity of the Moon’s
Surface at Radio Frequencies SF
12.3.2 The Coefficients of Reflection and of Transmission at an Interface
Between Two Nonconductors Sly
12.4 Total Reflection at an Interface Between Two Nonconductors 520
Example: Light Emission from a Cathode Ray Tube 525
Example: The Critical Angle and Brewster’s Angle 526
12.4.1 Demonstration of the Validity of Snell’s Law, of the Laws
of Reflection and Refraction, and of Fresnel’s Equations
in the Case of Total Reflection Si
12.4.1.1 The Wave Numbers ki, and ky, for the Reflected
Wave 528
12.4.1.2 The Wave Numbers ky, and ky, for the
Transmitted Wave 528
12.4.1.3 The Amplitudes of E and H in the Reflected and
Transmitted Waves 529
12.4.1.4 The Poynting Vector for the Transmitted Wave 531
12.5 Reflection and Refraction at the Surface of aGood Conductor SB
Example: Communicating with Submarines at Sea 535
Example: Standing Waves at Normal Incidence 535
Example: Transmission of an Electromagnetic Wave Through a
Thin Sheet of Copper at Normal Incidence 536
12.5.1 Demonstration of the Validity of Snell’s Law, of the Laws of Reflection
and Refraction, and of Fresnel’s Equations at the Interface Between
a Nonconductor and a Good Conductor 540
12.5.1.1 The Wave Numbers ke; and ke, for
Refraction in a Good Conductor 541
12.5.1.2 The Amplitudes of E and H in the Reflected and
Transmitted Waves 542
12.6 Radiation Pressure at Normal Incidence on a Good Conductor 543
Examples 546
12.7. Reflection of an Electromagnetic Wave by an Ionized Gas 547
12.8 Summary 549
Problems all
Xvi CONTENTS
Ils)
PROPAGATION OF
ELECTROMAGNETIC WAVES III Spy
Guided Waves
13.1 Propagation in a Straight Line 557
13.1.1 TE and TM Waves 560
13.1.2 TEM Waves 561
13.1.3 Boundary Conditions at the Surface of Metallic Wave Guides 565
13.2 The Coaxial Line 566
13.3. The Hollow Rectangular Wave Guide 568
13.3.1 The TE Wave 569
13.3.2 Internal Reflections 575
13.3.3 Energy Transmission Si
13.3.4 Attenuation 578
13.4 Summary 582
Problems 585
14
RADIATION OF ELECTROMAGNETIC WAVES ae3s.
14.1 Electric Dipole Radiation 595
14.1.1 The Scalar Potential V 597
14.1.2 The Vector Potential 4 and the Magnetic Field
Intensity H 598
14.1.3 The Electric Field Intensity E 601
14.1.4 The Average Poynting Vector and the Radiated Power 603
14.1.5 The Electric and Magnetic Lines of Force 605
14.1.6 The KX Surface 607
14.2 Radiation from a Half-wave Antenna 611
14.2.1 The Electric Field Intensity E 611
14.2.2 The Magnetic Field Intensity H 614
14.2.3 The Average Poynting Vector and the Radiated Power 614
14.3. Antenna Arrays 616
14.4 Electric Quadrupole Radiation 620
14.5. Magnetic Dipole Radiation 423
14.5.1 The Potentials V and A 424
14.5.2 The E and H Vectors 624
14.5.3 The Average Poynting Vector and the Radiated Power
14.6 Magnetic Quadrupole Radiation 426
14.7. The Electric and Magnetic Dipoles as Receiving Antennas
14.8 The Reciprocity Theorem 629
14.9. Summary 633
Problems 635
CONTENTS
XVli
APPENDIXES 641
A. Conversion Table 643
E. Waves 670
E.1 Plane Sinusoidal Waves 670
E.2. Waves on a Stretched String. The Differential
Equation for an Unattenuated Wave 674
E.3 Solution of the Differential Equation for an Unattenuated Wave
by the Separation of Variables 676
E.4_ Reflection of a Wave on a Stretched String at a Point
Where the Density Changes from p; to p» 677
E.5 Waves on a Stretched String with Damping. The Differential
Equation for an Attenuated Wave 679
E.6 Solution of the Differential Equation for an Attenuated Wave
by the Separation of Variables 681
E.7 Wave Propagation in Three Dimensions 682
E.8 Wave Propagation of a Vector Quantity 683
E.9 The Nonhomogeneous Wave Equation 683
ANSWERS 685
INDEX 699
~~”
7
ro‘ee Gi Vv
wi 6
Preface
CONTENTS
So as to reduce the mathematical requirements, we have included a chapter
on vectors (Chapter 1), a discussion of Legendre’s differential equation
xXx PREFACE
PROBLEMS
This second edition contains an extensive collection of problems, many of
which are drawn from the current literature of Physics and Engineering. Such
problems should serve as a clear indication that electromagnetism has a
wealth of applications in a wide variety of fields. They should favor the
transferability of what has been learned, and they should develop creativity.
Few persons will spend the time and energy required for solving all of the
longer problems. Those left unsolved should provide some profitable reading.
Double daggers (1) indicate those problems for which hints are given.
It is suggested that the problems be solved on a computer when this is
feasible.
The units used are those of the Systéme International d’Unités (designated
Sl in all languages), adopted in Paris at the 1960 Eleventh General Conference
on Weights and Measures. This system, which is based on the meter, kilo-
gram, second, ampere, kelvin, and candela, is essentially the same as the
Giorgi, or MKSA, system in the field of electromagnetism.
The exponential function for periodic phenomena can be either exp (jwf)
or exp (—jwf), since the real part of both of these functions is cos wt. We
have used the positive exponent, which is preferable at this level; otherwise,
an impedance becomes R — /X instead of the conventional R + jX.
ACKNOWLEDGMENTS
The authors owe much to Gilles Cliche, to Gaétan Marchand, and to Paul
Carriere who performed the numerical calculations and assisted in the design
of the figures for the first edition. Several other persons assisted in the
preparation of this second edition. Francois Lorrain, Luce Gauthier-Labonté,
and Gilles Labonté all checked the complete manuscript and proposed in-
numerable improvements. Francois Lorrain and Guy Basque performed the
numerical calculations required for the new figures. Jean and Michel Barrette
checked the answers of many of the problems, Roland Savage drew the
one-hundred-odd new figures, and Angele Elias typed over a thousand pages
of text with admirable care. Despite the invaluable assistance of all these
xxii PREFACE
MATHEMATICAL SYMBOLS
Approximately Real part of z Re z
equal to awe Imaginary part of z Im z
Proportional to ac Modulus of z |z|
Factorial n n! Decadic log of x log x
Exponential of x e7, exp (x) Natural log of x In x
LIST OF SYMBOLS XXV
VECTORS
We shall discuss electric and magnetic pneu etin terms of the yields of
A vector can be specified by its components along any three mutually perpen-
dicular axes. In the Cartesian coordinate system of Figure 1-1, for example,
the vector A has components 4,, A,, Az.
The vector A can be uniquely expressed in terms of its components
through the use of unit vectors i, j, k, which are defined as vectors of unit
magnitude in the positive x, y, z directions, respectively:
A= Ajit Ajj + Ak. (1-1)
The vector A is the sum of three vectors of magnitude A,, A,, A., parallel
to the x-, y-, z-axes, respectively.
The magnitude of A is
of the second and by the cosine of the angle between the two vectors. In
Figure 1-2, for example,
A-B = ABcos(¢ — 6) (1-5)
It follows from this definition that the usual commutative and distribu-
tive rules of ordinary arithmetic multiplication apply to the scalar product:
A-B = B.-A, (1-6)
and
A-(B+ C)=A-B+A4-C. (1-7)
From the definition of the scalar product it follows that
fi=1, jj=Hl, kek =1, (1-8)
g-k=0, kt=0, ij = 0. (1-9)
Then
A-B = (Axi + A,j + Azk)-(B.i + B,j + Bk), (1-10)
ards ated Bs. (111)
It is easy to check that this result is correct for two vectors in a plane, as in
Figure 1-2:
A-B = ABcos(¢ — 6) = ABcos¢cosé+ ABsingsin#é, (1-12)
= A,B, + A,B. (1-13)
Example A simple physical example of the scalar product is the work W done
by a force F acting through a displacement s:W = Fs.
whose magnitude is the product of the magnitudes of those vectors and the
sine of the angle between them. We indicate the vector product thus:
AXB=C, (1-14)
where the magnitude of C is
C = |ABsin(¢ — 6)|, (1-15)
with » and 6 defined as in Figure 1-2. The direction of C is given by the right-
hand screw rule: it is the direction of advance of a right-hand screw whose
axis, held perpendicular to the plane of A and B, is rotated in the sense that
rotates the first-named vector (A) into the second-named (B) through the
smaller angle.
The commutative rule is not followed for the vector product, since
inverting the order of A and B inverts the direction of C:
AX B= —(B X A). (1-16)
The distributive rule, however, is followed:
Examples A good physical example of the vector product is the torque T pro-
duced by a force F acting with a moment arm r about a point O, as
in Figure 1-3, where T = r X F.
A second example is the area of a parallelogram, as in Figure
1-4, where the area S = A X B. The area is thus represented by a
vector perpendicular to the surface.
1.2 INVARIANCE
You have probably noticed that we have two definitions for the scalar prod-
uct: Equations 1-5 and 1-11.
The first definition clearly does not refer to a particular coordinate sys-
tem and depends solely on the magnitudes of the two vectors and on the
angle between then y that is independent of the coordinate system
A, = Ay. (1-35)
Dia ae ’ jj’ #K
SSN ei in Re el (1-41)
Roath Bis (ie beeen
The cross product as defined in Eq. 1-22 is therefore also invariant.
We shall often be concerned with the rates of change of scalar and vector
quantities with both time and space coordinates, and thus with the time and
space derivatives.
The time derivative of a vector quantity is straightforward. In a time At,
a vector A, as in Figure 1-8, may change by AA, which in general represents
a change both in magnitude and in direction. Since AA has components A4A,,
AA,, and AA,,
AA = AA,i + AA,j + AA. (1-42)
On dividing AA by Ar and taking the limit in the usual way, we arrive at
the
definition of dA/dt:
Figure 1-8. A vector A, its incre-
ment AA, and their components.
dA me CAGE
AN) (2)
a At : (1-43)
24 A Ad u PAA Je ’
eo At Oa
Lehn
Sidaes ik athe
ae © (1-45)
The time derivative of a vector is thus equal to the vector sum of the time
derivatives of its components.
Example The time derivative of the position r of a point is its velocity v, and
the time derivative of v is the acceleration a.
OY =aor
ae Ox oh:
ay dy. :
(1-47)
SeTies
ay ,
(1-48)
A = gradf=Vf. (1-50)
In three dimensions the operator V is defined as
(1-51)
va ue a si a ah, (1-52)
and
(1) Its components at any point are the rates of change of the function
along the directions of the coordinate axes at that point.
(2) Its magnitude at the point is the maximum rate of change of the
function with distance.
(3) Its direction is that of the maximum rate of change of the function.
(4) It points toward Jarger values of the function.
ae OM aOt ae Om on Gees :
Vomias! Ws ay Jigs oma at ate ag eer ee SoEE):
The scalar quantity f is of course invariant.
Let us check whether V is invariant under a rotation of axes. We have
oe
ay = COS 6 ee
a7 cn
Sin 6 ay? 7
(1-57)
Thus
0 0 ; ta]
ae cos 6 =; — Sa (1-58)
Similarly,
0 : 0 0
ay = Sin ox? + cos @ ay” (1-59)
CA
oz ue oe
Using these last three equations, together with the equations of transforma-
tion for the unit vectors i, j, k, we do find Eq. 1-55.
We have shown that the V operator is invariant and therefore that the
gradient Vf is also invariant.
13
(1-62)
It should be obvious that this integral is invariant since the scalar product
A-da is invariant.
For a closed surface bounding a finite volume, the vector da is taken to
point outward.
Example Let us consider fluid flow. We define a vector pv, p being the fluid
density and v the fluid velocity at a point. The flux of pv through
any closed surface is the net rate at which mass leaves the volume
bounded by the surface. In an incompressible fluid this flux is always
equal to zero.
Fie (4.4 as
ue) ay dz, (1-63)
0A, dx ‘
The minus sign before the parenthesis is necessary here because, 4,i being
inward at this face and da being outward, the cosine of the angle between the
two vectors is —1. The net outward flux through the two faces is then
APDio, = ( (1-66)
OA 0A, 0A,
u )dr.
Ox 1 oy nk OZ
for each of the infinitesimal volume elements in the finite volume and the total
outward flux is
0A, , OA 0A
Biottot = | Co+ !
eS): a, :
(1-68)
iz da- . a
AW te fe
[(C43 Browns
4 EVE|
f y - zs —
/ V-Z
A dr ‘
(1-70)
-
This is the divergence theorem. Note that the left-hand side involves only the
values of A on the surface S, whereas the right-hand side involves the values
of A throughout the volume 7 enclosed by S. This is also called Green's
theorem.
If the volume + is allowed to shrink sufficiently, so that V-A does not
vary appreciably over it, then
As we have seen, the divergence is the outward flux per unit volume as the
volume 7 approaches zero.
The integrals
b b b
[ +a [ axar and fra.
evaluated from the point a to the point b on some specified curve, are ex-
amples of line integrals.
In the first one, which is especially important, each element of length dl
on the curve is multiplied by the local value of A according to the rule for the
scalar product. These products are then summed to obtain the value of the
integral.
Example The work W done by a force F acting from a to 6 along some speci-
fied path is
W = [? Feat,
a
(1-73)
where both F and dl must of course be known functions of the
coordinates if the integral is to be evaluated analytically. Let us
calculate the work done by a force F, which is in the y direction and
has a magnitude proportional to y, as it moves around the circular
path from a to 6 in Figure 1-12. Since
Feaipj and dla ueieew 7,
a, [Fa » ikkydy = sia (1-74)
fara =0 (1-75)
17
The circle on the integral sign indicates that the path of integration is
closed.
Let us calculate the value of the above integral in the more general case
where it is not equal to zero.
For an infinitesimal element of path dl in the xy-plane, and from the
definition of the scalar product,
A-dl = A, dx + A, dy. (1-76)
Thus, for any closed path in the xy-plane and for any A,
Now consider the infinitesimal path in Figure 1-13. There are two con-
tributions to the first integral on the right-hand side of the above equation,
one at y — (dy/2) and one at y + (dy/2):
There is a minus sign before the second term because the path element at
y + (dy/2) is in the negative x direction. Therefore
then
ieee
0 0 0
vx43 | ee (1-85)
A eden
he quantity g; is then its z component.
1.6 LINE INTEGRAL AND CURL 19
We shall calculate the curl from Eq. 1-88. For (V X v): we choose a
path parallel to the yz-plane. In evaluating
f v-dl
around such a path we note that the contributions are equal and
opposite on the parts parallel to the z-axis, hence (V X v); = 0.
Likewise, (V X v). = 0.
For the y-component we choose a path parallel to the xz-plane
and evaluate the integral around it in the sense that would advance
a right-hand screw in the positive y direction. On the parts of the
path parallel to the x-axis, v-dl = 0 since v and dl are perpendicular.
20
whereas at (x + Ax)
f vedl —c Ax Az
(1-93)
ea ee Sead Aca ia
Calculating V X v directly from Eq. 1-85,
toy ok
Gy ye) te)
V <i) = ax ay az aa al 1 (1-94)
OD Omecs
which is the same result as above.
V but lead to more complicated expressions for the gradient, the divergence,
and the curl.
Equation 1-87 is true only for a path so small that V X A can be considered
constant over the surface da bounded by the path. What if the path is so large
that this condition is not met? The equation can be extended readily to arbi-
trary paths. We divide the surface—any surface bounded by the path of
integration in question—into elements of area day, daz, and so forth, as in
Figure 1-15. For any one of these small areas,
A,
_ of
as ax’ A,
=a of
oy
_ of
A; rei az
:
(1 98)
and
2 2
vp oeLi i+ ee
Le, (1-100)
then
2 2 EY)
ViVi = Vi = Sat ot (1-101)
The divergence of the gradient is the sum of the second derivatives with
respect to the rectangular coordinates. The quantity V-Vf is abbreviated to
Vf, and is called the Laplacian of f. The operator V? is called the Laplace
operator.
The Laplacian is invariant because it is the result of two successive in-
variant operations.
Example We shall show in the next chapter that the Laplacian of the electro-
static potential is equal to zero in regions where there is zero space
charge.
S(% y, Z) = 4, (1-103)
in which q is a constant. This equation determines a family of surfaces in
space, each member of the family being characterized by a particular value
of the parameter g. An obvious example is x = qg, which determines the sur-
faces parallel to the yz-plane in Cartesian coordinates.
Consider now the three equations
Ail, ys Z) — Cal
which are chosen so that the three families of surfaces are mutually perpen-
dicular, or orthogonal. A point in space can then be defined as the intersection
of three of these surfaces, one of each family; the point is completely defined
if we state the values of gi, qo, gs corresponding to these three surfaces. The
variables qi, go, gs are called the curvilinear coordinates of the point, as in Fig-
ure 1-16.
Let us call di an element of length perpendicular to the surface q. This
element of length is the distance between the surfaces gq: and q, + dq in the
infinitesimal region considered. The element of length d/, is related to dq, by
the equation
dh, = h dg, (1-105)
in which /, is, in general, a function of the coordinates q, q2, qs. Similarly,
dl, = h, dq, (1-106)
dl; = hg dq. (1-107)
In the Cartesian system of coordinates, /,, 4), hz are all unity.
The unit vectors u;, 11, Ww, are of unit length, normal respectively to the
Gis 925 3 Surfaces, and oriented toward increasing values of these coordinates.
They are chosen so that uw. X m = wy.
The orientation of the three unit vectors depends, in general, on the point
of space considered. Only in rectangular coordinates do they all have fixed di-
rections.
The volume element is
dr = dl, dl; dl; = hy hy h3(dqy dq, dq;). (1-108)
We can now find the q’s, the h’s, the elements of length, and the elements
of volume for cylindrical and spherical coordinates.
z-axis, measured from the xz-plane in the right-hand screw sense; and z is
the distance from the xy-plane. The q’s are thus p, ¢, z in this case.
At the point P there are three mutually orthogonal directions specified
by three unit vectors: p; in the direction of the perpendicular from the z-axis
extended through P; yg; perpendicular to the plane containing the z-axis and
P in the direction corresponding to increasing y; and z in the positive z
direction. These unit vectors do not maintain the same directions in space as
the point P moves about, but they always remain mutually orthogonal.
The vector describing the position of P is
rs pmb ea | (1-109)
Note that the angle ¢ does not appear explicitly on the right-hand side; it is
given by the orientation of the unit vector py.
Elements of length corresponding to infinitesimal changes in the coordi-
nates of a point are important. If the coordinates ¢ and z of the point P are
kept constant while p is allowed to increase by dp, P is displaced by dr =
dp pi. On the other hand, if p and z are held constant while ¢ is allowed to
increase by dy, then P is displaced by dr = pdg ¢1. Finally, if p and ¢ are
held constant while z is allowed to increase by dz, then dr = dz x. For arbi-
trary increments dp, dg, dz,
1 x p if
q2 J 2 0
43 Zz Z 2)
hy 1 1 1
hy 1 p r
h, 1 1 r sin 6
ul i P1 ry
V1 i 71 A;
wi k zy 91
IE
a hy aq ct Uk 3s eee
uj, Se hy qo V1 ae hy 0q3 U1, (1-117 )
where wy, V1, Ww, are the unit vectors as in Figure 1-16.
In cylindrical coordinates,
ogy ae ON ea (1-118)
On the z-axis, p and sin @ are both zero and neither of these expressions
is valid.
We must ask ourselves again whether or not these expressions are in-
variant. Let us just check the one for cylindrical coordinates.
Let us first check whether it is invariant to a change from cylindrical
coordinates to Cartesian coordinates. You can do this yourself by using the
transformation equations
(1-120)
(1-121)
You will find that Vf in cylindrical coordinates does transform into the proper
expression for Vf in Cartesian coordinates, so that Vf is invariant to a change
from cylindrical to Cartesian coordinates. The inverse is equally true.
Now what about invariance to a change from one system of cylindrical
coordinates S; to another system of cylindrical coordinates S,? The gradient
is also invariant under this transformation for the following reason. Call C,
and C, the corresponding Cartesian systems. We have just shown that we have
invariance in going from S, to C;. Now we also have invariance in going from
C; to Cy, as we showed earlier in Section 1.4, and invariance in going from
C, to Sp. It follows that the gradient is invariant if one changes from S; to S».
The gradient expressed in spherical coordinates is similarly invariant to
a change from a spherical system of reference to either a Cartesian system, or
to a cylindrical system, or to a different spherical system.
element; /,, /», hz are the h values at the center. Then the outward flux
through the left-hand face is
== — (4.
(4.- 24d)
ee] (p,
lee shed)
tee] ie (, _ an
dai
ae ke a
It must be remembered that /. and h3 may be functions of gq, as well as Aj.
If we neglect differentials of order higher than the third, then
By a similar argument,
0 da
d®p = Ahh; dgz dqz + aa:(Ayhyh3) ne dqo dqs (1-125)
for the right-hand face. The net flux through this pair of faces is then
0
d®rp = aa (Ajhohs) dqi dgz dqs. (1-126)
1 y
The same calculation can be repeated for the other pairs of faces to find
the net outward flux through the bounding surface. Dividing by the volume
of the element then gives the divergence. Since we have considered only
differentials up to the third order, we have already passed to the limit r— 0,
and
Substituing the h’s and q’s gives the divergence in cylindrical coordi-_
(1-129)
aL aA
aa . male (7? sin 0A,) + 25
STA — rae], (1-130)
a¢ sin 0A,) + 2
These expressions are again not valid on the z-axis where p and sin @ are
zero.
where the path of integration must lie in the surface defined by qi = constant,
and where the direction of integration must be related to the direction of the
unit vector uw; by the right-hand screw rule. For the paths labeled a, b, c, d in
Figure 1-22, we have the following contributions to the line integral:
(b) (4+
get
5% ae 2)(is- coas) dq,
va
Then, adding up these four terms and neglecting the higher order differentials
because we are interested in the limit S — 0,
1 0 as
F)
V =
Mars) hohs dqz dgs fee(Ashs) digo ds 0q3 (Ashe) de den
| NESS
ot? ito 0
= lols E (Ashs) = Ags (Ash) | (1-134)
Corresponding expressions for the other two components of the curl can
be found either by proceeding again as above or by proper rotation of the
indices. Finally,
hyuy hov, haw
1 0 0 0 : 2
Va eA
Iyhzhs3|0g. 9q2 9s (1-135)
hy A, hy A» hg As
Ay pAw -A,
1 OmmnO) 0
V XA andlor 36 ay CD)
Aw TA Frsil 0A.
1.10 SUMMARY
V- 0 : cs) : ) ?
se Cape T 927°
(1-51)
and the gradient of a scalar function f is written
i naw (1-52)
eas ose
The gradient gives the maximum rate of change of f with distance at the
point considered, and it points toward larger values of f.
The flux ® of a vector A through a surface S is
Q = [Ada (1-62)
Ss
over a specified curve is the sum of the terms 4-dl for each element dl of the
curve between the points a and b and, for a closed curve C which bounds a
surface S, we have Stokes’s theorem:
A, A, A;
is the curl of the vector function A. The above surface integral is evaluated
over any surface bounded by the curve C.
The Laplacian is the divergence of the gradient:
ViaOf ae ue
,54+
oF, ,
& (1-101)
The Laplacian of a vector in Cartesian coordinates is defined as
V?A = V?A,i + V2A,j + V2A_ k. (1-102)
In cylindrical coordinates, Figures 1-17 and 1-18,
r = pp, + zm, (1-109)
dr =dppit+pdegitdzn, (1-110)
dr = p dp de az, (1-112)
V a p p ¢ 4
f = a oe AF Fi a5 az? (1-129)
Pl pi Fi
iL|)@) 0 0
V xX A = alas Sden de I E
(1-136)
Ay pas Ay
. NS) feSR RIES of
VFaf = one~ | ap: = ae at a7 : 3
(1-140)
These last four equations are meaningless on the z-axis, where ps0:
1.10 SUMMARY 35
dr = r’ sin
6 dr d6 do, (1-115)
moh 1 of lof
Ai easeat ae (-119)
no 0A,
V4 === =
oa ae eea
rain 0g" Ga)
ry 10, rsin 6g;
1 ay 8} 0
re eee | BED)
A, YAg rsin 6A,
i
ZeOl SO) COLO 0; mm Oy. S
1 of =
ep r or ay or? a alt) ver 06? ipr?sin? @ dg?” CRE)
These last four equations are also meaningless on the z-axis, where sin 6 = 0.
In other than Cartesian coordinates, VA is defined as follows:
VA=V(V-A)-—VXVXA.
We shall have occasion to use the following operations:*
AOL AE
ERETAME
V(4A-B) = (B-V)A+ (A:V)B+BX(V XA+AXW XB);
V-(fA) = (Vf)-A4+f(V-A);
* See J. Van Bladel, Electromagnetic Fields (McGraw-Hill, New York, 1964), Appendixes
1 and 2, for an extensive collection of vector identities and theorems.
VECTORS
36
when the gradient is calculated at the field point, with the same unit vector;
| wan I |faa:
[x nar=- [4x da
7 8
where S is the surface that bounds the volume r.
PROBLEMS
NOTE: Double daggers ({) indicate problems for which hints are given at the end of the
Problems section.
1-5. Let a and b be two unit vectors lying in the xy-plane. Let a be the angle a
makes with the x-axis, and let 8 be the angle b makes with the x-axis, so that
a = cosai+sinaj and b = cos8i+ sinGj.
Show that the trigonometric relations for the sine and cosine of the sum
and difference of two angles follow from the interpretation of a-b anda &X b.
1-7. Show that the distributive rule applies to the vector product (Eq. 1-17).t
1-8. Show that a X (b X ce) = b(a-c) — c(a-b).
1-10. A gun fires a bullet at a velocity of 500 meters/second and at an angle of 30°
with the horizontal. Find the position vector r, the velocity vector v, and the
acceleration vector a of the bullet, ¢ seconds after the gun is fired. Draw a
sketch of the trajectory and show the three vectors for some time f.
1-11. Let r be the radius vector from the origin of coordinates to any point, and
let 4 be a constant vector. Show that V(4-r) = 4.
ee k=12. Phe vector r is directed from P’(x’, y’, z’) to P(x, y, z).
= (a) If the point P is fixed and the point P’ is allowed to move, show that the
gradient of (1/r) under these conditions is
A
1 ry
vs©) is
! — = —
where r; is the unit vector along r. Show that this is the maximum rate of
change of 1/r.
rQ--8
(b) Show similarly that, if P’ is fixed and P is allowed to move,
r r
1-16. It is suggested that V-(A X r/r’) is identically equal to zero as long as the
vector A is a constant.
Is this correct?
AS
af, of
yao aes ay’ A, Se
=—Z Ae
Dies
x ie A, ee
x ay y ae
1-18. Show that V X (f4) = (Vf) X 4+ f(V X A), wheref is a scalar function
and 4 is a vector function.
1-19. Show that V-(A X D) = D-(V X A) — A(V X D), where A and D are any
two vectors.
VECTORS
38
1-20. One of the four Maxwell equations states that V x E = —0B/dt, where E
and B are respectively the electric field intensity in volts/meter and the mag-
netic induction in teslas at a point.
Show that
|f Beal
is 20.0 microvolts over a square 10.0 centimeters on the side when B is
2.00 10-*¢ tesla and normal to the square.
1-22. Show, by differentiating the appropriate expressions for r, that the velocity
r in cylindrical coordinates is
ppi + pegi + 2m,
while in spherical coordinates it is
rr; + 160, + rsin Ogi.
1-23. The vector 4 is everywhere perpendicular to, and directed away from, a
given straight line; that is, in cylindrical coordinates, A. = A, = 0. Calculate
the net outgoing flux for a volume element, and show that
eet
p dp
. A fluid rotates with an angular velocity w about the z-axis. The direction of
rotation is related to that of the z-axis by the right-hand screw rule.
(a) Find the velocity v of a point in the fluid, and show that
VX v= Zook:
1-25. The gravitational forces exerted by the Sun on the planets are always directed
toward the Sun and depend only on the distance r. This type of field is called
a central force field.
Find the potential energy at a distance r from a center of attraction when
the force varies as 1/r?. Set the potential energy equal to zero at infinity.
[ vrar
= [faa,
wheref is a function of (x, y, z) and S is the surface bounding the volume 7.{
. Show that
[wxda=- faxda,
where 4 is an arbitrary vector and S is the surface bounding the volume r.t{
Hints
1-7. Write out the components in Cartesian coordinates.
1-28. You can solve this problem without having to write out the equations in full.
1-29. Multiply both sides by a constant vector (scalar product) and use the identity of
Problem 1-14.
1-30. Multiply both sides by a constant vector (scalar product) and use the identities
of Problems 1-6 and 1-19.
CHAPTER »,
ELECTROSTATIC FIELDS I
Electrostatic Fields in a Vacuum
It has been found experimentally that the force between two stationary elec-
tric point charges Q, and Q, (a) acts along the line joining the two charges,
(b) is proportional to the product Q,Q,, and (c) is inversely proportional to
the square of the distance r separating the charges.
If the charges are extended, the situation is more complicated in that the
“distance between the charges” has no definite meaning. Moreover, the pres-
ence of Q, can modify the charge distribution within Q,, and vice versa,
leading to a complicated variation of force with distance.
We thus have Coulomb’s law for stationary point charges:
ae (2-1)
41
where the force F,, is measured in newtons; the charges Q, and Q), in
coulombs (the magnitude of which will be defined in terms of magnetic
interactions in Chapters 6 and 7); and the distance r, in meters. The quantity
47 appears in explicit form so as to simplify other equations that are used
much more extensively than Coulomb’s law. The constant ¢ is called the
permittivity of free space:
where the factor of 9 is accurate to about one part in 1000; it should really
be 8.988.
Example The Coulomb forces in nature are enormous when compared with
the gravitational forces, which are given by
42 ELECTROSTATIC FIELDS I
MIN
F = 6.67 X 10" ~~ 1. (2-4)
the held or
tne (
2 at
E = lim —- (2-6)
Q’>0
Consider a test point charge Q’ that can be moved about in an electric field.
The work W required to move it at a constant speed from a point P; to a
point P, along a given path is
The negative sign is required to obtain the work done against the field. Here
again, we assume that Q’ is so small that the charge distributions are not
appreciably disturbed by its presence. ‘
If the path is closed, the total work done is
Now the term under the integral on the right is simply dr/r? or —d(l/r).
r has the
The sum of the increments of (1/r) over a closed path is zero, since
line integral
same value at the beginning and at the end of the path. Then the
in moving a point charge Q’ around any
is zero, and the net work done
is fixed, is zero.
closed path in the field of a point charge Q, which
QO but by
If the electric field is produced not by a single point charge
onding to each in-
some fixed charge distribution, the line integrals corresp
44 ELECTROSTATIC FIELDS I
dividual charge of the distribution are all zero. Thus, for any distribution of
fixed charges,
el =, (2-11)
_E=-vv,! (2-13)
where V is a scalar point function, since V X VV = 0
as in Figure 2-2.
Note that the electric field intensity E determines only differences between
the potentials at two different points. When we wish to speak of the electric
45
(2-17)
It will be observed that the sign of the potential V is the sam e as that of Q.
4 7 T 1]
electri
| tential V aS well
The set of points in space that are at a given potential defines an equi-
potential surface. For example, the equipotential surfaces about a point
charge are concentric spheres. We can see from Eq. 2-13 that the electric
field intensity FE is everywhere normal to the equipotential surfaces (Sec-
tion 1.4).
46 ELECTROSTATIC FIELDS I
We have so far been thinking in terms of point charges. Now electric charges
are usually distributed over macroscopic bodies, which are composed of
positively charged nuclei and negative electrons. This brings up two questions.
(a) First of all, can we calculate the field owtside an electrically charged
body by assuming that the charge distribution inside the body is continuous?
If our assumption is valid, we can calculate the field by integrating over the
charge distribution. If our assumption is not valid, then we must find some
other form of calculation.
It is, in fact, usually appropriate to treat the discrete charges carried by
nuclei and electrons within macroscopic bodies as though they were con-
tinuously distributed within these bodies. Even the largest nuclei have
diameters that are only of the order of 10-4 meter. Nuclei and electrons
are therefore so small and so closely packed compared to the dimensions of
ordinary macroscopic objects that we may define an average electric charge
density p, measured in coulombs/meter®, as AQ/Ar’, where AQ is the net
charge within Ar’.
The volume Ar’ may be assumed large enough to render the fluctuations
in AQ with time, or from one Ar’ to a neighboring one, negligible. At the
same time the volume Ar’ may be assumed small enough to permit the use of
integral calculus.
Thus, instead of summing the potentials of a large number of individual
charges, we may integrate over a continuous distribution p(x’, y’, z’). An
element of charge p dr’ contributes at a point P(x, y, z) outside dr’ a potential
Wiehrmaie (2-18)
tos lope ee
idee p dr’
igs Are Jo Tr aes,
The integral is evaluated over all regions where the net charge density p is
not zero.
2.5 GAUSS’S LAW 47
Gauss’s law relates the flux of E through a closed surface to the total charge
enclosed within the surface. Consider Figure 2-3, in which a point charge Q
is located inside a closed surface S at a point P’. We can calculate the flux
of the electric field intensity E through the closed surface as follows. he
flux of E through the element of area da is
oh) Q nda
=.= (2-20)
UN 2
Figure 2-3. A point charge Q
located at P’ inside a closed
surface S. The total flux of the
electric field intensity E through
the surface S is Q/€.
Ee aay
Are
(2-21)
where dQ is the element of solid angle subtended by da at the point P’.
To find the total flux of E, we integrate over the whole surface S:
If the point charge Q were outside the surface at some point P”, the
solid angle subtended by the surface S at P’’ would be zero.
The situation remains unchanged if the surface is convoluted in such
a way that a line drawn outward from P’ cuts the surface at more than one
point. The total solid angle subtended by the closed surface is still 47 at an
inside point P’, and is still zero at an outside point P”’.
If more than one point charge resides within S, the fluxes add alge-
braically, and the total flux of E leaving the volume is equal to the total
enclosed charge divided by e. This is Gauss’s law.
Gauss’s law provides us with a powerful method for calculating the
electric field intensity FE of simple charge distributions.
If the charge enclosed by the surface S’ is distributed over a finite
volume, the total enclosed charge is
Q= |pdr’, (2-23)
where p is the charge density and 7’ is the volume enclosed by the surface S’.
Then
2.5' GAUSS’S LAW 49
Pe aes (2-26)
- 25 \Lead } (2-27)
J
Also, according to Coulomb’s law,
y= [ 4reor
2%, (2-28)
where the integral is evaluated over the surface of the shell. Equating
the two values of V and remembering that o is O/47a’, we obtain a
purely geometrical relation concerning a sphere of radius a and a
point P at a distance R from its center, as can be seen by canceling
the Q’s:
Os (ag okele (2-29)
4iregR * 4ra’ An eor
This isPoisson’s equation. Ina region of the field where the charge density p
is Zero,
electrostatic field is conservative, and that we can describe the electric field
intensity in terms of a potential V: E = —VYV.
Finally, (a), (b), (c), and the principle of superposition led us to Gauss’s
law V-E = p/e. When we expressed this in terms of the electric potential V,
we found Poisson’s equation V?V = —p/«, and Laplace’s equation V*°V = 0
for the special case where p = 0.
2.7 CONDUCTORS
ict
Coulomb’s law applies within conductors, even though the net field is
zero. Then Gauss’s law V-E = p/e, which is a consequence of Coulomb’s
law, must also be valid within conductors. Now, under static conditions,
FE = O inside a conductor, and hence the charge density p must be zero.
As 2 corollary gigeiaeeiehic bit0S Cosiclesec ea
At the surface of a conductor, the electric field intensity E must be
normal, for if there were a tangential component of E, charges would flow
along the surface, which would be contrary to our hypothesis. Then, accord-
ing to Gauss’s law, just outside the surface, E = o/«, where o is the surface
charge density.
2.8 CALCULATION OF ELECTRIC FIELDS
53
Example A hollow conductor has a charge on its inner surface that is equal
in magnitude and opposite in sign to any charge that may be en-
closed within the hollow. This is readily demonstrated by considering
a Gaussian surface that lies within the conductor and that encloses
the hollow. Since E is everywhere zero on this surface, the total-
enclosed charge must be zero.
(a) We can use Eq. 2-7 and integrate over the whole charge distribution.
We must of course keep in mind that the electric field intensity is a vector
quantity.
(b) Another way consists in writing down the potential produced by an
element of charge, integrating over the whole charge to obtain V, and then
calculating the electric field intensity from E = —VV. The potential calcula-
tion is generally simpler because the potential is a scalar quantity that can be
integrated more easily than the vector E.
(c) In some cases we can integrate either Laplace’s or Poisson’s equa-
tion, depending on whether the charge density p is zero or not, if the geom-
etry is simple enough, and if we have sufficient information about the field
to determine the constants of integration.
(d) We can use Gauss’s law if the charge distribution possesses a sym-
metry that ensures constancy of the electric field intensity over certain
imaginary surfaces in the field. When this method is applicable, it is usually
the easiest one.
(e) We shall find still another method in Section 4.7.
These five types of calculation are often inadequate, and we must then
use other methods, described in Chapter 4 and in Appendix B. Even then,
54 ELECTROSTATIC FIELDS I
where s is the distance from the volume element to the point P. The
axis along which @ = 0 can be taken to be the line OP. The element
of electric field intensity is written d’e,, since it is a third-order
differential.
It should be obvious, from the symmetry of the charge distribu-
tion, that E, must be radial. For example, while the charge element
shown in Figure 2-5 produces an electric field intensity d’e, that is
not along the radius OP, there is another symmetrically placed ele-
ment that produces a symmetrically oriented field of the same mag-
nitude, and the net result is a field along OP.
SB)
cos 6’ = 7 as (2-36)
Now we wish to eliminate sin 6’ d0’ from the expression for d°E,. We
can find sin 6’ d6’ in terms of 7’, r, s, by differentiating the above
equation. Here we must remember that r is a constant and that r’ is
taken as a constant when we integrate Eq. 2-34 with respect to 0’.
Thus we must differentiate Eg. 2-36, taking both r and r’ as con-
stants, and thus
one rr
(2-37)
56 ELECTROSTATIC FIELDS I
If we substitute Eqs. 2-35 and 2-37 into Eq. 2-34, and integrate,
)ds ar’,(2-38)
s=r+r’ , rP—r? r :
15;
reg ho tAbie r (1$55
BeArey ee ee (2-39)
3
ie 4ire, r”
where Q is the total charge (4/3)7R’p, and where nr; is the radial unit
vector directed outward.
The vector E, is directed outward along OP if Q is positive,
and inward along OP if Q is negative.
This result is the same as if the total charge Q were concentrated
at the center of the sphere.
eek
V,= tren (2-41)
r
The electric potential V,, like E,, is the same as if the total charge Q
were concentrated at the center of the sphere.
To find E,, we now calculate VV,. By symmetry, E, must be
radial, hence
emery 6 EC
Le eit Gla ae cage 1
Airey r?
ri, (2-42)
as previously.
(ee aay
E,= -4, (2-46)
where A is a constant of integration. We shall be able to determine
its value later on, after we have found E;.
2.8 CALCULATION OF ELECTRIC FIELDS Sy
477r°E, = 2 (2-47)
€
and again i
caAA
Ares r1. (2-48)
The electric field intensity thus increases linearly with r inside the
spherical charge distribution.
Yoian= (Og
2(4 eis a
_) :
(2-55)
Gaia) o Ps
Ae 4renR® 2 AreoR ( )
where the second term is the potential at the surface of the sphere,
and where the first term is the increase above the surface value for
the interior points.
Then
Ey= VV = a = (2-57)
O53, OViN 2 ee bs
x(r a) = en aa
Ee ese: (2-61)
i 3
or 3€
pe ae (2-62)
where B is a constant of integration.
It is intuitively obvious that E; cannot become infinite at the
center of a uniform spherical charge distribution; B must therefore
be zero, and
= = ri. (2-63)
We are now in a position to find the value of the constant of
integration A, which we found when calculating E, with the equation
of Laplace. Should not the two values we have found for the electric
field intensity E, one valid inside and one valid outside (Eqs. 2-39
and 2-52), be equal at the surface of the charge distribution atr = R?
How could they be different? According to Gauss’s law, they could
be different if we had a surface charge density as on the surface of a
charged conductor. But, by assumption, the charged sphere has a
uniform volume density p out to the radius R and there can be no
pR?[
2€
pR?|
3€ |
al is
ER = R = Ben (2-64)
so that
moh Gee ees
7 Ta eae eb)
and Eq. 2-46 does give the correct value for E,.
E, = s ry (2-67)
as previously.
Figure 2-7 shows E and V for our spherical charge distribution
of radius R as a function of the radial distance r.
The electric dipole shown in Figure 2-8 is one type of charge distribution that
is encountered frequently. We shall return to it later in this chapter and also
in Chapters 3 and 14.
The electric dipole consists of two charges, a positive and a negative
charge of the same magnitude, separated by a distance s, which is small
compared to the distance r to the point P at which we require the electric
potential V and the electric field intensity E.
At P,
y_ 2 (Z ") (2-68)
Ameo \71p =a
where
r
= Vas
[1+(5) + £cos |bes (2-70)
=ay fae2 f+
\4p £0088)
r +38 (s+
\4r?2
2eose) — ---
i/
(2-71)
62
F : s
or, if we neglect terms of order higher than —>
r
ie = 3c0s? Gk
ie 1-— = cos 6 eT 5 (2-72)
Similarly,
F214
To
Xcose + 5-8 y)F—, (2-73)
k=me apaele 23
cert cos 6, (2-76)
fe oy
SS sre 0g a
‘Id intensity of a dipole thus falls off as the cube of the distance.
The equation for the lines of force of a dipole can be found by con-
sidering Figure 2-9, which shows an element dl of a line of force. Since the
two vectors are parallel, the components ofE and of dl are proportional and
rdo_E& _ sing
dr E, 2cosé (2-79)
dr _ 2d(sin@)
6)
r sind en)
r= Asin? 6. (2-81)
This equation determines the family of lines of force for an electric
dipole. The constant A is a parameter that varies from one line of force to
another.
Figure 2-10 shows lines of force and equipotential lines for an electric
dipole. Equipotential surfaces are generated by rotating equipotential lines
around the vertical axis.
Figure 2-10. Lines of force (arrows) and equipotential lines for the dipole of Figure
2-8. The dipole is in the vertical position at the center of the figure with the
positive charge very close to and above the negative charge. In the central region
the lines come too close together to be shown.
V = i (2 — 22 + °), (2-82)
i 8)
The ratios r/ra and r/r, can be expanded as previously, except that s re-
65
P(r,0,¢)
Let us consider an arbitrary charge distribution of density p(x’, y’, z’) occupy-
ing a volume 7’ and extending to a maximum distance rfax from the origin of
coordinates O. We select O either within the volume or close to it. Such a
distribution is illustrated in Figure 2-12.
pdr’
V= ibAap: (2-87)
where 7’ is the distance between the point of observation P and the position
P'(x', y’, z') of the element of charge p dr’:
al ih | EP a CRRA
tl
eatYayttazLF)
IPS Va ill oeER
15 PI | ax’ ae 3 ay’ 32 w\ (=) ale ’ (2-89)
where the subscripts 0 indicate that the derivatives are evaluated at the origin.
In squaring the second bracket, we consider the factors x’, y’, z’ to be con-
stants. The same applies to all the terms that follow.
Now
3 (5) - - m5 meee
” =
(2-90)
and
Oval x if
E (=) |,7a =) a r. (2-91)
where / = x/r is the cosine of the angle between the vector r and the x-axis.
The other first derivatives with respect to y’ and z’ are given by similar
expressions.
The second derivatives of the third term on the right-hand side of Eq.
2-89 are calculated similarly, the substitution of x’ = 0, y’ = 0, 2’ = 0,
r’’ = r being made only at the end.
The result of the calculation is that at the point P
1 p dr’ Le, ; np dr
y= |)
1 r 4rre9
+ [Ao Bey ee a
pdr’
4+ 4 GP — 1x? + § Gm? — 1y* + 4 Gn? — 12?) T+ +++, (2-92)
68 ELECTROSTATIC FIELDS I
where /, m,n are the direction cosines of the line joining the origin to the point
of observation P.
Let us now examine the various terms in succession. The first term is
merely the electric potential that we would have at P if the whole charge were
concentrated at the origin. It is called the monopole term and is zero only if
the total net charge is zero. If the charges are all of the same sign, then it is
the most important term of the series, since it decreases only as 1/r.
The second term varies as 1/r?, as does the electric potential of a dipole.
Let us find the value of this term when the charge distribution is simply a
dipole, as in Figure 2-8. The integral must then be evaluated over the two
charges Q and —Q situated at z’ = s/2 and z’ = —s/2 respectively. The
result is the value of V for the dipole, as in Eq. 2-74. The second term V2 of
Eq. 2-92 can thus be taken to be a dipole term. In fact,
mee Lil i
ee 4rreor” Ces
if we set
= face iar
= T wi
(2-99)
V3 = =
Aas [3mn Pyz + 3nl pez + 31m Pry
0)
distribution. Let us calculate the average electric field intensity at the interior
point directly. To do this we shall first calculate the average E for a single
point charge.
Figure 2-13 shows a point charge Q located at a distance r’ from the
center of a sphere of radius R. The z-axis is taken to be along the line joining
the center of the sphere and the charge Q. By symmetry, the average field
over the volume of the sphere must be along the z-axis.
We wish then to calculate
where 7 is the volume of the sphere. This integral can be separated into two
parts, one over the spherical shell between the radii 7’ and R, and one over
the sphere of radius 7’.
We can show that the integral over the outer volume is zero by
using an argument somewhat analogous to that used on pages 57 to 61 to
find the electric field inside a uniform spherical charge distribution. In Figure
2-13 the solid-angle element dQ intercepts the volume elements dr; and dry
in the shell. Since the value of E, decreases as the square of the distance from
Q, whereas dr increases as the square of the distance (see Eq. 2-49), their
product remains constant. However, at dr, E, is positive, whereas at dry E; is
negative; thus the two contributions to the integral cancel. The same is true
‘app. ELECTROSTATIC FIELDS I
for the whole outer shell, the contributions to the integral canceling in pairs.
To calculate the integral of E, dr over the inner volume, which is equal
to the same integral over the complete volume 7, we shall use spherical polar
coordinates and set the origin at the position of the charge Q. Then, at the
point P,
= Apepie oe 6, (2-113)
and
g=2r 6=7 r!! = — 2r’cos6
/py / i qm €08 8 (r'” sin 6 dr’”’ do dy). (2-114)
e e g=0 @= a/2 r/=0 Areor
In this integral, 7’’ varies from zero to —2r’ cos 6, which is the distance from
Q to the surface of the inner sphere in the direction 6. The negative sign comes
from the fact that cos 6 is negative. Then 6 sweeps from 7/2 to z and ¢ from
0 to 27. Thus
ih
Ear = — oes cos’é sin 6 dé, (2-115)
T €0 a/2
Eaves!
AregR® (118)
This average electric field intensity has been calculated for a single charge
Q on the z-axis. For an arbitrary charge distribution the field is the super-
position of the fields due to the different charges, according to the theorem of
superposition. Then E is again as above, with Pp equal to the dipole moment
of the arbitrary charge distribution within the sphere of radius R.
the other charges has some definite value. Each charge thus has associated
with it a definite potential energy, either positive or negative, and the system
as a whole has a potential energy that we shall calculate.
We assume that the charges remain in equilibrium under the action of
the electric forces and of restraining mechanic forces.
Let Q;, Qo, Q3, ---, On be the N charges; let 712 be the distance between
Q, and Qs», r13 the distance between Q, and Q;, and so on, as in Figure 2-14.
Now let Q; recede to infinity slowly, so that the electric and the mechanical
forces are always in equilibrium. In this way there is no acceleration and,
therefore, no kinetic energy involved. The other charges remain fixed.
The decrease W, in the electric potential energy of the charge is equal
to the work done by the electric forces. This is the product of the charge Q,
and the potential V; produced by the other charges at the original position of
Q,;. Then
w= 2 (24 24 2), (2-119)
Arey \ Nie ri3 Yin
Note that all the charges except Q; are represented in the series of terms
within the parentheses.
Now that Q,; has been removed, let Q») recede to some point infinitely
distant from Q;. The decrease W, in the potential energy of Q» is
be removed without any change in energy, since it is left in a zero field once all
the other charges have been removed.
The total potential energy of the original charge distribution is then
W= Wit Wit Wse+---+ Wy, (2-121)
a eae =)
lad ~ae ee)
+2 ( 0 ): (2-122)
Let us now rewrite the array of terms within the parentheses, adding in,
to the left of and below the diagonal line of zeros, terms that are equal to their
counterparts on the other side of the diagonal. Then every term of the series
appears twice, and
Pi
Tr€0
(o +2424
Vio ri3
24... 4 Ov)
V4 MN
pi (S+0 +2424
Té9 \ Foi 93 V4
+ 2)
ron
Thus the first line is Q,V1, the second line is Q.V2, and so forth, so that
(2-125)
“We en not igken into account the self energy of the individual charges,
that is, the energy that would be liberated if each one were allowed to expand
to an infinite volume.
2.14 POTENTIAL ENERGY OF A CHARGE DISTRIBUTION 715
The reason the factor } appears in the above equation should be clear
from the reasoning we have used to arrive at it. It is that the potential at the
position of a given charge at the time it is removed to infinity is less, in general,
than the potential at the same point in the original charge distribution. On
the average, the potential at the time of removal is just one-half the potential
in the original charge distribution.
You will note that this energy W, which does not include the energy
required to assemble the individual charges themselves, can be either positive
or negative. It is the energy required to assemble already existing point
charges. For example, if we have two positive point charges Q; and Q», then
the positive charge Q, is situated at a position where the potential Vi, due to
the other charge Q», is positive, so that Q,V, is positive. The product QoV,
is also positive and the potential energy of the charge distribution
QO1Q>/4eor? is positive. However, if we have a positive charge Q, and a nega-
tive charge Q», then V; is negative and Q,V; is negative. The other product
Q2V. is similarly negative, and W is negative. Note also that, for a single
point charge, the above W is zero.
For a continuous electric charge distribution of density p(x’, y’, 2’), we
simply replace Q; by p dr’ and the summation by an integration:
W= tf Vp dr’. (2-126)
This operation is misleading, however, because we have now included the
energies required to assemble the individual macroscopic charges. In fact, as
we shall see in the next section, this integral is always positive. It is evaluated
over any arbitrary volume 7’ that contains all the charges in the system.
If we have conducting bodies carrying surface charge densities o, then
the stored energy is
W= at aV da’, (2-127)
w=ov,! (2-128)
where Q is the net charge carried by the conductor and V is its potential.
16 ELECTROSTATIC FIELDS I
(2-129)
ron Bel5
and thus its stored energy can also be written as
The capacitance is the electric charge that must be added per unit increase in
potential.
One might well ask why the capacitance C should be a constant for a
given conductor, as we have implied above. This can be shown as follows. The
isolated conductor carries a charge Q, and the electric potential in the region
surrounding it is V(x, y, z). At the surface of the conductor the charge density
iS é£ (Section 2-7), or — & times the rate of change of V in the direction normal
to the surface. Since the potential V satisfies Laplace’s equation V*V = 0
(Eq. 2-32), and since this equation is linear, any multiple of V is also a
solution. Let us therefore increase V everywhere by some factor a. Then,
since E = — VV, the surface charge density will also increase by a, and the
total charge Q on the conductor will increase likewise. The charge on an
isolated conductor is therefore proportional to the potential, and C is a
constant.
Imagine now that we have two uncharged isolated conductors. If a
charge Q is transferred from one to the other, a potential difference AV is
established between them and, by definition, the capacitance C between the
conductors is O/AV.
The unit of capacitance is the coulomb/volt, or farad.
The potential energy W can be related to the electric field of the charge
distribution as follows.
According to Poisson’s equation,
p= — aV?V (2-131)
at every point in the field. If we substitute this into Eq. 2-126, then
since E = |V V|. The volume 7’ need not be infinite now, but only large
enough to include all regions where E differs from zero. This integral takes
into account the self-energies of the charges; it is always positive, and it is
obviously not zero for a single charge.
The potential energy W of Eq. 2-126 is equal to the W we have deduced
from it above and is therefore also positive for all charge configurations.
Equation 2-137 shows that the energy associated with a charge distribu-
tion, that is, the energy required to assemble it, starting with a configuration
in which the charge is spread over an infinite volume, may be calculated by
associating with each point of the field an energy density
OME
Zz = 7 E 24joules/meter’.3 (2-138)
A
718 ELECTROSTATIC FIELDS I
In the next section we shall arrive at this same result by another method.
We can therefore calculate the energy stored in an electric charge distri-
bution, either in terms of the charge density p and the potential V, or in terms
of the electric field intensity E.
Thus
Oz
= . 2-141
87regR ( )
= © f ar, (2-142)
Pan
aa e 4rr?
(25) dr = SreR ce (2-143)eS
which is the same result as above. We have integrated from R to
infinity, since there is zero E inside the sphere.
Figure 2-15. The local electric charge density at the surface of a con-
ductor gives rise to oppositely directed electric field intensities ¢/2¢€ as
shown by the two arrows on the left; the other charges on the conductor
give rise to the field ¢/2¢) shown by the arrow on the right. The net
result is ¢/€) outside, and zero inside. The vector n is a unit vector
normal to the conductor surface, and it points outward.
electric field intensity due to the local surface chage density must be o/2¢
inwards and o/2¢) outward.
The element of charge o da itself therefore produces exactly half the total
field at a point outside, arbitrarily close to the surface. This is reasonable, for
the nearby charge is more effective than the rest.
Now if o da produces half the field, then all the other charges must
produce the other half, and the electric field intensity acting on o da must be
o/ 2€.
The force dF on the element of area da of the conductor is therefore
given by the product of its charge o da multiplied by the field of all the other
charges:
dF = 5-0da, (2-144)
and the force per unit area is
dF “ ig
da 5s eg 2h)
Now this force is just equal to the energy density that we found in the
preceding section. We should really have expected this for the following
reason. Let us imagine that the conducting bodies in the field are disconnected
from their power supplies. Then, if the electric forces are allowed to perform
mechanical work, they must do so at the expense of the electric energy stored
in the field. Now imagine that a small area a of a conductor is allowed to be
80 ELECTROSTATIC FIELDS I
pulled into the field by a small distance x. The mechanical work performed is
ax times the force per unit area. It is also equal to the energy lost by the field,
which is ax times the energy density. Then the force per unit area must be
equal to the energy density.
If we use the same type of argument for a compressed gas, we find that
its energy density is equal to its pressure.
‘We have therefore found another and a more simple demonstration of
the fact that we can calculate the energy stored in an electric field by inte-
grating eE?/2.
This second demonstration is in some ways preferable to that of Section
2.15. First, itis purely local; it does not require that the field be integrated
over an infinite volume. As you will remember, we had to perform such an
operation to eliminate the first integral in Eq. 2-136. Second, it relies solely on
Gauss’s law and on the fact that the force on a charge Q situated in an electric
field Eis QE. Both of these postulates are valid even if the field is not constant.
The proof of Section 2.15, on the other hand, is valid only for electrostatic
fields.
It is important to note that
Example , Electric forces are never very large. For example, even for E = 108
volts/meter, the electric force is only 4.4 newtons/meter?,
or 0.45
kilogram/meter?, or about 2 ounces/foot?. See Problem 3-23.
if the plates are moved closer together by a small distance ds, the
electric field intensity E remains unchanged, according to Gauss’s
law. The volume between the plates decreases, however, and hence
the energy in the field also decreases. Since we must have conserva-
tion of energy, the energy lost by the field has gone into the work
F,,, ds against the mechanical forces F,,, holding the plates apart. Then
and
€ o? Q?
Nis NIN
2 = he, S= 2€,S ' 2-14
as)
Since the electric force F, is in equilibrium with the mechanical
force, it must be of equal magnitude and oriented so as to pull the
plates together.
2.17 SUMMARY
where r is the distance between the charges and r; is a unit vector pointing
from Q, to Q». This is Coulomb’s law. We consider the force Fy as being the
product of Q» by the electric field intensity due to Qa,
E, — 2 Let (2-5)
or vice versa.
82 ELECTROSTATIC FIELDS I
Vaxeh =, (2-12)
hence
Rien v4 (2-13)
where
Vyasa (2-31)
€0
then follows from the differential statement of Gauss’s law and from the
definition of the potential.
When the charge density is zero we have Laplace’s equation
Vv = 0. (2-32)
The above is summarized schematically in Figure 2-17.
The electric field produced by simple charge distributions can be calcu-
lated in four different ways: (a) by evaluating dE for an element of charge
and integrating over the complete charge distribution, (b) by calculating V
in the same manner and then calculating its gradient, (c) by integrating
Poisson’s or Laplace’s equation, or (d) by using Gauss’s law. The fields due
to more complex charge distributions can be calculated by other methods
described in Chapter 4 and in Appendix B.
The potential produced by an arbitrary charge distribution at a distance
r > Tmax is the same as that produced by a point charge, a point dipole, a
point quadrupole, and so forth, all situated at the origin, where the monopole
83
Coulomb’s Law
Potential V
Poisson’s Equation
\
|Laplace’s Equation
Gauss’s Law
Figure 2-17.
carries the total charge of the distribution, the dipole has the dipole moment
of the charge distribution,
W = al Vp dr’ (2-126)
OF as
W = 3/ E? dr’. (2-137)
In the first integral, r’ must be chosen to include all the charge distribution, and
in the second it must include all regions of space where £ is nonvanishing. The
assignment of an energy density eE’/2 to every point in space therefore leads
to the correct potential energy for the whole charge distribution.
Finally, electric forces on conductors can be calculated by either of two
methods. We can either (a) utilize the fact that the force per unit area is
o?/2e = e&?/2, where E is the electric field intensity near the surface of the
conductor, or (b) equate the work performed by the electric forces to the work
84 ELECTROSTATIC FIELDS I
PROBLEMS
2-1. Show that e, is expressed in farads/meter, that E is expressed in volts/meter,
and that the volt is a joule/coulomb.
(3-2Xa) Calculate the electric field intensity E that would be just sufficient to bal-
~~ ance the gravitational force on an electron.
(b) If this electric field were produced by a second electron located below
the first one, what would be the distance between the two electrons? (The
charge on an electron is —1.6 X 107-19 coulomb and its mass is 9.1 & 107*!
kilogram.)
3./It is known that the maximum electric field intensity which can be maintained
in air at NTP is about 3 X 10° volts/meter. This determines the maximum
surface charge density.
If a spherical body is charged to half this maximum surface charge
density and placed in a vertical electric field of 1.5 & 108 volts/meter, what
is the largest solid sphere that can be supported by the electric field ?Choose
any material you wish for the sphere.
Assume that it is nonconducting and that its charge is uniformly dis-
tributed over its surface.
2-4. A charge + Q is situated at (—a, 0, 0) and a charge —2Q is situated at (a, 0, 0).
Is there a point in space where E = 0?
(25. A thin infinite conducting plate carries a surface charge density c. Show that
a one-half of the electric field intensity E at a point situated z meters from the
surface of the plate is due to the charge located on the plate within a circle
of radius V3 z.
2-6. Show that the potential due to two parallel line charges of opposite polarity
is (A/27€) In (71/re), where r; is the distance from the point considered to the
negative line, and rz is the corresponding distance to the positive line.
If your answer turns out to be 0/0, try another method.
2-11. (a) Show that the electric field intensity near a flat sheet of charge having a
uniform surface density o is ¢/2¢.
(b) Deduce the value of the electric potential.
2-14. Two infinite parallel plates separated by a distance s are at potentials 0 and Vo.
(a) Use Poisson’s equation to find the potential V in the region between the
plates where the space charge density is p = po(x/s). The distance x is meas-
ured from the plate at zero potential.
(b) What are the charge densities on the plates?
. You can use the property demonstrated in Problem 2-15 to calculate the
potential distribution within a square, two of whose adjacent edges are main-
tained at 100 volts while the other two are maintained at 0 and at 50 volts.
(a) First draw a square grid of 36 points. Of these, 20 will be on the edges
of the square, and 16 will be in the interior.
(b) Guess carefully the potentials at the interior points and write them on your
grid. This is your first approximation.
(c) Now you can correct your first approximation as follows. Start at an
interior point, near one corner of the square. Let us call this point P. Calculate
the sum ofthe potentials at the four nearest points and subtract four times the
potential at P. This number is called the residual at P. Show this number on
your grid, preferably in a different color, and repeat the calculation for the
other interior points.
To obtain a second approximation, select the points where the residuals
are the largest and add to your original estimates one quarter of the residuals.
Round off the potentials to the nearest volt. This makes the potential distribu-
tion satisfy more closely the condition found in Problem 2-15.
(d) Now calculate the new residuals and repeat the operation until the re-
siduals are everywhere less than 2, corresponding to corrections of less than
one-half volt.
You might wish to solve this problem on a computer.
(e) Sketch in some equipotentials.
This is a good illustration of the relaxation method of calculation, which
we largely owe to R. V. Southwell, and which is described in many books
on applied mathematics. The method is applicable to a wide variety of prob-
lems, and not only to the solution of Laplace’s equation.
2-17, A charged conductor has a small-diameter hole in a region where the surface
charge density is oc.
Show that the electric field intensity in the hole close to the surface of the
conductor is o/2¢.
2-18. The electric field intensity at the surface of a real conductor falls from its ex-
ternal value FE, to zero within the conductor in a finite distance 56,
1 Fi 6
Show that the surface charge density ¢ = if p(x) dx = e£,.
PROBLEMS
87
2-20. (a) Calculate the electric field intensity in volts/meter at the surface of an
iodine nucleus (53 protons and 74 neutrons).
(b) Find the electric potential at the center of the nucleus.
Assume that the charge density is uniform and that the radius of the
nucleus is 1.25 X 107!°41/8 meter, where 4 is the total number of particles
in the nucleus.
(2A) A spherical charge distribution is given by
p=m(1-2), <a)
as) (Ca)
(a) Calculate the total charge QO.
(b) Find the electric field intensity E and the potential V outside the charge
distribution.
(c) Find £ and Vinside.
(d) Show that the maximum value of E is at (r/a) = 0.745.
(e) The above charge distribution applies roughly to light nuclei. Draw
graphs showing p, E, and V as functions of r/a for calcium (atomic number
20), assuming that pp = 5.0 & 102° coulombs/meter* and a = 4.5 femtometers
(1 femtometer = 10~' meter).
2-23. Show that the dipole moment of an arbitrary charge distribution is independ-
ent of the choice of origin, provided the net charge in the distribution is zero.
88 ELECTROSTATIC FIELDS I
2-24. Compute the dipole moment of a spherical shell with a surface charge density
co = 0, cos 6, where @ is the polar angle.
2-25. Calculate the potential for a dipole exactly and indentify the quadrupole and
octupole terms.
2-26. A charge Q is uniformly distributed inside a cube of side a. Show that the
dipole and quadrupole moments of the charge distribution are zero.
2-29. Find the capacitance per unit length of a capacitor consisting of a pair of
infinite coaxial cylinders having inner and outer radii a and 4, respectively.
2-31. Two capacitors of capacitance C, and C, have charges Q; and Qs, respectively.
(a) Calculate the amount of energy dissipated when they are connected in
parallel.
(b) How is this energy dissipated?
2-32. A capacitor consisting of two concentric spheres is arranged so that the outer
sphere can be separated and removed without disturbing the charges on either.
The radius of the inner sphere is a, that of the outer sphere is 6, and the charges
are O and — Q respectively.
(a) If the outer sphere is removed and restored to its original form, find the
increase in energy when the two spheres are separated by a large distance.
(b) Where does this extra energy come from?
PROBLEMS 89
2-34. According to the theory of relativity, a particle at rest has an energy myc®,
where mp is the particle’s mass and c is the velocity of light (Section 5.15).
Imagine that this is the electrostatic energy of the electron.
Find the radius of the electron (a) if the charge e is distributed uniformly
throughout its spherical volume, and (b) if the charge is distributed uniformly
r its spherical surface.
Gaye have shown in Section 2.16 that the force of attraction between the plates
of a parallel-plate capacitor is (1/2)e9E2S when the plates are disconnected
from the voltage source. Check this result by repeating the calculation for the
case where the capacitor is connected to the source. The force should be the
same.
You will find that one half of the energy supplied by the source performs
mechanical work and that the other half increases the energy in the field.
This is a general result.
2-38. Find the time required for the plates (mass mp per unit area) of a parallel-
plate capacitor to come together when released from a separation Xo, (a) when
the plates are charged with a charge density o and then insulated, and (b) when
the plates are maintained at a constant potential difference V.
2-39. A variable capacitor consists of two thin coaxial cylinders of radii a and 5,
with (6 — a) <a, free to move with respect to each other in the axial direc-
tion.
90 ELECTROSTATIC FIELDS I
(a) Using energy methods, compute the magnitude and direction of the force
on the inner cylinder when it is displaced with respect to the outer one.
(b) Explain how this force arises.
|
2 ia)Imagine the following simple-minded high-voltage generator. A parallel-plate
capacitor has one fixed plate that is permanently connected to ground, and
N
one plate that is movable. When the plates are close together at the distance
_s the capacitor is charged by a battery to a voltage V. Then the movable plate
is disconnected from the battery and moved out to a distance ns. The voltage
on this plate then increases to nV, if we neglect edge effects. Once the voltage
has been raised to nV, the plate is discharged through a load resistance.
(a) Verify that there is conservation of energy.
(b) Can you suggest a rough design for such a high-voltage generator with a
more convenient geometry?
2-42. Normal seeds can be separated from discolored ones and from foreign objects
by means of an electrostatic seed-sorting apparatus that operates as follows.
The seeds are observed by a pair of photocells as they fall one by one inside
a tube. If the color is not right, voltage is applied to a needle that deposits a
charge on the seed. The seeds then fall between a pair of electrically charged
plates that deflect the undesired ones into a separate bin. One such machine
can sort peas at the rate of 100/second, or about 2 tons per 24-hour day.
(a) If the seeds are dropped at the rate of 100/second, over what distance
must they fall if they must be separated by 2 centimeters when they pass be-
tween the photocells? Neglect air resistance.
(b) Assuming that the seeds acquire a charge of 1.5 X 10-° coulomb, that
the deflecting plates are parallel and 5 centimeters apart, and that the poten-
tial difference between them is 25,000 volts, how long should the plates extend
below the charging needle if the charged seeds must be deflected by 4 centi-
meters on leaving the plates? Assume that the charging needle and the top of
the deflecting plates are close to the photocell.
Hint
2-37. Do not forget to take into account the energy supplied either by or to the battery.
CHAPTER 3
ELECTROSTATIC FIELDS I
Dielectric Materials
Dielectrics differ from conductors in that they have no free charges that can
move through the material under the influence of an electric field. In dielec-
trics, all the electrons are bound; the only motion possible in the presence
of an electric field is a minute displacement of positive and negative charges
in opposite directions. The displacement is usually small compared to atomic
dimensions.
A dielectric in which this charge displacement has taken place is said to
be polarized, and its molecules are said to possess induced dipole moments.
These dipoles produce their own field, which adds to that of the external
charges. The dipole field and the externally applied electric field can be compa-
rable in magnitude.
In addition to displacing the positive and negative charges, an applied
electric field can also polarize a dielectric by orienting molecules that possess
a permanent dipole moment. Such molecules experience a torque which tends
to align them with the field, but collisions arising from the thermal agitation
of the molecules tend to destroy the alignment. An equilibrium polarization
is thus established in which there is, on the average, a net alignment. >
This is, in fact, a simplified view of dielectric behavior because many
solids are not made up of molecules but of individual ions which interact with
a large number of other ions. Sodium chloride is an example of such a solid.
We shall limit most of our discussion to dielectrics that are made up of
molecules that are either distorted or oriented, or both distorted and oriented
in an electric field.
In a polarized dielectric each molecule acts as an electric dipole of moment
92 ELECTROSTATIC FIELDS II
p. We shall treat these dipoles both from a molecular and from a macro-
scopic point of view: we shall consider the fields of individual dipoles, as well
as those of large numbers of dipoles.
In the end we shall deal with dielectrics entirely from the macroscopic
point of view, but we must justify this procedure by first examining carefully
the molecular aspect.
3 STHES PLECTRICSPOLARIZATIONS?
Let us first consider nonpolar dielectrics, that is, dielectrics whose molecules
have zero permanent electric dipole moment. We shall not have to consider
polar dielectrics until Section 3.9.
The electric polarization P is the dipole moment per unit volume at a
given point and is a macroscopic quantity. If p is the average electric dipole
moment per molecule in a small volume r, and if NV is the number of molecules
per unit volume, then
abr (3-1)
It will be recalled from Problem 2-23 that the dipole moment of a charge
distribution is independent of the origin of coordinates, provided the net
charge in the distribution is zero, as it is in molecules.
The volume + over which we define P must be chosen large enough to
render the fluctuations in P, from one instant to the next or from one rt toa
neighboring one, negligible. Although on the average the volume 7 contains
Nr molecules, we may expect, from statistical considerations, that there can
be fluctuations of +(Nr)!/ in the number Nr. The actual number of molecules
per unit volume thus fluctuates between N — (N/r)!2 and N + (N/r)!2, or
by +100 (1/Nr)!”? %. This restriction on the size of 7 is ordinarily unim-
portant. Take the case of water, for example, where N Y 3 X 1028 molecules/
meter’. Even in a submicroscopic cube measuring 100 angstroms or 10-8
meter on the side, the statistical fluctuation in P is only about one half of one
percent.
Figure 3-1 shows a block of dielectric material with a dipole moment P per
unit volume, P being a function of position within the dielectric. Let us
calculate the electric potential V which the dipoles in the dielectric produce at
a point P outside. This potential can then be added to that produced by all
93
the other charges in the system to give the total potential. The negative
gradient of this total potential at the point P is the electric field intensity at
that point.
At the point P(x, y, z) the potential dV due to the dipole P dr’ situated at
(en ez.).15
1 P-r,
dV
Aney Fr?
dr’, (3-2)
pracd i!|P-v" =)
Arey |7 r
a. (3-4)
This integration can be carried out analytically only for simple geome-
tries and only if P is a simple function of position within the dielectric.
Furthermore, P is never known apriori, although we can often deduce it from
measurable quantities. We shall nevertheless assume, for the time being,
that P is known. This integral can be put into a more interesting form by
transforming it through the vector identity
I P I V oP
V== Det
—— (vae . )dr ’ Aicet i -— dr dr, 3-6
(3-6)
ELECTROSTATIC FIELDS II
94
We can use the divergence theorem to express the first term as a surface
integral; then
Now both integrals in Eq. 3-7 involve a 1/r dependence and are multiplied by
1/4ré. We thus have the following remarkable result: these two terms are
exactly the potentials that would result from surface and volume charge
distributions having densitites
(3-8)
and
(3-9)
respectively, where n is the unit outward normal vector at the surface of the
dielectric. Thus
en til oda’ 1 pdr’
ie af. rp T Are, i" x ale
We have omitted the prime on V since it is obvious that the divergence must
be rmence at thea where the volume charge density is py.
» diele herefore be replaced by the bound chargedistributions —
OD — — re re ee field outside the dielectric. We shall see
later that we can use the same procedure to find a macroscopic potential at
points inside the dielectric.
We can demonstrate with the aid of Figure 3-2 that the bound electric
charge densities o, and p, represent actual accumulations of charge. Let us
first consider the surface density o,. We imagine a small element of surface da’
inside the dielectric. Under the action of the field, an average charge separa-
tion s is produced in the molecules. Positive charge crosses the surface by
moving in the direction of the field; negative charge crosses it by moving in
the opposite direction. For the purpose of our calculation, we may consider
the positive charge to be in the form of point charges Q and the negative
charge to be in the form of point charges — Q. Furthermore, we may consider
95
Figure 3-2. Under the action of an electric field E, which is the resultant of an
external field and of the field of the dipoles within the dielectric, positive and
negative charges in the molecules are separated by an average distance s. In the
process a net charge dO = NQs-da’ crosses the surface da’, N being the number
of molecules per unit volume and Q the positive charge in a molecule. The vector
da is perpendicular to the shaded surface. The circles indicate the centers of
charge for the positive and for the negative charges in one molecule.
the negative charges to be fixed and the positive charges to move a distance s.
The amount of charge dQ that crosses da’ is then just the total amount of
positive charge within the imaginary parallelepiped shown in Figure 3-2. The
volume of this parallelepiped is
dr’ = s-da’, (3-11)
and
dQ = NQs-da’, (3-12)
where JN is the number of molecules per unit volume and Qs is the dipole
moment p of a molecule. Then
dQ = P-da’. (3-13)
If da’ is on the surface of the dielectric material, dQ accumulates there in a
layer of thickness s-n. Since the thickness of the layer is of the order of the
dimensions of a molecule, we may treat the charge as a surface distribution
with a density
dQ
= = P-n. (3-14)
The bound surface charge density o, is thus equal to the normal component
of the polarization vector at the surface.
We can show similarly that —V-P represents a volume density of
charge. The net charge that flows out of a volume r’ across an element da’ of
96 ELECTROSTATIC FIELDS II
its surface is P-da’, as we found above. The net charge that flows out of the
surface S’ bounding 7’ is thus
O= | Pda’,is
(3-15)
and the net charge that remains within the volume r’ must be — Q. If p» is the
volume density of the charge remaining within this volume, then
eal ig (3-18)
We refer to o, and p, as either bound, polarization, or induced charge
densities, as distinguished from the free or conductible charge densities o; and
py. Bound charges are those that accumulate through the displacements that
occur on a molecular scale in the polarization process. The other charges are
“called free charges. The conduction electrons in a conductor and the electrons
injected into a dielectric with a high-energy electron beam are examples of —
free charges.*
oulomb’s law applies to any net accumulation of charge, regardless of—
a
other matter that may be present.
*The concepts of free and bound charges, and even the concept of polarization, are
based on the assumption that the dielectric is composed of molecules. If there are no well-
defined molecules, the values of os», oy, and P become rather arbitrary. See, for example,
Edward Purcell, Electricity and Magnetism, Berkeley Physics Course Volume 2 (McGraw-
Hill, New York, 1965) p. 344,
3.3. ELECTRIC FIELD AT AN INTERIOR POINT 97
0
| san = - 2 f var (3-19)
Using the divergence theorem on the left and substituting the value of p, on
the right from Eq. 3-18,
0 oP
[ven dr = 2 [ve (eye |v : ap OT (3-20)
We have put the d/d¢ under the integral sign because it is immaterial whether
the derivative with respect to the time is calculated first or last. Since 7 is any
volume, the integrands must be equal and the polarization current density 1s
P
j= -.amperes/meter?. (3-21)
It is often necessary to know the electric potential V and the electric field
intensity FEat points within a dielectric. We may wish to know, for example,
the potential difference between two charged conductors separated by a
dielectric.
The electric field intensity at a particular time and at a particular point
within the dielectric is a rapidly fluctuating quantity, both with regard to
position and to time. For example, it reaches enormous magnitudes at the
surfaces of nuclei. There are also large variations in its direction: on one side
of a nucleus it is in one direction, and on the other side it is in the opposite
direction. Thermal agitation of the molecules also produces large time
fluctuations at any given point.
The macroscopic electric field intensity FE, which is what we really wish
to know, is the space-and-time average of this electric field intensity.
The macroscopic electric field intensity E is a slowly varying function of
the coordinates and is independent of the time in the static case. It is this
electric field intensity that we shall integrate in order to calculate potential
differences.
We shall show that the part of this macroscopic field that originates from
the dipoles within the dielectric can be calculated from the bound charge
distributions o, and p, discussed above, exactly as in Eq. 3-10. This is not
intuitively obvious.
The reason the type of calculation used for the exterior point might not
be valid at an interior point is that, in the immediate neighborhood of the
point considered, there may be large numbers of molecules for which our
98
expressions for the dipole field are not valid, since Eq. 3-2 requires that the
distance between the dipole and the point considered be much larger than
the dimensions of the dipole.
To calculate the macroscopic electric field intensity E inside a dielectric,
let us consider Figure 3-3, which shows a small imaginary sphere of radius R
centered on P. The surface S” of this sphere divides the dielectric into two
volumes, 7’’ outside and 7’” inside. The potential at P may be calculated for
all the polarized material outside S’” by treating P as an exterior point, as in
Section 3.2; however, the molecules within S” are too close to P to be treated
in this fashion. The volume r’”’ is macroscopically small, and the macroscopic
quantities E, P, and p, = —V-P do not vary significantly from one side to the
other.
Thus the macroscopic electric field intensity that the dipoles of the
dielectric produce at the point P is the sum of two terms, which we shall
calculate separately:
E=E" +E", (3-22)
where r; is the unit vector in the direction from the element of volume dr’’ to
the point P, and r is the distance from dr”’ to P. Proceeding similarly for the
surface charges,
}
Vie L ov da’ I py dr”
f I
ely oy da”
oy da’’
The first cos @ factor comes from o», and the second comes from taking the
100 ELECTROSTATIC FIELDS II
component of the electric field intensity in the axial direction, since symmetry
requires the resultant field to be in that direction. On performing the integra-
tion over the whole surface S’’, from 6 = 0 to 6 = zm, we find that the third
integral of Eq. 3-23 is equal to P/3¢e0.
ier ae
where WN is the number of molecules per unit volume and p is the average
dipole moment per molecule.
Thus the contribution of the molecules within S” to the macroscopic
electric field intensity E at P is equal in magnitude and opposite in direction
to the field of the bound charges on the surface S” (the third integral of Eq.
3-23), and
E = B+ Ms 4rreq
/ HE
r? Arey
| pe
pe
The second integral is calculated only over that portion of the dielectric
that is outside the spherical surface S”’. Of course, it would be simpler to
calculate the second integral if it extended over the complete volume 7’ of the
dielectric, but this would add the term
1 Db dr!"
Tite
Are) Jan?
Since the bound charge density does not vary significantly over the small
sphere 7’’’, this integral is merely the electric field intensity at the center of a
uniform spherical charge distribution. For every element of charge giving an
element of electric field intensity in one direction, there is a symmetrical
element of charge giving a field in the opposite direction, and this integral is
zero. The second integral of Eq. 3-26 can therefore be extended to the
com-
plete volume 7’ of the dielectric, and
_ 1 Ob da’ 2 l Pb dr’
That part of the electric field intensity that is due to the dielectric itself
can therefore be calculated at points inside the dielectric just as it is calculated
outside, namely by replacing the dielectric by a space charge distribution
—V-P and a surface charge distribution P-n.
=— Arey Js’ r
where p; is the free charge density,p, is the bound charge density, r is the
distance between the source point where the total charge density is py + p, and
the field point where V and E are calculated, r; is a unit vector pointing from
the source point fo the field point, and r’ is any volume enclosing all the
charges.
Gr)
Jas)
Ses
Reiners,
cai oes Figure 3-5. The field of an indi-
p=Qs vidual dipole in an electric field.
where p is the dipole moment of the molecule and r,, is its volume.
If we further set 7, ~ 1/N, then
Evo E+ es (3-32)
3
This expression for the local field is at best approximate, owing to the
approximations that have gone into it. Although we assumed a spherical
shape for the molecule, we set tm = 1/N. This means that all of space is filled
with molecules, which contradicts the assumption of spherical shape. None-
theless, dielectric behavior predicted from the above expression for the local
field agrees surprisingly well with the experimental data for many substances.
Fortunately, our dielectric theory does not require that the expression
for the local field be quantitatively accurate; our conclusions will all be valid
if only
104 ELECTROSTATIC FIELDS II
eter eb ‘, (3-33)
0
where JN, is the number of molecules of the first type per unit volume, with the
average dipole moment pi, N, is the number of the second type per unit
volume, and so on. The theory is not otherwise modified.
Selb re CURIGsSUSGEPTIBIV
ERY Sy,
We can now find a relationship between the electric polarization P and the
macroscopic electric field intensity E. To do this, we consider the electric
dipole moment p induced in a single molecule. The magnitude of p depends
on the local field E,,.: the positive and negative charges are separated under
the action of the local field until the restoring force, which is an internal
electrostatic force arising from their separation, just balances the local field
LOrce:
In most dielectrics the molecular charge separation is directly propor-
tional to, and in the same direction as, the local field. Thus
P = AB, = a (B+b =) 0
(3-35)
where the constant a is known as the molecular polarizability.
Dielectrics which show this simple dependence of polarization on local
field are said to be linear and isotropic. Many commercially important
dielectrics are homogeneous, as well as linear and isotropic. We shall desig-
nate such materials as Class A dielectrics and shall confine our attention to
them for the time being.
The dipole moment per unit volume in Class A dielectries is
P = No :
t= (Nob/e) al!
= ex, (3-38)
3.6 THE DIVERGENCE OF E. THE ELECTRIC DISPLACEMENT D 105
We have found that the electric field due to the atomic and molecular dipoles
of a polarized dielectric can be calculated from the bound charge densities and
from Coulomb’s law, just as for any other charge distribution. Let us investi-
gate the implications of this fact on Gauss’s law, which is a direct consequence
of Coulomb’s law.
Gauss’s law relates the flux of the electric field intensity E through a
closed surface to the total net charge Q; enclosed within that surface:
VB =e (3-41)
a ees (3-42)
106 ELECTROSTATIC FIELDS II
V-E = +(/—V-P),
0
(3-43)
or
V -(ek + P) = py. (3-44)
The vector esf + P is therefore such that its divergence depends only on the
free charge density p;. This vector is called the electric displacement and is
designated by D:
D = ok + P. (3-45)
Thus
V-D = py. (3-46)
In integral form Gauss’s law for D becomes
[Deda = |evar
S T
(3-47)
and the flux of the electric displacement D through a closed surface is equal
to the free charge enclosed by the surface.
From the definition of the electric displacement,
ee (3-48)
and the electric field intensity E inside a dielectric is the sum of two fields:
D/e associated with the free charges, since
Ve (7)
D(a eeePy"
(3-49)
%
Np On = 12
P=
: Pa
ait field intensity
Electric =oe =eg
= = = |]
Electric displacement ——>—__—__
obtain in the particular field. Figure 3-6 shows the vectors p, FE, D and the
electric charge densities o; and o, for a parallel-plate capacitor.
In writing down the expressions for V-E and V-P, we have implicity
assumed the existence of the space derivatives of E and P, These derivatives
do not exist, of course, either at a point charge or at the interface between two
media. In such cases, we must revert to the integral form of Gauss’s law:
Veh (3-54)
€,€0 €
Then
4 pe eal (3-55)
Er€Q €
e,
—l
P=D—6E=
€;
D, (3-57)
and since the dielectric is homogeneous (c, is independent of the coordinat
es)
by hypothesis,
V-eP= € — 1 1
( )v-p = (1- =|Pf, (3-58)
€; ,
3.7 CALCULATION OF ELECTRIC FIELDS INVOLVING DIELECTRICS 109
or
re (*és‘) (3-59)
The total charge density
for linear and isotropic dielectrics. Note that p;/e is equal to p;/¢, but that no
such relation exists for the surface charge densities a; and a; + oa». Indeed, oy
and o, are unrelated. For example, a dielectric carrying zero of acquires a oy
when it is placed in an electric field. (However, see the second example that
follows.)
110 ELECTROSTATIC FIELDS II
= =. = Ge 5, (3-63)
where S is the area of one plate. The capacitance is therefore in-
creased by a factor e, through the presence of the dielectric.
The measurement of the capacitance of a suitable capacitor
with and without a dielectric provides a convenient method for
measuring a relative permittivity e,.
SeWere 1 \
eeeTee
(3-66)
€r €;
and
iS = (3-67)
as in Eq. 3-60.
Since e, > 1, the bound surface charge density o, and the free
surface charge density o; have opposite signs. The bound charge
density 1s also smaller in magnitude than the free charge density by
the factor (e, — 1)/e,. This relationship is always true whenever a
Class A dielectric is in contact with a conductor, or carries a free
surface charge density.
The free charge density a; on a conductor in contact with any
dielectric is always equal to the electric displacement D just inside
the dielectric. This follows from Gauss’s law, as in the case of the
parallel-plate capacitor. Furthermore, if a conducting surface is
introduced into a dielectric so as to coincide with an equipotential
surface, then the free surface charge density a; on the conductor at
a particular point is equal in magnitude to the electric displacement
D in the dielectric at that point.
= mre (3-68)
ee~ Aarecer?
3-69
Ce
Fey€; jee
én 4nr
(3-70)
At the outer surface of the sphere,
¢-—1 Q
op = Pen =P= aa ne (3-71)
Fas (3-77)
=Pa
eee (3-78)
which is to be compared to Eqs. 3-68 and 3-69 for the field inside the
sphere. At the surface of the sphere the electric field intensity is dis-
continuous, the magnitude just outside the surface being e, times as
large as the magnitude just inside the surface. The difference is due
to the bound charges, which produce an opposing field within the
dielectric.
(a) (b)
Figure 3-9. (a) Bar electret polarized uniformly parallel to its axis.
(b) The E field of the bar electret is the same as that of a pair of circular
plates carrying uniform surface charge densities of opposite polarities.
114
Figure 3-10. Lines of E (solid) for the bar electret of Figure 3-9. Lines of
D are shown broken inside the electret; outside, they follow the lines
of E.
between the electrodes and the dielectic. The free and the bound
charges have opposite signs. The sample is then cooled to room tem-
perature without removing the electric field,
In the presence of an external electric field the polarization P
inside an electret is related to the total E as follows:
P= x.0E + Po (3-79)
where Py is the permanent polarization. Then
D = @E + P = e(1 + x,)E + Po, (3-80)
= «E+ Pp. (3-81)
Since there are no free charges inside the material, p, = 0 (from
3.8 THE CLAUSIUS-MOSSOTTI EQUATION 115
Eq. 3-60) and the electric field of an electret can be calculated solely
from the surface charges.
As an exercise, let us see how one can calculate the field of the
bar electret illustrated in Figure 3-9a. We assume that the polariza-
tion P is uniform. We also assume that the surface density of free
charges oy is zero.
The electric field intensity E both inside and outside is the same
as that of a pair of parallel circular plates carrying charge densities
+P as in Figure 3-9b. This field is shown as solid lines in Figure 3-10.
The lines of D are identical to the lines of E outside the electret
since D = ¢,E there. Inside the electret, D is «JE + P with E and P
pointing in approximately opposite directions. If the electret were a
thin sheet polarized in the direction normal to its surface, the E inside
would be P/e, and D would be zero. However, inside a bar electret,
€,E is smaller than P, the result that D points in the direction of P,
and E and D point in approximately opposite directions, as in Figure
3-10.
ie
Evo = E + (3-82)
3€
As we saw, this equation is not necessarily correct, in that the factor multi-
plying P/e) may differ from 3, but 3 is of the right order of magnitude.
If we set b = 3, then from Eq. 3-36,
ji (E x) (3-83)
3€
or
ee ,
e(e, — IE = Na ( te —— ) (3-84)
and
Glee Ne. (3-85)
—> €; +. 2} 3€
the molecular
where N is the number of molecules per unit volume and a 1s
polarizability.
y.
* Sections 3-8 to 3-10 can be omitted without losing continuit
116 ELECTROSTATIC FIELDS II
where N; and a; are the appropriate quantities for the ith type of molecule or
atom.
Equation 3-86 is known as the Clausius-Mossotti equation. It relates the
relative permittivity «, to the mass density of a material, since
N= ArNa, (3-87)
where p is the mass density, V4 is Avogadro’s number, and M is the molecular
weight. In MKSA units, p is expressed in kilograms/meter*®, NV, is the number
of molecules in one kilogram molecular weight, or 6.02 X 10%, and M is the
kilogram molecular weight (for example, the kilogram molecular weight of
oxygen is 32 kilograms). Then
Me— Na
a= ay. (3-88)
p&+2 3¢€
SO ePOCARSDIELECTRICS
field. However, the local field exerts a restoring force between collisions and,
on the average, effects a net alignment, and consequently a net dipole moment
P per unit volume.
To compute this net polarization, we consider a unit volume containing
N dipoles. In the absence of an external electric field, the dipoles are oriented
at random, and, at any instant, dN are oriented at angles lying between @ and
6 + dé with respect to a given direction. The fraction dN/N is merely the ratio
of the solid angle corresponding to the angular interval dé to the total solid
angle 47:
dN 2xrsin@dé _ sinédé
(3-90)
Nite ain ee? ae
If the dipoles are subjected to a local field, the solid-angle elements are no
longer equally probable, since a dipole has a potential energy W, which de-
pends on its orientation relative to the field. It is shown in statistical me-
chanics that if a large number of molecules are in statistical equilibrium, the
number possessing a particular energy W is proportional to exp (—W/kT),
where k is the Boltzmann constant 1.381 X 10-* joule/kelvin and T is the
absolute temperature in kelvins.
In the present case the dipoles whose axes lie in the range between @ and
6 + dé, measured from the direction of the local field, all possess an energy
W = —p- Ej. = —PpE to. cos 8. Then the number of dipoles per unit volume
lying in the interval 6 to 6 + dé is
(3-92)
2! PEt.
u= kT ’
and where the constant C must be chosen so that the total number of mole-
cules in a unit volume is NV:
na ~~ 1 rs. DENce % il
= Np (coth u ‘)= Np (cotn kT =) (3-98)
SS Pe geen :
Pa Nola eae ~ul Coy
Np
ANE
ee
ogy
OD)
ie
(1= =)
6
= u|
Al (3-100)
~py Npu
OE _ Np
TE Boe :
(3-101)
Thus, when pE\o. < kT, the polarization P in a polar dielectric is proportional
to the local field, and the dielectric is linear. We found earlier in Eq. 3-36 that
the same applies to nonpolar dielectrics. When pE\oc < kT the susceptibility
of polar dielectric is inversely proportional to the temperature.
From a practical point of view, the only feature that distinguishes a polar
from a nonpolar dielectric is the temperature dependence of the susceptibility
and, therefore, of the relative permittivity.
120
ea em a a (a Pe a Ee ee ol
0 1 2 3 4 5 6 7 8 9 10 11 1:
yx PE toc
kT
Figure 3-12. The Langevin function.
=
Pea («a LB
ia) Ejiocs :
(3-102)
where
ieee JED
(3-103)
as in Eq. 3-32 or, for a linear dielectric (Eqs. 3-38 and 3-53),
at he
peas oe (0T Fz) Golly
Multiplying by the molecular weight M and dividing by the mass density
p, aS we did in discussing the Clausius-Mossotti equation, we find a new
expression for the molar polarization that is valid for polar dielectrics:
_ iM geal NG P
OM a hace (0+ i) (3-107)
This is known as the Debye equation. It is similar to Eq. 3-88, except for
the term p?/3kT, which comes from the alignment of the polar molecules.
In principle, the Debye equation can be used to determine both the
121
NANG 1
(P| [Ge ae 2
For a given magnitude of E, both the magnitude and the phase of p are
functions of the frequency, first because of the various polarization processes
that come into play as the frequency changes, and also because of the exist-
ence of resonances. Thus, since e, is a function of the frequency, e, is strictly
definable only for a pure sine wave.
In many substances the relative permittivity decreases by a large factor
as the temperature is lowered through the freezing point.
Pee egEy J
(3-109)
where the subscripts i and j represent the three coordinate directions Kaa es
All three components of P depend on all three components of E, with different
susceptibilities for each. The susceptibility x, thus has nine components and is
a tensor. Actually, there are only six independent components, and, if the
coordinate axes are properly chosen, these six components reduce to three.
The relationship between P and Eis still linear but is more complicated than
for isotropic dielectrics.
For anisotropic dielectrics, Eq. 3-44 still applies:
where e¢,;; is a tensor which, again, has six independent components. The re-
lationship between D and E is still linear, and the general features of the
discussion of Class A dielectrics are valid.
If the relative permittivity is a function of position in the dielectric, then
a volume density of bound charge may exist when there is no corresponding
volume density of free charge (see Eq. 3-60). Demonstration of this fact will
be left as an exercise.
W= 5| Var (3-113)
124 ELECTROSTATIC FIELDS II
where the volume r is any volume which includes all the free charges in the
system. Now p; = V-D, and, if we use the vector identity of Problem 1-14,
W= 5 ome, (3-117)
where 7 is any volume that includes all the points where D and E differ from
zero. This expression is independent of the type of dielectric present in the
system and is completely general.
As for free space, we may define an energy density
dw 1
aaa 7 D-E, (3-118)
Similarly,
pe
= OE, dE, dE,
(De )
dz
Then
F= (p:V)E. (3-127)
Ea ee) (3-130)
oy Ox
a (3-131)
OZ Ox
aC =) VP=2 : ly ( E) (3-133)
OE, OF,
e ~ |e >|OB,
ar (3-135)
a Ere Ew dp = a eee
In R. (3-136)
and
r V
(3-137)
i= 27rep 2 p In (R2/Ri)
You will be able to show in Problem 3-19 that this force can be
larger than the gravitational force by a few orders of magnitude.
We found in the example in Section 2.16 that the force between the
plates of an air-insulated parallel-plate capacitor is (o?/2¢€)S, where
a is the surface charge density and S is the area of one plate. We can
perform a similar calculation for a pair of plates immersed in a liquid
dielectric. In this case, if the spacing s between the plates decreases
by ds, the work done against the mechanical forces F,, holding the
plates apart is again F,, ds and
Fn Soy
= 5 BS =
= 6,©0
5B'S, (3-140)
is
and, for a given electric field strength E, or for a given voltage differ-
ence between the plates, the electric force is €, times /arger than in air.
We also have that
D: ae QO? 1 Q?
lal = eli — = -
2e 2 2e - 26S &, Zen (3-141)
and, for given charges +-Q and —Q on the plates, the electric force
1s ¢, times smaller than in air.
129
3.14 SUMMARY
Jo = oeamperes/meter’?. (3-21)
The Jocal electric field intensity Eo. is the space-and-time average of the
electric field intensity acting on a particular molecule:
P
Ew. = E+ 6 = (3-33)
0
Vika (3-41)
130 ELECTROSTATIC FIELDS II
where the total charge density p, is the free charge density p, plus the bound
s of
charge density py. This is one of Maxwell’s four fundamental equation
electromagnetism.
The electric displacement
D = wE + P, (3-45)
= e(1+ x.)E = eek = cb, (3-52)
a al LP (3-46)
or
P=N; P E kT
(cothoes
kT ‘ 7) : 3
(3-98)
PROBLEMS
131
where p is the permanent dipole moment of the molecule; and for real
polar
dielectrics, in which there are both induced and permanent dipoles, we have
the Debye equation:
mn M eo 1 = Na Pp
oe (ea) = 3 (2+): CR
In the presence of dielectrics the potential energy of a charge distribution
is given by an integral similar to that for free space:
1
W= 5 | War (3-113)
l
= [on dr. (3-117)
dw FG1
7 (D-E), (3-118)
T=pXE, (3-121)
while the force is
F =(p-V)E. 127)
The force per unit volume is (P-V)E, or (4)(€ — €)VE°.
Charged conductors are subjected to electric forces that depend on the
nature of the dielectric in which they are immersed. In a liquid dielectric the
dielectric decreases the forces by a factor of ¢, if the charges are kept constant,
and it increases the forces by a factor of e, if the voltages are kept constant.
PROBLEMS
3-1. A sample of diamond has a density of 3.5 grams/centimeter® and a polariza-
tion of 10-7 coulomb/meter?.
(a) Compute the average dipole moment per atom.
(b) Find the average separation between centers of positive and negative
charge. Carbon has a nucleus with a charge +6e, surrounded by 6 electrons.
3-2. Show that the bound charge density at the interface between two dielectrics
1 and 2 that is crossed by an electric field is (P; — P2)-n. The polarization in 1
is P, and is directed into the interface, while the polarization in 2 is P, and
ELECTROSTATIC FIELDS II
132
points away from the interface. The unit vector n is normal to the interface
and points in the direction from 1 to 2.
3-3. A large block of dielectric contains small cavities of various shapes that may
be assumed not to disturb appreciably the polarization.
(a) Show that, inside a needle-like cavity parallel to P, E is the same as in the
dielectric. ;
(b) Show that, inside a thin crack perpendicular to P, E is e, times larger than
in the dielectric. .
(c) Show that, at the center of a small spherical cavity, E is P/3¢0.
. The space between the plates of a parallel-plate capacitor with plate separation
s and surface area 5S is partially filled with a dielectric plate of area S and of
thickness t < s.
Show that the capacitance is
€oS
s — [(e, — 1)t/e,]
3-5. A dielectric sphere of radius R contains a uniform density of free charge py.
Show that the potential at the center is
26.+1 p,R*
2€, 3€,
3-6. Draw graphs of D, E, V as functions of r for a point charge at the center of a
dielectric sphere as in the second example on page 111. Set Q = 1.00 X 1079
coulomb, R = 2.00 centimeters, «¢, = 3.00.
Show also the curves of D, E, V in the absence of the dielectric sphere.
(a) Calculate the volume and the surface density of bound charge.
(b) Calculate the volume density of free charge.
(c) Calculate the potential inside and outside the sphere.
(d) Sketch a curve of potential versus distance from r = 0 tor =~.
3-8. A conducting wire carrying a charge \ per unit length is embedded along the
axis of a circular cylinder of dielectric. The radius of the wire is a; the radius
of the cylinder is b.
(a) Show that the bound charge on the outer surface of the dielectric is equal
to the bound charge on the inner surface, except for sign.
(b) Show that the net charge along the axis is \/e, per unit length.
(c) Show that the volume density of bound charge is zero in the dielectric.
3-9, The relative permittivity of the dielectric between the plates of a parallel-plate
capacitor varies linearly from one plate to the other. If €,, and €,) are the val-
ues at the two plates, where €,) > €,1, and if the plate separation is s, show that
the capacitance per unit area is
€0 (€2 — €,1)
Sain (€,2/ €r1)
133
Back plate
Diaphragm
Metallized surface
—. Figure 3-15,
3-10. (a) If the space between two long, coaxial cylindrical conductors were filled
with a dielectric, how would the relative permittivity have to depend on the
distance p from the axis in order that the electric field intensity be independent
of p?
(b) What would be the volume density of bound charge?
“Eeie-d
Are
1 1 ib ff i ry
énl1 €r73 r2 \Ero €r1
3-12. An electret has the form of a thin circular sheet of radius R and thickness f¢,
permanently polarized in the direction parallel to its axis. The polarization P
is uniform throughout the volume of the disk.
Calculate E and D on the axis, both inside and outside the disk.
The surfaces of the film carry both free and bound charges, the free
charges being predominant. The net charge density has a time constant that
can be as large as a few thousand years at room temperature.
(a) Calculate the induced charge densities on the electrodes, under steady-
state conditions, for the case of a 0.006 millimeter thick Mylar (e, = 3.0) film
and a 0.013 millimeter air gap. Assume constant free charge densities of
10-4 coulomb/meter? on the film surfaces and neglect permanent bound
charges. Neglect edge effects.
(b) Find a differential equation for c; when the diaphragm oscillates at an
angular frequency w.
. The dipole moment of the H,O molecule is 6.2 X 10-%° coulomb-meter. Find
the maximum electric polarization of water vapor at a temperature of 100°C
and at a pressure of 760 millimeters of mercury.
3-16. (a) Show that a nonhomogeneous dielectric can have a volume density of
bound charge in the absence of a free charge density.
(b) Calculate py in this case.
3-17, (a) Show that the maximum energy stored per cubic meter in a parallel-p
late
capacitor is ea?/2, where a is the dielectric strength of the insulator (maximu
m
electric field intensity before breakdown),
(b) It is suggested that a small vehicle could be propelled by an
electric motor
fed by charged capacitors. Comment on this suggestion, assuming
that the
only problem is one of energy storage. A good dielectric to
use would be
Mylar which has a dielectric strength of 4000 volts per mil
(0.001 inch) and
a relative permittivity of 3.2.
PROBLEMS 135
3-18. A dipole of moment p is lined up with the z-axis at the origin of coordinates.
A second dipole of moment p is centered at the point (a, 0, a) and is pointed
toward the origin. Calculate the force on the second dipole.
3-19. (a) Calculate the force per cubic meter on the dielectric of a coaxial cable
whose inner conductor has a radius of 1 millimeter and whose outer conductor
has an inner radius of 5 millimeters. The dielectric has a relative permittivity
of 2.5. The outer conductor is gounded, and the inner conductor is maintained
at 25 kilovolts.
(b) Show that the electric force near the inner conductor is about 300 times
larger than the gravitational force if the dielectric has the density of water,
namely 10% kilograms/meter?.
where Vy is the voltage difference between the inner and outer electrodes, and
A =
(Gig De PU = a
fay):
Chaar)
(b) Under what conditions is A positive? Sketch a curve of A as a function
of p, for a given value of f.Can you explain qualitatively the shape of this
curve? '
(c) The drift velocity » is given by Stokes’s equation:
Fret = Omnar,
where 7 is the coefficient of viscosity of the fluid.
136
— D
| a4 ‘ Figure 3-16.
3-23. One author states that he can attain fields of 4 < 10° volts/centimeter over a
0.1 inch gap in purified nitrobenzene (¢, = 3.5) and that the resulting electric
force on the electrodes is then 36 pounds per square inch, or more than two
atmospheres. Is the force really that large?
3-24. Electrostatic clamps are used for holding work pieces while they are being
machined. They utilize an insulated conducting plate charged to several
thousand volts and covered with a thin insulating sheet. The work piece is
placed on the sheet and grounded.
One particular type operates at 3000 volts and is advertised as having a
holding power of 30 pounds per square inch. If the insulator is Mylar
(ce, = 3.2), what is its thickness?
Hints
3-21. (a) Assume that +Q and —Q are point charges, and calculate the force exerted
on them by the molecule.
(b) You will have to calculate an approximate value for Q.
3-22. (a) Remember that, if the plate separation is increased by ds, the extra field is in air.
3-24. See hint for 3-22(a).
CHAPTER |
Up to this point our discussion of electrostatic fields has been limited to rather
simple charge distributions; we shall now develop methods for calculating
more complex fields.
Except for Section 4.1, this chapter deals with the solution of the differ-
ential equations V?V= 0, V7V= — p,/e,, VE = Vp,/e,. The first is of course
Laplace’s equation, while the other two are Poisson’s equations for V and E.
Now these equations are valid not only for electrostatic fields but also for
several other classes of phenomena. For example, the flow of heat in a
medium of thermal conductivity K obeys Poisson’s equation V°T = —q/K,
where T is the temperature and q is the thermal energy generated per unit
volume and per unit time. This chapter therefore has broad applications.*
We shall first discuss the continuity of various quantities at the interface
between two different media, and then we shall prove the uniqueness theorem,
according to which there is only one physically possible electric field that can
satisfy both Poisson’s equation and a given set of boundary conditions. Then
we shall illustrate the method of images, and finally we shall discuss at con-
siderable length the solution of Laplace’s and Poisson’s equations in rec-
tangular and in spherical coordinates.
If you are interested in this type of calculation, you can work through
Appendix B, which describes a method that is widely used for calculating
two-dimensional electrostatic fields when the volume charge density p is zero.
ceSections 4.1 and 4.2 are fundamental, but the rest of this chapter can be omitted if
time is lacking; it is not required for what follows.
139
4.1.1. Potential
The potential must also be zero at infinity if the charge distributions are
of finite extent, and it must be constant throughout any conductor as long
as the electric charges are at rest.
arbitrarily close to it. The boundary carries a free surface charge density oy.
If the area Sis small, D and o; do not vary significantly over it, and according
to Gauss’s law the flux of D emerging from the flat cylinder is equal to the
charge enclosed:
At the boundary between two dielectric media the free surface charge
density o; is generally zero and then D, is continuous across the boundary.
On the other hand, if the boundary is between a conductor and a dielectric,
and if the electric field is constant, D = 0 in the conductor and D, = o; in
the dielectric, o; being the free charge density on the surface of the conductor.
Eyl = Ex = 0, (4-3)
or
Eu a Ey. (4-4)
4.1. CONTINUITY oF V, D,, E, AT INTERFACES
141
tan; é5 cc
The larger angle from the normal is in the medium with the larger relative
permittivity.
For the case of an electret, see Problem 4-2.
The electric field is therefore subjected to several boundary conditions
at an interface, but these conditions do not all have the same degree of gener-
ality. The condition that V be continuous is perfectly general. Equation 4-2
D; = D, — D,, (4-13)
and
V-D; = V-D, —V-D, = 0 (4-14)
at every point. On the surfaces of the conductors V3 = 0 since both V; and
V, reduce to the specified boundary values.
We now use the vector identity
where the surface integral on the left is evaluated over all the surfaces that
bound the volume r. Let us take this volume to be the volume external to the
conductors, extending to infinity in all directions.
The surface integral is then to be evaluated over the surfaces of the
conductors and over an imaginary sphere of infinite radius. Since the quantity
V; is zero on all these surfaces, this portion of the integral is zero. To evaluate
the integral over the sphere of infinite radius, we consider the integral over a
finite sphere and let its surface recede to infinity. Both V, and V; must fall off
as 1/r at sufficiently large distances, since all the charge in the system will
appear as a point charge from a distance large in comparison to the dimen-
sions of the charge system. Then V3, the difference between V, and V;, must
also fall as 1/r. Now D; must fall off as VV3, or as 1/r?. Since the area S over
which the integration is performed increases as 7’, the whole integral falls off
as 1/r and approaches zero at infinity. The left side of the equation is thus
zero.
The first term on the right is also zero, since V-D; = 0 at every point. We
are thus left with the second term on the right, which must be identically equal
to zero. Thus
i(D,-E;) dr = 0. (4-17)
the same on the surfaces of the conductors, they must be the same everywhere.
Therefore V, = Vi, and there is only one possible potential V.
We have therefore shown that the solution of the Poisson equation for
given boundary conditions is unique, as long as D-E is positive throughout
the dielectric material in the system.
4.3 IMAGES
The method of images involves the conversion of an electric field into another
equivalent field that is simpler to calculate. It is particularly useful for point
charges near conductors: it is possible in certain cases to replace the con-
ductors by one or more point charges in such a way that the conductor
surfaces are replaced by equipotential surfaces at the same potentials. Since
the boundary conditions are then conserved, the electric field thus found is
the correct one for the region outside the conductors.
10
(a) . (b)
Figure 4-4. (a) Point charge Q near a grounded conducting plane.
(b) The conducting plane has been replaced by the image charge —Q
to
calculate the field at P.
4.3 IMAGES
145
In the region to the left of the plane the two point charges
must there-
fore give the proper solution for the point and the plane.
The charge
— Q is said to be the image of the charge Q in the plane.
The potential V at a point P, whose coordinates are r and
0, as
in Figure 4-4b, is given by
4reV = 2 — £, (4-19)
where
r! = Vr? + 42? — 4rD cos 6.
(4-20)
The components of the electric field intensity at P are given by the
components of VV:
OV Q Q(r— 2D cos 6)
4reE, = —4reo SiC ya ? (4-21)
The lines of force and the equipotentials are shown in Figure 4-5.
The induced charge density o’ on the surface of the conducting
plane is readily found from the normal component of the electric
field intensity at the conductor, since
E, = 0'/€0. (4-23)
4
<r.XN
,
Faex
In this particular case the surface charge density is negative, and the
electric field intensity points to the right at the surface of the con-
ducting plate.
From Figure 4-6, r = r’ at all points on the plane and
,_ QD
CS a (4-25)
remove the conductor and try to find the position and magnitude of
an “image” charge Q’, as in Figure 4-7b, that will make the potential
zero on the spherical surface.
It is clear from the symmetry of the problem that if such a
charge exists it must lie on the line connecting Q and the center of
the sphere. We begin by making the potential zero at the points P;
and P:. Then
Q jaan.
(Dieaye bynes sy
Q
WET wi
oo
CED
:
(4-28)
g=-50,
peed, +4 = —-. (4-29)
-29
bee 16
Es
V4
(4-33)
where ave, Pees 2 £
= VD? + r+ 2Drcos 8, (4-34)
and, at r = a,
4a (D? 6)?"
If we integrate this density over the surface of the sphere, we
get the total induced charge:
charge, since the fields outside the sphere are identical in either case.
Figure 4-9 shows lines of force and equipotentials for a point
charge in the vicinity of a grounded conducting sphere.
uniform distribution calculated from QO and Q’, and where o”’ is the
uniform distribution calculated from (Q,; — Q’).
(a) = (b)
Figure 4-10. (a) Conducting sphere carrying a charge Q near a grounded
conducting plane. When Q is positive, the induced surface charge
density ao’ is negative. The plane is assumed to be infinite. (b) The field
outside the sphere and to the left of the conducting plane is calculated
by successive approximations by using the image charges Q1, Qo,...,
etc., and —Qi, —Qy,..., etc.
4.3. IMAGES 151
OQ» = rQ,,
r2
r3
Os = aS aye, ere an (4-39)
(1 =a ae i=)
a
Qs= —____—_—__" r3
ig,
Ce) (1- —) E Bis rT
~ al
and so on. The equipotentials for D = 3a are shown in Figure 4-11.
The total charge on the sphere is
= a
4treoa
’
(4-41)
and the capacitance between the sphere and the plane is
Ct DRONES Toit)
4rea(1 +r-+---). (4-42)
~V O1/4reoa
Distance Distance
from Center from Center
Charge of Sphere Charge of Sphere
Or 0
a a/2Dy OQ aA ep
we
Sa) 1 — (a/2D)*
a NEE EEE EEE
152
2.0
LES
1.0
0 0.10 0.20 0.30 0.40 0.50
-
This field is the same, both in magnitude and in direction, just inside
and just outside the dielectric surface, as in Figure 4-13.
From Gauss’s law, the normal component of the field of o, is
a,/2€. This field is directed away from the boundary if a, is positive
(QO negative), and into the boundary if a» is negative (Q positive).
Therefore, just inside the dielectric surface, the normal com-
ponent of the electric field intensity is
Eqs = QO D
eo (4-43)
This quantity is positive if E,,; points outward.
where the unit vector n is normal to the surface and points outward,
and
Q D ob
Oh = = Eq Grae.) Fe (s? + D3? a 5° | (4-45)
ie = han (4-46)
Oe 2(er Ge
The induced charge density o, and Q have opposite signs, as
expected.
At this stage we could calculate the electric potential and field
intensity at any point, either within the dielectric or in free space,
by using Coulomb’s law, integrating over the o, distribution, and
adding the contribution from the point charge Q. But this is not the
simplest way to deal with this field. We shall find instead a set of
image charges that will satisfy the boundary conditions.
To find these charges, we confine our attention to the boundary
and write down the normal components of the resultant electric
field intensity for a point just inside the dielectric, E,;, and just out-
side, En.. From Eqs. 4-43 and 4-46,
Eno
=
= es a ee
2€, OD
- € + , 4reo(s? + D3? (4-50)
It will be observed that the normal component of D is con-
tinuous across the boundary:
= ee
E>,
_(e mz 1)
(4-52)
Figure 4-14. (a) If the dielectric is replaced by the image charge
—(e, — 1)Q/(e, + 1), the field is unaffected outside the dielectric. (b) If
the dielectric is extended to both sides and 2e,Q/(e, + 1) is substituted for
Q, the field is unaffected inside the dielectric.
(= 1)
OS =e Q (4-53)
(e, + 1)
at the image position inside the dielectric and (b) a single charge
D\c Eng ~ Oe
Oo, Oa— —- é-+1 Q, (4-54)
the same medium 1 on both sides of the interface, the original charge
Q, and a charge
;
eae en
en a Erg
4-55
2 Sa eas g (
The methods that we have considered until now for the calculation of electric
fields are useful only in special cases. We must find more general methods
for solving Poisson’s equation,
4.4 SOLUTION OF LAPLACE’S EQUATION IN RECTANGULAR COORDINATES 157
V2 Vacate: Ge (4-57)
To begin with, we shall confine our attention to electric fields with zero
space charge density, and we shall deal with Laplace’s equation,
(4-58)
functions have a number of general properties, of which we shall use the
following one. If the functions V;, V2, V3, --- are solutions, then any linear
combination 4,V; + A,.V.+ A3;V3; + --- of these functions, where the A’s
are arbitrary constants, is also a solution. This can be readily demonstrated
by substitution into the original equation.
It is usually possible to find solutions of Laplace’s equation that will
satisfy required boundary conditions by separating the variables. For ex-
ample, in Cartesian coordinates, we can usually find a solution of the form
V = X@YO)ZO), (4-59)
where X(x), Y(), Z(z) are respectively functions only of x, y, z. We can then
fit boundary conditions by adding a series of such solutions multiplied by
suitable coefficients. The uniqueness theorem assures us that the solution
found in this way is correct.
We can find the form of the functions X(x), Y(yv), Z(z) by substituting
V = X(x)Y(y)Z(z) into Laplace’s equation. Then
(4-60)
where we have written total instead of partial derivatives, since each one of
the X, Y, Z functions is a function of a single variable. Dividing through by
XeZe
Now since the second and third terms are independent of x, and since the
three terms must add to zero at all points, the first term must also be independ-
ent of x. It is therefore constant in value and
Similarly,
ye ieee een: t
158 ELECTROSTATIC FIELDS III
ay,
qn (4-67)
We have substituted k2 for C; and —k? for GC. so as to eliminate
square roots in the solution. The choice between C; and C, as the
negative constant is immaterial; the boundary conditions will force
us to the same final solution in either case.
We solve Eq. 4-67 by setting
Y = Asinky + Bcos ky, (4-68)
where 4 and B are arbitrary constants. This can be easily verified
by substitution. Our value of V must satisfy the boundary conditions
vV=0 (y = 0,» = d), (4-69)
ax ni \?
and
X = Genrz/b + He-nrz/b, (4-75)
where G and H are arbitrary constants. We can again verify this
solution by substitution. The condition that V->0 as x > re-
quires that G = 0.
Altogether then,
V(0, y) = ay Coin
a (4-78)
n=1
y
iE Vy sin ans dy -[ 2BC,, sin sin oe dy. (4-79)
160
V/Vo
nt tg . sas
ieCG. ob 5 dy = jo. oe ae (4-81)
Thus the only term of the infinite series on the right-hand side
of Eq. 4-79 that differs from zero is the one for which n = p. A se-
quence of functions possessing this property is said to be orthogonal.
Combining Eqs. 4-80 and 4-81,
=>
[see
Bt AT
it wis odd, (4-82)
if n is even.
We can now write down the potential V at any point (x, y, z):
Vo
AV, ~. I ine eh
V(x, y, Zz) = (4-83)
T noe>Sess b
The successive terms in the series become progressively less impor-
tant, both because of the factor (1/n) and because of the exponential
function. The degree of approximation that is achieved at x = 0
with one, three, ten, and one hundred terms of the series is indicated
in Figure 4-17. At y = 6 the first term alone gives a good approxi-
mation. The equipotentials are shown in Figure 4-18.
161
Vr [sin
0
™b dy = (Ay + Br) e,
2
(4-88)
and
2 Vi
Piaey | if is odd,
(4-89)
0 if n is even.
We can find another relationship between A, and B, from the
boundary condition at x = a: from Eq. 4-86,
V. ——= »»
.
(deel?
—ntu
++ Byeot=l)
na
sin+. ANY
=. (4-90)
0 if n is even,
and, from Egs. 4-89 and 4-91,
4 (Vy — Voeral
A, = nw :Sad a ) (4-92)
g, = 2 (fake ) (4-93)
4e-nta/b V2 — Vie nralb
163
e shall restrict ourselves here to fields with axial symmetry, that is, to
fields where V is independent of the angle ¢. Then
164 ELECTROSTATIC FIELDS III
anol 1 ai vy 2 ri
ar (" ean Bg et cae)
As in Cartesian coordinates, we seek solutions in which the variables are
separated and set
Vir, 0) = R(r) 0), (4-96)
where R is a function of r only and © is a function of @ only. Substituting
V = RO into Eq. 4-95,
a (aR Ra zy 7 :
oe (" =) "sie a8 (sina) = oot)
and dividing through by RO,
1d (dR 1 d dO
Rae a) Pea ee Cy
We have written total instead of partial derivatives, since R and © are each
functions of a single variable.
Since the second term is independent of r, the first term must also be
independent of r. The first term must therefore be constant:
ied (ooh as
ale ele oe
and then
Le ep ade ;
© sin 6 dé (sing a) = = arn
since the sum of the two constants must equal zero.
Let us examine the R equation first. Multiplying both sides by R and
differentiating the term between parentheses, we obtain
? os tty kR = 0. (4-101)
The solution of this equation is of the form
R = Ar” +
B
(4-102)
prt
d
WB (sino
6 a)
—4 + nin + 1)sin@®= 0 (4-104)
4.5 |SOLUTION OF LAPLACE’S EQUATION IN SPHERICAL COORDINATES 165
b= COS 0. (4-105)
Now, for any function f(y),
df_ fie
dear ik ae _ae
5 = -(1—- eae off (4-106)
That is, Eq. 4-107 remains unchanged when the number n’ is substituted
for n. Hence
P_+1(cos 6) = P,{cos 6), (4-111)
fied by substitution in Eq. 4-94. Since we are looking for solutions of the form
indicated in Eqs. 4-112 or 4-113, it follows from the latter equation that
Po(cos 6) = 1. (4-115)
We use a prime on the symbol P because the polynomials that we derive here
differ from the Legendre polynomials by constant factors, as we shall see
later. They are nevertheless solutions of Eq. 4-107. Substituting © =
Pi(cos 0) = 1 and n = 0 into Eq. 4-107 does in fact solve it.
Having found P(cos 6), how can we find Pi(cos 6) and all the other
polynomials corresponding to integral values of the index n in Eq. 4-107?
We shall do this starting with V;, but first we must know that any partial
derivative of a solution of Laplace’s equation with respect to any of the
Cartesian coordinate variables is also a solution. This is easily demonstrated
by substituting 0V/dx in Laplace’s equation and remembering that the order
of differentiation in partial derivatives is immaterial.
Let us therefore find the negative partial derivative of V; with respect
LOW
aC
eis Gor
waa (4-116)
where
See (2 + y+ 22 = Sarcaee
OZ
(4-117)
OZ iP ;
We therefore have a new solution of Laplace’s equation:
V, = CF (4-118)
Comparing once again with Eq. 4-113, we see that
ee Cos 0 a) z (3 cos* #— 1)
V3 as (c ) - (c4) C A > (4-121)
Z jee Oz
n P,,(cos 0)
0 1
1 cos 8
2 3 cos? 6 — 4
3 5 cos’ @— 3 cos
4 85 cost @ — 1 cos? 0 + §
a 63 cos’ @ — 8° cos* 6 + 4,° cos 8
168 ELECTROSTATIC FIELDS III
i ae
1 rcosé f= COSiU,
2 4$r%(3 cos? @ — 1) 3r_*(3 cos? 6 — 1)
3 =4r%(5 cos? 8 — 3 cos @) 3r_*(5 cos* 6 — 3 cos A)
4 4r'(35 cos! 6 — 30 cos? 6 + 3) tr_°(35 cos! 6 — 30 cos? @ + 3)
5 4r*(63 cos> 6 — 70 cos? @ + 15 cos @) 4r (63 cos® @ — 70 cos? 6 + 15 cos 6)
V= 2d,int
CE COS 0) \ 22d,Brr-°*» P,(cos 8). (4-125)
Hey 0 if m # n,
/ : P,,(cos 0) P,(cos 6) d(cos 6) = ieeea
ifm=n. (4-126)
(2n + 1
This property of orthogonality of the Legendre polynomials is important in
evaluating the coefficients 4, and B,, of Eq. 4-125.
a
Figure 4-22. Lines of force (indicated by arrows) and equipotentials for
a conducting sphere in a uniform electric field. The lines of force are
normal at the surface of the sphere, and there is zero electric field
intensity inside. Observe that the field is hardly disturbed at distances
larger than one radius from the surface of the sphere. The origin is at the
center of the sphere and the polar axis used in the calculation points to
the right.
= 1
)Pm(cos 9) d(cos 8)
io 2,/af A,,a"P (COS
According to Eq. 4-126, the only nonvanishing terms are those for
for which n = m and
170 ELECTROSTATIC FIELDS III
+1
()) = Algae fi
eeP?(cos 8) d(cos 0) + B,a~ “TY /= P?(cos @) d(cos 6),
(4-131)
= Aaa (525)
2n+1
+ Barss (5 5)
2n+1
(4-132)
Thus
Ba A ae (4-133)
For r — © the potential V is — E,r cos @. All the terms involving
inverse powers of r go to zero, and
Inspection of this equation shows that the only term that is not zero
on the right-hand side is that for which n = 1. We can show this in a
formal manner by multiplying both sides by P,,(cos #) and integrat-
ing from cos 6 = —1 to cos @ = +1. By either method
Ae ee (4-135)
and all the other A,,’s are zero. Then all the B’s are also zero except
153
By = —A,a* = E,a’. (4-136)
Finally, at any point (r, 6),
3 3
VGx0) = —E,rcos6 + E, 4% sok
r
= —E, (1— 5) ros 6
(4-137)
and
OV 2a
E, = Gane E, (1= =) cos 6, (4-138)
10V i \e
Ee = Soe = —E, (1= “;)sin é. (4-139)
Dp = 4reE,a’ (4-141)
located at the center, the field outside the surface previously occu-
pied by the sphere remains unchanged. We shall examine the image
aspect of this field in a problem at the end of this chapter.
(b) We could also have determined the field quickly from Eq.
4-125 by a less formal method. We must have the term — Fr cos @ to
fit the condition at infinity. No other function with positive powers
171
P(r,0)
B cos6
V = —£E,rcos@ +
poe
(4-142)
We finally set B = E,a* to make V = Oatr = a. Our solution satis-
fies both Laplace’s equation and the boundary conditions; thus, it
is the correct solution, according to the uniqueness theorem.
(c) There is still another method of calculating this same field
that will add to our understanding and that will further illustrate
the use of Legendre polynomials. Consider Figure 4-23. As indicated
previously, the potential at any point (r, @) arises from two charge
distributions: (1) that which produces the electric field intensity
E, and which resides on electrodes situated far away, and (2) that
which is induced on the surface ofthe sphere. This latter distribution
is unknown, and we denote it by o(6’). We use a prime on @ to dis-
tinguish it from the polar angle for a point (r, @) outside the sphere.
At the general point (r, @) the total potential from these two sources
must be of the form shown in Eq. 4-125.
Now it is possible to compute from Coulomb’s law the potential
at a point P on the axis @ = 0 at a distance r = z from the center
of the sphere:
teo(0 ee 0! do" (4.143)
V = —£&,z+ ;
Atréo
172 ELECTROSTATIC FIELDS III
where a is the radius of the sphere and s is the distance from a point
on the sphere to P as in Figure 4-23:
SS ee 9
s=V24
@ — 2azcos 6’ = Zot + 5 cos 6. (4-144)
='+4
+=© P,(cos 6’) ae~Ps(cos Oy) + -+., (4-146)
as long as z > a.
We have already seen that any continuous function of the polar
angle @ can be expanded as a series of Legendre polynomials. Thus
a(6’) = by + b:Pi(cos 6’) + boP2(cos 6’) + -+-, (4-147)
where do, b;, --- are constants. Equation 4-143 then becomes
2
Yai Baris 0
itie [bo + biP\(cos 6’) + beP2(cos 6’) + +++]
—
A, = —E,, (4-150)
and that
b,, a” +2
(4-151)
~In+1 ee
To evaluate the 5,’s, we use the fact that the potential is zero at
~ |= a. Substituting the above coefficients into Eq. 4-125 and setting
r="a:
= —E,aP,\(cos @) +78
wae—« Pileos 6) + 36,7(cos @) + -
(4-152)
which must be true for all @. Thus both the term boa/eo, which is
4.5 |SOLUTION OF LAPLACE’S EQUATION IN SPHERICAL COORDINATES 173
Writing V, for the potential outside the sphere and V; for that
inside,
where e, is the relative permittivity of the sphere. These are the sec-
ond and third boundary conditions discussed above. Therefore
B,P,(cos @) seB»P(cos
= 6) rae
—E,aP\(cos 0) + on =
By. (4-162)
a
By P
Be Coa?,+ +, (4-164)
2B
Ei ae = = —€,Or, (4-166)
3B,
4 = —2e,Cra. (4-167)
These sets of equations lead to the following values for the coeffi-
cients:
Bo —— C=SS 0, (4-168)
iy 1 |a(S 7— 5) seake3 :
(4-169)
175
Figure 4-24. The field near a dielectric sphere (e, = 3) ina uniform electric
field. The lines of D (indicated by arrows) crowd into the sphere as
shown, with the result that D is larger inside than outside. Since there is
no free charge at the surface of the sphere, the lines of D neither origi-
nate nor terminate there, and they are continuous across the boundary.
The equipotentials spread out inside, corresponding to a Jower electric
field intensity E. The electric field intensity E is discontinuous at the
surface, and the density of lines of force is lower inside than outside. As
in the conducting sphere, the field is hardly disturbed at distances larger
than one radius from the surface. The field inside is uniform. The origin
is chosen at the center of the sphere and the polar axis used in the
calculation points to the right.
BE
Cragg (4-170)
B,= C, =10 (n > 1). (4-171)
Thus
WAC -|1 a (2 =5 7 E,rcos0,
e- -1\a’
(4-172)
and
Vir, 0) = s 5)Eencoss—=
-(- mg = 5)E,z.
-(< 7 oz. (4-173
(4-173)
We have as yet dealt only with solutions of Laplace’s equation, since we have
concerned ourselves only with regions where the electric charge density p, = 0.
When p, ~ 0 we must find a solution of Poisson’s equation VV = —p;/é
(Eq. 3-42) which is consistent with the given boundary conditions. We have
already seen in Section 4.2 that such a solution is unique for a given p.
Oe
dx €
tsa (ce TO)
120
Figure 4-25. Curves for a plane p-n junction in silicon at 300 Kelvins. Density of donor and
acceptor atoms on either side No = Py = 107 meter~, thickness f = 200 microns, or
2 X 10-4 meter. Curves (a) and (b) show respectively the densities of donor and acceptor
atoms as functions of x. The x-axis is perpendicular to the junction and the origin is at
the center of the junction. Curve (c) shows the net density of fixed charges. Remember
that donor atoms give fixed positive charges and mobile electrons, while acceptor atoms
give fixed negative charges and mobile positive holes. Curve (d) shows the potential V as
a function of x. There exists a permanent contact potential between the two sides of the
junction. Note that V varies roughly like P — N. The potential is positive on the side
that has a positive net fixed charge (n-type material) and negative on the other side. This
is because thermal agitation forces the free electrons from the ” side to diffuse into the
p side, while it forces holes from the p side to diffuse in the other direction. The electric
field intensity E = —dV/dx is shown in the next curve, (e). Since it is negative, it points
away from the n-type material, which contains a majority of fixed positive charges. This
electric field limits the diffusion of electrons and holes across the junction. The net
charge density p, curve (f), is proportional to minus the second derivative of V. It has
the same sign as V and as the fixed charge density P — N. Finally, curves (g) and (h)
show the numbers ” and p of mobile negative and positive charges. Curve (f) for the
charge density p is related to curves (c), (g), (h):
p=e(P—N+p-—Rn).
Mads
(d) — 0.064
E ete
ee eS Se
—120 —100 —80 —60 —40 —20 D0 OO eS LOO ml e20)
Sol Microns
L
(e) =80
ae ! | |
ae 80 100 120
—60 —40 —20 0 20 40 60
—120 —100 —80
Microns
(h)
180 ELECTROSTATIC FIELDS III
We can find the Poisson equation for E as follows. In Section 2.3 we found
that the curl of E is zero. Now, from Eq. 1-143,
VXVXE= —-WE+ V(V-E) = 0, (4-181)
and. thus
VE = V(V -EB), (4-182)
(eeAre /
J7
Aad)
if
(4-184)
We have added a prime to the del to stress the fact that it contains derivatives
with respect to the source point (x’, y’, 2’).
This equation relates the electric field intensity to the gradient of the
total charge density. Note the negative sign and the first power of r in the
denominator. It is valid whatever the nature of the media that are present in
the field, as long as the gradient is definable.
The more usual expression for the electrostatic field intensity of a volume
distribution of charge is a consequence of Coulomb’s law and was found in
Eq. 3-30:
poe ea ae (4-185)
Although the two integrals for E are equal, the integrands are obviously un-
equal, since there exists no general relationship between p, and V’p, at a point
in space. Indeed, the two integrals are equal only if they extend over ail the
charge distribution.
It will be shown in Problem 4-30 that, as a result of the equality of these
two integrals,
4.8 SUMMARY
In this chapter we have dealt with electric fields that cannot be calculated
easily by the methods of Chapter 2.
We first established the conditions that must be satisfied by the potential
V, by the normal component D, of the electric displacement, and by the
tangential component £, of the electric field intensity at the boundary between
two media: all three must be continuous across any boundary that does not
carry a surface charge. If there is a surface charge, then D,, is discontinuous
but V and E£, are still continuous.
We then demonstrated the uniqueness theorem according to which, for a
given set of boundary conditions, there is only one possible electric field. This
theorem is of great practical importance: if we can somehow find a potential
V (x, y, Zz) that satisfies both the boundary conditions and Poisson’s equation,
then we know that we have found the correct potential.
The method of images can sometimes simplify the calculation of electric
fields that involve conducting surfaces and dielectrics. For example, the field
of a point charge Q near an infinite conducting plane is the same as if the
plane were replaced by a charge —Q at the position of the image of Q. The
force on Q is of course the same in both cases. The method of images gives
the correct field only outside the region where the image is situated. In some
cases there is an infinite set of images.
We then solved Laplace’s equation for a number of different boundary
conditions requiring either Cartesian or spherical coordinates. Solutions of
Laplace’s equation are called harmonic functions. We found such solutions by
separating the variables: in Cartesian coordinates we set
n=1
b
182 ELECTROSTATIC FIELDS Ili
are called Fourier series; an arbitrary boundary condition V(y) can be sat-
isfied by such a series. The value of the C,, coefficients is found by multiplying
both the boundary condition V(y) and the Fourier series by sin (pry/b),
integrating from 0 to 6, and utilizing the orthogonality property of the se-
quence of sin (nry/b) functions:
ie ii p # n,
(B C, sin nny
|= sin =7 dy of
ATTY b
= ]On5 if 4-8]
pep ar
The separation of variables in spherical coordinates transforms Laplace’s
equation into two ordinary differential equations:
Hed
Pe
Months
dr n(n + 1)R = 0, (4-101, 4-103)
where u = cos 6.
The first equation is readily solved by functions of the type
R(r)= Arm + —
ae (4-102)
P, (cos 0) = a
Jnl a(cos (cost) —
6)" (cos? 6 — 1) I -12
(4-124)
+1 2 Li 201. 9%. 0,
ie P,,(cos 0) P,,(cos 6) d(cos 0) = | (4-126)
if m =n,
183
Figure 4-26.
WE
_ Vos + ps) _ Vow (4-183)
€0 €0
(4-186)
PROBLEMS
in the atmosphere
4-1. It is found experimentally that the electric field intensity
ter and that it points
near the surface of the Earth is about 100 volts/me
184 ELECTROSTATIC FIELDS III
4-4, Various devices, such as Van de Graaff particle accelerators for example,
have high voltage electrodes maintained under pressure in a metal tank. Let
us assume that the electrode is spherical and that it has a radius r;. The elec-
trode must operate at a voltage V with respect to the tank, which has a radius
r, and which is grounded.
We focus our attention on the electric field intensity at the surface of the
electrode, for that is where it is highest, and we try to optimize r; and fro.
(a) If we disregard the cost of the tank, then its optimum radius will be that
for which the electric field intensity in the insulating gas will be minimum, for
this will permit using a minimum gas pressure. What is this optimum value
of ro?
(b) In actual practice, cost, weight, and space requirements limit ry. One must
therefore optimize the ratio r./r by varying ri. Show that the E at the surface
of the high-voltage electrode (r = ri) has a minimum value of 2V/r; when
i), = Dp,
(c) Can you explain qualitatively why there should be an optimum conditio
n?
(d) To determine whether the optimum condition is critical or
not, plot the
PROBLEMS 185
value of E at r = r; for an actual case where rz was 0.483 meter and for values
of r; ranging from 0.1 to 0.4 meter.
(e) What range of values of 7; can be tolerated if one can tolerate an electric
field intensity 10% higher than 2V/r,?
(f) Calculate 2V/r, for V = 5 X 10° volts?
4-5. Perform the same calculations as above for a cylindrical geometry. You
should find that, when the radius r. of the outer cylinder is fixed, the electric
field intensity at the surface of the inner cylinder has a minimum value of
V/r, when r; = r2/e, where e is the base of the natural logarithms.
4-6. According to the uniqueness theorem, the Poisson equation V?V = —p,/€o
can have only one solution if the potential V is determined at the boundaries
of the field.
Show that two solutions can differ at most by a constant if the normal
component of VV is determined everywhere at the boundaries.
4-7. When the space between the plates of a parallel-plate capacitor is filled with a
dielectric e, it has a capacitance of C farads. If the dielectric is replaced by a
material whose resistivity p is much smaller than that of the electrodes, the
resistance between the electrodes is R ohms.
(a) Show that RC = pe, neglecting edge effects.{
(b) Show that this result also applies to cylindrical and spherical capacitors.
(c) Show that it applies to any pair of electrodes submerged in a medium
whose resistivity p is much smaller than that of the electrodes.
You should be able to show that the field is unaffected by the conduc-
tivity, with the above restriction.
One important application of this fact is the electrolytic plotting tank,
which is used for plotting electric fields in two, and in some cases three,
dimensions.
4-8. Ekctrified dust particles are ejected into the atmosphere and form an elon-
Fr ~ gated cloud of approximately cylindrical form at an altitude of 40 meters in a
3 meter/second wind. The current feeding the cloud is 80 microamperes.
Calculate the resulting electric field intensity at the surface of the Earth
directly under the cloud. Assume the Earth to be flat and conducting.
4-10. (a) Show that the force of attraction or repulsion between a point charge g
and a conducting sphere of radius R carrying a charge Q 1s
q {Q+(R/D)q _- Rq .
Arr, D? DED ae
where D is the distance from g to the center of the sphere.
(b) Show that the force is attractive when q and Q are of the same sign, for
Opes Bie
7 ORY
4-11. In the last example in Section 4.3 we calculated the surface charge density
induced on the surface of a block of dielectric by a point charge Q at a dis-
tance D in front of it.
Show that this surface charge density and Q give the correct electric
potential V (a) at the foot of the perpendicular drawn from Q to the dielectric
surface, and (b) at a distance D behind the boundary.
4-12. Show that there exist solutions of Laplace’s equation that are of the form
X(x) + Y(y) + Z(@).
4-13. Calculate the potential at x = O for the electrodes of Figure 4-16, using the
first five terms of the Fourier series. Perform the calculation for y = 0.1 5,
0.2 6b, and so on.
4-14, Use the fact that 1/r is a solution of Laplace’s equation to show that
4-17. Two point charges +@Q and —Q are situated on a diameter of a conducting
sphere of radius a at distances D > a to the right and to the left, respectively.
(a) Show that the image charges constitute a dipole of moment 2a*Q/D?
at the center of the sphere.
(b) Now let D and Q approach infinity in such a way that Q/D® remains
constant. Superpose the fields of + Q and of the dipole to find the field outside
the sphere,
PROBLEMS
187
4-18. Show that a sum of terms of the form r” sin n@, where n is any positive or nega-
tive integer, is a solution of Laplace’s equation in two dimensions.
4-20. The axial electric field intensity E, on the axis of the accelerating tube in a
particular type of ion accelerator is given approximately by
FE, = En + kz’,
where z is measured from the center of the tube along its axis. The azimuthal
component £E, is zero.
(a) Show that the radial electric field intensity in the neighborhood of the
axis is —kzp, assuming that the charge density is zero.
(b) Draw a rough sketch of the lines of force in a plane that contains the axis.
(c) What is the maximum charge density that can be tolerated if the value
calculated above for the radial field is to be accurate within 5 % at the ends of
the tube? The accelerating tube is 1.00 meter long, E,9 is 7.5 X 10° volts/meter,
and k is 1.00 * 10° volts/meter’.
where f(r) is the magnitude of the radial force of attraction exerted on the
particle.
(a) Are the orbits stable when there is zero space charge?
(b) Now assume that the gas pressure is zero and that the space charge arises
solely from the orbiting electrons. Qualitatively, how will this space charge
affect the electric field intensity? Will it eventually lead to unstable orbits?
(c) Assume that, in the presence of gas, the radial electric field intensity is
constant. Are the electron orbits again stable?
It is possible to use essentially the same device for transporting charged
particles over considerable distances. In this case the particles are captured
into spiral orbits at one end of the tube and travel to the other end where they
are detected. The distance of travel is limited only by the quality of the vac-
uum. One particular guide that was used for extracting alpha particles from a
target inside a reactor had a length of 6.3 meters and a diameter of 7.5 centi-
meters. The central wire was maintained at —30 kilovolts.
4-22. In 1959 Lyttleton and Bondi* suggested that the expansion of the Universe
could be explained on the basis of Newtonian mechanics if matter carried a
net electric charge.
Imagine a spherical volume of astronomical size containing un-ionized
atomic hydrogen of uniform density », and assume that the proton charge
€» = —(i + y)e, where e is the electron charge.
(a) Show that, for y > 10718, the electrostatic repulsion becomes larger than
the gravitational attraction and the gas expands.
(b) Show that the force of repulsion on an atom is then proportional to its
distance R from the center and that, as a consequence, the radial velocity of
an atom at R is proportional to R. Assume that the density is maintained
constant by the continuous creation of matter in space.
(c) Show that the velocity » = R/T, where T is the time required for the radial
distance R of a given atom to increase by a factor of e. This time T can be
taken to be the age of the Universe.
(d) In the Millikan oil-drop experiment an electrically charged droplet of oil
is suspended in the electric field between two plane horizontal electrodes. It is
observed that the charge carried by the droplet changes by integral amounts
within an accuracy of about 1 part in 10°,
Imagine that the proton charge e, is equal to —(1 + y)e, where e is the
charge of the electron. Can you show that the Millikan oil-drop experiment
leads us to believe that y is less than about 10-17?
. Electrostatic precipitation is used extensively for the elimination of dust par-
ticles from industrial gases, for example, for eliminating fly-ash from the
smoke of coalfired electric power plants. This method was developed by
Cottrell at the beginning of the century. In this process the dust particles are
charged by the ions formed in a corona discharge, and they then drift in the
electric field to the electrodes, where they are deposited.
Op
V-J> + anes = e(G = R),
of
where n and p are, respectively, the number of electrons and the number
holes per unit volume, and where J, and J, are, respective ly, the electron
current density and the hole current density.
is a
The number of recombinations R per unit time and per unit volume
concentra tions n and p, while the number G depends
function of the carrier
on the temperature, on the illumination, etc.
and collected
4-26. In the vacuum diode, electrons are emitted from a hot filament
Show that, for an idealized diode having two plane parallel
at the anode.
electrodes, the electron current density is
J = 2.34 X 10-*V3/2/s? amperes/meter’,
190 ELECTROSTATIC FIELDS III
where Vy is the potential difference and s is the distance between cathode and
anode.
This is the Child-Langmuir law. It is valid only for the plane parallel
diode and for electrons emitted with zero velocity. More generally,
J = kV3”,
where K is a constant, for both ions and electrons and for any geometry, as
long as the current is space-charge limited, and as long as uw is negligible at the
source. ft
4-27. Let us see how current flows through an ionized gas. Two large parallel elec-
trodes of area S are separated by a distance s which is small enough to render
edge effects negligible. Pairs of ions are created throughout the gas between
the electrodes at the constant rate of m) pairs/meter? second. Let us set
nt and n~ to be the numbers of positive and negative ions per cubic
meter, respectively ;
vt and v-, the velocities of the positive and negative ions;
J+ and J- the currents due to the positive and negative ions;
e, the absolute value of the electric charge of an ion;
the x-axis to be perpendicular to the plates, in the direction of the electric
field, with the origin at the positive plate. With these conventions v*
is positive and v~ is negative.
Show that the total current density
J = entv* — en-v~ = enox + en(D — x) = emD,
and that the charge density
as x — D
Pa eno yt y-
4-28. The thrust produced by a rocket motor is equal to m’v, where m’ is the mass
of propellant ejected per unit time and v is the exhaust velocity. Rocket engi-
neers therefore strive to make m’ as small as possible by increasing v.
One way of achieving large values of » is to eject a beam of charged par-
ticles, as in Figure 4-27, which shows a schematic diagram of an ion motor.
The propellant is ionized in the ion source and is ejected as a positive ion
beam at a velocity corresponding to the accelerating voltage V. Electrons are
injected into the beam to prevent the rocket from charging up.
(a) The current / of positive ions in the beam is carried by particles of
mass m
and charge ne, where e is the electronic charge.
Show that the thrust is
F = [(2Vm/ne)}??.
191
(b) What is the value of F for a 0.1 ampere beam of protons accelerated to
50 kilovolts? Ion motors are used in outer space where small thrusts are re-
quired to correct either the attitude or the trajectory of satellites.
(c) Show qualitatively that the thrust is independent of the shape of the field
or of the presence of space charge in the acceleration region and that it is
always given by the above formula.
(d) If we call W the power JV spent in accelerating the beam, show that
Thus, for given values of W and m’, the thrust is independent of the
charge-to-mass ratio of the ions. Or, for a given W, F is inversely proportional
to v. The last expression shows that, for a given power expenditure W, it is
preferable to use heavy ions carrying a single charge (n = 1) and to use as low
an accelerating voltage V as possible.
(e) Sketch a graph of the voltage along the axis, inside and outside the motor,
(i) when the electron source is on, and (ii) some time after it has been turned
off. Assume that the body of the rocket is connected to electrode B.
(f) If the electron source is turned off, and if the beam current J is one ampere,
how long will it take the body of the rocket to attain a voltage equal to the
accelerating voltage if V is 50 kilovolts? Assume that the rocket is spherical
and that it has a radius of 1 meter. At that point the motor ceases to operate
because the ions follow the rocket.
4-29. A sphere of electric charge has a density p which is a function of the radius
as in Figure 4-28.
(a) Use Gauss’s law to show that, at a distance r > B from the center of the
sphere,
a
Po_ (8? + a°\(6 + a).
12e0r
192
(b) Show that one arrives at the same result by using Eq. 4-184.
Note that the result is unaffected as (8 — a) — 0.
(c) Does this value of E make sense when a = 6?
1 Put ot = = : [ee
Aire, Je 9 4rr€ i}
r;
Figure 4-28,
i7 Vv’ (£)
ir
dr' = 0.t
4-31. Show that
i)v’ x (2) ar’ =0
for any finite charge distribution.
Hints
4-15. Choose the z-axis along P and use Legendre polynomials to calculate V. You can
determine the values of all the coefficients from the properties of the field without
performing any integration.
. Show that this V obeys Laplace’s equation in cylindrical coordinates and that it
gives the correct boundary conditions.
Solve Poisson’s equation, setting the charge density p equal to J/u, where u is
the velocity of the electrons and (1/2)mu? = eV.
If your calculation involves (—J)'/2, disregard the minus sign and consider
simply the magnitude of the current.
Assume that the electrons are emitted with u = 0.
Set dV/dx = 0 at the cathode. This means that there is an unlimited supply of
electrons available at the cathode and that the current is limited by the electron
space charge, and not by the nature of the cathode.
RELATIVITY I
The Basic Concepts
We have now studied at quite some length the fields of stationary electric
charges. At this point we can either go on directly to the fields of moving
charges, or we can leave the subject of electromagnetism aside, for the mo-
ment, and study the basic concepts of relativity. The reason for this digression
is that magnetic fields result from relativistic transformations of electric fields.
The longer path is more interesting, as always, but it may not be the
better one. There is no doubt that relativity throws much light on the nature
of magnetic fields. It also leads directly to the fundamental equations of
electromagnetism. But Chapters 5 and 6 do not replace the more conventional
approach that begins with Chapter 7. Selecting one path or the other is a
matter of time and personal taste.
Therefore, if you wish to go on with electromagnetism without delay, you
can omit Chapters 5 and 6, and go directly to Chapter 7 without losing conti-
nuity.
The present chapter is devoted to the basic concepts of relativity, with
little reference to electrical phenomena except near the end. In the next
chapter we shall utilize these concepts, first to reveal the origin of magnetic
fields, and then to establish several fundamental relations that we shall later
rediscover without using relativity.
We shall keep our discussion of relativity as simple as possible. In fact,
the only mathematical requirement for this chapter is elementary differential
calculus and the vector analysis of Chapter 1.*
* For a more detailed introduction to the basic concepts of relativity, see Edwin F.
A. Wheeler, Spacetime Physics (W. H. Freeman and Company, San
Taylor and John
Francisco, 1966).
194
We could start directly with the Lorentz transformation, which forms the
basis of special relativity, but this transformation is so contrary to everyday
experience that we shall first demonstrate the inadequacy of the more obvious
Galilean transformation.
We have assumed that the two coordinate systems coincide at the time t = 0.
This set of equations constitutes a Galilean transformation.
Now, although the Galilean transformation is self-evident, and although
it is completely compatible with classical mechanics, it is not generally valid,
as we shall see, either for mechanical or for electromagnetic phenomena. We
shall have to use a more complex transformation that will reduce to Eqs. 5-1
for everyday mechanical phenomena.
Both classical mechanics and the Galilean transformation break down when
velocities approach the velocity of light. Many examples can be given.
* The film ‘‘The Ultimate Speed; an Exploration with High Energy Electrons” by
W. Bertozzi demonstrates this phenomenon clearly. It was produced by Educational
Development Center, Newton, Mass.
196
* See American Journal of Physics, 37, 342 (1963). See also the film “Time
Dilation;
an Experiment on Mu-mesons” by F. Friedman, D. Frisch, and J.
Smith, produced by
Educational Development Center, Newton, Mass.
197
eee ce OC
200
Then laser B should oscillate as if its length were yLz, and the beat
frequency should be
Tom fn | 1
Vy = Yep—
Pa =
2 Ley e Ty :
(5-6)
If the complete set-up is now rotated through 90°, the beat
frequency should change by
NOG (elas —3)-(& go ES) 7
Avy = 2 UN ey Lay’ Lay’? Lay J’ Sot
Bee 7 al 1 1 1K peg1 2)
> EtE JG Jee : =o
where L is the average length of the lasers, and where the term (U/c)?
is obtained by expanding 1/7 using the binomial theorem.
(2) 9
Thus
Avy, ; 0 2
The experiments of Trouton and Noble, and of Townes et al. had two points
in common. Both were designed to detect effects of the velocity of the Earth
through the ether, and both gave negative results. Many other experiments
that had the same objective also gave negative results.*
It is on the basis of such evidence that Einstein proposed in 1905 the
fundamental postulate of relativity, which can be stated as follows: it is
physically impossible to detect the uniform motion of a frame of reference from
observations made entirely within that frame.
This postulate is quite clear in itself, but it is so fundamental that we
shall state it in another way to emphasize its meaning. It means that any
experiment gives precisely the same result, whether it is performed in refer-
ence frame | or in reference frame 2, or whether it is performed in a standing
or in a moving vehicle, as long as it is not accelerated.
Then, if we are given some physical law like Fy = mq which is valid in 1,
there exists an identical law F, = mod, which is valid in frame 2. In fact, there
must exist a transformation, different from the Galilean transformation, that
renders all the laws of nature identical in frames 1 and 2.
We shall develop this concept of the invariance of a physical law in the
next section, using classical mechanics and the Galilean transformation as an
example.
The Galilean transformation leaves the laws, and hence the phenomena, of
classical mechanics unaltered in going from frame | to frame 2, and inversely.
In other words, the laws of classical mechanics are invariant under a Galilean
transformation. Classical, or prerelativistic, mechanics disregards the effects
that we shall discuss later, which become prominent when velocities become
comparable to the velocity of light.
Let us write down the equations for a simple experiment that the
passenger performs, first when the train is stopped, and then when
the train is moving at a constant velocity U. We shall see that the
equations of motion in reference frame 1 are of the same form as
those in frame 2.
The passenger is given a mass mm on which he exerts a known
force F in the direction of the track by means of a calibrated spring.
As previously, reference frame 1 is fixed with respect to the ground
and 2 is fixed with respect to the train. We assume that the law
F = ma applies in frame 1, and we deduce the corresponding law
in frame 2.
When the train is stopped we can assume that the two frames
overlap; then x; = x. and
@ d2
F=m aia ae a5 (5-10)
Famay
2
(5-12)
with respect to the ground, and
2
ea ape? (5-13)
fe kee UE Sy salts aa Vh
1 Tt = (erp T= (O/ory”
Yi = yo, yo =), (5-14)
41 = 2, 22 = Z1,
er oT= a acy
ae to +o (U/C?) Xe ee (Kh (0/ ce") ’
Vi
eeee SS
= Teen
205
y2
Xo 0 x5 = lo X4
t—=10) t,=0
(a)
Vi
SP paps |
x, =0 Xie lov |1
t; =0 t, =0
(b)
Figure 5-8. (a) A ruler is fixed parallel to the x-axis in reference frame 2 and, for
an observer in that frame, it extends from x2. = 0 to x. = 4, and hence it has
a length /. (b) At the time 4; = 0 observer 0, in reference frame 1 notes the
positions of the two ends of the same ruler. He finds that it extends from x; = 0
to x1 = //y, and hence that it is shorter than J). This is the Lorentz contraction.
Lengths parallel to the y and z axes are unaffected.
h = yo + Vtr), (5-15)
according to the Lorentz transformation. We must find fy.
Now we know that 4 = 0 and x2 = / for o2’s reading at the right-hand
end. Then
b= = (0/e yh, (5-16)
and
h = ybl1 — (0/eP] = hl — (0/c)]!? = h/y. (5-17)
Thus the ruler that is fixed in reference frame 2 and that has a proper
length /) according to observer 0, appears shorter by a factor of 1/y to observer
01, as in Figure 5-8b. In other words, a ruler moving in the direction of its
length at a velocity U relative to the observer appears to be shortened by the
factor 1/y = [1 — (U/c)*]!”. This is the Lorentz contraction. It is independent
of the sign of 0. Both observers agree on the correctness of this figure; their
disagreement bears on the validity of the measurements. Observer 0; maintains
that his measurement is valid because he has observed the positions of the two
ends simultaneously, namely, at ¢; = 0. But observer 0, maintains that, on the
contrary, 0;’s measurements were not simultaneous—that 0, first noted the
position of the right-hand end at the time 4, = — (U/c*)l and Jater noted the
position of the left-hand end.
Of course the Lorentz contraction would be precisely the same if the
ruler were fixed anywhere else on the x» axis.
If the ruler is fixed along the x-axis in reference frame 1, observer 0, finds
it shortened by the same factor 1/y, as in Figure 5-9,
Thus 0; tells 0, that o’s meters are too short, and Q tells 0; that o,’s
meters are also too short! This is not really contradictory,
because the two
comparisons are not really the same. They involve two different
pairs of
measurements,
This is a general rule: the proper length of an object, measure
d in a certain
direction, is always LONGER than the same length measur
ed in a frame of
reference moving in that particular direction.
207
What if the ruler moves relative to the observer in the direction perpen-
dicular to its length? It then appears to have its proper length hh.
if the cube were rotated through the angle 6 = arc sin (U/c). (If the
velocity U of the cube were equal to the velocity of light c, the photo-
graph would only show the left-hand face!)
Of course, the object would have the same appearance to an ob-
server as it does on the photograph, because the eye, like the camera,
registers photons that arrive together at a given instant.
It will be noticed that the appearance of the object is not exactly
the same as if it were rotated, because the face CDFE remains
normal to the line of sight.
0; x; 0, x}
Figure 5-14. The situation here is the inverse of that illustrated in Figure 5-13. The
phenomenon now occurs at the origin O, instead of O., observer 0; measures its
duration with a single clock, and 0 uses two clocks. The time interval measured
by the moving observer is again longer than 7) by the factor y.
point on 2 with a single clock, then o, finds that his own time intervals
are longer than those of 0 by the factor y = 1/[1 — (0/e)?]1/2,
Now, would 0, arrive at the same conclusion if he used a single
clock and /ooked at o,’s moving clock, as in Figure 5-15? We shall
assume that 02’s clock, which is situated at the origin O, of reference
frame 2, is moving away from 0, who stays at O;. This means that Oz
is to the right of O,, or that tf, is positive.
Let us first imagine that observer o, has a set of identical syn-
chronized clocks along his x-axis. As 0»’s clock goes by each one of
these, the relation 4; = yfy holds.
But 0; is at the origin O; of his system, and the light from o,’s
clock takes some time to reach 0. Thus 0, takes his readings later
than 4.
Suppose 0; reads a time fy on o»’s clock. What time is it on o,’s
clock? Let us call this time t{. Then ¢{ is the above t; plus the time
required for light to travel the distance Uf; between o»’s clock and
ors:
f=nt—=1+@/olr, (5-20)
BGO) :
“ft — @/on” ey)
—
1+ (v/e) 2
Fest to > fo. (5-22)
takes into account the time taken by the signal to arrive to him, then
he finds the previous result 4) = yh.
What if 02’s clock is moving toward 0,? Then the origin O» is to
the left of O,, and both 4 and fy are negative. Therefore the time f{
at O, when 0, reads f2 on 0»’s clock is
|ur| Uh
ti =e ty +- c th = é ’ (5-23)
T,
a(eueiays
= E a (v/c) ie > Ties (5-25)
fa rye
= 1/T, = ELee
+(0/e)a)" Torah (5-26)
l +- CU/c)
fa= Fess = is (5-27)
An interesting feature of these equations is that they are always
valid, whether it is the source or the detector that moves. This is
because there is no way ofknowing which one it is that moves, accord-
ing to the fundamental postulate of relativity. The relative velocity
0 is a positive quantity in all three equations.
213
If the relative velocity forms a right angle with the line joining
the source to the detector, we have either the first or the second
equation below, depending on the reference frame in which the
angle is measured:
5.9 SIMULTANEITY
Imagine that observer 0; sees two events A and B that, to him, occur at the
same x-coordinate x4; = Xg1, and, at the same time, t4; = fs:. He maintains
that A and B are simultaneous. Are they also simultaneous for 02?
From the Lorentz transformation,
Then two events that occur at the same value of x and that are simultaneous
for one observer are also simultaneous for another observer. Note that the two
events need not occur at the same place, because the y’s and the z’s can be
different.
What if the events do not occur at the same value of x? If they are
simultaneous for 01, then t4; = fz, but since x4; # Xm, they are not simul-
taneous for the other observer 0».
214 RELATIVITY I
In fact, the time interval between the two events, as seen from reference
frame 2 is
tre — tar = YW(U/c? X41 — Xz), (5-30)
and it can be either positive or negative, depending on the sign of x41 — Xz.
We have just seen that two events that are simultaneous in frame | are not
simultaneous in frame 2, unless they occur at the same x. We can go even
further: the order in which two events occur can be different in different
frames because
tee — tao = y[(tar — tar) — (0/C?xn1 — Xa1)], 6-31)
and the signs of tz. — t42 and tp; — tai can be different.
This is a disturbing result indeed: it appears to violate the principle of
causality, according to which a cause necessarily occurs before its effect.
For example, imagine that observer 0; throws a ball in the direction of
the x-axis, and the ball, after a flight of a few seconds, breaks a windowpane.
The Lorentz transformation surely cannot imply that the series of events,
starting with the throwing ofthe ball and ending with the strewn broken glass,
would occur backwards in time for certain observers.
Let us imagine two events A and B. Event A occurs at the origins O, and
O, at the moment they coincide. Event A is the cause of B, which occurs at
Xp at a Jater time fs), and in frame 2, at xp. Event B cannot occur before A
in frame 2. Then fs: must not be negative. Event A could be the throwing of
the ball, and event B the breaking of the glass. Event A causes event B
through some device, in this case it is the ball, which propagates in some way
the signal from A at a velocity 2, in frame 1. Then
nv < c. (5-34)
Let us assume that », and U are both positive. We have already seen in Section
5-6 that ‘0 can be as large as, but never larger than c. So the above inequality
must be satisfied even when U = c. Then
5.11 TRANSFORMATION OF A VELOCITY 215
m1 SC. (5-35)
If 2, and © are both negative, we arrive at the same result, by symmetry.
In other words, a signal can never be propagated at a velocity larger
than the velocity of light. Otherwise, the principle of causality would be
violated, and certain observers would perceive some effects before their causes.
If 02 observes that an object has some velocity 2», in the direction of the x-axis,
what is the velocity of this same object according to observer 0;? (The velocity
V2; need not be constant.)
The velocity 14,, by definition, is dx,/dt, and
dx dx; dt
V2 = dh = diy dt; (5-36)
where
dx;
pra= VS
dts
Pres + Ute)= ¥ (22 + V). (5-37)
Then
Sige (0/2)x] = ¥ [1 — (0/c?) m2],
2 = 177 (5-39)
and
Ue = V(vx + V)[1L — (0/c?)rr2]. (5-40)
Putting both 2, terms on the left-hand side and simplifying,
Vor ave : (5-41)
U2 =
1 + (v2 0/c’)
Voz as eee,
i Lan (diz U/c*) (5-42)
216 RELATIVITY I
Example This result makes the velocity of light equal to c in both systems.
For example, if the moving object is a photon in frame 2, ve, is equal
to c, and 2, is also equal to c, for any value of U! Or, if 0 = c, then
Viz = c for any value of v2;.
What if the velocity is not along the x-axis? Let us find a relation between
_ nd i
dy»
(5-43)
hy ris a dt,
Paade
The relation between 2, and v2, will be of the same form, by symmetry.
We proceed as above:
=
dy dts
———— =
dt.
4, — = Vo == Ziel -4
Vly dty dt, Voy dt, Voy y[1 (V/c )r1 i (5 5)
Then
Viy
*y == ST — (0/0) .
aie
5-46
dy = oe (5-47)
afl + (@x20/2)]
Similarly,
Vo
UT Ne =
0/2
Ee
Bey
5-48
Vy = Voy : = Viy
= TF On0/OP
Qoz
pas)
Table 5-3 gives the six transformation equations.
Qrz
ar:=
y*[1 + (02,0/c*)]?
peerage Sey = ans
Ay ALL +E(sd/e))? + od
il A Voz0 }
ie =ALL + Cn0/e9 2
Zz =. Sa
a ES ‘ze a
cod
eS aaa a Zz
> Air
Qo
~ [1 — (@1e0/c?)]*
1 if 01,0 }
ey Al ae O/e) oe Sn
1 if 0120
ag, =
v= owen oe |
We have seen that relativity requires that lengths, times, velocities, and
accelerations be transformed according to rules that are different from those
218
—VW Vv =
Figure 5-17. This idealized experiment permits us to find a relation between mass
and velocity. Observer 0, slides two equal masses A and B with equal and opposite
velocities along the x-axis. They collide and rebound elastically. We conclude
from observer 02’s measurements that m = m/[1 — (v/c)?]*/*, where » is the
velocity of the mass with respect to the observer.
If we multiply the first of these two equations by VU and then add it to the
second, we find that
mg Ute
a OL: (5-55)
Observer 0; on reference frame 1 observes a moving object and finds that its
mass is 7. Observer 0, in reference frame 2 observes the same object, and he
finds a mass 7». What are the two relations between m, and my»?
‘For observer 0, the mass is
Mo
m
SS
re
SS
(v/cpp?
Se
(5-60 )
where 7m is the mass of the object when it is at rest with respect to the observer,
and 2 1s the velocity of the object with respect to reference frame 1. Thus
Mo
™ = To +b + yey Se
Now, from Eqs. 5-41, 5-47, and 5-48,
= moc! 1 bs a+e5t ob
Lae Se
(5-66)
ys
I moc? + 5mon" + ;Mo a ~- SURES (5-67)
The term (1/2)mpv? is the kinetic energy of classical mechanics. The term
5.16 THE FOUR-VECTOR r 221
mic” is an energy that is associated with the rest mass and that
is called the
rest energy.
The quantity mec? is the relativistic energy &, and the difference mc: —
MC?
is called the relativistic kinetic energy. The relativistic kinetic energy is equal
to (1/2)mv? for 0? « ce.
Se OmUP erhOUR-VEBOLOR Tr
Note that the word invariant is not synonymous with the word constant.
In fact, if the point a were fixed in frame 1 and b were fixed in 2, r,, would be
a function of the time. The distance r,,, is said to be invariant because it has
the same numerical value in both frames.
222 RELATIVITY I
Now you can easily show that, with the Lorentz transformation, ra, is
not an invariant. However, there does exist a corresponding quantity that is
invariant under a Lorentz transformation. Imagine that a flash of light is
emitted at O, at the moment when the two origins coincide. The light prop-
agates in all directions at the velocity c in both systems, and
2 2 2 2 2
aa = ams == Bla aN a ( (5-70)
1
The light arrives at (%4, yi, 21) at the time 4 and at the point (x, yo, Zz) at the
time f,. Thus, in this case,
p=mv (5-76)
1 hEaG oar
It is to be understood that v is the velocity of m with respect to the observer.
In reference frame 1,
Pi = m1, (5-77)
and it has the three components
Coordinates x y z t
Corresponding variable Dz Py Dz &/c?
Ve nt Ol ey erie
9
must be an invariant.
& hy tw fh
p , (5-94)
aes chee leet,
where / is Planck’s constant 6.626 X 10-34 joule second, v is the fre-
quency, c is as usual the velocity of light in a vacuum 3 X 108
meters/second, #4 = h/2r = 1.05 X 10-84 joule-second, w = 27»,
and X is the wavelength divided by 27.
225
At the very beginning of the next chapter we shall need to transform a force
from one reference frame to another. This is fairly simple now that we know
how to transform a momentum. We use the relation
_ dp
Lire (5-95)
Then
Ppl
xX, a
as tye
dt, Piz =
|
dtp [vl Pes To U(E2/¢ »)
2 \e
dt,
'
(5 97)
AD 2 vd 1
= —— 5-98
o dto He‘ Cc dto fs yl + (vor /c?)]’ ( )
d&
X + (0/c) oF
= S, (5-99)
1 + (02.0/c?)
since d&/dt, is the rate at which the energy & = mc” builds up under the
action of the force F,. Finally,
ae Reta ae
+ y[1 + W20/e?)]
vA 41
* ¥f1 + Q2x0/c?)]
Sy2 po
1 C2 7 M10 (lyAh,
ly4l lzH1
Y;
Woy ’
> yf = (@u0/c?)]
wz,
vA :
~ y[l — @n0/c?))
5.19 TRANSFORMATION OF
AN ELEMENT OF VOLUME
discuss charge invariance. After that, it will be a simple matter to deduce the
required transformations.
An element of volume at rest in reference frame 2 has a volume dro
for observer 02. What is its volume for 0,? Let us assume that the element of
volume is a small cube of volume /j in 2.
The element of volume is affected by the Lorentz contraction only in the
direction of the x-axis. Then, for 0;,
dr ==
aL — (Oye
Ln)
Gas ee
1 + (v2,0 /c?) 2
:
(5 107)
- dro .
(5-108)
Al + @x0/c?))]
This transformation is really the same as for a length parallel to the x-axis and
moving at a velocity v2, with respect to reference frame 2.
The inverse transformation is
dr
dr. =
¥(1 = @20/e)]
; (5-109)
where the element of volume now has a velocity 2, with respect to frame 1.
We have assumed that the velocity in reference frame 2 is along the
x-axis. What happens if it has components along the y- and z-axes? According
to the Lorentz transformation, y; = y2 and z, = Ze. Therefore distances
perpendicular to the x-axis are the same in both frames, and the y and z
components of velocity have no effect on the ratio dr;/dr2. Equations 5-108
and 5-109 are thus valid for any value of x, and of ws.
From the experience that we have gained to date with relativistic transfor-
mations, we might venture to guess that an electric charge of Qy) coulombs
that is stationary with respect to reference frame 2 would appear to carry
either Qyy or Qo/y coulombs for an observer in frame 1.
This is in fact wrong. Electric charge is invariant, and a charged body
carries the same electrical charge for all observers.
Possibly the most direct demonstration of the invariance of charge is the
fact that the charge-to-mass ratio e/m for a charged particle moving at a
velocity v is found experimentally to agree with the law
a ee (5-110)
e e
No
varies with the velocity, as in Section 5-13. This relation is found to apply in
particle accelerators up to the highest energies attained to date.
Another convincing demonstration of the invariance of charge is the
fact that a metal does not acquire an electric charge when it is heated or
cooled, despite the fact that the average kinetic energy of its conduction
electrons is much /ess affected than that of its atoms.* One might expect, at
first sight, that the extra charge would be negligible, since the change in
velocity is small. However, as we saw in Chapter 2, the electric charges in
matter are enormous, and it is only because their fields cancel perfectly at all
temperatures that ordinary matter remains macroscopically neutral.
Example Ten kilograms of copper contain about 102° atoms and 1.5 x 107
coulombs of conduction electrons. If the charges on the atoms could
be modified by only one part in 10! by heating or cooling, the total
charge in ten kilograms of copper would be 1.5 < 10-8 coulomb.
If the copper were spherical, it would have a radius of about 6.5
centimeters and, assuming that the extra charge migrated to the
surface, the copper would have a potential of about 2 kilovolts.
Such electrostatic effects have never been observed.
a=dn = Qdn
(5-111)
where dr; is the volume of the element of volume as seen by 0, and dr» is the
volume of the same element as seen by 0. Thus
* See for example, Charles Kittel, Introduction to Solid State Physics, Third Edition
(John Wiley, New York, 1966) p. 209 ff.
230 RELATIVITY I
pi = eee
T1
= pxy[1 + (220/c’)]. G-112)
The inverse relation is
p= ae
drs
= pry[1 — (uz0/c?)). (G-113)
The quantity pode, is the product of the electric charge density and of the
velocity of these charges in the direction parallel to the x-axis, as seen by
observer 05. This is analogous to the product pv, which we used in Section 1.5
for the mass flux in water. The product p., is the electric charge flowing per
second and per square meter at the point considered, in the direction of the
x-axis, or the electric current density /2, in amperes/meter? as observed by 0»:
hoe = poz. (5-114)
Similarly,
Siz = pie. (5-115)
Then our two equations for p; and p, can be rewritten as follows:
= pry[1 + (22,0/c?)]
/ 2
It +Vor@,0/2)’
+ 0) me
(5-119)
If the charges were fixed with respect to frame 2, Jo, would be zero and dis
would be Up.
We have therefore found transformation equations for the electric charge
density and for the electric current density parallel to the x-axis. To be com-
plete, we require corresponding equations for Jiy, Joys Jz, Jo. We need only
find a relation between Jj, and Jy,:
and, by symmetry,
Jig = Joa (5-124)
See Table 5-8.
5.21 TRANSFORMATION OF AN ELECTRIC CHARGE DENSITY 231
Coordinates ie y Zz t
Corresponding variable Af As ie p
Components of r 36 y Z jet
Components of J Ak: If, Ap jcp
0 0 ) 0
ay : OY2 Oye 7 ay
paoO22
OZ;
eee
OZ, OZ
0 0 0 0 0 0
2a Man Deeikpo tral
By analogy with the four-vector (x, y, Zz, jet), we expect to find that the com-
ponents of the four-dimensional operator (.) (occasionally called “‘quad’’),
which is the equivalent of the V operator, will be 0/dx, 0/dy, 0/dz, (1/jc)(0/dt).
This is in fact correct. The transformation equations for the partial derivatives
will indicate to us which are the proper components:
— = oe (5-128)
:~ - (5-129)
Coordinates Ss y Zz t
Corresponding operators 0/0x 0/dy 0/dz —(1/c?)(0/02)
Components of r x, y zz Jct
Components of [] 0/0x 0/dy 0/dz (1/jc)(0/0t)
or a divergence
_ 0a; e da, , | da;
(UratS sy 7 =U az oe aie (5-133)
The scalar product of (_] by itself is called the d’Alembertian and is written (_}?:
2—
0? 0? 0 =
16? =
ego ay Ton” Gar eae
Sy 1 2
Aas (5-135)
Also
faye eeClotVie (5-136)
Setting the d’Alembertian equal to zero gives the wave equation corre-
sponding to a phase velocity equal to the velocity of light c.
I V+ n (5-138)
This is the Jaw of conservation of charge. It says that, in any given frame of
reference, electric charge is neither created nor destroyed. The law of conser-
vation of charge can also be written as
()-J = 0. (5-141)
This law must not be confused with charge invariance. Charge invariance
means that the electric charge carried by an object is independent of the
velocity of the object with respect to the observer. In other words, that the
charge is the same in all frames of reference.
5.24 SUMMARY
We constantly refer to the two reference frames of Figure 5-1 (or 5-7), where
U is the velocity of reference frame 2 with respect to reference frame 1—in the
positive direction of the common x-axis. It is also the velocity of reference
frame | with respect to 2, in the negative direction of the common x-axis.
According to classical mechanics, it is possible to transform one set of
coordinates into the other by means of the obvious relations
Te oop ae
af af
(a) The velocity is along the line joining the source to the detector
ae Reson (5-25)
Nile (O/o)| 4" (5-26)
where fy is the frequency measured at the detector, and f, is the frequency
measured at the source. The velocity U is taken to be positive when the source
recedes from the observer.
(b) The velocity is perpendicular to the line joining the source to the
detector
Sa = rf, (source) leu Pe: tyf. (detector). (5-28)
236 RELATIVITY I
Vy = Poy ’ Oi = os
© YES @e0/c)] ~~ yl (eG /e*))
Voz Viz
Viz
ey =,
+ On0/A) , 3
=™ ~ afl — (z0/2)]
\2z ae f 9 .
ALT On0/2)]*
l ==; fi
\do — = Vo,2y U 7 1
OS La Gate 1 2 2 ea
a, =
1 V9, 0 |
Q, = in aarp ae =. PS
{ iG -- Vor VU ay J
: [1 == (dex /c?)]?
Qo. = Air
22 fl a (v420/c)]*”
cae 4 Ao
Vy O
rg |
uN
l a fa : aV120 .
Q2, = or
(20 / ec) \ s 5 Cc? — 4,0 ay ,|
y [1 =
Pa ee ee
l — w/o” 5-59
ee
where mp is the rest mass.
TRANSFORMATION OF A MASS.
RELATIVISTIC ENERGY.
RELATIVISTIC MOMENTUM.
p=mv= Mo
(5-76)
[I — @o/cp}?”
TRANSFORMATION OF A MOMENTUM AND OF A RELATIVISTIC ENERGY (SECTION
5.17).
Piz = Y[P2x + U(E2/c’)] Pr = [Pre — U(Ei/c*)],
Py = Pr Py = Py
Piz = Px P22 = Pu
1/2 = y[(S2/2) + (0/C)por] &o/c? = y[(Ei/c?) — (0/c)pis]
Moyes 2 yj oa
yl + (ox0/c’)) > yl = (@20/e)]
Zi a Zi Ai
~ fl + Cac0/2))
dr, =
dr»
(5-108)
y[1 + (e20/c*))"
dry =
dr
(5-109)
vL1 = (20/2)
pi = [2 + (0/C’\oe], D i)
ll yer — (0/c*) Az].
oO a 0 2 =9 0 0 Re
Ox, SW E (w/e) } cs v[& + (o/c
od Bi ie
dV1 dy.’ OVe oy,
GEE AG 6aaa
0%, OZ OZ» 7 Oz,
PROBLEMS
239
Components:
oleic
dx dy dz jc dt
Four-dimensional gradient:
Four-dimensional divergence:
anh = da x da y da z 1 da t x 25
Ox ap oy a Oz ie or O19)
Four-dimensional Laplacian, or d’Alerzbertian:
fe (v:Ea
COL
(5-136)
LAW OF CONSERVATION OF CHARGE. It is found experimentally that, in any
given frame of reference, electric charge is always conserved:
V-J=——%0ot (5-140)
or
O-J =0. (5-141)
PROBLEMS
Several of the following problems are adapted, with permission, from a book by
Edwin F. Taylor and John A. Wheeler.*
5-1. For what value of U does the value of y differ from unity by 1%?
5-2. Draw a graph of y as a function of U. Try to find “‘the’’ best way of conveying
the manner in which y depends on V.{
5-3. Two events occur at the same place in the laboratory and are separated in
time by 3 seconds.
(a) What is the spatial distance between these two events in a moving frame
with respect to which the events are separated in time by 5 seconds?
(b)What is the relative speed of the moving and laboratory frames?
5-4. Three men, A, O, and B, ride on a train moving at a velocity VU. A is in front,
O is in the middle, and Bis at the rear. A fourth man, O’, stands beside the
rails. At the moment O passes O’, light signals from A and B reach O and O’.
Both O and O’ are asked who emitted his light signal first. What do they
answer ?
. Observers 0; and 02 repeat the measurement of the length of the ruler (Section
5.7), but in a different way. Observer 0» fixes to the ends of his ruler a pair of
flash bulbs that can project the shadow of the edge of the ruler on a photo-
graphic plate in reference frame 1. Midway along the ruler, he sets up an elec-
tronic circuit that can send pulses in both directions to flash the bulbs simul-
taneously (according to him).
(a) What will be the distance between the edges of the ruler on the photo-
graphic plate?
(b) How can 0; account for this result?
5-6. A straight line passing through the origin O, of reference frame 2 forms an
angle a» with the x-axis.
(a) Calculate the value of a, as measured by an observer 0; on reference
frame 1.
(b) What is the value of a; when the velocity U of reference frame 2 with re-
spect to 1 approaches c?
5-7. A physicist is arrested for going through a red light. In court he pleads that
he approached at such a speed that the red light appeared green to him. The
judge, a graduate of a physics class, changes the charge to speeding and
fines the defendant one dollar for every kilometer per hour he exceeded the
speed limit of SO kilometers per hour. What is the fine? (Agreen & 5.3 X 1077
meter, Area & 6.5 X 1077 meter).
5-8. The radio galaxy 3C295 has a red shift of 46°%. The astronomers mean by
this that the observed wavelength is 1.46 times the wavelength of the same
radiation produced in the laboratory.
(a) Calculate the radial velocity of this galaxy.
(b) Some quasars have red shifts of 200°%. What is their radial velocity?
5-9. The twin, or clock, paradox can be illustrated as follows. On their twenty-first
birthday, Peter leaves his twin Paul behind on the Earth and goes off in a
straight line for seven years of his time at a speed of 0.96c. Peter then reverses
direction and returns at the same speed.
(a) What are the ages of Peter and of Paul at the moment of reunion?
(b) Peter and Paul, expecting this strange result, performed the following
experiment during Peter’s trip. They both observed a distant variable star
whose light alternates from dim to bright at a frequency f when observed from
the Earth. The variable star is ina direction perpendicular to Peter’s trajectory.
They of course both counted the same number of pulsations during the trip.
Use the expression for the Doppler shift to verify the difference in age between
Peter and Paul at the end of the trip.
5-10. It has been observed that some quasars exhibit light fluctuations with a period
of about one day.
Can you infer an upper limit for their size?
PROBLEMS 241
5-11. The Lorentz transformation implies that the relative velocity U of two frames
of reference cannot exceed the velocity of light c. We have also shown that a
mass and a signal cannot exceed the velocity of light. Discuss the following
cases.
(a) A very long straight rod, which is inclined at a small angle @ in relation to
a horizontal axis, moves downward at a velocity v.
What is the speed of the point of intersection of the lower edge of the
rod with the axis?
Can this speed be greater than c?
Can it be used to transmit a signal?
(b) The same rod is initially at rest with the point of intersection at the origin.
The rod is struck a downward blow at the origin with a hammer,
Can the motion of the point of intersection be used to transmit a signal
at a velocity greater than the velocity of light?
(c) A powerful laser is rotated rapidly about a vertical axis.
Can the azimuthal velocity of the beam exceed the velocity of light?
Can the beam be used to transmit a signal between two points at a veloc-
ity greater than c?
(d) The manufacturers of some oscilloscopes claim writing speeds in excess
of the speed of light.
Is this possible?
. We have found six formulas for calculating the velocity components in one
frame when the velocity components in the other frame are known. Show that
they can be written in the following vector form:
v= [v2 ==) == (vw, /y))/01 ak (vo: V/c?)].
. Light moves more slowly through a material medium than through a vacuum,
its phase velocity » being c/n, where v is the index of refraction of the medium.
If now the medium itself moves at a velocity U « c with respect to the labo-
ratory, show that the phase velocity of the light with respect to the laboratory
is approximately
1
¢ ++ 1b) (1ee +)
n n
. Aring of electrons rotates about its axis of symmetry and also moves parallel
to itself in the direction of this axis.
If v, is the axial velocity of the ring with respect to the laboratory, and
if vp is the azimuthal velocity of the electrons in the reference frame of the
ring, show that, with respect to the laboratory,
(a) 6? = 62 + 63 — 6:63,
(b) y = Ye.
a= ov
a2 ak Cc
5-16. Show that the ratio m14/m1pz of Eq. 5-55 is always positive.
Sell7/, Show that the two values of the ratio m4/m, given in Eqs. 5-55 and 5-56 are
equal.
5-22. A cosmic-ray particle, which may be assumed to be a proton, has been ob-
served to have an energy of 16 joules.
(a) Calculate its mass in micrograms.
(b) How long would this particle take to cross our galaxy (diameter 10° light
years), as measured by a clock moving with the proton? Express your answer
in seconds (1 year + 7 X 10’ seconds).
5-23. The Stanford Linear Accelerator is used to accelerate electrons up to energies
of 40 GeV (40 10° electron-volts).
(a) Calculate the mass of an electron that has the full energy. How does this
mass compare with that of a proton at rest (mass 1.7 X 107-27 kilogram). An
electron at rest has a mass of 9 X 10°54! kilogram.
(b) What is the length of the accelerator in the reference frame of an electron
that has the full energy? The length of the accelerator, as measured on the
ground, is 3000 meters.
(c) What time would be required for such an electron to go from one end
ofthe accelerator to the other, (i) in the laboratory frame, (ii) in the electron’s
frame of reference?
(d) What is the velocity ofthis electron, according to its own measurements?
5-24. It is shown by the example in Section 8.9 that the net outward force on an ion
PROBLEMS 243
that is at the periphery of an ion beam is smaller than the electrostatic force
of repulsion by a factor of 1/72, U being the velocity of the ion.
Show that, if the kinetic energy of the ion is equal to its rest energy, the
electrostatic force of repulsion is reduced by a factor of 4.
5-25. Visible light can be transformed into high-energy gamma radiation in the
following way.
Head-on electron-photon collisions are produced by reflecting a ray of
light from a laser backwards on a high-energy electron beam. Let us assume
that the photons have an energy hy of 2 electron-volts and that the electrons
have a kinetic energy of 6 X 10° electron-volts.
Let us disregard the electron recoil, for the moment. In the reference frame
of the electrons, the incident photons are Doppler-shifted to hv’ >> hv, and then
reflected forward at the same energy hv’. With respect to the laboratory, there is
a further shift to hv” >> hy’.
Calculate the final photon energy, using conservation of both momentum
and energy in the laboratory frame.
5-26. Since the thrust produced by a rocket motor is equal to the product m’v, where
m’ is the mass of propellant ejected per second and 2 is the exhaust velocity,
the ultimate rocket would transform its propellant into radiation and eject
photons backward at the velocity of light. The weight of the propellant would
then be minimum.
(a) Show that the power-to-thrust ratio W/F for a photon motor is c.
(b) Show that the thrust F is c(dM/dt), where M is the mass of the propellant
remaining at the time f.
Note that these two relations are independent of the frequency: the
source of radiation need not be monochromatic.
According to this last equation a photon rocket burning one gram of
matter per second would have a thrust of 30 tons. The difficulty is to trans-
form an appreciable fraction of the propellant mass into radiation, as the
following example will show.
(c) An ordinary flashlight has a capacity of about 2 ampere hours at about 2
volts. If a flashlight is switched on in outer space, show that its terminal veloc-
ity will be of the order of 10~4 meter/second.
(d) Show that, in the process, the flashlight loses about one part in 10' of its
original mass.
5-27. Let us investigate some of the conditions under which interstellar travel would
be possible.
(a) First, time should be contracted by, say, a factor of 10. Then y = 10.
Show that v/c must then be 0.995.
(b) Imagine a space ship equipped with a photon motor. The motor annihi-
lates the fuel and produces a beam of light directed backwards. This type of
motor consumes the least amount of fuel; see the preceding problem.
You can find the fractionf of the initial mass that remains after the ship
has attained a velocity of 99.5% c by writing out the equations for the conser-
vation of mass-energy and for the conservation of momentum, Take into
RELATIVITY I
account the mass-energy and the momentum of both the rocket and the
radiation.
You should find that f = 0.05.
The space ship must later be braked to a stop, and this requires 99.5%
of the remaining mass. At the end of the return trip we are left with a frac-
tion f4, or 6.25 & 10~°, of the initial mass.
(c) It has been suggested that, in principle, the rocket could collect and burn
‘interstellar matter which has a density of about 1 atom of hydrogen per cubic
centimeter.
Calculate the mass of gas collected during one year if the rocket sweeps
out a volume 1000 square meters in cross-section at the velocity of light.
Is the interstellar gas a useful source of fuel?
recoilless process much more strongly than it absorbs gamma rays of any
nearby energy. The excited nuclei thus formed re-emit the 14.4 KeV radiation
in random directions some time later. This process is called resonant scat-
tering.
(e) If now the source is moved toward the absorber at a velocity v, what must
be the value of » if the absorber is to see gamma rays shifted in frequency by
3 parts in 10%, corresponding to one resonance line width?
(f) What happens to the counting rate of a counter placed behind the absorber
when the source is moved (i) toward the absorber, (ii) away from the absorber
at the same speed?
(g) If a 14.4 KeV gamma ray, emitted without recoil by an Fe®’ nucleus,
travels 22.5 meters vertically upward, by what fraction will its energy be
reduced ?
(h) An Fe*’ absorber located at this height must move with what speed and
in what direction in order to scatter these gamma rays by recoilless processes ?
R. V. Pound and G. A. Rebka found that Av/y = (2.56 + 0.26) 10735
in this case.
to show that the three vectors F, v, a are coplanar, where a is the acceleration.
2a E 1 ue]
Pi I ey if oF Cc? P25
Viz V0 2
p2 ll ve [1aia ae i
5-33. Imagine that electric charge is not invariant and that Q = Qo[1 — (0/e)7i"2)
(Remember that charge is in fact invariant, according to all experiments per-
formed to date.) The charge Qp is that measured by an observer moving with
the charge, and Q is the charge for an observer moving at a velocity U with
respect to it.
(a) If the electrons in a given sample have an average energy of 100 electron-
volts, what percentage increase in their charge must we expect if their velocity
increases by 1%?
(b) Would this be enough to produce an observable effect?
5-34, Calculate v, 6, y for electrons in matter, assuming that they have kinetic
energies of 10 electon-volts.
in refer-
5-35. We have seen in Section 5-21 that the charge and current densities
246 RELATIVITY I
Hints
5-2. You might think of using either a semilog or a log-log plot. You might also think
of plotting several graphs, or of plotting y — 1, etc.
5-29. (a) Call the initial rest mass of the excited nucleus mz and its final rest mass 1.
Then, immediately after the emission of the gamma ray, the nucleus has a mass m7’
and a velocity uw. Eliminate both m’ and u. If there were no recoil, the gamma-ray
energy would be AE = (1) — mj)c2. Show that
hy = AE{1 — (AE/2myc?)}.
5-35. Remember that, in a conductor, we have fixed positive charges and mobile negative
electrons. Treat both sets of charges separately, and then calculate the net charge
density and the net current density in 1.
CHAPTER 6
RELATIVITY II
The Electric and Magnetic Fields
of Moving Electric Charges*
* This chapter is based on Chapter 5. If you have not studied Chapter 5 or its equivalent,
proceed to Chapter 7.
248 RELATIVITY II
Xvi
&
a Xve2 —
= QaQoxe
Arreo(x3 ic yaya? (6-2)
y — 4mevy(x3 + p25?”
Zu = 0. (6-4)
249
iG, me
= rekv¥QaQox
BE ae ype (6 a >)
ON
|) 6-10
The product 4:ec?, which appears in the last term, occurs in many
calculations and is defined to be 10’:
Aregc? = 10’. (6-11)
This, in effect, defines the coulomb as the unit of charge, because the units for
the other variables F, U, x, y are already defined from elementary mechanics.
In fact, the constant ep which appears in Coulomb’s law is best determined by
first measuring the velocity of light c and then using the above equation.
It is customary to write that
We now wish to obtain a more general expression for the field of a charge
moving at a constant velocity. To do this we shall use two charges as in
Figure 6-5: the force exerted on Q, will give us the field of Q, at the position
of Qs. The charge Q, is situated at O. and thus moves at the constant velocity
Vi with respect to the laboratory (reference frame 1). The charge Q, has some
unspecified velocity v,, also with respect to the laboratory, The velocity vp
need not be constant.
We calculate again the force exerted by Q. on Qz, first in reference frame
2, and then in reference frame 1.
252
(a) (b)
Figure 6-4. The two charges Q, and Q, again move at the same uniform
velocity U. In (a) the electric and magnetic forces are in opposite di-
rections. In (b) there is only an electric force.
X= Q.QsX2
Areor3
vu Q.Qrvy2
Arers
) Zug = 2226-16)
4rers
We again use the transformation of Table 5-7, except that now Uns = 0:
:
|f
Q.0s | Vo
Je
Yunbl = Arerey |\ Toe are (6-18)
Q.Q0> | Z2
ZnS F =
(6 19)
Areoray 1 f OTe |
Now we must transform the coordinates and the velocities on the right-hand
side. Let us start with the brackets:
253
Xe+
PGiVo2yVSah y DS Vpoz U0 A,
C2 + Vpe2V C2 + Vp00V
= ¥(% — Vb) + yUoyV0/c?)y1 + y(0m120/c?)z1, (6-20)
We have again omitted the subscript 1 for simplicity. We have made only one
assumption: we have assumed that the velocity © of Q, is constant. Now
re = yx — Of)? + y? + 2, (6-26)
ay ren Ul) aye 8), (6-27)
= (1 — B? sin? 6), (6-28)
254
This is the electric field intensity E and the magnetic induction B at the point
(x, y, z) and at the time ¢ due toa charge Q moving at a constant velocity Vi
and passing through the origin at the time t = 0.
We have arrived at this result by calculating the force on a second moving
255
Figure 6-7. Figure (a) shows the force F, exerted by Q, on Q,. Figure
(b) shows the new r; and the new @ which we use to calculate F,. It
is found that F, # —F,, contrary to what one would expect from ele-
mentary mechanics.
charge, first in a reference frame moving with Q, and then in the reference
frame of the laboratory.
What is the force exerted by Q; on Q,? According to elementary me-
chanics, F, should be equal to — F,. However, our experience warns us to be
cautious, so let us calculate Fy.
We can find F, from the value of F; given in Eq. 6-29. In this equation U
is the velocity v, of Q, and
0.0% 9 ,
(6-35)
y= Aregy?r2(1 — 6? sin? 6)3/2 {r: + (1/c?)uv X (va X rij.
=
B = 4/C, y= pe (6-36)
(a) (b)
Figure 6-8. (a) Typical line of E, and (b) typical lines of B for a charge
QO moving at a constant velocity U, as seen by a stationary observer.
The electric field is radial. The lines of B are circles centered on the
trajectory and perpendicular to it.
Thus
The first and second terms within the braces give, respectively, the
electric and magnetic components of Fy.
It is obvious that F, ~ —F,. The electric components are exactly in
opposite directions, but their magnitudes are not the same because of the 6’,
6’, y’ terms, and the magnetic components are completely different.
There is a difference between F, and F, that is worth noting. The
expression for F, is valid if, or v,, is constant. That for F,, on the contrary,
is valid if v, is constant.
The fact that F, ~ —F, is not peculiar to electric phenomena. It is a
purely relativistic effect that would be observed with any type of force.
Note that the electric field of Q is radial: the lines of force of E are straight
lines that converge on Q as in Figure 6-8. The fact that the electric field is
radial comes as a surprise. Imagine that you have fixed a light source to Q,
and that you move with Q,. You will not see Q, where it is at the moment
when you look at it, but where it was when the light that reaches your eye left
Qu. At the time ¢ you will see Q, where it was at the previous time t — (r/c), r
being the distance traveled by the light that arrives at the time ¢. You must
B=0.85 ae B=0.90
Figure 6-9. Lines of force for a charge Q moving along the diameter of an
imaginary sphere. The dots show where lines emerge from the sphere at the
instant when the charge is at its center. The density of the dots is a measure of
the electric field intensity. The total number of dots is the same in all six figures,
so as to satisfy Gauss’s law (Section 6.8). Note how the field shifts to the region
of 0 = 90° as the velocity increases. For U0 = c the field is all concentrated at
() = OOF,
2.0
0.4
! ll | | fae re ml
cara
WOE ee Ee
0° 30° 60° 90' 120° 150° 180°
Figure 6-10. The electric field intensity E of a moving point charge as a function
of the polar angle @ of Figure 6-6, for seven values of 8 = U/c. The observer is
stationary and sees the charge moving at the uniform velocity 0. For 8 = 0 the
field is isotropic. It is hardly disturbed at 8 = 0.25. As the velocity increases the
field increases near 6 = 90° and decreases both ahead of the charge (near @ = 0)
and behind it (near 6 = 180°). At extremely high velocities most of the electric
field is concentrated near 9 = 90°. These curves explain qualitatively the validity
of Gauss’s law for moving charges: as the velocity increases, the flux of E shifts
from the regions where 06 ~ 0 and ~ 180° to 6 = 90° and the total flux of E
remains constant. (Then why are the areas under the curves not equal?) Note that
the electric field is always symmetrical about 90°. This means that there is no
way of telling, from the shape of the field, whether the charge is moving to the
right or to the left. The vertical scale gives E divided by Q/4ze0r?.
therefore look, not in the direction of the charge, but some distance back; at a
given moment, the light rays originating from Q, are not radial lines converg-
ing on Quz.
But the electric field is radial—as if the information concerning the
position of the charge traveled at an infinite velocity! Actually, it is only when
the velocity of the charge is constant that its electric field is radial; if the
charge is accelerated there is a retardation effect, and the field is not radial.
We have assumed implicitly that the velocity of the charge has been equal to
6.2 FIELD OF A CHARGE MOVING AT A CONSTANT VELOCITY 259
The second term on the right in Eq. 6-30 gives the magnetic force on
Q,. The fact that this force is of the form Q,v, x B, has three important
consequences.
(a) The magnetic force can exist, in a given frame of reference, only if
the charge Q, has a velocity with respect to that particular frame.
(b) The magnetic force on Q, is independent of the component of v,
that is parallel to the B of Q,.
(c) The magnetic force is always perpendicular to the velocity v,, and it
does no work.
Then, if & is the relativistic energy mc? of a particle of mass m which
carries the charge Q,
re
dt
any | Peg) See (6-38)
It is only with an electric field that one can increase or decrease the kinetic
energy of a charged particle; a magnetic field can deflect a charged particle,
but cannot change its speed.
yQ
ER ace = , (6-39)
: 4rreor?
in the radial direction,
_ MoyYQV
Bee I Arr? (6-40)
ry 1.6 X10"
9.1 X 10" xX 9 x 108 +1=20X104% (6-42)
Then
Re peisz ip eee Liles
= 0.29 volt/meter,
Ome
(6-43)
6.3 TRANSFORMATION OF ELECTRIC AND MAGNETIC FIELDS 261
We have just seen that an electric field in one reference frame becomes both
an electric and a magnetic field in another reference frame. There exists a
general rule for transforming electric and magnetic fields, and we shall be able
to deduce it by again transforming the force on a moving charge.
A charge Q has a velocity v; in a region where there is an electric field
intensity E, and a magnetic induction By, all with respect to reference frame 1.
Then
F, = O(F, + vu X B)). (6-45)
To find FE; and B; in a reference frame moving at the velocity Ui with respect
to the first, we shall transform F, into F,. We should find that
F, —= O(E, + (2D) x B,), (6-46)
ae Y; — (6-51)
Ye = TT — (uix0/e)]
ie yl — (0/c?)|
AN i (6-52)
We must substitute into these equations the above values of X;, Yi, Zi, and
then set
262 RELATIVITY II
a
i Le
SY CE
a,
GyEy
Voz a V
6-53
hues)
ty apne
=T+eee(sd)
yee, 6-54
oe
eee ee (6-55)
y[1 + (v2°0/c?)]
as in Table 5-2.
We start with Y, because it is relatively simple to calculate:
Bi-S = BE =, (6-64)
and that
E,-B, —— E,- By. (6-65)
6.3 TRANSFORMATION OF ELECTRIC AND MAGNETIC FIELDS 263
Therefore both
B- — and E-B
are invariants.
Figure 6-12. Figure (a) shows a parallel-plate capacitor as seen in its own
reference frame. The electric field intensity inside it is E2. For a sta-
tionary observer who sees the capacitor moving to the right at a velocity
U, the field is that of Figure (b): the electric field intensity is y times
stronger and there is a magnetic induction y(‘O/c?)E2 in the direction of
the z-axis.
264 RELATIVITY II
B=VXA, (6-77)
where the quantity A is the vector potential at x, y, z, t. The vector potential
is expressed in webers/meter. We shall show that, when the field is due to a
single charge Q moving at a constant velocity U along the x-axis,
_ Ho yOUVI
(6-78)
4a [ye UD sy? oe
265
La wed!
Yo
- Ox
0 0
oy
0
dz
eye
OAM
az
0A
oy e
6-85
ee
enone
= Gn [yx 7x Ut)? =. y? -f- z2|3/2
as required.
266 RELATIVITY II
as in Figure 6-14.
We shall demonstrate this general result for the special case of a single
charge moving at a constant velocity.
B = 0.85
B = 0.90
Figure 6-15. Equipotential surfaces V = Constant for a point charge Q moving
either to the right or to the left. The equipotentials near Q are not shown because
they are too close together.
268 RELATIVITY II
The electric potential at the point (x, y, z) and at the time ¢ due to our
charge Q is
Oe ;
uot th (6-91)
y? + 27]?
ou 4rreg [yx — Ut)? +
It is again sometimes more convenient to write V in the following form:
= ! Q: . (
6-92 )
Are r(1 = Be sin2 6) 1/2
Note the analogy between the formulas for V and A. Figure 6-15 shows
equipotential surfaces for six values of 8.
We can prove the validity of Eq. 6-90 by substituting our expressions for
V and A:
eV. OV ee ey, 0A
Gi. Say ede ar Sak
E=
».
[y?(x =
7yQ
Ut)? + y? — Bene |ge1 Ge — UNE ; + yj‘ + 2
=OAr yx — Opi) (6-94)
_ _yQl(x — vii + yj + 2k) (6-95)
Srey — Oi? + y? + 2]
which is the electric field intensity E as given in Equation 6-75.
Vena mie
= Goes (6-96)
‘
Qf —1 B2 sin 6 cos 6 \
~ 4areg LL — B? sin? 2"! * 721 — B? sin? 972 91| ’
(6-100)
6.6 TRANSFORMATION OF THE ELECTROMAGNETIC POTENTIALS V AND A 269
‘YQ
VWV=- aL at 0 = 90°.é (6-101)
We can find the value of 04/dt at 6 = 90° from Eq. 6-79 with-
out performing any calculation. As the particle sweeps by the point
1 centimeter away from its trajectory, r is minimum at 6 = 90°. The
angle @ changes from 0 to z, but sin? 6 has the same value at corre-
sponding angles on either side of 6 = 7/2. Then, as a function of
the time, A is maximum at the instant considered when 6 = 7/2,
and our 04/0dt is zero. Then, for this particular case, E is simply
equal to VV, and this agrees with the result we obtained in the fore-
going example (page 266).
6.6 TRANSFORMATION OF
THE ELECTROMAGNETIC POTENTIALS
V AND A. THE FOUR-POTENTIAL A
is an invariant.
Let us first check whether the transformation equations of Table 6-2 for
A and V are consistent with the fact that Ai, = F2, as in Eq. 6-63. To trans-
form
we substitute the values of 0/dx, and 0/dt, from Eqs. 5-127 and 5-130, and we
substitute the values of V; and A ;, from Table 6-2:
Similarly,
— 9s = fo + core)
Oh 2 — SE — (w/e El, Ot» J
6-106)
Upon simplifying, we do find that
OV ene ‘
Eis = OXs Oty 7s Foz, (6 107)
as required.
We shall not show that the transformation equations for the other
components of E and B are satisfied, but you might wish to try one or two
more.
V2 = —* Vow (6-110)
Then
ub if Ik
1 dj@
B =— y0/c%)2) — Vosew esl
ans =
(6-112)
» OO @
= 7(0/c) Esk, (6-113)
again as in the example on page 263.
You can easily check that A? — (V2/c?) is the same in both
frames of reference, and that it is therefore invariant.
The integrals for the vector potential A and for the electric potential V are so
similar that one suspects the existence of some general relation linking them
together. Such a relation does exist and it is called the Lorentz condition:
V-E= %. (6-116)
€0
This was Eq. 3-41 and is called Gauss’s law. It is one of Maxwell’s equations.
We shall now show that it applies to the E of Eq. 6-31. Gauss’s law applies in
fact to all electric fields.
272
8 ar sina
Jo (1 — B? sin? 63/2 (6-118)
iE:am Arey?
oe OR le sin 6 dé
= Dey? Jo [2 00870 + Uy] (6-119)
Bae a dx
J—1 [82x? + (1/72) ]8 (6-120)
2evy?
==€0 (6-121)
Gauss’s law therefore applies to the E of point charges moving at
constant velocities.
= 1— 3yz 3yz
{r yap pas % (6-123)
where we have set
i= la Ut) oye? (6-124)
This is the second of Maxwell’s equations:
|V-B=0 | (6-125)
oB
VXE=—-5, ‘
(6-127)
0 Od
pip heEe eo2
f real | Bde atrp (6-128)
274 RELATIVITY II
where Sis any surface bounded by the curve C, and where ® is the magnetic
flux linking the curve C.
Let us verify Eq. 6-127 by substituting the values of E and B from
Eqs. 6-33 and 6-34:
VXE=|— — =} (6-129)
yO [dO Zz 0 yah =o
(VX Be = tre Lay 1 (6-130)
az 1
TON Oo ULC a 2 i
(V X E), = rele 4 3? ax [ 3/2) (6-131)
10 (es Ue x ar =e,
(6-132)
Arey | ai
wo NGL) oe Ve
<r Ate [ 52 (y tay (6-133)
while
OB wovQ 3y?0(x — V2)
(—zj + yk). (6-137)
Ot = 4r [ ee
We can see that Eq. 6-127 is satisfied since
VX B= wp @ ae a) (6-139)
where J, is the current density at the point P, where the electric field intensity
is E and the magnetic induction is B, or
OE :
f Bat = w | @ + € =) -da = Molt. c iy
C Ss ot
The meaning of the index m, for matter, will become apparent later in this
section. Let us disregard it for the moment. The current /; is the total current
linking the curve C.
We shall be able to deduce this equation from the fact that Gauss’s law
is invariant under a Lorentz transformation.
Charges that are stationary at P in reference frame 1, in which E and B
are measured, contribute nothing to V X B. Let us disregard them. Imagine
that the density of moving charges at P is p and that these charges move at a
velocity Ui. Then
J = pvi. (6-141)
€0 Y€0
or
F
aa F)
ee F) i aep 6-144
OX2 |oe a OV, Ey Ar OZ» E, Ep ( )
We can deduce an equation for the field in frame 1 by using the equations
of transformation of Table 5-10 for the partial derivatives, and those of
Table 6-1 for the components of E:
0 Vd 0 0
(2+35 ) E; bY 5, Bs Ub,
UB) ~(E. ++ UB,) UBy) == p/ven, , ((6-145 )
+ ¥ —(£,
V-E== €0
i)EdeAS €0
V-B=0 [x da = 0
S
OB 0 fees)
Vy, fra --3 | pda = ar
We have again omitted the subscripts 1. But Gauss’s law also applies in
frame 1, and V-E = p/e. Then, dividing by V,
Maxwell’s equations are grouped together in Table 6-3. You will be able to
show (Problem 6-18) that they are invariant under a Lorentz transformation.
These are the four fundamental equations of electromagnetism. We shall
have many occasions to discuss them, and especially to use them, throughout
the remaining chapters. We shall not therefore say more about them for the
moment.
6.12 MAXWELL’S EQUATIONS 277
Pp
ers (6-150)
27 e0p
where the coefficient of Q is the electric field intensity at a distance p,
from an infinite line charge of density \, coulombs/meter. Note that
p is the radial distance to Q and not a charge density.
The negative charges are stationary in frame 2 and, similarly,
Y= os
AnQ | (6-151)
Both the charge Q and the distance p, which is perpendicular to the
x-axis, are the same in frames 1 and 2. Then, from Table 5-7,
\n2O
¥, = 2rrenp
y[1 — @,0/c’)], (6-152)
where 0, is the x component of the velocity of Q.
Now the linear charge density inside the wire is simply the vol-
ume charge density multiplied by the cross-sectional area of the
wire, the latter quantity being the same in both frames. Thus ),,»
must transform like a volume charge density, as in Eq. 5-117:
\n2 — [An ry (0/c?)In], (6-153)
Y== Hol
oe Qd;. (6-157)
=,ee
a ?1. (6-158)
f,Beale h (6-160)
Now, by symmetry, B must be azimuthal as in Figure 6-17, from
what we know about the magnetic field of a single moving charge
(Figure 6-8). Then the line integral must be simply 2rp B and we
have again Eq. 6-158.
This magnetic field results from a relativistic transformation
of the electric field of the moving electrons. This is indeed surprising
because the drift velocity 0 of the electrons in a conductor is only
of the order of 10-4 meter/second, or one foot per hour, or 3 X 1078
times the velocity of light. Then y is equal to unity within one part
in 1025! Relativistic effects could hardly be expected at such small
velocities.
280 RELATIVITY II
(*:)
Uz
i (5xLOSEam) x 1 2~ 10-*!mad coulomb/meter.
/
(6-162)
2
Therefore the bracket in Eq. 6-155 is the sum of two terms, the
first one of which is about 102° times larger than the second. But the
first one is perfectly cancelled by \,, and we are left with only the
second.
Of course the force on the second wire is appreciable for the
simple reason that it also contains an enormous number of conduc-
tion electrons.
SG ee eas (6 167)
roe 61 = B sin2 ae ri,
i
Bp — bo A= BY), di sin 8
~ 4x PL — Bin? ye? P*
(6-168)
The electric field intensity is not zero! The electric fields of the
moving electrons and of the stationary electric charges do not can-
cel, although they are of different signs. This result is apparently in
contradiction with the previous section. As we shall see below, the
E for a short length of wire is positive for a certain range of 0, and
negative elsewhere; it turns out that, once integrated over an infinite
straight wire, the two contributions cancel perfectly. Integrating the
above B over an infinite straight wire also yields the same result as
in the previous section.
Let us deduce approximate values for E and B by setting 6? < 1.
The approximations will be extraordinarily good because, as we have
SCelly O72 LOn-o hus
282 RELATIVITY II
Thus E is directed outward for sin? @ < 2/3 and inward for larger
angles. The explanation is simple: the E of the positive charges
points outward and is the same in all directions, while the E of the
negative charges points inward and is slightly larger near 6 = 7/2,
slightly smaller near 0 = 0 and r.
Since J = \,(— UV)= ApV, where J is the current in the positive
direction of the x-axis, and since (1 /e9c”)= po,
bol al ee +o ) S
ER Aap? o(1 3 sin 6) ry. (6-171)
Similarly,
Ree fed —89(1 +3 8 sin? 6)sin 6
(6-172)
Ce caer (6-174)
ae Ho Tdl x Yr)
BeHl ee
dh} (6-175)
This is the Biot-Savart law. We shall return to it in the next chapter.
= bo Jadla X 11
= (Anv dh)Vy X eee ge (6-177)
i 4a
BY 17, ES ae
(6-178)
We shall use this result at the very beginning of the next chapter.
To find the force d?Fy, exerted by I,dl, on I,dl., we exchange
283
subscripts a and b, and change the sign of r;. Contrary to what one
would expect from elementary mechanics, the result is not equal to
—d?F x, unless dl, and dl, are parallel. This purely relativistic effect
was mentioned earlier in Sections 5.18 and 6.2.
6.13 SUMMARY
where E and B are respectively the electric field intensity and the magnetic
induction at the point occupied by the particle. This is called the Lorentz
force.
By investigating the force between two moving particles we were able to
show that the field of a charge Q moving at a constant velocity V is given by
Q
~ Aregy?r?(1 — B? sin? 6)3/2 oe
(6-31)
HoQVD sin 6
(6-32)
ae 4ry2r(1 — B? sin? 9)32 ©”
where r and @ are defined as in in Figure 6-2 and where 6 = V/c,
Teer
The permeability of free space is defined as
It follows that
oc CB? and E-B
are both invariant.
The magnetic induction B is related to the vector potential A as follows:
B= Yas (6-77)
where
ee. yOUI
4m [yx — Ot)? + yp? + 22}? (6-78)
_ Ho Ovi
4x r(1 — B? sin? 6)?” Said.
for a single charge Q moving at the constant velocity Vi.
Similarly, the electric field intensity E can be related to both the vector
potential A and the scalar potential V:
0A
i —V ¥ here ar (6-90)
-
6.13 SUMMARY 285
il @) a @ @ I @ 1 0?
Vv ag Tae ’ ’ > 2 bean ee ere
O ( jc =) (2 dy Oz jc >. Me c? Or
where
Vow v2 :
Arely(x —U1)? yy? 271 Oe
Q
~ 4rer(1 — B? sin? 6)!”
again for a single charge Q having a constant velocity Vi.
The three components of A and V are the components of the four-
potential
A = (A, jV/c), (6-102)
and hence A;, A,, A,, V/c? transform like x, y, z, ¢ (Section 6.6):
is an invariant.
The Lorentz condition provides us with a relation between V and 4:
or
(a2 An—=0, (6-115)
SE aa 2 Vie Ea ee,or’
OE
V-B=0, VX B= wo(Jn + 050):
B= ol $1. (6-158)
2p
This result constituted a striking demonstration of the fact that the magnetic
fields associated with conduction currents result from an exceedingly small
relativistic correction to the e/ectric field of the moving electrons.
For an element of wire dl carrying a current /,
_ wo ldl Xr
ior fr?
(6-174)
The magnetic force exerted by an element of wire dl,, carrying a current J,, on
an element of wire dl,, carrying a current J;, is
PF, ae Ko
ie fees dl, x
=
(dl, x ry)
(6-178)
The element / dl also has an electric field, but the resulting electric force
is either zero or negligibly small.
PROBLEMS
6-1. Show that the force exerted by Q, on Q;, when both charges move at the same
velocity U, as in Section 6.1, can also be written
6-2. Repeat the calculation of Section 6.1 for a pair of equal masses m, and my,
pulled together by a spring under tension exerting a force Kr.
(a) Show that, in frame 1,
yB?K(xu1 — Xa1)Vo1
6-3. Calculate the force, as observed in the laboratory system, between two elec-
trons moving side by side along parallel paths one millimeter apart if they
each have a kinetic energy of (i) one electron-volt, and (ii) one million
electron-volts.
6-5. Draw a curve of the magnetic induction B as a function of the time at a point
1 centimeter away from the path of a single electron having an energy of 10
GeV (10° electron-volts).
6-6. Show that the equation of motion of a particle of rest mass my and charge Q
can be written as
ym &= o(F+oxB-2uk)
a =vxXw
where
OB
Orn
is called the Larmor frequency, or cyclotron frequency.
. (a) Show that, if there exists a field E,, B, with respect to the laboratory, then
the field E,, B, seen by an observer moving at a velocity VU is
288 RELATIVITY II
B, = B, —— 0 X E,.
(b) Show that, if E and B are orthogonal, the field in reference frame 2 is
purely magnetic if
6-11. Given a field FE, B in the laboratory frame, under what conditions is it possible
to find (a) a reference frame in which either E or B is zero; (b) one in which
FE and B are parallel, say in the direction of the y-axis; (c) one in which E
and B are perpendicular?
(b) We now turn the capacitor so that its plates are parallel to the yz plane.
The capacitor is again situated in frame 2.
Show that e, is again the same in both reference frames.
6-13. Calculate E and B as functions of the time at a point 1 centimeter away from
the path of a 10 million electron-volt proton.
6-14. The capacitor of the examples on pages 263 and 270 moves in a direction
perpendicular to its plates.
Calculate the potentials on the plates, and the A, E, B inside the capacitor
in the reference frame of the laboratory.
6-15. The magnetic induction inside a long solenoid of radius R that has N’
turns/meter and carries a current J is uo N’/. The magnetic induction outside
is zero (example on page 315).
(a) Calculate the values of E, and B,, both inside and outside such a solenoid,
as measured by a stationary observer, when the solenoid moves at a velocity
© in a direction perpendicular to its length.
Assume that the axis of the solenoid is the z axis and that U = Vi.
(b) Now calculate this same field by transforming the potentials.
First show that, in the frame of the solenoid, the vector potential
Bhy LB :
Ay = —— yoi + 5 xl
E, 1 = —VV, -—
BV
1
eA
6-16. The solenoid of the preceding problem moves at a velocity VU in the direction
of its length.
Calculate E; and B, both inside and outside the solenoid, as measured
by a stationary observer.
290 RELATIVITY II
6-17. Verify that the Lorentz condition applies to the potentials 4 and V of Egs.
6-78 and 6-91.
6-18. Show that Maxwell’s equations are invariant under a Lorentz transformation.tf
6-19. Calculate the drift velocity of conduction electrons in copper from the follow-
ing data:
J = 1 ampere/millimeter?;
Density, 9.
6-22. A beam of particles of mass m and velocity v has a radius r. The charge density
may be assumed to be constant throughout the cross-section of the beam.
Show that the outward force on an ion situated at the periphery of the
beam is
IQ I y2
>
9
21 eqrv Cc?
where J is the current in the beam and Q is the magnitude of the charge on an
ion.
6-23. A certain linear accelerator produces a beam of electrons having a kinetic
energy of 100 million electron-volts. The beam is pulsed, each pulse lasting
25 picoseconds (25 X 10-!? second) and containing 3.0 X 10° electrons. At
the exit of the accelerator the beam has a diameter of 5.0 millimeters.
You are required to investigate the divergence of the beam and, in par-
ticular, to calculate what distance a pulse could travel before blowing up toa
diameter of 2.0 centimeters. The beam travels inside a grounded conducting
tube that has an inside diameter of 10 centimeters.
One may assume that the charge density inside each pulse is uniform.t
(a) Calculate the length, diameter, and linear charge density of a pulse at the
exit of the accelerator in the reference frame of the electrons.
(b) To calculate the divergence of the beam, we shall first calculate the di-
vergence in the frame of the electrons, and then we shall transform to the
laboratory system under (c) below.
You should have found under (a) above that, in the moving frame, the
pulses are long cylinders of charge; so let us calculate the radius of a cylinder
as a function of the time.
PROBLEMS 291
Call the initial radius at the exit of the accelerator ry and the radius of the
tube r;, and calculate the beam radius r as a function of the proper time fo.
To do this you can calculate r as a function of f) for an ion at the periph-
ery of the beam. Radial velocities are small compared to c, and the mass can
be taken to be the rest mass #9.
Neglect end effects.
The integral for r is evaluated as follows:
Tr
iL dii 7
ip U U
= / dr = i ?r du — ar f exp ur du.
ie u
To (in") To 0 0
0
ps exp (u2) du = — em tn
o Teen
where Xz is the linear charge density and f. is the time, as measured in the
moving frame.
(c) Now transform to the laboratory system and draw a curve of r as a func-
tion of the distance x; traveled by a pulse in the laboratory, for values of r
ranging up to 10 millimeters.
You should find that the pulse has to travel about 400 meters before its
radius becomes equal to 10 millimeters. (We have neglected scattering by gas
molecules.)
(d) We have assumed that the radial velocity dr/dt, in the frame of the elec-
trons is small compared to c.
Was this assumption justifiable?
(e) We have also assumed that the charge density was uniform inside each
pulse.
Was this assumption really necessary?
(f) We shall see in Problem 8-30 that a continuous beam diverges in the same
manner as above.
Why is this?
Hints
6-22. Calculate the charge per unit length and the force in the reference frame of the ions,
and then transform to the reference frame of the laboratory.
6-23. Use the relation Potential Energy + Kinetic Energy = Constant, setting the po-
tential energy equal to zero at r;, and the radial kinetic energy equal to zero at ro.
CHAPTER 7
MAGNETIC FIELDS I
Steady Currents
and Nonmagnetic Materials
The next three chapters will deal with various aspects of magnetic fields. We
shall start by studying in some detail the properties of two vector quantities,
namely, the magnetic induction B and the vector potential A, which are used
to describe magnetic fields. We shall limit ourselves to steady currents and
nonmagnetic materials. We shall then discuss varying currents in Chapter 8,
and magnetic materials in Chapter 9.
We already know that there are several types of electric current.
(a) There are first the usual conduction currents through good conductors
such as copper, which are due to the drift of conduction electrons.
(b) There are also the conduction currents in semiconductors, which are
due to the drift of either or both conduction electrons and holes (see the
example on page 176).
(c) There are electrolytic currents.
(d) The motion of ions or electrons in a vacuum—for example, in an ion
beam or in a vacuum diode—gives convection currents.
(e) The motion of macroscopic charged bodies also produces electric cur-
rents.
All these currents are associated with the motion of free charges, and they
are the ones we shall be thinking of in this chapter. We shall use the symbol J;
for the current density associated with free charges, and J or J; for the corre-
sponding current.
(f) The polarization current density dP/dt in dielectrics (Section 3.2.2) is
associated with bound charges, and we shall assume for the moment that it is
zero.
7.1. MAGNETIC FORCES 293
(g) There are two other types of electric current, which we shall study
later, in Chapters 9 and 10. These are (a) equivalent currents in magnetized
substances and (b) displacement currents associated with changing electric
fields.
If you have not studied Chapter 6, simply disregard references to it in this
chapter and in all following chapters.
where F,,, is the force exerted by current J, on current J,, and where the line
integrals are evaluated over the two circuits.
This is the magnetic force law. The vectors dl, and dl, point in the
direction of current flow, r; is a unit vector pointing from dl, to dl, and r is
the distance between the two elements dl, and dl,. The force is measured in
newtons, the currents in amperes, and the lengths in meters. This law is in
agreement with Eq. 6-178.
The meaning of the double integral is as follows. We choose a fixed
element dl, on circuit 6 and add the vectors dl, X (dl, X 1r)/r? corresponding
to each element dl, of circuit a. We then repeat the operation for all the other
elements dl, of circuit b and, finally, calculate the over-all sum. In general,
this integration cannot be performed analytically. We then divide the circuits
into small finite elements and evaluate the sum numerically.
As we saw in Section 6-1, the constant po is defined as follows:
We now show that the double integral of the first term on the right is zero.
First
ffa(dly-r) ¥ f il fBets (7-4)
hi a b
Now, on the right-hand side, the second integral is zero for the following
reason. It is the integral of dr/r? around a closed curve, circuit b. It is therefore
the integral of (1/r?) dr with identical upper and lower limits of integration,
and it is zero.
7.2 THE MAGNETIC INDUCTION B. THE BIOT-SAVART LAW 295
We are thus left with the double integral of only the second term for the
triple vector product dl, X (dl, X 11), and
es Pes
re EE pp r:(dl,-dl,)
err ars (7-5)
Therefore, F., = —F%., since the unit vector r; is directed toward the circuit
on which the force is to be calculated, with the result that it is oriented in one
direction for F, and in the opposite direction for Fa.
Despite the fact that the above integral for F,, is simpler and more sym-
metrical than that of Eq. 7-1, it is not as useful. The reason is that, with the
above integral, the force cannot be expressed as the interaction of current b
with the field of current a. We can perform such an operation on Eq. 7-1,
however, since
is the magnetic induction due to circuit a at the position of the element dl, of
circuit b. As usual, the unit vector r; points from the source, to the point of
observation: it points from the element dl,, to the point where B is calculated.
The magnetic induction is expressed in fes/as or in webers/meter?:
dF =TdlX BL (7-10)
296 MAGNETIC FIELDS I
re a
Pie at) ee
Ee ae
(7-11)
The integration is carried out over any volume 7’ that includes all the currents.
The subscriptfon the current density J; refers to the fact that the currents we
are considering in this chapter are due to free charges. We assume, as we do
throughout this chapter, that J; is not a function of the time and that there are
no magnetic materials in the field.
Can this integral be used to calculate B at a point inside acurrent-—
carrying conductor? Since r is the distance between (a) the point of observa-
‘tion where B is measured and (b) the point where the current density is J;,
we are led to expect that the contribution of the local current density will be
infinite because of the 1/r? factor. The integral does not, in fact, diverge; it
does apply within current-carrying conductors. This can be seen by analogy
with electrostatics, where the same problem arises in calculating the electric
field intensity FEinside a charge distribution. The components of B and of E
both vary as 1/7’. Since those of FEdo remain finite within a charge distribu-
tion, those of B must also remain finite within a current distribution. We
assume that both the charge density and the current density are finite.
As in electrostatics, where we used lines of force to describe an electric
field, we can describe a magnetic field by drawing Jines of B that are every-
where tangent to the direction of B.
Similarly, it is convenient to use the concept of flux, the flux of the mag-
netic induction B through a surface S being defined as the normal component
of B integrated over S:
v= | Badu (7-12)
S
The flux @ is expressed in webers.
=
bol
—
4rp
+7/2
di
J—1/2 take dose ab
=
bol
2mp a
(7-14)
as in Eq. 6-158.
The magnitude of B thus falls off inversely as the first power of
the distance from an infinitely long wire and is in the direction per-
pendicular to a plane containing the wire. The lines of B are circles
lying in a plane perpendicular to the wire and are centered on it.
The force is attractive if the currents are in the same direction, andit
_is repulsiveifthey are in opposite directions.
This equation is the basis of the international definition of the
ampere: two long parallel wires separated by a distance of one meter
exert on each other a force of 2 X 10-7 newton per meter of length
when the current in each is one ampere. It is assumed that the diam-
eters of the wires are negligible compared to their separation.
bolal
Us, = dq p28 0; (7-18)
hence
bol 27a Lola?
Tee) oe Sermon Ce
The magnetic induction is maximum in the plane of the ring and
drops off as z3 for z? > a?.
where nv is the number of carriers per unit volume, v is their average drift
velocity, and Q is the charge on one carrier. The reason for this relation is
that the total charge flowing per second is the charge on the carriers that are
contained in a length » of the wire.
Then the force on the element dl is
ndadlQv X B,
and the force on a single charge Q moving at a velocity v in a field B is
Ov X B.
This force is perpendicular both to the velocity v and to the local magnetic
induction B.
More generally, if there is also an electric field E, the force is
the Hall field. The net transverse force on the mobile charges inside
the bar is then zero as in Figures 7-5c and 7-5d.
Let us assume that the conduction is due to holes carrying
charges +-e, e being the magnitude of the electronic charge, and that
there are p holes per cubic meter. When the transverse drift has
stopped,
Aa, = ef = evB. (7-21)
Now
I = Jab = (pev)ab, (7-22)
and
.-—. = pea
973 (7-23)
If the conduction is due rather to conduction electrons carrying
a charge —e, the Hall field has the opposite polarity, as in Figure
7-5b.
Note that in both cases the carriers are swept down by the mag-
netic field.
Although the Hall effect is in fact more complex than we have
assumed it to be, our value of V is nonetheless approximately correct.
The Hall effect is commonly used for measuring magnetic fields and
for studying conduction phenomena in semiconductors.
Q »
ee
Regt! i
Fee
Ay
Lge
vi \
Rt (7-28)
wo | Ie Xn
“i 4r Jy rr
dr’, (7-30)
Jv xtet-@xn-i x8), om
where the first term on the right is zero because J; is a function of the source
point x’, y’, z’, while the del operator involves derivatives with respect to
the field point x, y, z. The second term on the right is also zero because
i J k
Ae uae ©. 2 a a
WEDS eye) VE agers Ox oy dz ee
(x = xr — yYrt @ Zr
Then
a (7-34)
This equation follows from the definition of B given in Eq. 7-30. In
Chapter 6 we also showed that it is a consequence of Coulomb’s law and of
the Lorentz transformation. The fact that V-B is zero means that there can-
not exist sources of B.
The net flux of magnetic induction through any closed surface is equal
to zero since
(7-35)
7 mlb ey EC LTORSPOTBNTVALUTA
B=] 4r ae
v (2) x dea
lf
(7-37)
from Problem 1-12. Using the vector identity of Problem 1-18,
B= ef Ve (7-39)
4r J r
A=ryAeA beeiia dr :
(7-42)
aloes
nol
cfal (7-43)
Note that the vector potential A is not uniquely defined by the above
integrals. Indeed we can add to these integrals any quantity whose curl is
zero without affecting the value of B in any way. Since we have no reason
to add such a term for the moment, we shall use these integrals for defining
A, and we shall defer this question until later.
Note also that B depends on the space derivatives of A, and not on A
itself. The value of B at a given point can thus be calculated from A only if
A is known in the region around the point considered.
AV ee (7-44)
all in the same direction. From the fundamental definition of the
curl in terms of a line integral (Eq. 1-88), and from the azimuthal
symmetry of the field, V X A = B is in the azimuthal direction
around the conductor.
For an infinitely long conductor, dA is proportional to d/// for
large values of r, since r = /, and A tends to infinity logarithmically.
However, the fact that a function is infinite does not necessarily
mean that its derivatives are also infinite; that is, B can be finite even
though A is infinite.
Let us calculate A and B for a current of finite length 2L, and
then we can let L go to infinity. Referring to Figure 7-8,
pol fL dl .
= oe
Lol f
(ne 1+ 1+ 4) — inp} wa in
_, boll nz,
(7-47)
ys
0) Oar << IE
Figure 7-8. An element / dl of a current J
in a long straight wire produces an element
of vector potential dA at the point P.
Pi pPi 1
Ke) fe]
lt Palle Ge (7-48)
0 OQ. =A;
B, = 0, (7-49)
Bass (7-50)
OA,
By = — ap (7-51)
Figure 7-9 shows two long parallel wires that are separated by a dis-
tance R and that carry currents J of equal magnitude but in opposite
directions. To calculate 4 and B we begin with wires of finite length
2L, use Eq. 7-47 for a single wire, and add the two vector potentials
together:
iy peess 2L
Ave aes (in —In ) (7-53)
2n Pa Ps
307
In this case we can let L tend to infinity before computing the curl,
and
Pe eee
4r pz
a ee
Qn Pa
(7-54)
With the Cartesian coordinates shown in Figure 7-9,
B, SL OA,
ay me Hol
oe
E my)
3 = 0ill,
| 7-56
(7-56)
The line integral of 4-dl around a closed curve C is equal to the magnetic
flux @ linking C:
We have shown that the magnetic induction is always equal to the curl of
the vector potential -X A. We shall now show that
VX B= poly, (7-61)
again assuming a steady state and the absence of magnetic materials.
In terms of A,
VXB=VXVXA=V(V-4)—V2A (7-62)
from Section 1.9.6.
We have already shown in Section 6.7 that V-A is proportional to the
time derivative of the electric potential V, and we shall also show this again
in Section 10.3. Then, with the above assumptions,
V-4=0, (7-63)
For the second term, we have from the definition of A that
the contributions from all such sources in the volume 7’, which includes all
points at which J; exists. The volume 7’ may include the field point P, where
r = 0, as will be shown below.
Since J; is not a function of the coordinates of P, we can write the integral
as
Df) =. eo | ay 2 ()
van 1 ar’ 2
(7-65)
Now, by differentiation of
1 1
(7-66)
r (@—xP+0—-y¥ +E- zy)”
we find that V7(1/r) = 0 if r # 0. There can thus be no contribution to the
integral from any element dr’, except possibly if P and P’ coincide and r is
zero.
To investigate the integral at r = 0 we consider a small volume enclosing
a point P, where we wish to calculate V*A, situated inside the current distribu-
tion.
We take the volume so small that J; does not change appreciably within
it; J; may then be removed from the integral:
where dQ is the element of solid angle subtended at the point P by the element
of area da. Since the surface S’ completely surrounds P,
We can express the above result in integral form by integrating the normal
component of V X B over an arbitrary surface S:
Then, using Stokes’s theorem, we transform the left side of this equation into
a line integral around the closed path C, which bounds the surface S, and
the line integral path. Again, we are limited to nonmagnetic materials and
to static fields.
In many cases the same current crosses the surface bounded by the
integration path several times. With a solenoid, for example, the integration
path could follow the axis and return outside the solenoid. The total current
crossing the surface is then the current in each turn multiplied by the number
of turns, or the number of ampere-turns.
The circuital law can be used to calculate B when it is constant along
the path of integration. This law is thus somewhat similar to Gauss’s law,
which is used to compute E when E is constant over a surface.
J; w Ble (7-78)
wR?
Outside the conductor, B is azimuthal and independent of ¢,
so that, according to the circuital law,
3
3
B
5
pe)
oO
= =
cs
0 5 10 15 20 25
p (millimeters)
Figure 7-14. The vector potential A for a length of wire 2 meters long
and | millimeter in radius carrying a current of 1 ampere.
T or
ae iy (7-83)
eo Ter
where L. is the half-length of the conductor. Then
Hol ( 4P? a
A(p) = In +1 (op < R). (7-84)
4n R? R?
The logarithmic term is constant for a given wire and for a given
current, whereas the term p?/R* increases as the square of the radius
p inside the conductor. Figure 7-14 shows the value of A both inside
and outside a wire that is 1 millimeter in radius and 2 meters long
and that carries a current of 1 ampere.
We have assumed that the conductor is nonmagnetic.
The only term of importance inside the parenthesis is p?/R*, since
the others are independent of the coordinates. One may therefore set
A equal to
_ Hol p?
Aq R*
on ween a
Ne
posdbunnedds |
NINN
Hf
NE a
Figure 7-15. Toroidal coil of square cross section carrying a current J.
The figures shown by broken lines are paths of integration for B.
* See The Feynman Lectures on Physics (Addison-Wesley, Reading, Mass., 1964) Vol II,
Section 15-5, for a discussion of the quantum-mechanical aspects of this question.
7.7 AMPERE’S CIRCUITAL LAW 315
")
_ Hol
spree (7-88)
3. Inside the solenoid.
(a) B, = 0 inside because the line integral of B, d/ over a circle
of radius p, say the top edge of the small cylinder shown in Figure
7-16, is 27pBy, and this must be zero according to Eq. 7-75 because
there is no current enclosed by the path.
(b) Considering now path c in Figure 7-16, and remembering that
B, = 0 both inside and outside, and that B, = 0 outside, we see
that B,s = uoN’Is, and
Bosna (7-89)
The magnetic induction inside a long solenoid in the region
remote from the ends is therefore uniform and equal to po times the
number of ampere-turns/meter.
eB serene (7-93)
+1/2 any
Lo a*N'l dz
= ——__—___, 7-93
We shall now show that a small loop of wire of area S, situated at the origin
in a plane perpendicular to the z-axis and carrying a current J, produces a
magnetic field
bom .,
Vi ie “a Sin 6, (7-97)
Bei) (7-98)
where m = IS is called the magnetic dipole moment of the loop. Note that
this B field is similar to the E field of an electric dipole (Section 2.9 and
Figure 2-10).
We shall start by calculating
pol A dl
HM a (7-99)
for a circular loop, as in Figure 7-20, and then we can generalize to any shape
of loop.
Since the dl vectors have no z component, 4 can only have x and y
components. A little thought will show that, for any given value of r’, we
have two symmetrical dl vectors whose y components add and whose x com-
ponents cancel. Then we need only calculate the y, or azimuthal, component
of A. If ais the radius of the loop, then
7 Te 0
La lie
(7-100)
where g; is the azimuthal unit vector in spherical polar coordinates.
We must now express 7’ in terms of r and of y. First
r2 = p? + @ — 2arcosy, (7-101)
" 2 A 1/2
= (1te ee tons v) (7-102)
le |
i lie
Eo | wet7a
a
T we7. COS Y;
a
(7-103 )
if we make the assumption that a<_r. Since r/r’ & 1, we have substituted
r for r’ in the two correction terms on the right-hand side.
320
x y
Figure 7-20. Circular loop of wire carrying a current J. The vector potential 4 at
the point P is directed in the azimuthal direction.
To replace the cos y term by a function of y, we use the fact that the
scalar product r-a can be expressed either as
(xi + zk)-(acosei+asing j) = xacos¢ (7-104)
or as racos y. Then
cosy = =COS 9, (7-105)
if ie ax
ae 1 — > + 72 608 ¢, (7-106)
and
fol ae a ax
As Ae . a COs ~ 1 — ip + Pp COS ¢ dp 71; (7-107)
B, _= Mo
4° 2m
= cos 8, :
(7-112)
m 1 pr x Il, (7-115)
where r is the vector from the origin to the current element / dl, and where
the integral is evaluated over the circuit. Figure 7-21 shows the vector m for
a simple loop.
The concept of magnetic dipole moment can be further generalized to a
current distribution J; in a volume r:
Ti — a xX Jy dr. (7-116)
322 MAGNETIC FIELDS I
7.9 SUMMARY
We are concerned in this chapter with the magnetic fields that are due to the
motion of free charges. As a rule, we think in terms of conduction currents
in wires.
The magnetic force exerted by a circuit a carrying a current J, on a circuit
b carrying a current J; is
ee See fp dl, X ee x ri) (7-1)
where r; is the unit vector pointing from the element dl, to the element dl).
The constant yo is arbitrarily chosen to be exactly 44 X 10-7 newton/ampere?.
The magnetic interaction is best thought of as the interaction between the
current J, and the magnetic field of J,, or inversely. The magnetic induction of
current J, is
dl,
B, = 1, foun (7-8)
and is expressed in fes/as or in webers/meter®. For continuous distributions
of current,
=| B-da. (7-12)
S
Q(E + v X B).
This is the Lorentz force.
Starting from our definition of B, we then showed that
V-B= 0. (7-34)
Then
Be Vek: (7-41)
where
A= _ ual
ie falz (7-42)
is the vector potential. The line integral is evaluated over the circuit carrying
the current J.
If there are only conduction currents,
VX B= wd), (7-61)
and then
PROBLEMS
QQ» 2)1/2+
reper: See:
Now calculate this same force using Coulomb’s law and the magnetic
force law of Section 7.1, substituting Q,0 for J, and Q,V for J;.
Does the magnetic force law give the correct result in this case?
The magnetic force law should never be used to calculate forces between
single particles.
7-3. An electric arc is 5.00 centimeters long and carries a current of 400 amperes in
a direction perpendicular to a magnetic field of 5O gauss.
Calculate the magnetic force exerted on the arc.
7-4. Calculate the force due to the Earth’s magnetic field on a horizontal wire 100
meters long carrying a current of 50 amperes due north (magnetic). Take the
magnetic field to be 0.5 & 10~4 tesla at an angle of 70° with the horizontal.
7-5. (a) Find the current density necessary to float a copper wire in the Earth’s
magnetic field.
Assume the experiment to be done at the Earth’s magnetic equator in a
field of 10~4 tesla. The density of copper is 8.9 grams/centimeter?.
(b) Will the wire become hot?
The resistivity of copper is 1.7 * 10-8 ohm-meter.
7-6. Show that the total force is zero on a closed circuit carrying a current Jina
uniform magnetic field.
7-7. Show that the magnetic force on a coil is zero in regions where J; is parallel
to B.
. Show that the magnetic induction B can also be written in the form
where 7’ is any volume that encloses all the currents J;, and where the operator
V’ is evaluated at the source point x’, ’, 2’, where the current density is oF
7-9. It was stated in the example on page 277 that the magnetic induction due to
the current flowing in a straight wire is given by ol/2mrp within 1 °%, if p is
less than 7% of the length of the wire, and for points near the middle of the
wire. Is this correct? Neglect the field of the rest of the circuit,
7-10. Figure 7-22 shows part of a coil whose cross section has the shape of a pair
of intersecting circles.
(a) In what general direction is the field oriented in the region between the
conductors?
(b) One author states that the magnetic induction is uniform in the
hollow.
Is this true?
325
Figure 7-22.
7-11. (a) A straight, flat conductor of width 2a carries a current J. Show that
a Mol
oe 4ra™
The Mol ro
Dire 4ra us r
in the first quadrant, when the coordinate axes are chosen so that the edges
of the conductor are situated at x = --a, and when the current flows in the
direction of the positive z-axis.
The distance from the point where B is measured to the edge at x = a is
r;. The angle @ is that between r; and r, and is positive in the direction from
r; tO ro.
(b) Calculate B,, B,, B at a distance of 26.0 centimeters from the axis of the
conductor at an angle of 72.0° from the x-axis in the first quadrant for a strip
10 centimeters wide carrying a current of 5.76 amperes.
(c) Find the magnetic induction B (i) at an external point in the plane of the
strip at a distance D from its axis and (ii) at an external point in the plane
perpendicular to the strip and at the same distance D from its axis.
(d) How do the results of (i) and (ii) compare when D > a? Explain.
ide Along straight conductor has a circular cross section of radius R and carries
a current J. Inside the conductor, there is a cylindrical hole of radius a whose
axis is parallel to the axis of the conductor and at a distance 5 from it.
Show that the magnetic induction inside the hole is uniform and is equal
to
Mobl +t ——-
27(R? — a’) ‘Lt
. Use the Biot-Savart law to compute the magnetic induction B at the center
of a square current loop carrying a current /.
where a and a» are the angles subtended at the point by a radius R at either
end of the solenoid.
For example, if the coil has a length 2Z and if the point is situated at a
distance x from the center,
L—x aoe es L-x ;
COS a, =
[R?+(L — x)” Co [R?+(L + x2
(b) Plot a curve of 2B/yoIN’ as a function of x/R for L = 10R and for values
of x/R ranging from 0 to 15.
Note how quickly the field drops off outside the solenoid.
7-15. A solenoid has an inner radius R;, an outer radius Ro, and a length L.
(a) Show that the magnetic induction at the center, when the solenoid carries
a current J, is
_ pont poe+ (a? + B12
where
Ba en {ae
_® _
a, PIR
l= Vn = 2rn(a? — 1)8R}.
(c) Fabry has shown that, at the center of any solenoid,
BNA
B=G EF
734 Pe
i Magnetic
et field
Figure 7-23.
7-17. (a) Compute the period and radius of curvature of the path of an electron
moving in a plane perpendicular to the Earth’s magnetic field. Assume
B = 107% tesla,
e = 1.6 X 10719 coulomb,
m= 9.1 X 10-* kilogram,
electron energy = 3000 electron-volts,
one electron volt = 1.6 < 107! joule.
*(b) Repeat the calculation for an electron having a kinetic energy of 3.0
million electron-volts.
7-18. In the parallel-plate magnetron, the cathode and the anode are flat parallel
plates, and a magnetic field is applied in a direction parallel to the plates.
Electrons are emitted from the cathode at essentially zero velocity.
If the anode is at a distance s from the cathode, and if it is held at a poten-
tial V with respect to the cathode, show that for
eB*s?
fo 2m
no current will flow to the anode. The magnetic induction is B, e is the
magnitude of the electronic charge, and m is the mass of an electron.{
The parallel-plate magnetron is apparently not a useful device, but certain
types of crossed-field amplifiers use a fairly similar geometry.*
7-20. Figure 7-23 shows one type of rocket motor that utilizes the magnetic force.
An arc A ionizes the gas, which enters from the left, and the ions are blown
into the crossed electric and magnetic fields. If the current flowing between
* See, for example, A. F. Harvey, Microwave Engineering (Academic Press, New York,
1963), Chapter 12.
328 MAGNETIC FIELDS I
the electrodes C and D is J, then the thrust F developed by the motor is BJs,
if the values of B and s are assumed to be constant throughout the thrust
chamber, and if end effects are neglected.
If m’ is the mass of gas flowing through the motor per unit time, and if v
is the exhaust velocity, then F is m’v, and the kinetic power communicated
to the gas is
Vo=21 0 = Sho BI
If the efficiency of the motor is defined as
1 Wet Wo
where Wp is the power dissipated as heat between C and D, show that
n = ==
1 Lea =>
1
—_———_ =>
1
———’
2m Dy 2E
italy Fe eatery May
where go is the electrical conductivity, t is the volume of the plasma in the
crossed fields, J is the current density, and E is the transverse electric field
produced by the electrodes C and D.
7-21. When a direct current flows in a long straight conductor, the current density
may. be assumed to be the same throughout the cross section, unless the cur-
rent is subjected to a magnetic field originating elsewhere. Let us see how this
comes about.
(a) Calculate the magnetic force on a conduction electron drifting at the sur-
face of a wire 1 millimeter in radius and carrying a current of 1 ampere.
Set the electron velocity equal to 10-4 meter/second.
(b) Calculate the radial electric field intensity required to cancel the mag-
netic force at a radius r inside the wire, and find the corresponding volume
charge density.
(c) How many extra electrons are required per meter to supply this field, and
how does this compare with the number of conduction electrons per meter?
(d) Is it reasonable to assume that the current density is uniform throughout
the cross section of the wire?
7-22. A solenoid has a uniform turn density, except near the ends, where extra turns
are added to obtain a higher magnetic induction than near the center.
Show qualitatively that, if the axial velocity is not too large, a charged
particle that spirals along the axis in a vacuum inside the solenoid will be
reflected back when it reaches the higher magnetic field.
The regions of higher magnetic field are called magnetic mirrors.
Viegs. In 1949 Fermi proposed the following mechanism, now called Fermi accel-
eration, to explain the existence of very high energy particles in space.
Imagine a clump of plasma traveling in space at some velocity v,i. The
plasma carries a current, and thus has a magnetic field. Imagine now a par-
ticle traveling in the opposite direction at a velocity —2,i, both », and v, being
positive quantities. The particle is deflected in the magnetic field and acquires
a velocity vi, where v, is being also a positive quantity.
PROBLEMS
329
Set ». = v2 = c/2. Show that, after reflection, the y of the particle has
increased from 1.15 to 2.69.
See also Problem 5-25.
7-24. A sheet of dielectric situated in the x-y plane has a velocity vi in a region where
the magnetic induction is Bk. Find the polarization P and the surface charge
densities o,.
7-25. A current J flows in a wire bent around a square form measuring 2a meters
on a side.
(a) Calculate the vector potential 4 and the magnetic induction B along an
axis passing through the center of the square and parallel to one of the sides.
(b) Draw curves of A and B as functions of distance from the center of the
square when J is 1.00 ampere and a is 10.0 centimeters.
7-26. (a) Calculate the axial component of the vector potential A at the center of a
helix of 2N turns, of radius R, and of length 2H, carrying a current J.
(b) Show that the result is the same as that for a single wire of length 2H
along the side of the helix that carries a current J.
Why is this so?
(c) Is it possible to use this result to calculate the axial component of B at the
center ? :
(d) Show that the axial component of B at the center of the helix is
LolN
RES es
7-27. /A conducting sheet carries a current density of X amperes/meter.
(a) Show that, very close to the sheet, the magnetic induction due to the cur-
rent in the sheet is woA/2 in a direction perpendicular to X.
*(b) Show that the B given in the example on page 263 in the case of the
moving capacitor agrees with the above result.
7-29. A circular loop whose axis is coincident with the z-axis carries a current J.
(a) Calculate 0B,/dp at a point on the axis, (b) B, in the neighborhood of the
axis, (c) 0B,/dp at a point on the axis, and (d) B, in the neighborhood of the
axis.
7-30. Show that the permeability of free space uo can be defined as follows: if an
infinitely long solenoid carries a current density of 1.0 ampere/meter, then
the magnetic induction in teslas inside the solenoid is numerically equal to po.
330 MAGNETIC FIELDS I
7-32. A magnetic bottle is a field which has the property of containing a plasma.
One type of magnetic bottle operates as follows. An evacuated tube sey-
eral meters long and about half a meter in diameter is enclosed in a long sole-
noid that produces a uniform magnetic field. An accelerator injects electrons,
which spiral inside the tube, forming in effect a second winding whose field
opposes that of the solenoid. Magnetic mirrors (regions of stronger magnetic
field; see Problem 7-22) at both ends of the tube reflect the electrons, causing
them to spiral back and forth along the length of the tube, thus increasing
the magnetic field of the electrons which becomes larger than that of the
solenoid.
(a) Sketch a graph of B as a function of radius.
(b) Sketch lines of B in a plane containing the axis of the solenoid.
(c) Show, qualitatively, that the magnetic field prevents the electrons from
straying away from the current sheet.
(d) The electrons ionize the residual gas in the tube. Sketch the path of a
low-energy proton formed by the collision of one of the high-energy electrons
with a molecule or an atom of hydrogen.
2Ver 2
Bo a oe M.
a (gh
(c) What is the numerical value of this magnetic induction for a sphere 10.0
centimeters in radius, charged to 10.0 kilovolts, and spinning at 1.00 x 104
turns/minute.
(d) Show that the dipole moment is
47R3M k,
where k is a unit vector along the axis and is related to the direction of rotation
by the right-hand screw rule.
(e) What is the dipole moment of a sphere as in (c) above?
(f) What current flowing through a loop 10.0 centimeters in diameter would
give the same dipole moment?
PROBLEMS 331
7-34. Anelectron revolves in a circular orbit with an angular momentum 2!/24 about
a fixed proton. The constant # = 1.05 < 10~*4 joule-second is Planck’s con-
stant divided by 27.
(a) Calculate the magnetic moment in terms of #.
(b) Calculate B at the position of the proton.
Hints
7-8. Use the results of Problems 1-12 and 1-30. The current density J; must be zero
everywhere on the surface of r’.
7-12. Imagine that the hole is filled with a conductor carrying current of the same density.
Use the circuital law to find B. Then imagine that another current of the same density
but of opposite direction is superposed in the space occupied by the hole. Superpose
the two magnetic fields.
7-18. Use the Lorentz force and remember that its magnetic component does no work.
Thus the kinetic energy of an electron is equal to the potential energy it has lost
in the electric field. If the y-axis is normal to the plates and if the origin is at the
surface of the cathode,
dia?
> =e ie
5 ys)
7-19. Use the conservation of energy and Newton’s second law in the form Torque = Rate
of Change of Angular Momentum. Use polar coordinates.
MAGNETIC FIELDS II
Induced Electromotance
and Magnetic Energy
fpk-dl = 0, (8-1)
Thus the work performed by electrostatic forces is zero when a charge moves
around a closed path.
Equation 8-1 is not applicable if the path is linked by a changing mag-
netic flux. Let us consider a simple path having the shape of a loop. Then, if
® is the magnetic flux linking the loop,
if the direction in which the flux is taken to be positive is related to the direc-
333
tion in which the line integral is evaluated according to the right-hand screw
rule, as in Figure 8-1. This is the Faraday induction law; the line integral of
E-dl around the path is called the induced electromotance.
If there are no sources in a circuit, the current is equal to the induced
electromotance divided by the resistance of the circuit, exactly as if the electro-
motance were replaced by a battery of the same voltage and polarity. The
induced electromotance is expressed in volts, and it adds algebraically to the
voltages of the other sources that may be present in the circuit.
The above equation can be rewritten as
where S is any surface bounded by the path of integration chosen for the line
integral. This path can be chosen at will, and it need not lie in conducting
material.
If the path ofintegration is not just a simple loop, the surface over which
the magnetic induction must be integrated to obtain the flux linkage can be
complicated. The procedure for a solenoid is illustrated in Figure 8-2.
If the lines of B are parallel to the axis of the solenoid, they cross the
spiral ramp once for each turn of the solenoid, and the total flux crossing
the spiral surface, of the flux linkage, is N times the magnetic flux crossing the
surface corresponding to a single turn. The lines of B are then said to link all
N turns. If the lines of B are not parallel, the flux linkage varies from turn to
334
turn and the electromotance induced in the solenoid is equal to minus the
rate of change of the total flux linkage.
In general, therefore, the quantity ® in Eq. 8-2 must be interpreted as
the flux linkage. Flux linkage is measured in weber-turns.
from the magnetic force law discussed at the beginning of the last chapter.
Transformer-induced electromotance is observed in a rigid, fixed circuit
when it is linked by a variable magnetic flux. The distinction between motional
and transformer electromotance is artificial in some cases, in that different
observers would identify them differently. They would always agree, however,
on the electromotance. This will be shown in Appendix C.
If the flux linkage © increases, d®/dt is positive, and the electromotance is
negative, that is, the induced electromotance is in the negative direction.
On the other hand, if & decreases, d®/dt is negative, and the electromotance
is in the positive direction. >
irectionof the induced current is always such that it produces a
marnete field that opposes, to a greater or lesser extent, the change in flux,
depending on the resistance in the circuit. Thus, if ® increases, the induced
current produces an opposing flux. If @ decreases, the induced current pro-
duces an aiding flux. This is Lenz’s law.
x x
x
B
x
x x
= —wuB. (8-5)
We have assumed that the resistance of the loop is so large that the
current flowing through it has a negligible effect on the value of B.
The right-hand side is the area swept by the wire per unit time
multiplied by the magnetic induction B. It is thus the magnetic flux
swept per unit time as in Eq. 8-2.
If we had assumed that two or more sides of the circuit moved
in the magnetic field, the end result would have been the same; the
induced electromotance is always —d®/dt in such cases.
iepos S
Ot
(8-8)
8.2 THE INDUCED ELECTRIC FIELD INTENSITY 337.
This is one of the four Maxwell equations. The only other one that we
found without using relativity, was Eq. 3-41. All four equations were discussed
in Chapter 6.
Equation 8-8 is a differential equation that relates the space derivatives
of E at a particular point to the time rate of change of B at the same point.
The equation does not give the value of EF, unless it can be integrated.
(glvxas (8-9)
then, from Eq. 8-8,
VXE= -20 x 4), (8-10)
aA
or
vx (Eae=) =. (8-12)
The term between parentheses must equal a quantity whose curl is zero,
namely a gradient. Then
0A
Sle er VV. (8-13)
For steady currents, A is a constant and Eq. 8-13 reduces to Eq. 2-13.
The quantity V is therefore the electric potential of Section 2.3, which is also
known as the scalar potential.
Equation 8-13 is the general expression for EF, which we have already
found in Section 6.5. This equation states that an electric field intensity can
arise both from accumulations of charge, through the —VV term, or from
changing magnetic fields, through the —0A/dt term.
We could also have found Equation 8-12 in the following way. According
to the Faraday induction law, the induced electromotance for any given path
1S
db d
pe. dl SSS
i ap fpfi {-dl
* 5 (8-14 )
338 MAGNETIC FIELDS II
f ( as “) a0. (8-15)
We have used a partial derivative under the integral sign because we require
the time derivative of A at a given point on the path. Thus
(8-17)
and a similar relation exists for wire b. The flux through the loop is
thus
ii Tot+w h dpa ro+w
bie = (p chan / ites) (8-18)
To Pa rb Pb
fan ae dl
—ttdn[Mee]
ro(Fa + w)
The fact that the line integral is negative for a positive d//dt indicates
aan
that the induced current /’ is in the negative direction with respect
to B. Since B is directed into the paper in Figure 8-4, J’ must flow
in the counterclockwise direction. This is in agreement with Lenz’s
law: the current J’ produces a magnetic field that opposes the in-
crease in ®,
Let us now use Eq. 8-13 to calculate this same electromotance
from the time derivative of the vector potential 4, V being equal to
zero. From Section 7.5, A is parallel to the wires. Choosing the up-
ward direction as positive,
_ tol
A, = es In ae (8-21)
_ Kol
wu(nts) iy a
along the left- and right-hand sides of the loop, respectively. Thus
oan
Mo dl Tp
Se ee -2
Ey, 2m dt % Ce)
Ho dl (0+ *) 8.24
~ Qa dt ue eo! ( )
oe poh dl (rx + w) .
PEdi= oF alBees | oP}
Ho Talo WwW
8-25
Our differential form of the Faraday law, Eq. 8-8, was limited to systems at
rest. We shall now consider moving systems.
340
We return to the integral form of the Faraday law, Eq. 8-2, and consider
the rate of change of magnetic flux through a surface bounded by a path C
that moves in some arbitrary manner from C, to C; in the time df, as in
Figure 8-5. A given point P on the path C moves with a velocity u through
a region where the magnetic induction depends both on the coordinates and
on time.
The rate of change of flux through the circuit is
+ ar | V X (u X B)-da, = 0, (8-32)
and, finally,
db
posi)wryil
big xX (u X B) ; aa + OB
a peed
oF da. (8-33)
5B da.
[vy ad er Foe [Ir Yatinein = oe (8-35)
Since this equation must be valid for any surface S bounded by any curve C,
the integrands must be equal at every point, and
f Edt = i,
S
(FGh)-da = |s OB
sda.
Ot
men (ea)
= —BwS cos 6 cos wt, (8-40)
where S = Wh is the area of the loop.
For a square loop 10 X 10 centimeters? normal to a magnetic
field varying at a frequency of 60 hertz with a maximum value of
0.01 tesla (100 gauss), the induced electromotance has a frequency of
60 hertz maximum value of about 38 millivolts/turn. At any
mome
The only contributions to the integral are on the vertical sides, since
(uw X B) is perpendicular to dl along the top and bottom parts of
the integration path. Hence
Figure 8-7. Two circuits a and b. The flux ®,, shown linking 4 and
originating in a is positive. This is because its direction is related by the
right-hand screw rule to the direction chosen to be positive around b.
= Pb Aural, (8-48)
as Hola dl, ,
= fp(a f = ) dl, (8-49)
a ee ee (8-51)
where
Bas Ho dl,: dl,
Ma = ae pp rs (8-52)
is the mutual inductance between the two circuits. This is the Neumann equa-
tion.
Similarly, the flux @,, linking circuit a is
The mutual inductance between two circuits is one henry when a current
of one ampere in one of the circuits produces a flux linkage of one weber-turn
in the other.
The electromotance induced in circuit b by a change in the current J, is
fr-at tee
= —8 = Carpe
— m4 (8-55)
Similarly, the electromotance induced in circuit a by a change in the current
I, is
f beatee _ den
oe ay,abyAl (8-56)
8.4.2 Self-inductance
A single circuit is of course linked by its own flux as in Figure 8-8, and
wp nl @57
where L is called the self-inductance of the circuit and depends solely on the
346
db
f ral Sage dl (8-58)
It was shown in the example on page 315 that the magnetic induction
inside a long solenoid, neglecting end effects, is constant, and that
B = woN'l, (8-59)
8.4 INDUCTANCE AND INDUCED ELECTROMOTANCE 347
Radius/Length K
0 1.00
0.2 0.85
0.4 0.74
0.6 0.65
0.8 0.58
1.0 0.53
ik 0.43
2.0 0.37
4.0 0.24
10.0 0.12
NI
che ane TR?, (8-60)
2m Jr—w/2 ?P
ae
aR,
= ww In biees (8-65)
348
ins
~ 3
Figure 8-9. Toroidal coil of square cross section and mean radius R.
and
a polV2w 2R a “| rs
ee 6 ees . —
For R>w,
ZR RS (ew w ’
2R—w (174 wa )(4 + am) eolh
wy l+RaWw 8-
(8-68)
2R ww :
In [et] a (8-69)
and
Bolas
Poop (R> w). :
(8-70)
Figure 8-10. Coaxial solenoids. The two radii are taken to be approx-
imately equal.
RN,
ae tae (8-71)
and the mutual inductance is
N,®, R2N,
Vp Me)
Ih
UsHe
(8-72)
We can also calculate the mutual inductance by assuming a
current J; in coil 6. Then
_= Saye
Kom R°*NoI
sab
Pra (8-73)
This flux links only (/,//,)N. turns of coil a since B falls rapidly to
zero beyond the end of a long solenoid, as we saw in Problem 7-14.
Then the mutual inductance is
i [,NaPoa
a
a bot R2NLNs
(8-74)
as previously.
It is paradoxical that a varying current in the inner solenoid
should induce an electromotance in the outer one, since we have
shown (example, page 315) that the magnetic induction outside a
long solenoid is zero. The explanation is that the induced electric-
field intensity at any given point is equal to the negative time deriv-
ative of the vector potential at that point and that the vector poten-
tial A does not vanish outside an infinite solenoid, despite the fact
that B = V X A does.
We can actually calculate the vector potential just outside the
inner solenoid from the mutual inductance. If we consider the direc-
tion of the current /, in the primary as positive, then the electro-
motance induced in the secondary is — MdlI,/dt, or
pot R?N GN, dl,
Taare dt’
and the induced electric field intensity —d0A/dt is 2rRN, times
smaller:
350 MAGNETIC FIELDS II
= Ot = = PH.
as,
aki
8-75
CMB):
Integrating, we have the vector potential at any point close to the
solenoid:
yh me ip (8-76)
in the azimuthal direction. This can be shown to be correct by calcu-
lating the line integral of A-dl as in Section 7.5.1.
= ® Ls Safe:
La i (8-78)
called the coefficient of coupling between the two coils, can have values
ranging from —1 to +1. If the two single-turn coils coincide, the absolute
value of k is unity and its sign is the same as that of M.
Although this simple reasoning, which is based on single-turn coils, does
not apply to the coils which one uses in practice, Eq. 8-82 remains true and
the maximum possible value of k is unity.
We have seen in Section 8.4.2 that the self-inductance of a circuit tends
to infinity as the wire radius r tends to zero. The mutual inductance, however,
does not tend to infinity if one or both of the circuits is filamentary. The
reason is that the intense flux very close to a thin wire does not link another
8.5 ENERGY STORED IN A MAGNETIC FIELD 351
circuit some distance away. That is, as the radius of the wire tends to zero,
L— «,k-—0, and the mutual inductance remains about the same.
k a aoe (uomR?NN»)/la
(8-84)
(LaLs)¥? (oN 2a R2/1a)"! (ugN 3 R2/1,) 2
“sGy". (8-85)
It was assumed in calculating M that /, was shorter than, or equal
LOM
To find the energy stored in a magnetic field, we shall calculate the energy
supplied by a source to an isolated circuit, as the current density increases
from zero to some value J;.
We shall assume that Ohm’s law applies to the conducting medium.
Then, at any given point outside a source,
ip = ae. (8-86)
where a is the conductivity of the medium, expressed in mhos/meter, at that
point. This is Ohm’s Law.
This electric field intensity is the sum of two terms: —VV, produced by
accumulations of charge on the terminals of the source and on the surfaces
of the conductor, and —dA/dt, due to the changing magnetic field. In a wire,
the surface charges adjust themselves so that the total EF is along the axis as
in Figure 8-11.
Inside a source there is also a third electric field E,, which comes from
the local generation of energy and
the circuit is equal to the voltage supplied by the source. The source supplies
to the element of volume dt a power
dw 0A
“dt = . (E-J + Ot y J) dr, (8-89)
2
= [at [4 > J; dr. (8-90)
= 0 Ok
The first term on the right gives the power spent in Ohmic or Joule
losses (Problem 8-19), while the second gives the rate at which work is done
by the source against the induced electromotance. This latter work is that
which must be done to establish the magnetic field, and it is the one that
concerns us here.
Writing W,, for the energy stored in the magnetic field,
dW,
fe _ edie
[ad (8-91)
where 7 is again the volume occupied by the currents.
Wn = :uol2N’21R2, (8-94)
8.5 ENERGY STORED IN A MAGNETIC FIELD 353
dW,
Sade pst ld 8-100
dé Qu at (|RE a!
Setting W,, = 0 when B = 0,
Wu = | B? dr. (8-101)
354 MAGNETIC FIELDS II
The quantity W,,, is the work that must be done to establish a magnetic field
in terms of the magnetic induction B, either in free space or in nonmagnetic
matter.
We2uo /c-) B? dr = 2
pol?NUR?. (8-103)
The surface integral vanishes again as in Eq. 8-98, and when vy X B = Mods
(Section 7.6),
l
Le = 3 fs dr, (8-106)
where r is any volume that includes all regions where J; is not zero.
It is convenient to assign an energy density
dW, _ 1
Fi ey 5 Wy A) (8-107)
Example In the case of the long solenoid, the current is distributed, not over
a volume, but over the surface of the solenoid with a density \ =
N’'T. The vector potential A is parallel to J and is given by Eq. 8-76.
Then
Wn = ;/ NA da = 5 pol?NIR?. (8-108)
1 1
Va sip A-al = 5/8, (8-109)
from Eq. 7-60. The positive directions for J and ® are related according to
the right-hand screw rule.
Wn = 5Lr (8-111)
For two circuits carrying currents /, and /h,
] l
Wn = 2 [Pa “- wy I,®,, (8-1 12)
where ®, and , are the total fluxes linking circuits a and b respectively and
consist of contributions from both circuits:
Thus
1 1 1
Wr = 5Ube oe 5)TPy_ + 5)IPreg 0)TyPap. (8-115)
But
Di a Mi,, Pay —_ MI,, (8-116)
and therefore
The first two terms on the right are the self-energies due to the interaction
of each current with its own field, whereas the third term is the interaction
energy due to the mutual inductance.
Example Returning once more to the long solenoid, its stored energy is
Wn = LP, (8-119)
and, using the value of L given in Eq. 8-62, we again find that
Ve 5 Hol*N"IR?. (8-120)
We x | ey een (8-121)
2140 * 2
L= ae / B? dr, (8-122)
Mol Jj
We assume that the frequency is low enough to ensure that the cur-
rents are distributed uniformly throughout the cross sections of the
conductors, and we neglect end effects for simplicity.
We shall calculate successively the magnetic energies per unit
357
= OF bolo Hol?
=. -124
Wm all (ue MOE (G12)
(b) In region 2,
Ba Bol, (8-125)
27p
2
Ve ae (8-126)
4r a
(c) In region 3, the current within a circular path of radius p is
that in the center conductor, namely J, less that part of the current
that lies in the outer conductor between the radii b and p. Thus
LEN [7Se (2, es= )}
a (8-127 )
_ bol (S = a (8-128)
orp Xce —ob4
and 3¢2 — b
Wa =
wolf
Ar
cf
Ee — 5)? In b
. c¢ 3c2— B i 8-129
4(c? — BD?
(d) In region 4, B = 0 and W,,4 = 0.
Finally, the inductance per unit length is
(8-130)
Li =
nt 2(Wima a0 Wa
7=e Wns + Wma)
The second term within the braces is normally the most important;
it gives the inductance associated with the magnetic energy in the
annular region between the conductors.
Although the magnetic force law stated in Eqs. 7-1 and 7-5 was the starting
point for Chapter 7, we have not really used it to calculate the magnetic
force between two circuits, because of the difficulty of evaluating the line
integrals.
We are now in a position to find other expressions that are more con-
venient to use because they are based on mutual inductance. Although mutual
inductance is just as difficult to calculate, because it too involves a line
integral (Eq. 8-52), it is easily measurable.
We shall use the same method as in the example on page 78, where we
calculated forces on conductors. We assume a small virtual translation of
one coil, without any rotation, and then apply the principle of the conserva-
tion of energy:
Work done by the sources = Increase in magnetic energy
+ Mechanical work done.
For simplicity we disregard Joule losses. We also assume that the displace-
ments are made infinitely slowly so as to avoid taking kinetic energy into
account.
We consider two loops carrying currents J, and J, in the same direction
as in Figure 8-13. Since the forces shown in the figure indicate that the loops
tend to move toward each other, the coils can be kept fixed in space only if
the magnetic forces are balanced by equal and opposite mechanical forces.
The virtual displacement can be made in any convenient way: we may,
for example, assume that either the currents or the flux linkages are kept
constant. Whatever the assumption, the result must be the same, since the
force acting between two fixed circuits obviously has some single definite
value. We had a similar situation in the example on page 78 and in Problem
2-37, where we found the force on a capacitor plate from the conservation of
energy, assuming first that the plates were insulated, and then assuming that
they were connected to a battery. The forces were found to be the same.
359
Figure 8-13. The two loops carry currents J, and J,. Typical lines of B originating
in a are shown linking b. Observe that the elementary forces dF have a compo-
nent in the direction of coil a, with the result that the total force F is attractive.
To calculate F, coil b is assumed to be displaced parallel to itself by a distance
dr into the position shown by the broken line, and the mechanical work done,
F-dr, is found from the principle of conservation of energy.
where ®,, is the flux originating in b and linking a, and similarly for Pas.
The mutual inductance M and the fluxes ,, and ®,, all increase in this case,
so that dW,,, is a positive quantity, and the stored magnetic energy increases.
The mutual inductance is positive here.
Consider next the work done by the sources producing the currents in
the loops. In loop 6, ®,» increases, and the induced electromotance is in
such a direction that it would, by itself, produce a magnetic field opposing Ba.
360 MAGNETIC FIELDS II
aM
Fave = Inte ae (8-139)
Fy-dr = Ar
tale D
aJb
(Pe
lk
(8-142)
In the process of moving coil 6 by a distance dr, both dl, and dl, remain
unaffected, and the first d under the integral signs therefore operates only
on the 1/r factor. If we now define r; as a unit vector pointing from a to b
in the direction of dr,
a(*) =-9=-"%,
l ~ har 2 _nidr
(8-143)
q
and
Since the term dr on the right is independent of both dl, and dl,, it can be
removed from under the integral sign, and, finally,
oete
Nias = ri(dl, + dly) 8-145
Idk fp 2 : ( )
This is exactly Eq. 7-5, which is an alternative form of the magnetic force
law, Eq. 7-1. We have therefore rediscovered the law that was the starting
point for our discussion of magnetic fields in Chapter 7.
M=
(ND®oo _ (NiDPov (8-146)
ie | ae
where N; and Nj are the numbers of turns per meter on the two sole-
noids. Or, setting S to be their cross-sectional area,
BMoNENIDS1 = HoNNelSI
M : : (8-147)
= woNiN, SI, (8-148)
OM Ir
are 2 KoNLN;S, (8-149)
1 z
(8 151)
Wan = —
20 i B 2 dr,
1 : :
(8-152)
== Pita [Bi(ta — SI) + Biro — SD) + (Ba + Bs)2S/],
“95,
ce © Bir» + 2B.BpS)), (8-153)
1 2 y,
F = poNINESIalp. (8-155)
Note that the magnetic energy W,, is a function of / because
the magnetic energy density depends on the square of B. If it de-
pended on the first power of B, W,, would not be a functiom of J/,
and the force would be zero.
or
Bat Bt = Constant, (8-160)
and
dB, + dB, ;+ By ;dl = 0. (8-161)
i a, dl, (8-162)
CB ee pare
and a symmetrical expression for dB. Substituting then these two
equations into Eq. 8-158, and remembering that 7, is /,S while 7,
is /,S, we find again an axial force
F= usBABS. (8-163)
Ho
r= 1h, (8-164)
a (32). (8-165)
~ (FP). (8-166)
= - (2) (8-167)
where I’ is the torque and © is the flux linking the circuit.
In an isolated circuit the current flows in its own magnetic field and is there-
fore subjected to a force. This is illustrated in Figure 8-16: the interaction
of the current J with its own magnetic induction B produces a force that has
the effect of extending the circuit to the right.
Magnetic forces within an isolated circuit are calculated by using either
the magnetic force law or the principle of conservation of energy. In the case
of Figure 8-16, it is not practical to use the former method because it would
be too difficult to calculate the values of B and J inside the moving wire. The
energy method is therefore indicated. See Problem 8-28.
Figure 8-16. Schematic diagram of a rail gun. The battery A charges, through a
resistance R, the capacitor C, which can be made to discharge through the line by
closing switch S. The role of the capacitor is to store electric charge and to supply
a very large current to the loop for a very short time. If side D is allowed to
move, it moves to the right under the action of the magnetic force F’, Compare
this figure with Figure 8-3.
366 MAGNETIC FIELDS II
m= = Mol
See OF, (8-170)
8-170
a es
IreaR ‘
(8-171)
i=meet!
ye (8-172)
Figure 8-17. The pinch effect. In an ion beam the individual ions are
subject to an outward electrostatic force F, and to an inward magnetic
force F,,. At low velocities F,, << F, but, as the velocity v approaches
the velocity of light c, F,, approaches F,. If the electrostatic repulsion is
cancelled by mixing ions of the opposite charge in the ion beam, Fn
acts alone and the beam contracts. This phenomenon is often observed
in positive ion accelerators. Residual gas in the path of the ion beam is
ionized by impact, and the resulting low-energy electrons are trapped
in the positive beam while the low-energy positive ions drift away. Thus,
if the vacuum is not too good, the positive space charge in the beam
can be reduced and the focusing improved. This phenomenon is known
as the pinch effect, or as gas focusing.
8.10 MAGNETIC PRESSURE 367
ee a
IQ
185, = QmeoRv a w €oMov”), (8-173)
- fe v?
~ QarepRv ey te)
The net force tends to zero for v > c. This result is valid even if
Vv SC.
If the kinetic energy of the particle is equal to its rest energy
(5 X 10° electron-volts for an electron and 10° electron-volts for a
proton), then the net force F, — F,, is smaller than F, by a factor
of four (Problem 5-24).
In practice, a vacuum is never perfect. Let us say the ions are
positive. If their energy is of the order of tens of electron-volts or
more, they ionize the residual gas, and the resulting positive ions drift
away from the positive beam. The low-energy electrons, however,
remain trapped in the beam and neutralize part of its space charge,
thereby reducing F,. This type of pinch effect is often called gas
focusing.
Except for the ion beam of the last example, we have considered as yet only
the magnetic forces acting on conducting wires. If, instead, we have a current
Sheet, then it is appropriate to think in terms of magnetic pressure.
Let us imagine a flat current sheet carrying \ amperes/meter and
situated in a uniform tangential magnetic field B/2 due to currents flowing
elsewhere, as in Figure 8-18a, with A normal to B. If \ is very small, the
magnetic field is not disturbed appreciably by the current sheet and the force
per unit area is AB/2.
We now increase \ until it produces an aiding field B/2 on one side and
an opposing field B/2 on the other side, as in Figure 8-18b. Then, from Prob-
lem 7-27, \ = B/o, and the force per unit area, or the pressure, is B?/2,0.
We have arrived at this result for a flat current sheet, but the same applies
to any current sheet when B = 0 on one side.
This value of the magnetic pressure permits us to calculate the energy
density (Sections 8.5 to 8.5.4) in still another way. Imagine that the current
sheet moves back infinitely slowly by a small distance x. Then the work
performed by the magnetic pressure on a small area a of the current sheet
is ax(B?/2u9). This work is supplied by the sources that maintain B constant,
and it is equal to the increase in the energy stored in the field (Section 8.7.1),
368
(a) (b)
Figure 8-18. (a) A current sheet carries a very weak current density of \ ampere
per meter of its width and is situated in a uniform magnetic field B/2. (b) The
total magnetic field is that of figure (a), plus that of the current sheet. Selecting
so that it just cancels the magnetic field on the left-hand side of the sheet, we have
a total field B on the right.
10
log P 8
(newtons/ meter
or ’
joules/meter) 6
Ll ee | 1 =
2 3
log B (teslas)
energy density in an electric field from the electric force exerted on the
surface of a conductor. The remarks we made at that time concerning this
approach to the energy density in a field are equally valid here.
Figure 8-19 shows the magnetic pressure, or the magnetic energy density
B?/2yuo as a function of B.
since B is polN’.
8.11 SUMMARY
This chapter is concerned with (a) the nonconservative electric fields asso-
ciated with time-dependent magnetic fields, (b) the energy stored in magnetic
fields, and (c) the forces exerted on current-carrying circuits.
For a closed circuit,
db
fpral ear (8-2)
d
= -4 | Baa, (8-3)
where E is the total electric field intensity, including the induced part. The
directions in which the line integral is evaluated and the flux ® is taken to be
positive are related according to the right-hand screw rule.
This is the Faraday induction law, and the line integral is the induced
electromotance.
The negative sign indicates that the induced electromotance tends to
oppose the change in flux. This is Lenz’s law.
Induced electromotance 1s observed whenever the magnetic flux linking
a circuit changes with time, whether the circuit moves relative to the mag-
netic field, or whether a fixed circuit is linked by a variable flux, as in a trans-
former.
In differential form, the Faraday induction law becomes
OB
V SE eee eat :
oF (8-8)
ee a (8-13)
where the first term is that part of E that is induced by changing magnetic
fields, and the second is the part that is produced by accumulations of charge.
For a conductor moving with a velocity u in a changing magnetic field,
Eq. 8-8 becomes
OB
VXE=-9 +7 x (xB) (8-36)
8.11 SUMMARY 371
where ®,, is the flux originating in circuit a and linking circuit b. It is found
that
bo dl,-dl,
Ma = A pp r 2 (8-52)
This is the Neumann equation. It shows that M depends solely on the geometry
of the two circuits, if there are no magnetic materials present. By symmetry,
oe (8-57)
dW,
SIMI as
OA
aeew, ; 8-9]
dt [% Jp ar Ce
If there are no magnetic materials present, and if V X B= oJ; is
applicable,
a 5| Aer (8-106)
372 MAGNETIC FIELDS II
where 7 is any volume which includes all regions in which J; is not zero.
We also find that
WV = 51®, (8-109)
where ® is the flux linking the current /. As usual, the positive directions for J
and © are related according to the right-hand screw rule. In terms of the
inductances L,, Ly, M for two circuits a and b,
ay (2s). (8-140)
= Ce *). (8-140)
i= (dW,, ), (8-166)
as well as by other expressions similar to the ones above for Fy», with x
replaced by 90.
Magnetic forces also occur within an isolated circuit because a current
flows in its own magnetic field.
PROBLEMS 373
PROBLEMS
CoE carriage runs on two rails on either side of a long tank of water equipped
or testing boat models. The rails are 3.0 meters apart and the carriage has a
maximum speed of 20 meters/second.
Calculate the maximum voltage between the two rails if the vertical com-
__ ponent of the earth’s magnetic field is 2.0 & 10-° tesla.
Fed conducting bar slides at a constant velocity u along conducting rails in a
/
{
~
region of uniform magnetic induction, as in Figure 8-3. The resistance in the
circuit is R and the inductance is negligible.
(a) Calculate the current J flowing in the circuit.
(b) How much power is required to move the bar?
(c) How does this power compare with the power loss in the resistance R?
8-3/ If a conductor is given an acceleration a, the conduction electrons are sub-
\ jected to inertia forces —ma.
Show that, if E’ is the equivalent total electric field intensity in the con-
ductor, then
ed
Ot
ee
e
es
This effect was predicted by Maxwell, and was observed for the first time
in 1916 by Tolman and collaborators. c
The inverse effect, namely the acceleration of a body carrying a variable
current was also predicted by Maxwell, and was first observed by Barnett and
others in 1930.
/In the beratron, electrons are held in a circular orbit in a vacuum chamber
by a magnetic field B. The electrons are accelerated by increasing the magnetic
flux linking the orbit.
Show that the average magnetic induction over the plane of the orbit
must be twice the magnetic induction at the orbit if the orbit radius is to re-
main fixed as the electron’s energy is increased.
8-5. A toroidal coil is fed a voltage V(t) by a variable power supply. The length
of the wire is /.
374 MAGNETIC FIELDS II
Discuss the electric field intensity E inside the wire and show that E is
not equal to V/1.
In Problem 8-18, we shall see that the magnitude of the extra term is
0A/ot.
8-6. Show that the electromotance induced in a loop rotating at an angular velocity
w in a magnetic field By sin wr, as in Figure 8-6, is
— BySw cos 2wt.
if = w(1 _ a@
es
where Q is the volume of fluid flowing through the tube in one second.
This result is in fact incorrect, as we shall see below.
(b) Sketch a cross section of the tube, showing by means of arrows of various
lengths the magnitude of u X B at various points inside the tube.
(c) The fact that w X B is not the same everywhere in the fluid will cause
currents to flow.
PROBLEMS 375
Sketch another cross section of the tube, showing the lines of current
flow.
The current drawn by the electrodes is negligible.
It turns out that this current causes an Ohmic drop that reduces the
voltage calculated under (a) above by one quarter, and that the correct volt-
age difference is
2Q0B
Ta
(d) Show that this is the voltage one would expect if the velocity u were uni-
form and equal to the average velocity, and if there were zero Ohmic drop as
above.
(e) We have neglected edge effects in the regions where the conducting fluid
enters into and emerges from the magnetic field.
Sketch lines of current flow in these two regions.
These currents further increase the Ohmic drop.
*8-10. Repeat part (a) of the preceding problem without restricting your calcula-
tions to nonrelativistic velocities.
You should find that
AT = ;mo? a.
*8-11. In one type of ion drag accelerator a ring of high-energy electrons rotating in
a magnetic field and containing a small percentage of protons is accelerated
as a unit along its axis of symmetry.
A useful analogy for explaining the accelerating force on the protons is
that of amarble in a saucer in a railway train. If the acceleration is sufficiently
gradual, the marble stays in the saucer and acquires the velocity of the train.
the force on the protons is of course due to the electric attraction of the
electrons.
376 MAGNETIC FIELDS II
[sy =I fb) = :
B3,
Thus, at a plane where B3, is 0.2 tesla, r3 is 4.6 X 5!/2 = 10.3 centimeters.
This process is called expansion acceleration.t
(f) Show that y. corresponding to the axial velocity obeys the following
equation:
Yor (3)
VY 22 at B3:
(g) The protons have the same y. as the electrons, and their transverse velocity
is negligible.
(h) Show that the protons have a kinetic energy of 1.2 < 10° electron-volts
in the plane where Bz, is 0.2 tesla.
The protons therefore acquire energies of the order of 40 times that of
the electrons. This is the major advantage of this type of accelerator.
The ring could also be accelerated in an electric field (electric accelera-
tion). For example, if, the ring were accelerated through a difference of poten-
tial of 10 million volts, the protons would acquire a kinetic energy of 400
million electron-volts. Heavy ions would gain even higher energies.
. A wire bent in the form of a circle of radius R is placed so that its center is
at a distance D = 2R from a long straight wire, the two being in the same
plane.
Show that the mutual inductance is 0.268 yoR.
PROBLEMS
377
8-13. Two long parallel rectangular loops lying in the same plane have lengths /,
and /, and widths w; and we, respectively. The loops do not overlap, and the
distance between the near sides is s.
Show that the mutual inductance between the loops is
Mol» | S++ We
M=
20 a Ww
(1 sae
if 4 < h, and if the loops have but a single turn. Neglect end effects.
8-14. The coefficient of coupling k between two single-turn coils was defined in
Section 8.4.3. In general, ka ¥ ky.
Show that
ka _ Ly
ky ibe,
8-15. Show that the self-inductance of a close-wound toroidal coil of m turns, and
of major radius R and minor radius r is
8-18. (a) Show that the magnitude of the vector potential at the surface of a close-
wound toroidal coil of circular cross section is /27r, where ® is the magnetic
flux inside the coil and r is the minor radius of the toroid.
(b) Show that the magnitude ofthe extra term found in Problem 8-5 is 0.4/0t.
8-19. Show that the power lost by Joule heating per cubic meter in a conductor is
EJ;, or J?/c, where J; is the current density and a is the conductivity.
£20, compare the energies per unit volume in (a) a magnetic field of 1.0 tesla and
(b) an electrostatic field of 10° volts/meter.
8-21. Calculate the magnetic energy in a toroidal coil in at least two different ways.
8-22. Consider n rigid fixed circuits, the ith circuit carrying a current J/;.
378 MAGNETIC FIELDS II
Show that the magnetic energy associated with the system can be written
as a sum of self-energy terms of the form (3) L;J?, plus a sum of mutual energy
terms of the form M,,/;J;, each pair of currents appearing once.
*8-23. In the examples on pages 263 and 270 we transformed the field inside a
parallel-plate capacitor.
(a) Show that the field energy in frame 1 is y(1 + 6”) times that in frame 2.
(b) How do the energies & and & compare in the case of the solenoid of
Problem 6-15?
8-24. A coil of resistance R and inductance L is fed by a de power supply.
(a) Show that the power expended by the source when the voltage changes is
db
PR+ é G LI?) weit Tier
(b) Show that this is in agreement with Eq. 8-89.
8-25. Electromagnetic levitation forces are sometimes used to support objects in
space. For example, if a metal object is situated near a coil carrying an alter-
nating current J, eddy currents will flow in the object and there will result a
repulsive force.
Show that the force in the direction of x is
if the object is allowed to move in the x direction, and if the effective induc-
tance of the coil is L.t
8-26. (a) Show that a current-carrying coil tends to orient itself in a magnetic field
in such a way that the total magnetic flux linking the coil is maximum.
(b) Show that the torque exerted on the coil is m X B, where m is the mag-
netic moment of the coil and B is the magnetic induction when the current
in the coil is zero.
8-27. Many satellites require attitude control to keep their instrumentation properly
oriented. For example, communication satellites must keep their antenna
systems on target. Attitude control requires a mechanism for exerting appro-
priate torques as they are required.
Attitude control can be achieved by means of coils whose magnetic fields
interact with that of the Earth.
(a) Show that the torque exerted by such a coil is
NIBA sin 0,
where WN is the number of turns, / is the current, B is the magnetic induction
due to the earth, A is the area of the coil, and @ is the angle between the earth’s
magnetic field and the normal to the coil.
(b) Calculate the number of ampere-turns required for a coil wound around
the outside surface of a satellite whose diameter is 1.14 meters. The torque
required at 6 = 5 degrees is 10~* newton-meter, and the magnetic induction
at an altitude of 700 kilometers over the equator, where the orbit of the satel-
lite will be situated, is 4.0 X 10-5 tesla.
PROBLEMS 379
8-28. Let us calculate the force on the movable link D in the circuit of Figure 8-16.
Instead of being a metallic rod, the link can also be an electric arc. The
device then accelerates blobs of plasma and is called a plasma gun. Such
plasma guns can be used as rocket motors.
We cannot solve this problem by integrating the magnetic force J * B
over the volume of the link, since neither J nor B are known, or even easily
calculable. Instead, we shall find the force by investigating the magnetic and
mechanical energies involved.
If we set x = 0 at the initial position of the link, then the inductance L
in the circuit is Lyp + L’/x, where L’x is the inductance associated with the
magnetic flux linking the rails.
During the discharge of the capacitor C we can neglect the battery and
its resistance R, since the current flowing in that part of the circuit is negligible.
(a) Show that the force F on the link D is (i2/2)L’, where /is the instantaneous
current flowing in the right-hand side of the circuit.
Call the resistance on that side R’.t
(b) Show that, during the discharge,
COmmUOdE
__— ee
dQ
ptt 2D =—a()s
. We have seen in the example on page 366 that an ion situated at the periphery
of an ion beam of radius r is subjected to a force
LO “)
27 €or CG
(a) Show that the divergence of the beam is given by the following equation:
1/2
dt ro
where ry is the radius of the beam at a point where the radial velocity is zero,
and where
en I
TEUIN
(8) vy
(
Sa
dr [ edo r ]1/2
— = In
dty TEM Fo
which we found in Problem 6-23 for a pulsed beam of electrons of charge e.
*8-31. Show that F,, is equal to F, (Figure 8-17) within 10% when the beam energy
(the kinetic energy of a particle) is at least equal to twice the rest energy of a
particle, approximately.
For electrons, this beam energy is 1 MeV, and for protons it is 2 GeV.
8-32. Check the graph of Figure 8-19 and show that magnetic pressure is equal to
one atmosphere when B is only about 0.5 tesla, or 5000 gauss.
8-33. (a) Show that magnetic pressure is about 4B? atmospheres, if B is expressed
in teslas.
(b) Draw a log-log plot of the electric force per unit area }¢9E? as a function
of E.
(c) Discuss the similarities and differences between the magnetic pressure and
the electric force per unit area.
8-34. A thin tube of wall thickness ¢ and inside radius R carries a current /.
(a) Show that the magnetic pressure is directed inward and is (uo/87?)(//R)?.
(b) Calculate the pressure in atmospheres for a current of 1000 amperes and a
tube of 1 centimeter radius.
8-35. Magnetic fields are used for performing various mechanical tasks that require
a high power level for a very short time.
For example, the magnetic pressure inside a coil can be used to crush a
light aluminum tube that acts as a shutter to turn off a beam of light or of
soft X-rays. When the coil is suddenly connected to a charged capacitor, the
induced electromotance d®/dt produces a large current in the aluminum tube,
which collapses under the magnetic pressure.
Let us calculate the pressure exerted on a conducting tube placed inside
a long solenoid. If the current 7 in the solenoid is increased gradually from
zero to some arbitrarily large value, the current induced in the tube is small
and the magnetic pressure is negligible. Let us assume that d//dr is so large
that the induced current in the tube maintains zero magnetic field inside it.
Then there is a magnetic field B only in the annular region between the sole-
noid and the conducting tube.
(a) Calculate the pressure on the tube for B = 1 tesla.
(b) What would be the pressure if the conducting tube were parallel = the
axis, but off center?
8-36. Extremely high magnetic fields can be obtained by discharging a capacitor
through a low-inductance coil. The capacitor itself must of course have a low
inductance. Such fast capacitors cost approximately one dollar per joule of
stored-energy capacity.
(a) Estimate the cost of a capacitor that could store an energy equal to that
of a 100 tesla (one megagauss) magnetic field occupying a volume of one
liter.
PROBLEMS 381
(b) Estimate the cost of the electricity required to charge the capacitors.
(c) Calculate the magnetic pressure in atmospheres, at 100 teslas.
ee eg ah
(. 8-37 Flux compression is one method of obtaining large values of B. For example,
ee cs
a magnetic field Bo is established inside a conducting tube of radius Ro, and
the tube is then imploded by means of an explosive placed around the tube.
Currents flow in the tube, and the internal magnetic pressure builds up until
it is equal to the external gas pressure.
(a) Show that, if the radius of the tube shrinks very rapidly,
; B = Bo(Ro/R)?,
where B is the magnetic induction when the radius has been reduced to R.
(b) Show that the surface current density in the tube must be 10° am-
peres/meter to achieve a field of 10° teslas.
(c) If the initial B is 20 teslas, and if the tube has initially a diameter of 20
centimeters, what should be the value of B when the tube is compressed to a
diameter of about 5 millimeters?
(d) ‘Calculate the resulting increase in magnetic energy, and hence the explo-
sive energy, required to compress the field. Assume that the cylinder is 20
centimeters long, and neglect end effects. 2,
(e)) Flux compression can also be achieved by means of a conducting piston
sKot axially into a solenoid.
If the radius of the solenoid is Ro, and if the radius of the piston is R,
show that the magnetic induction in the annular region between the piston
and the solenoid reaches a value of
Bo
- (Ri)
Hints
8-4. Relate the centripetal acceleration to the magnetic force acting on the electron, and
then find the condition that lets the linear momentum of the electron increase with
fixed orbit radius. Use Newton’s second law with the tangential force on the electron
given by the Faraday induction law. |
8-9. (a) If you find a AT that is too large by a factor of 2, you have probably made an
approximation that is not legitimate. The increase in kinetic energy AT is the work
done by the force Q(0A/d1).
8-10. When the magnetic field is constant, the mass of an electron remains constant and
the magnetic force BeV is equal to the centripetal force mv’/r, as for the nonrela-
tivistic case, except that m is the relativistic mass mo.
(e) Remember that the particles are submitted to a central force. What can you say
about their angular momentum?
(f) Use the fact that yo2y29 = y3zv3e = y, aS demonstrated in Problem 5-14, and
remember that v9 v39 & c in the reference frame of the ring.
8-29. Assume that the current is concentrated along the axes of the conductors. Can you show
that this gives the correct result?
* For a more general proof of this result, see Sir James Jeans, The Mathematical Theory
of Electricity and Magnetism, Fifth Edition, Cambridge University Press (1951), Paragraph
564.
CHAPTER Y)
Thus far we have studied only those magnetic fields that are due to the motion
of free charges. Now, on the atomic scale, all bodies contain spinning elec-
trons that move around in orbits, and these electrons also produce magnetic
fields.
Our purpose in this chapter is to express the magnetic fields of these
atomic currents in macroscopic terms.
Magnetic materials are similar to dielectrics in that individual charges
or systems of charges can possess magnetic moments, and these moments,
when properly oriented, produce a resultant magnetic moment in a macro-
scopic body. Such a body is then said to be magnetized.
If m is the average magnetic dipole moment per atom, and if N is the number
of atoms per unit volume, the magnetization is defined as
M=Nmg (9-1)
if the individual dipole moments in the element of volume dz considered are
all aligned in the same direction. If the moments are not all aligned, then the
magnetization is the net magnetic moment per unit volume. The volume dr
is subject to the limitations discussed in Section 3.1.
soitpolieatonPinia NNNSnSnon
9.2 MAGNETIC INDUCTION B
Ad ANCEXTERIORSPOINT
aes Mo m Pa al
: 4r
(9-2)
where m is the magnetic dipole moment of the loop, and where the distance
ris large compared to the size of the loop. The unit vector r; points, as usual,
from the source, which is the loop, to the point P where the field is calculated,
385
dA = (9-3)
and
eh ae ae
_ Mo [Mx
=e le \ciet
(2)ar, :
(9-5)
Ae= del
aia | Ae / Ho
4 M= dr’ ! + a i)Veh
: ’
dr’, (9-6)
A=_4real |a MXn,,,
ie da eae VXM
u |Pair ,
dr’, :
(9-7)
.
Figure 9-2. Ampere
provided the numerators represent a surface current density in the first term
and a volume current density in the second term. We therefore define an
equivalent surface current density
qweMixend 0-8
and an equivalent volume current density
=v xXM. (9-9)
Then the vector potential at P is
eM (9-12)
The origin of the equivalent surface current density A, can thus be understood
in terms of currents on the atomic scale.
9.3. MAGNETIC INDUCTION B AT AN INTERIOR POINT 387
(9-13)
Thus, if we know the magnetization M, we can find the equivalent
current densities. Then we can use the equivalent currents to calculate B
as if they were in a vacuum.
We have demonstrated the validity of this procedure only for points
situated outside the material, but the procedure is equally valid inside, as
we shall see in the next section.
We now calculate the macroscopic magnetic induction, or the space and time
average of the magnetic induction on the atomic scale, at an interior point in
magnetized material.
We proceed as we did with dielectrics and divide the material into two
regions, one near the point P and one farther away. The near region is a small
sphere of radius R as in Figure 9-3, and the far region is the remaining volume
of magnetized material.
ib to fe tae
- mf KO Ge +B Me (9-14)
Led is i nr Ss”
a”
389
(b)
Figure 9-4. The fields (a) of a magnetic dipole and (b) of an electric dipole.
Let us first compute the average magnetic induction produced over the
volume of a sphere by a small current loop inside it. For simplicity let us
place the loop of magnetic moment m at the center of the sphere, as in
Figure 9-5. By definition, the average field in the sphere is
where 7 is the volume of the sphere. Since we do not know the value of 4 in
the immediate vicinity of an atom, we transform the volume integral into a
surface integral, using the result of Problem 1-30:
B= af (n X A) da, (9-16)
eels!
390
A = (Re (9-17)
4r Pr
We can now evaluate the volume integral of Eq. 9-15 without knowing the
value of A inside, provided we know its value at the surface:
.
Be 4x Js
nee
qe
ag (9-18)
or, expanding the double vector product as in Problem 1-8,
_ Mo m(n-r;) — ri(m-m)
| Ba = wf R da. (9-19)
The first term in the numerator on the right is simply m. The second term is
rym Cos 6 and, by virtue of the symmetry, we take its component along m,
which makes it m cos? 6. Therefore
and the average magnetic induction inside a sphere of radius R due to the
single loop of magnetic moment m at its center is
(2/3)uom
——y
wom
=
‘ (A/3ycR® ow Deke
9.3. MAGNETIC INDUCTION B AT AN INTERIOR POINT 391
If the current loop is not situated at the center of the sphere, the average B
is unaffected, as you will be able to show in Problem 9-3.
Since there are many atomic current loops inside the sphere, each one
produces an average field as above, and the total average field due to the near
dipoles is
47R$
Wy
eas 3 NR!
K
= 3 HM (9-22)
if they are all aligned in the same direction.
This result is open to question, because some of the loops will be too
close to the spherical surface for the dipole approximation of Eq. 9-17 to be
valid. However, the error can be made arbitrarily small by increasing the size
of the sphere. Molecules are so small that it is easy to choose a microscopic
sphere that is large enough to make this error negligible. This type of problem
was encountered previously in Section 3.3.
The magnetization M may be taken to be uniform inside the small sphere
S’’, and we may assume that the B at the center is this average B we have just
calculated, or (2/3)u.M.
(9-23)
Let us evaluate the third term. This term gives the magnetic induction at
the center of the sphere due to the currents flowing on the surface S’’. The
equivalent current density \, on S” is M X n and is in the direction shown
in Figure 9-6. Thus, at the center of the sphere,
Ae
Sra
X11 (MXn)Xn
R2 3
(9-24)
' These currents give a magnetic induction that is in the direction opposite to
M, and we must therefore take the component of (M X n) X n in the direc-
tion of —M:
= — FM. (9-26)
392
Mxn
Now, if we extended the first integral to the complete volume +’, which
is 7’’ + 7’’’, we would be adding the integral of (J. X ri)/r? over the volume
7’ of the small sphere. This latter integral is zero, by symmetry, because J,
is constant throughout the small volume 7’”’. Finally,
Mo
pam f Xm
[ 2 dr 7, etwoos f
an eX
p da’,Y, (9-28)
| v-B=01| (9-29)
of Sections 6.9 and 7.4
This is one of Maxwell’s equations.
It also follows that the vector potential due to magnetized material is
9.4 THE MAGNETIC FIELD INTENSITY H. AMPERE’S CIRCUITAL LAW 393
a= pf
T Jr
dt vxrBeal M 4
is Js
Sees
i
(OD)
Now we have just seen that magnetized material can always be replaced by
its equivalent currents for calculating B. Consequently, if we have magnetized
material as well as steady currents,
vx (7_ u) = J (9-35)
The vecto: between parentheses, whose curl is equal to the free current density
at the point, is the magnetic field intensity
nex 029
Integrating over a surface S,
where C is the curve bounding the surface S, and J; is the current of free
charges linking the curve C.
currents, The term on the left is called the magnetomotance.
This is a more general form of Ampére’s circuital law (Section 7.7) in
that it can be used to calculate H even in the presence of magnetic materials.
‘tisrigorouslyvalidhoweser,onlyforsendyeurens we shall dea with
variable currents in the next chapter.
Sections 9.8.1.1 and 9.8.1.2 will illustrate the use of H and of Ampére’s
circuital law.
395
Now, since
MM = xl) (9-42)
B= p(H + M), (9-43)
then
where
© B= wl + Xm)H = pouHl= wH, (9-44)
T= Ltd (9-45)
The quantity yu is called the permeability. Both x,, and yu, are dimensionless
quantities.
Equation 9-43 is general, but Eqs. 9-44 and 9-45 are based on the
assumption that the material is both isotropic and linear, in other words,
that M is proportional to H and in the same direction. This assumption is
unfortunately never completely true in ferromagnetic materials. In a perma-
nent magnet, for example, B and H point in roughly opposite directions
(example, page 402). Even in so-called soft magnetic materials, such as soft
iron, B and H do not always point in the same direction, and when they do,
u, 1s by no means constant. It can vary by orders of magnitude, depending
on the value of H and on the previous history of the material (Section 9.6).
The quantities x,, and yu, are therefore not as meaningful as the corresponding
quantities x, and e, for dielectrics.
Bommonsandstonep 0 Salesian ihn aL.
In paramagnetic materials M is given by the Langevin equation 3-98
with p replaced by m and £,,, by B. As for dielectrics, the magnetic energy
mB is much less than kT at room temperature. For example, m is of the order
of 10-3 ampere-meter?, and, even in a magnetic field of one tesla, mB/kT ~
2.5 X 10-*. Expanding the exponentials as in Section 3.9.1,
Nm
Ms Te
Pe ae (9-46)
ue
M M — wNm
Fi a) Lh,ee) ce 9-4
> OS Tg ey 3kT CD
The magnetic susceptibility of paramagnetic substances is smaller than unity
by several orders of magnitude and is proportional to the inverse of the
absolute temperature.
396 MAGNETIC FIELDS III
Dad Jr ay oe Ho vy + M xn
da’. (9-53)
An Jy 4a} sr r
A =
The surface integral cannot be transformed like the volume integral, because
A; and M X n are unrelated; for example, a piece of soft iron placed in a
magnetic field acquires a magnetization M even though its A; is zero.
9.6 HYSTERESIS
& 1 J
100 200
(ae
meter
(teslas)
B
100 200
(222ce
meter
e =|
Figure 9-9. The shaded area gives the energy required, per unit volume,
in going from g to 4 on the hysteresis loop. The total energy per unit
volume required to describe a complete hysteresis loop is equal to the
area enclosed by the loop.
where S is the cross-sectional area of the ring, N is the number of turns, and
B is the average magnetic induction in the core. Also,
dw INSIdB dB
pee am ar aD oe
where / is the mean circumference of the ring, 7 = S/ is its volume, and
H = NI/I, as in Eq. 9-54. Thus
b
We = | HdB (9-58)
g
is the energy supplied by the source in going from the point g to the point b
in Figure 9-9. This integral corresponds to the crosshatched area in the figure
and is equal to the energy supplied per unit volume of the magnetic core.
400 MAGNETIC FIELDS III
When the current is in the same direction but is decreasing, the polarity
of the induced electromotance is reversed, according to Lenz’s law, with the
result that the energy
W,=+ i HaB b
(9-59)
is returned to the source.
Finally, the energy supplied by the source during one cycle is
where the integral is evaluated around the hysteresis loop. The area of the
hysteresis loop in tesla-ampere turns/meter or in weber-ampere turns/meter®
is therefore the number ofjoules dissipated per cubic meter and per cycle in
thecore:
Note that this energy loss is unrelated to the losses associated with the
eddy currents. Eddy-current losses, which are caused by changes in the mag-
netic field, can be minimized by using laminated materials, as in transformers,
by using powders dispersed in an insulator, or by using nonconducting ma-
terials. Hysteresis losses, however, can be minimized only by selecting a
material with a narrow hysteresis loop.
Example For the case illustrated in Figure 9-9, the area enclosed by the hyster-
esis loop is roughly 150 weber-ampere turns/meter®, or 150 joules
per cubic meter and per cycle, or 600 calories per cubic meter and
per cycle. This energy loss could be important if the material were
subjected to an alternating field.
Let us examine the continuity conditions that B and H must obey at the
interface between two media. We shall proceed as in Section 4.1.
Figure 9-10a shows a short cylindrical volume whose top and bottom
faces are parallel and infinitely close to the interface. Since there is zero flux
through the cylindrical surface, the flux through the top face must equal that
through the bottom, and
the interface and close to it. From the circuital law, Eq. 9-41,
fpnat = 7, (9-62)
ee | 9-65
from the continuity of the tangential component of H. Then
fi) = be M, (9-67)
Ko
from Eq. 9-36. (Remember that the relation H = B/y is valid only
403
in soft magnetic materials that are both linear and isotropic.) Near
the center of a long bar magnet, B ~ uoM and H ~ 0. For a short
bar magnet we can use the values of B that we calculated for a short
solenoid in the example on page 317, substituting again the mag-
netization M for the current density /N’. Thus, at the center of the
magnet,
B= woM sin On, = —M(1 — sin @,,), (9-68)
while at the center of one face
404 MAGNETIC FIELDS III
5 (a) . (b)
2
en 2
L (a)
Figure 9-12. (a) Bar magnet. (b) Bar electret. (c) At any point in space
the magnetic induction B due to the magnet is the same as that of a
solenoid of the same size carrying a current density N’J equal to M. The
number of turns per meter on the solenoid is N’. (d) Similarly, the
electric field intensity E due to the electret is the same as that of a pair
of disks occupying the positions of the end faces and carrying charge
densities +P. :
Indeed, the H field of the magnet can be calculated like the E field
of the electret if one assumes the existence of a magnetic pole density
equal to +M on the end faces.
the core to pass again through the coil. This constitutes the leakage flux that
may, or may not, be negligible. For example, if the toroid is made up of a
long thin iron wire, the flux at P is negligible compared to that near the coil.
Let us assume that the cross section of the toroid is large enough to
render the leakage flux negligible. Then, applying Ampére’s circuital law,
Eq. 9-41, to a circular path of radius r going all around inside the toroid,
_ BNI
== :
(9-76)
Then, taking R, to be the radius corresponding to the average value of B, and
R, to be the minor radius of the toroid,
= pmR3
® ig NI. (9-77)
The flux through the core is therefore the same as if we had a toroidal coil
(example, page 313) of the same size and if the number of ampere-turns were
increased by a factor of u,. In other words, for each ampere-turn in the coil
there are u, — 1 ampere-turns in the core. The amplification can be as high as
10° or more.
This equation shows that the magnetic flux is given by the magnetomo-
tance N/ multiplied by the factor
pars
2rR,y
9.8 MAGNETIC FIELD CALCULATIONS 407
which is called the permeance of the magnetic circuit. The inverse of the
permeance is called the reluctance. Thus
_ Magnetic Flux = Permeance X Magnetomotance (9-78)
a orR>
u io 27k, A;
(9-80)
Thus the corresponding quantities in electric and magnetic circuits are as
follows:
Current J Magnetic flux ®
Current density J Magnetic induction B
Conductivity Permeability u
Electromotance Magnetomotance
Electric field intensity E Magnetic field intensity H
Conductance Permeance
Resistance Reluctance
There is one important difference between electric and magnetic circuits:
9.8.1.1 Magnetic Circuit With an Air Gap. Figure 9-14 shows a circuit with
an air gap whose cross section is different from that of the soft-iron yoke.
Each winding provides N/J/2 ampere-turn.
We shall see that the magnetic flux in the air gap is equal to the mag-
netomotance NJ divided by the sum of the reluctances of the iron yoke and
of the air gap.
This is a general la
We assume that the leakage flux is negligible. As we shall see, this will
result in quite a large error.
408
NI =H Las AL (9-81)
where the subscript i refers to the iron yoke and g to the air gap. As we shall
see below, the requirement that the reluctance of the yoke be much smaller
than that of the gap makes H,;L; < H,L,. The path length L; in the iron can
be taken to be the length measured along the center of the cross section of the
yoke.
Now, according to Eq. 9-29, or 6-125, the net outward flux of B through
any closed surface must be equal to zero. This results from the fact that V-B
is always zero; indeed, this is one of Maxwell’s equations. Thus, if we neglect
leakage flux, the flux of B must be the same over any cross section of the
magnetic circuit and
B;A; = B,A,; (9-82)
where A; and A, are respectively the cross sections of the iron yoke and of
the air gap.
Combining these two equations,
L; L, =
BAG |+ = —+ |= NI, (9-83)
LA; Lo A g
Now we have two reluctances in series, L;/A, in the yoke, and L,/uoA, in
the gap.
You can now show that H,L; << H,L,, as in Eq. 9-81, if the reluctance of
the yoke is much smaller than that of the gap.
Since we have neglected leakage flux, this equation can only serve to
provide an upper limit for &.
Example If there is a total of 10,000 turns in the two windings, and if J= 1.00
ampere, 4; = 100 centimeters”, 4, = 50.0 centimeters?, 4, = 1,000,
L; = 90.0 centimeters, L, = 1.00 centimeter, then
104
p = pena! 0.9we ERE a [Oa a ee (9-85)
CS Omeel02
is 1.00 = 63 henrys ie
(9-88)
In this particular case the leakage flux is 70% of the flux in the gap.
In other words, the magnetic induction in the gap is not 1.3 teslas,
but only 1.3/1.7 = 0.77 tesla.*
* There exist empirical formulas for calculating leakage fluxes. See, for example, R. K.
Tenzer, Estimating Leakage Factors for Magnetic Circuits by a Simple Method, Electrical
Manufacturing, February 1957.
410
This is a good illustration of the fact that the flux of H over a closed
surface is not necessarily equal to zero (Problem 9.19). In fact, the ends of the
permanent magnets act as either sources or sinks of H. For example, the
lower end of the top magnet acts as a source of H whereas the top end acts
as a sink of H.
To calculate the magnetic flux we proceed as follows. Since there are
neither free currents nor displacement currents, the line integral of H-dl
around the magnetic circuit must be zero. Then
where the subscripts i, g, m refer, respectively, to the soft iron, the air gap,
and the permanent magnets. The magnetomotance H;L; is again negligible
if the reluctance of the iron is much smaller than that of the air gap.
9.8 MAGNETIC FIELD CALCULATIONS 411
Also, if the leakage flux is negligible, & is the same over any cross section,
and
Dies Ze Ee (9-92)
These two equations are similar to Eqs. 9-81 and 9-82, except that NJ is
replaced by HinLm. Thus
LA; bog
Remember that L; is the length of the path in the soft-iron yoke only,
and that the total distance around the circuit is L; + Lm + Ly.
The design of a magnetic circuit energized by a permanent magnet is
somewhat complicated by the fact that the position of the operating point
on the hysteresis curve (Section 9.6), and hence the value of H,,, depends on
the permeance of the magnetic circuit.
As a rule, one tries to use an operating point that makes the energy
product H,,B, maximum. The reason for this is the following. The purpose
of the magnetic circuit is to produce a magnetic field in an air gap, and the
1.0
0.8
3 x 107 0.6 @
5 3
isa)
2 0.4
iS
1 0.2
0.5
= (he Sy eas ee a
5 x 10° 4 3 2 1 0
H (ampere-turns/meter)
Figure 9-16. Demagnetization curve for Alnico 5. The energy product is maximum
for B = 0.8 tesla.
412 MAGNETIC FIELDS III
magnetic energy required is the volume of the air gap, A,Z,, multiplied by
the energy density $40B?. Now
1 1
Ane A, LB; =e (H,L,)(A,B,), (9-94)
2 Lo 2
Example Let us design a magnetic circuit that will produce a magnetic field
of 1.3 teslas in an air gap 1.00 centimeter long and 50.0 centimeters?
in area as in the previous example. We again neglect the leakage flux
and the percentage error will be the same as in the previous example.
With the alloy Alnico 5, the energy product H,,,B,, is maximum
Iees
— Hole _ Bola
= Tis a bol m (9-100)
Le x Ome
= 4r X 1077 X 3.6 xX 10! = (0.29 meter. (9-101)
Re
Ar
i ESE I
SLD (9-104)
We have used primes on the dels on the right-hand side because they operate
at the source point (x’, y’, z’). Note the analogy with the corresponding
equation 4-184 for E. The above equation relates B, not to the current density
J; + V X M itself, but to its cur]. We again have ¢ to the first power in the
denominator, but a positive sign. Of course the integral can be calculated only
if the curl exists.
You will remember from Section 9.3 that, for a volume distribution of
charge,
B= #™ i Jee Vado A, r
(9-105)
Since the two integrals for B are equal, it follows (Problem 9-22) that
2
Te boKwR?2?. (9-111)
= 5: moKoR*t273 (9-115)
as before.
9.9 SUMMARY
We have shown that B can be calculated correctly, both inside and outside
magnetic material, by treating the equivalent currents as if they were real
conduction currents flowing in a vacuum. For example, the B field of a cyl-
inder uniformly magnetized in the direction of its axis of symmetry is identical
to that of a solenoid carrying a current density IN’ equal to M.
The magnetic field intensity H is related to B and to M as follows:
B
H=——M
=, " -
(9-36)
or
B= u(H + M). (9-37)
This equation is to be compared with
D=aE-+P, (9-38)
which applies to dielectrics.
In terms of H, Ampére’s circuital law is stated as follows:
VxXH=J (9-39)
or
fa = I, (9-41)
C;
where / is the current of free charges linking the curve C. This is a more
general form of Ampére’s law (Section 7-7), in that it is valid even in the
presence of magnetic materials, but it is still rigorously valid only for steady
currents.
It is the custom to write that
M = xnH, (9-42)
where Xm is the magnetic susceptibility. Then
B= pll + xn) = vowH = vw, (9-44)
where uy, is the relative permeability and p is the permeability.
In linear, isotropic, and homogeneous materials,
At the interface between two media | and 2, the boundary conditions for
B and H are as follows:
Bu a Bro, (9-61)
If the media are linear and isotropic, then the B and H vectors are parallel,
bothin medium | and in medium 2, and
tan, bn
(9-66)
tan OP) Mn,
where the angles are those between the B vectors and the normal as in
Figure 9-10a.
The concept of magnetic circuit is widely used whenever the magnetic
flux is guided mostly through magnetic material. We then have a magnetic
equivalent of Ohm’s law:
PROBLEMS
where n is a unit vector that is normal to the interface and that points from
medium 2 to medium 1.
9-7. Draw a parallel between the relative permeability 4, and the relative permit-
tivity €,.
(9-8. Imagine a parallel-plate capacitor whose plates are charged and insulated.
~ (a) How are Eand D affected by the introduction of a dielectric between the
plates?
(b) Show a polar molecule of the dielectric oriented in the field.
(c) Is its energy maximum or minimum?
(d) Now imagine a long solenoid carrying a fixed current.
How are B and H affected by the introduction of a cylinder of ferro-
magnetic material inside the solenoid?
(e) Show a small current loop representing the magnetic dipole moment of an
atom, and show the direction of the current on the loop.
(f) Is its energy maximum or minimum?
9-10. A long wire of radius a carries a current / and is surrounded by a long hollow
iron cylinder. The inner radius of the cylinder is b and the outer radius c.
418 MAGNETIC FIELDS III
(a) Compute the flux of B inside a section of the cylinder / meters long.
(b) Find the equivalent current density on the inner and outer iron surfaces,
and find the direction of the equivalent currents relative to the current in the
wire.
(c) Find the equivalent current density inside the iron.
(d) Find B at distances r > c from the wire.
How would this value be affected if the iron cylinder were removed?
9-11. It was mentioned in Section 9.6 that the hysteresis loop for a given sample of
material can be obtained by using a ring-shaped sample, as in Figure 9-7.
Show that, if the ratio R/L (resistance/inductance) for the circuit of
winding d is large, and if the current in winding a is not changed too rapidly,
A® = RQ/N,
where A® is the change in flux, Q is the charge that flows through the circuit,
and N is the number of turns in winding b.
9-12. The peaking strip is a device that is used for measuring B. It consists of a fine
wire of magnetic material, oriented in the direction of B, and centered on the
axis of a solenoid as in Figure 9-17. A small pick-up coil of a few thousand
turns is fixed to the center of the wire.
The peaking strip is operated as follows. If the solenoid carries a direct
current that just cancels the field one wishes to measure, plus a small alter-
nating current, then the H on the axis of the solenoid is that due to the alter-
nating current alone, and the peaking strip goes through a hysteresis loop at
every cycle. The electromotance induced in the pick-up coil is N d®/dr, where
N is the number of turns in the coil and ® is the magnetic flux in the strip.
Now the right- and left-hand sides of the hysteresis loop (Figure 9-8) are
nearly vertical in certain materials, such a molybdenum permalloy, and the
electromotance then has two sharp peaks, one positive and the other negative,
which can be observed on an oscilloscope. If the oscilloscope sweep is syn-
chronized with the alternating current in the solenoid, the two peaks are sym-
metrical when the time average of H on the axis of the solenoid is zero, or
when the dec field of the solenoid exactly cancels the measured field.
The peaking strip has a rather limited range of applications, first because
of the large size of the solenoid: for maximum sensitivity the length of the
wire should be much larger than its diameter, because end effects reduce the
value of B inside it, and the length of the solenoid must be several times that
of the wire, again to reduce end effects. In practice, the solenoid is 10 to 20
centimeters long. Also, because, of limitations in the power that can be dissi-
pated in the solenoid, the peaking strip can be used only up to fields of a few
hundred gauss. Finally, the field of the solenoid can alter the B in the pole
pieces and thus change their permeability locally.
Calculate the peak value of the electromotance induced in the pick-up
coil under the following conditions.
Peaking strip diameter, 2.5 < 10~5 meter;
Number of turns in pick-up coil, 1000;
Maximum value of dB/dH, 0.8 tesla-meter/ampere-turn;
419
ee
Solenoid
Peaking strip
Pick-up coil
Figure 9-17.
. If m is the unit vector normal to an interface, show that n-V x (A, — A,) = 0
and that n x (V x A,/p, — V x A,/p,) = 9.
. Show that both the normal and tangential components of the vector potential
A must be continuous across the interface between two media if the currents
are constant.
eee thin disk of iron of radius a and thickness ¢ is magnetized in the direction
parallel to its axis.
Calculate H and B on the axis, both inside and outside the iron.
9-16. A coil of 300 turns is wound on an iron ring (u, = 500) of 40 centimeters mean
diameter and 10 centimeters? in cross section.
(a) Calculate the magnetic flux in the ring when the current in the coil is one
ampere.
(b) Calculate the flux when there is a gap of 1.0 millimeter in the ring.
9-19, We have seen in Sections 6.9 and 7.4 that V- B is zero. This equation is valid
even in nonlineai, nonhomogeneous and nonisotropic media.
Show that it does not follow that V-H is always zero.
9-20. Calculate A,, and L,, for a magnetic circuit, similar to that of Figure 9-15,
which will produce a field of one tesla in an air gap 1.00 centimeter? in area
and 2.00 millimeters long. Assume a leakage factor of 50%.
and that lines of constant u are orthogonal to lines of B and thus to lines of
constant A.
(d) Hence show how a magnetic field can be simulated in an electrolytic tank,
using conducting electrodes to simulate the pole pieces. The pole pieces are
assumed to have infinite permeability.
Note that this method is not applicable if there are current-carrying con-
ductors or permanent magnets in the field.
(e) Show that
Vie — 08
The relaxation method described in Problem 2-16 can therefore also be used
for plotting magnetic fields.
Hints
9-3. The dipole m has two components, one that is radial and another that is perpen-
dicular to the diameter passing through m.
You should be able to show that, for a radial dipole mz,,
a cy sin’ @ de
Be Se eacra Sera
where S is the distance between the dipole and the center of the sphere, and @ = 0
at the position of the dipole. The integral on the right is equal to 4/3 for any value
of S>
Similarly, you can show that, for the other component my,
4r e. Melee COSa On—ma 057 -K) COS :
= [ Bar=am. f,Tl + (S/R? = 2(S/R) cos 8) 82 S18 8 48.
MAXWELL’S EQUATIONS
At this stage we have found all four of Maxwell’s equations in Chapter 6, but
we have found only three of them without using relativity. These were Eqs.
3-41, 8-8, and 9-29. Our first objective in this chapter is therefore to deduce
the fourth one, namely the equation for the curl of B, without using the
methods of relativity. Then we shall reexamine all four equations as a group
and deduce from them two general, and rather surprising, properties of elec-
tromagnetic fields.
The reasoning that will lead us to Maxwell’s equation for the curl of B is
unfortunately rather involved (see Figure 10-11), but we shall learn many
things on the way.
We assume, as uSual, that e, u, « are not functions of the time.
The following definitions will be useful
a surface S’ enclosing a volume 7’, the net free charge escaping through S$’
during a time interval Ar must be equal to the net free charge lost by 7’ during
this same time interval. Thus
[ va =— |Mar, (10-3)
and, since this equation is valid for any 7’,
0
V-J; = a (10-4)
and
Bh PP OP igre ae Dt, :
There is also a fourth type of current that can exist even in a vacuum and
that we shall use in Section 10.6. The corresponding current density is «0H/dt
and
Pe re Oi ee, (10-8)
since the divergence of E is equal to p,/é, from Section 3.6.
Then the total volume current density is
and
V-J,= 0. (10-10)
The second and fourth terms in J, are often grouped together, and
Oke OP Bye aD
eo + cra 3 (eo + P)= an (10-11)
We have seen in Sections 3.3.2 and 9.3 that, for a finite charge and current
distribution as in Figure 10-1, the scalar potential V and the vector potential
Va : 1 / Ps avs P iy! dt l |i op
+ 4 P-n yp. (10-16)
Tey | I Atrey | r
Ami | AM
V
ge, SEM aa’, (10-17)
where
The integrals are evaluated over volumes 7’ and surfaces S$’, which
enclose all charges and currents. If the charge or current distributions extend
to infinity, these integrals do not converge, as a rule.
426 MAXWELL’S EQUATIONS
We have also seen in Sections 3.7 and 9.5.1 that, in a linear, isotropic, and
homogeneous medium e, p,*
rob fmare
dt [esha | Pe
co
yh -Lfds re Ho
Bf ye M X La r (10-19)
i a
If you have studied Chapter 6, you will recall that we found in Sections
6.4 and 6.5 that, for an isolated charge passing through the origin at a constant
velocity © along the x-axis,
yQ :
12" (10-20)
~ Are fyX(x — U1)? + yp? + zy
The potentials V and A are related to the field vectors E and B by the
equations
hea Vig ““ (10-22)
=v ocd. (10-23)
These are Eqs. 6-90 and 6-77, or 8-13 and 7-41 respectively. In the first equa-
tion, —VV is that part of E that is due to accumulations of electric charge,
whereas —0A/0dt is the part that is due to changing magnetic fields. Once V
and A are known, it is merely a matter of differentiation to find E and B.
Therefore the six components of E and B can be deduced from only four
quantities, V and the three components of A.
* Our reasoning with respect to A will not be quite consistent. In Section 9.5.1 we used
the relation
Jy = wlJy+ VX M)
to transform Eq. 10-17 into 10-19. Now this relation is valid only for steady fiekds. The more
general expression for uJ; contains a time derivative on the right, and it can be deduced
from Eqs. 10-94 and 10-99. It is as follows:
The integrals of Eqs. 10-18 and 10-19 have one serious fault: they do not take
into account the finite velocity of propagation of electric and magnetic fields.
For example, if the charge density changes in one region, the integrals imply
that V changes simultaneously throughout all space. This is of course in-
admissible (Section 5.10). The potentials V and A at the time ¢ cannot corre-
spond exactly to the charge and current distributions at that particular
moment, unless the charges are fixed in space. The analogy with astronomy
is obvious: one cannot see a Star as it is now, but only as it was thousands or
millions of years ago.
If we have charge and current distributions p; and J; in a vacuum, then
Sea | eile
vege. ace (10-24)
— bo [ LWrle
ankle fo de,y :
(10-25)
where the brackets indicate that the charge and current distributions are
those at the previous time t — (r/c). These are the retarded potentials.
From now on, square brackets will be used exclusively to identify retarded
values of charge density, current density, position, velocity, etc.
The expressions for V and A in the presence of matter are simple in one
particular case, namely that of an infinite linear, homogeneous, and isotropic
medium e, u. Then the surface integrals are zero, and
y=_ 1df
{| bbe
ee,gy, :
(10-26)
where p, and J; are now calculated at the retarded time ¢ — (r/v), v being the
velocity of electromagnetic waves in the medium.
The integrals are evaluated over any volume 7’ that encloses all the
charges and currents.
tude of the time delay, on the nature of the phenomena involved, and on the
time resolution required. The following two examples illustrate the impor-
tance of retardation for extended pulsating sources.
QO = Qe. (10-28)
Then
p= OQve™'s = poet, (10-29)
where
Po = Qos. (10-30)
For example, the charges could be situated on a pair of spheres
joined by a thin wire of negligible resistance and capacitance.
The upward current J flowing through the connecting wire is
+Q
(b)
Figure 10-2. (a) An oscillating electric dipole. The total charge is zero and
the vector s is oriented from —Q to +Q as shown. (b) The charges and
the current are functions of the time.
429
P(r,0,@)
and
Ih = jwQo, (10-32)
hs = jwpo. (10-33)
Is = jwp. (10-34)
where the numerators are the charges, as they appear at the point
P(r, 6, ¢) of Figure 10-3 at the time f.
Note that the charges — Q and + Q give scalar potentials which
differ not only in amplitude but also in phase. The amplitudes differ
because of the terms r, and r, in the denominators, and the phases
differ because of the r,/c and r;,/c terms in the exponentials.
Similarly,
A el ie
ell I, exp jw
; (:engl)‘)S.Ves :
(10-36)
S
ry Rr — 5 COS 0, (10-38)
then
rp F yescos¢
wo(t-2)wo(s-24+ ae (10-39)
fe w(: = \+s
+ —5 00S 6, (10-40)
and
vale Qo exp jw (:— *) exp G5
3K cos a) exp (1%
x cos 0)
ae 1— 2r
= cosé te > cos 6
(10-41)
where A = )/27m is called the radian length. Expanding the expo-
nential functions and the denominators of the two terms within the
braces as power series, and neglecting terms of the third order and
higher in s/r and s/X, we obtain
Po &Xp Jw (- *) x
é = i)cos @, (10-42)
AtreorA
This is the sum of the dA’s due to us elements J dl around the loop,
all retarded by the ae time ss et be
— Rewriting,
Moaly exp AG
jw _ df "exp j
A= —— COS ae oie (10- 44)
431
and
poaly expjw (1 3) 2
c aed. ayes il
0 are Oia us
A —, sll =, = de :
d 4rr i lie Td X ¢ 1)) arg es
(10-47)
Also, from Eq. 7-106,
ie a ax
\ ~l- Ppt =| a COS Y, (10-48)
Mo EXP Jwa = “) x
A= iaSEES (3as i)sin 6 g, (10-50)
ped AE “
SDHn (+i) exp jo (1= *), (10-51)
~ Ar
where
m = 1S (10-52)
is the magnetic moment of the loop. The magnitude of the vector S
is the area za? of the loop, and its direction is related to the direction
of current flow by the right-hand screw rule.
For zero frequency, w = 0, X >, and
A= He menZS (10-53)
as in Section 7.8.
In view of the fact that p; and J;, which appear in the integrals for V and A in
Eqs. 10-26 and 10-27, are related together by Equation 10-4 for the conserva-
tion of free charge, we may suspect the existence of an equation linking
together the potentials V and A.
We have already found such a relation for the potentials of a single
charge moving at a constant velocity, in Section 6.7.
We shall now show that a similar relation exists between the potentials
of Eqs. 10-24 and 10-25. We shall use the wnretarded potentials V and A in
order to simplify the calculation. This means that we shall assume that the
time delay r/v is negligible.
We calculate the divergence of A:
ep
Usd 4 checm oh i eee
: dr’. :
(10-54)
Since the divergence of A is calculated at the field point P(x, y, z), while the
integral is evaluated over the source points P’(x’, py’, z’) as in Figure 10-1, the
order of the del and of the integral sign can be inverted and
V-A=htFi .
a : Ji
: dr’., :
(10-55)
Now
]
WW u = avs eV . (10-56)
10.3. THE LORENTZ CONDITION 433
The first term on the right is zero since J; is a function of ey ee and not of
the coordinates x, y, z of the field point that are involved in the operator V.
Also, from Problem 1-12, V(1/r) is equal to —V’(1/r), and thus
Ve Ft
of = JVpg! (10-57)
Similarly,
ies
Wi Ot Ie.1 (10-58)
Then, combining these last two equations, we find that
1
ae ut= hoy ce (10-59)
and
: — 1 Ye Je i? ML / J; !
V A = a [ ; dr A ae “ ae ° (10-60)
The second integral is similar to the one we started from (Eq. 10-55),
except that now the divergence of J;/r is computed at P’ instead of at P. The
second integral is zero for the following reason. According to the divergence
theorem, it is equal to the integral of (J;/r)-da over the surface S’ bounding
the volume 7’. Now because, by definition, all currents are contained within
the surface S’, everywhere on S’ the vector J; is either zero or tangential, and
(J;/r):da is zero. Then
lid Ve ‘Ji
V-A dr’.! (10-61)
An jer 7
Now we have had to use a prime on the del operator that appears under
the integral sign because it was essential for us to discriminate between
derivatives with respect to x, y, z and derivatives with respect to x’, y’, z’. But
the V’-J; term is really the same as the V- J; which appears in Eq. 10-4. In this
latter equation we were only concerned with the source point, and a prime on
the del would have been superfluous. Then, from Eq. 10-4, V’-J; is —0p;/dt
and
aera SideAr /J, CLICr i (10-62)
Now + is not a function of the time, since it is the distance between the two
fixed points P and P’. We can therefore remove the 0/d¢ operator from under
the integral sign and
V-A
. oy € ome
fae
SS pr=a f (10-63)
-63
OV
———— aay —— 9, (10-64 )
from Eq. 10-26.
434 MAXWELL’S EQUATIONS
The relation
[Baa si 1) (10-69)
over any closed surface S. If we choose S to be a spherical surface
concentric with the capacitor, then B-da is B,da. By symmetry, B,
must be independent of 6 and ¢, so that 4m7r?B, is zero. Therefore B,
is also zero. But, again by symmetry, By and B, must also be zero,
and thus B = 0.
Also, from Section 8.1.1,
OB
VXE= ark (10-70)
and, at a radius r,
Inet Re G 1 )
mE — (- — — }: 10-74
“é 4iro ee Wo) ( )
According to the Lorentz condition,
OV The“Re (1 1 )
V rey =_ — J ———— -———_— (O)9/
: “Hor *P 4raRC \r Ro MeUr ie)
or, setting the coefficient before the parenthesis on the right-hand
side equal to K,
Kote ie (; 1
2 Dp (r°A,) ~ K -= ——)>
R, (10-76 )
OMe r2
—ar (r°A,) — K(r ——)>
R. 0-77
(10-77)
a (5-3):
eh eae I TEN GE VE Ati)
Finally we must justify our assumption that
OA, chee ay eee
Ot RC? 40 \2 = are
is negligible compared to dV/dr in Equation 10-71. First, we note
that the above parenthesis is of the order of unity. Now, since
(eu)—'/2 & velocity of light in the dielectric (Section 11.4), RC/(eyu)!/?
is the distance traveled by light in one time constant RC. This,
squared, is vastly larger than the r? of Equation 10-71, under our
assumption that the dielectric is only slightly conducting.
It was in Section 3.6 that we found the first of our four Maxwell’s equations,
namely
(10-81)
where p, is the total charge density p; + p,. Although this result was arrived
at in discussing electrostatic fields, it is in fact general, as one can see from the
discussion in Section 6.8.
For a linear, homogeneous, and isotropic dielectric, we have Eq. 3-54,
VE="%, (10-82)
and
0A py
Vilarvpe
( V Sle
=) : g
(10-83)
Rewriting,
é 5 = 4 ts |
Se
Is (10-84)
10.6 THE CURL OF B 437
AI cat ee (10-85)
This is the nonhomogeneous wave equation for V. At points where
py = 0 we have the usual wave equation (Appendix E); the phase velocity is
1/(eu)"?.
If py is not a function of the time, neither is V and, in a homogeneous
linear and isotropic dielectric,
Vy ee Pt (10-86)
€ €
which is Eq. 3-55.
The solution of the nonhomogeneous wave equation 10-85 for an in-
finite medium e, u (not functions of x, y, z, f) is the retarded scalar potential
of Eq. 10-24 (Appendix E).
We can find the wave equation for A by analogy with that for V.
The vector equation 10-25 for A is equivalent to three scalar equations
for the three Cartesian components of A, each one of these equations being
similar to the single scalar equation for V. Then the three components of A
must satisfy three scalar differential equations similar to Eq. 10-85, and,
combining them together,
0°A
V2A — eu OP = —pJy. (10-87)
(Cie DHbeCUR OF Bb
We have already found the curl of B in Section 9.4, but that equation is
strictly valid only for steady currents. We have also found the curl of B in
438 MAXWELL’S EQUATIONS
Section 6.11, but we were not then familiar with the equivalent volume
current density V X M.
It will not be difficult to find the correct value of V X B, now that we
have the Lorentz condition and the wave equation for A.
In terms of the vector potential A,
VX B=VXVXA= —-V*%A+ V(V-A). (10-89)
Substituting now the value of V*A from the wave equation 10-87 and the
value of V-A from the Lorentz condition 10-65,
aD
Ve ere (10-94)
The term dD/dt is called the displacement current density (Section 10.1).
This is another of Maxwell’s equations. Let us rewrite it in terms of E and
B, and without the e and the p:
Or
1 OE
7) x B = af
me oe is Lod ms (10-99)
10.7 MAXWELL’S EQUATIONS 439
ab
or We
ai mara (10-103)
Let us group together the four Maxwell equations, which we found suc-
cessively as Eqs. 10-81, 7-34, 8-8, and 10-99:
440 MAXWELL’S EQUATIONS
J; = cE, (10-109)
te)
ai YB) = (0, (10-112)
The quantity V-B is therefore independent of the time at any point in space.
We can set the divergence of B equal to zero everywhere if we assume that, for
each point of space, it is equal to zero at some time, either in the past or in the
future. Under this assumption, Eq. 10-106 can be deduced from Eq. 10-107.
Equations 10-106 and 10-107 are sometimes called the first pair of Maxwell’s
equations.
Similarly, taking the divergence of Eq. 10-108,
oe = ae’a
v. OL Pode (10-113)
If we assume that there is conservation of charge, as we did in Section 10-1,
per
OK 1 Op,
hs, -114
Or €9 Ot al )
ay
0
apes~ (6) Pt
10-115
Or hy ) Ot €0 ( )
SE eng os (10-116)
where C is some quantity that can be a function of the coordinates, but that
is independent of the time. If we further assume that, at every point of space
at some time either in the past or in the future, both V- EFand p; are simultane-
ously equal to zero, then the constant of integration C must be zero, and we
are left with Eq. 10-105. Under these two assumptions, Eqs. 10-105 and 10-108
are therefore not independent. They form the second pair of Maxwell's
equations.
If we write out p, and J,, in full, Maxwell’s equations become
442 MAXWELL’S EQUATIONS
VX H- ee 27h, (10-124)
All three sets of equations (10-105 to 10-108, 10-117 to 10-120, and 10-121
to 10-124) are valid in nonhomogeneous, nonlinear, and nonisotropic media.
In the remaining chapters we shall be concerned mostly with fields where
E and H are sinusoidal functions of the time. In such cases we can replace the
time derivatives by factors of jw (Appendix D). Setting also D = eE and
B= wl,
a ete (10-125)
Via 20), (10-126)
TSC l= 0. (10-127)
UxCH jack =], (10-128)
therefore equal to 1/e times the net charge inside, as illustrated in Figure 10-6.
This is Gauss’s law (Sections 3.6 and 6.8).
Integrating Eq. 10-106 in a similar manner, we find that the outward flux
of B through any closed surface S is equal to zero (Sections 6.9 and 9.3):
i B-da = 0. (10-130)
S
or, if we use Stokes’s theorem on the left and invert the operations on the
right, the surface S being fixed in space,
fpeed
. =Clc 2a |Beda
. =
dd
= (10-132)
-132
Then the electromotance induced around curve C is equal to minus the rate
of change of the magnetic flux @ linking C, as in Figure 10-8. The positive
directions for B and around C are related according to the right-hand screw
rule (Section 8.1).
Finally, if we also integrate Eq. 10-108 over an area S bounded by a
Cunver@.
dE
f eat = wf (J.+ € ) \ da = jel. (10-133)
G s ot
If we perform the same operation on Eq. 10-124,
10.8 DUALITY
Let us consider one field where the electric field intensity is E and the
magnetic induction is B. Then the above equations are necessarily satisfied.
Let us now consider a different field E’, B’ such that
E’
—KB = —KpH, (10-139)
H’ I +KD = +KeE. (10-140)
Substituting E’ and B’ into Eqs . 10-121 to 10-124, and assuming that e and u
are independent of x, y, z, t, we find that Maxwell’s equations also apply to
the primed field:
V-D' =0, (10-141)
V-B’ =0, (10-142)
/
Vie Hi - = 0, (10-144)
Figure 10-10. Pair of symmetrical fields. Lines of E are solid and lines of
B are hatched.
446 MAXWELL’S EQUATIONS
Example If we have a purely electric field {E(x, y, z), B = 0}, then it is possi-
ble to produce a purely magnetic field {B’(x, y, z), E’ = 0}, where
B’ is the same function of x, y, z as E, within a constant factor. The
inverse is also true. Experimentally, of course, one of the two fields
can be more difficult to obtain than the other.
Example We have found in Section 2.9 that the E field of an electric dipole is
given by
je. __l
= awe 2p
i cos 6, (10-147)
Ey =Ju PRLS
lee sin¢ 6, (10-148)
E, = 0, (10-149)
as long as r is much larger than the length of the dipole. Then, in
Section 7.8, we showed that the B field of a magnetic dipole is given
by
B== dq
Be 2M
pe 08 6. 0, (10-150)
By Bs
= tq Ra
ps SD 6, %
(10-151)
By = 0, (10-152)
if r is again much larger than the dipole.
The analogy is obvious: these two fields are mathematically
identical, within a constant factor.
We again consider two distinct fields, E,, B,, and E,, By, which are not
necessarily related as in Section 10.8. According to the principle of super-
10.9 LORENTZ’S LEMMA 447
WC HB 3c BY gk (Jn+ Fey
P
x M,)
Ore
+ yok, - (J a5 air WS M,): (10-155)
ot
If, moreover, the medium is linear and isotropic, P is e9x.£, and the
above time derivatives also cancel. Finally, if the point considered is not
inside a source and if Ohm’s law applies, J; = cE and the J; terms cancel.
We are then left with the V X M terms. If the medium is nonmagnetic, then
M is zero and
Wik «Bhs XB) = 0 (10-156)
If the medium is magnetic, we can perform a similar calculation using
H’s instead of B’s and Eq. 10-124. This gives
Example If the a field is purely electric and the 4 field is purely magnetic, then
V-(E, X HM) = 0, (10-158)
subject only to the above limitations.
This surprising statement is easily shown to be correct by ex-
panding the divergence:
V(X Hy) = AW XE.) — EW X ii), (10-159)
OB, ODs\ _
= —H,: “ar E.- (Jn45 oe) =(0, (10-160)
In regions where p; and J; are both zero, we have the homogeneous wave
equation
VeE — ew * = 0. (10-169)
VE=V", (10-170)
Note that the velocity v of the E wave is (eu)!/? (see Appendix E), so that the
terms between brackets are taken at the time ¢ — (r/v), if E is calculated at
the time f. It will be shown in Problem 10-19 that the first term on the right is
—VYV, while the second is —dA/dt, so that Eis —VV — 0A/dt, as usual.
We can find the nonhomogeneous wave equation for B which corre-
sponds to Eq. 10-168 by multiplying Eq. 10-124 by uw and then taking its curl,
assuming that u is independent of the coordinates. Then
p= +T |
J}qf
Ly Elis (10-179)
for an infinite medium e, u. This is the curl of A, as will be shown in Problem
10-18.
See also Problem 7-8.
10.11 SUMMARY
Vey ee (10-4)
There is also conservation of bound charge, and, setting
oP
Jn tT Vv RM (10-6)
V-In = — = (10-7)
OE
he=Utont+2auxM, (10-9)
t Ot
)
= J; + - +VXM, (10-12)
and
VJ, = 0. (10-10)
Ey [i bee
V=7- a dr’, (10-26)
ge k /
A=H |esWikar, (10-27)
where [p;], and [J;], are the free charge and current densities at the source
point P(x’, y’, z’) at the time t — (r/v), r is the distance between P’ and
451
Retarded V Retarded A
Eq. 10-26 Eq. 10-27
oATE
Nonhomogeneous wave equation for A
Eq. 10-87
Figure 10-11.
the point P(x, y, z) where V and A are calculated at the time ft, and » =
1/(eu)"/? is the phase velocity of electromagnetic waves in the medium.
The Lorentz condition relates together the electromagnetic potentials V
and A: ~
VeAt eu - = 0. (10-65)
When Vis a function of the time, the six components of the field vectors E and
B can be obtained from the three components of A.
The nonhomogeneous wave equations for V and A in a homogeneous
medium e, » are as follows:
2V OV fe, pr
VV — ew WEES
ap ; ly
(0-85)
The reasoning that led us to the equation for the curl of B is shown
schematically in Figure 10-11. This is one of Maxwell’s equations:
1 OE
V x B eee C2
aegetend
at Lod ms (10-99 )
or
452 MAXWELL’S EQUATIONS
oD
Vier, Hise dla (10-94)
V-E == we (10-105)
€0
(10-106)
(10-107)
1 OE
VXB ria aae Hodm5 (10-108)
or
Vikan €0
(aE), (10-117)
V-B=0, (10-118)
OB
V xX E+ ae
—_
= (10-119)
1 OE oP
V x B Se
C2 Ot wo
(+o a +0 x u)-
M); (10-120)
Or
oD
V xX H cams
F FsPS
di (10-124)
or, for fields which are sinusoidal functions of the time in a medium e, Ml,
0:
ikE-d us
=»
(10-129)
E
ikB-da = 0, (10-130)
) Hdl d®
tees: (10-132)
PROBLEMS
\
| 10-. What is the relaxation time for a semiconductor having a rather low conduc-
tivity of about 100 mhos/meter and a rather large relative permittivity of 15?
454 MAXWELL’S EQUATIONS
10-2. ‘Two plane parallel electrodes are separated by a plate of thickness s whose
conductivity o varies linearly from ao near the positive plate to a) + a near
the negative plate.
(a) Calculate the space charge density p; when the current density is J;.
(b) Calculate p; near both plates for a = 1.00 * 10° mhos/meter, a) + a =
2.00 10’ mhos/meter, J; = 1.00 ampere/meter?, e, = 3, s = 1.00 centimeter.
10-3. Show that the nonhomogeneous wave equations for V and A are compatible
with the conservation of charge.
10-4. Calculate the magnetic induction inside and outside a large flat conducting
plate if the current density J decreases linearly with depth z inside the plate,
J being equal to Jo(1 — az). The plate thickness is 1/a.
10-5. Deduce Eq. 2-5 for the electric field intensity due to a point charge Q, start-
ing from Maxwell’s equations.
10-6. Write out Maxwell’s equations in terms of E and H only for a nonhomo-
geneous medium in which e, and yu, are functions of the coordinates.
BS VA:
where V and A are the scalar and vector potentials, respectively.
Rewrite Maxwell’s equations in terms of these potentials, for linear
homogeneous media.
10-8. At very high frequencies currents are limited to the region close to the surface
of a conductor and there is essentially zero electric and magnetic field in the
interior.
(a) Setting E, and B, to be the tangential electric field intensity and magnetic
induction in the conductor, show that
OE,
@z
_ _ 0B.Or
where the z-axis is normal to the surface of the conductor and points outward,
E, is taken to be positive in the direction of the x-axis, and B, in the direction
of the y-axis.
(b) Show that, outside the conductor, B is tangential to the surface, or very
nearly so.
(c) Show that B is related to the surface current density A by the equations
B= pA X n) and B= pod,
where nm is a unit vector normal to the surface of the conductor and pointing
outward.
(d) Does this result depend on the way in which the current varies with depth
inside the conductor?
(e) Can there be a normal component of E outside?
PROBLEMS 455
where
EX
v= ile
is a velocity perpendicular to B.
This means that the Lorentz force tends to make the component of the
velocity that is perpendicular to the magnetic lines of force equal to the ve-
locity v.
(c) Show that
OB 1
—
Ey = M6 Ee SE ap ——V-B:
456 MAXWELL’S EQUATIONS
In liquid metals 1/opo is of the order of unity. For example, mercury has
a conductivity of about 10° mhos/meter, and 1/cpo is about 10/47.
yea"tt —(4y)
(3 ey.
where the new terms are set between parentheses. The quantities V and A are
the usual scalar and vector potentials, which are related to E and B as in
Section 10.2:
E=-VV-—- ae
Ot
B=V XA.
The other two equations of Maxwell for V X E and V-B would remain un-
changed.
Lyttleton and Bondi suggested that the constant /, which has the dimen-
sions of a length, would be of the order of the radius of the Universe. The
new terms would therefore be negligible in all but cosmological problems.
(b) If these modified Maxwell equations were correct, would V and A be
measurable, in principle?
Remember that, with the above equations for E and B, it is only the rates
of change of V and A that affect E and B.
(c) Write out the equation for the conservation of the total charge (Section
10.1) in its modified form, assuming that charge is everywhere created at the
constant rate of g coulombs/meter*-second.
(d) Would the Lorentz condition (Section 10.3) still be valid?
(e) Now set A = A’r, where 4’ is a constant, and assume V to be a constant.
Show that
B= 0, E= 0, Jm = (q/3)r, pe = OV /P.
Assuming that the velocity of the outward flow of matter is the same as
that of the charge, namely J,,/p:, it follows that the radial velocity is propor-
tional to r, which is consistent with the linear velocity-distance relation ob-
served by astronomers: v = r/T, where T +3 X 10!” seconds is the Hubble
constant.
—_— ——— e
Pt oe?
Ee
oa (2 oi
(g) Show that, if this theory is correct, then Q ~1 hydrogen atom per
2 X 10!* meters?/second.
10-12. (a) Starting from an electromagnetic field characterized by the field vectors
E and H, utilize the duality property of electromagnetic fields four times to
find successively the fields FE’, H'; Et, H¥; Eiii, Hii: and Ev, Hiv,
(b) Assuming that E and H are respectively parallel to the x- and y-axes, show
the 10 vectors on 5 separate diagrams.
(c) Compare the fields when the constant of proportionality K is equal to the
velocity of light in the medium, that is, 1/(eu)!/*.
10-13. (a) Show that Maxwell’s equations for free space are invariant under the
transformation
E’ = Ecos6é+ cBsin 8,
B’ = —(E/c) sin@ + Bcos@.
The transformation E’ = —KB, H’ = KD of Section 10.8 and the trans-
formation E’ = —E, B’ = —B are special cases corresponding to 6 = 7/2
and to 6 = a respectively.
(b) Show that the energy density
=
1 1
2 e9H 2 + ea
oun B 2
and the Poynting vector FE < H for a plane electromagnetic wave (Section
11.1.1) are also invariant under this transformation.t
* Proc. [EEE 55, 1238 (1967). See also D. J. Epstein, Proc, IEEE, 56, 198 (1968).
458 MAXWELL’S EQUATIONS
10-16. Imagine various pairs of electromagnetic fields in free space, and verify
whether Lorentz’s lemma applies.{
10-17. In the example on page 434 we showed that the leakage current in a spherical
capacitor does not produce a magnetic field. There are two other ways of
demonstrating this.
(a) Show that V x A = 0.
(b) Use the fact that
pe | Cs dr’.
Gr }_,
SF
r
Hints
PROPAGATION OF
ELECTROMAGNETIC WAVES I
Plane Waves in Infinite Media
In this chapter we shall study the basic aspects of the propagation of electro-
magnetic waves in infinite media; reflection and refraction phenomena will
be the subject of the next chapter, after which we shall be prepared to study
guided waves in Chapter 13. We shall not deal with sources of electromag-
netic waves until Chapter 14.
Let us start the present discussion with propagation in free space, and
then take up three types of media: dielectrics, conductors, and ionized gases.
Dielectrics, which are defined as nonconductors, can be either magnetic or
nonmagnetic.
At this stage you would be well advised to work through Appendixes D
and E, unless you are familiar with the exponential notation for representing
the cosine function, and with wave propagation.
Maxwell’s equations impose no limit on the frequency of electromagnetic
waves. To date, the spectrum that has been investigated experimentally is
shown in Figure 11-1. It extends continuously from the long radio waves to
the very high energy gamma rays observed in cosmic radiation. In the former
the frequencies are about 104 hertz and the wavelengths about 3 X 10 meters;
in the latter the frequencies are of the order of 10% hertz (and higher), and
the wavelengths of the order of 3 & 10~' meter (and shorter). The known
spectrum thus covers a range of 20 or more orders of magnitude. Radio,
light, and heat waves, X-rays and gamma rays—all are electromagnetic,
although the sources and the detectors, as well as the modes of interaction
with matter, vary widely as the frequency changes by orders of magnitude.
The fundamental identity of all these types of waves is demonstrated by
460
!
Log, / (meters)
5 4 3 2 1 O —1 —2 -—3 —4 —5 —6 —7 —8 —9—10 -11—12-—13-14
-15-16
(theo edeee teeter heehee th eben See es ae Sp eS oe eee
|
|
Logi, f (hertz)
| eNO) AM A Sle aie ils UG IG, ile iS), BIN PAL Bek pie Os!
—— Gamma rays
~ Visible X-rays
light
Ultra
violet
ae
HESTON
mies
ty
as
BS
2 Infra
= red ee
Let us start with the relatively simple case of plane sinusoidal waves propa-
gating in a vacuum in a region infinitely remote from matter. Then Maxwell’s
11.1 PLANE ELECTROMAGNETIC WAVES IN FREE SPACE 461
These are two remarkable results. We have deduced from our investi-
gation of the basic electromagnetic phenomena: (a) the possibility of the
existence of electromagnetic waves and (b) the velocity of such waves in free
space.
The above expression for c is in itself remarkable. It links three basic
constants of electromagnetism: the velocity of an electromagnetic wave c,
the permittivity of free space ¢, which we first met in Section 2.1 while dis-
cussing Coulomb’s law, and the permeability of free space po, which enters
into the magnetic force law of Section 7.1.
It will be remembered from Sections 6.1 and 7.1 that the constant uo was
defined arbitrarily to be exactly 4r X 10~7 henry/meter. The constant e can
thus be deduced from the measured value for the velocity of electromagnetic
waves,
c = 2.9979 X 108 meters/second: (11-8)
The permittivity of free space ¢) can also be determined directly from measure-
ments involving electrostatic phenomena. The measurements lead to the above
value within experimental error, thereby confirming the theory.
For a plane electromagnetic wave propagating in the positive direction
of the z-axis, E is independent of x and y,
ae De OZ
(11-10)
462
Figure 11-3. The E and H vectors for a plane electromagnetic wave traveling in
the positive direction along the z-axis. (a) The variation of E and H with z ata
particular moment. The two vectors are in phase, but perpendicular to each other.
(b) The corresponding lines of force as seen when looking down on the xz-plane.
The lines represent the electric field. The dots represent magnetic lines of force
coming out of the paper, and the crosses represent magnetic lines of force going
into the paper. The vector E X H gives the direction of propagation.
The E and H vectors are perpendicular and oriented in such a way that
their vector product E X H points in the direction of propagation, as in
Figure 11-3. The E and H vectors are in phase, since E,/H, is real, and they
have the same relative magnitudes at all points at all times.
The electric and magnetic energy densities are in phase and are equal,
since
teok?
+ ie (11-18)
2
i
[ex
Lee -da =
de
0 f[ (cE? | mol?
2{ (2 + 5 )ar, -22
(11-22)
The integral on the right is the sum of the electric and magnetic energies,
according to Sections 2.15 and 8.5.1. The right-hand side is thus the energy
lost per unit time by the volume 7 and the left-hand side must be the total
outward flux of energy through the surface S bounding r.
The quantity
“S=EXH (11-23)
11.1 PLANE ELECTROMAGNETIC WAVES IN FREE SPACE 465
is cal
This is about ten times larger than the magnetic induction between
the pole pieces of a powerful permanent magnet.
* See The Feynman Lectures on Physics (Addison Wesley, Reading, Mass., 1964), Volume
II, Section 27-4 for an interesting discussion on energy density and energy flow in electro-
magnetic fields.
466
/ === —— /7
j\3- —/ gs
NGS eet wz ‘
ee
———— \ 7
owes SSS \ i
ee————
ae
Figure 11-5. The electromagnetic wave emanating from the horn sweeps
through the imaginary cylinder. The amplitude gradually decreases at
the source, and the intensity, represented by the Poynting vectors, is
larger on the right than on the left, as in Figure 11-6.
whereas, at a distance z,
ie(E X H)-da
Figure 11-6. Sine wave from a source whose amplitude decreases linearly
with time. The wave travels to the right.
ee a)
468 PROPAGATION OF ELECTROMAGNETIC WAVES I
CH oH
WH = a TCP i le EP cae 0. (11-40)
In both equations the second term on the left comes from the displacement
current, while the third comes from the conduction current. The two equa-
tions are identical, except for the fact that there is no magnetic equivalent of
the electric charge density p;. These differential equations describe an atten-
uated wave (see Eq. E-55 in Appendix E).
Let us first dispose of the charge density that appears in Eq. 11-39. We
consider again a plane wave propagating in the positive direction of the
z-axis. Then all derivatives with respect to x and to y are zero, so that
aE we et”
O Kate
toes On :
(11-41)
Then
Vv (Me 0 nt) ae 0° i -49
e Oz (2 I gz" Hak tt
and the wave equation for E becomes
a” ‘ ‘ 0° 0 : :
a7 (£.i + B,J) — ple ait +o a (44 + Ej + Bk) = 0. (11-43)
@E, dE,
€ pate ere Bt (11-44)
with the result that E,, if it exists, must be of the form
Substituting the values of E and H from Eqs. 11-50 and 11-52, we find that
—kh+ ww = 0. (11-56)
k= oe) (11-57)
is real and there i attenuation.
The phase velocity in nonconductors is therefore Jess than in free space and
th is
(11-59)
nee (11-60)
Note however that n and e, are both functions of the frequency. We have
discussed briefly the variation of ¢, with frequency in Sections 3.10 and 10.2.1.
Since tables of n are usually compiled at optical frequencies, whereas «, is
usually measured at much lower frequencies, such pairs of values cannot be
expected to correspond.
Also, from Eq. 11-53,
H, = ()"z (11-61)
11.4 PROPAGATION OF PLANE ELECTROMAGNETIC WAVES 471
and
1/2
H = Ayexp j(wt — kz)j = () Ey exp j(wt — kz)j. (11-62)
——_—_
eB = bat (11-63)
The total instantaneous energy density is thus eE? or »H?, and the average
total energy density is eEtms, Or uHims. The average value of the Poynting
vector is
] ] €\1/2 5
Sav = g) EoHy k = zu Eo k, (11-64)
€\12 1 A
— Mu Bens k —- (cu)! Chae k, a 1-65)
= ucE2,; k, (11-66)
1/2
= 2.66 X 10-3 (=) E>, k watts/meter?. (11-67)
The average value of the Poynting vector is equal to the phase velocity wu
multiplied by the average energy density.
Example If the 20 gigawatt laser beam of the example given at the end of Sec-
tion 11.1.1 passes through a glass whose index of refraction is 1.6,
then «!”? is 1.6, u, is unity, and E) is smaller than in air by a factor
of 1.61, or by 1.26. Then & in the glass is 1.7 < 10° volts/meter.
To calculate By we can use the fact that the Poynting vector
(3)E, Hp is the same in the glass as in the air; since E is smaller in
the glass by a factor of 1.26, By must be /arger by the same factor.
You can check that this makes the electric energy density equal
to the magnetic energy density (Eq. 11-63).
We must now solve the two wave equations 11-39 and 11-40 for a wave
traveling along the z-axis, with p, = 0. Substituting again Eqs. 11-50 and
11-52, we find that
where Xp = Ao/2m7 = C/w is the radian length for a wave of the same angular
frequency w propagating in free space.
The quantity jwe/o is the ratio of the displacement current density dD/dt
to the conduction current density J; = cE. We shall call the magnitude of
this ratio the © of the medium:
aD)
Ot
\aeE
Ot we Ge
SO A ae er (11-70)
For nonconductors, Q—~ «©. For common types of conductors, o is of the
order of 107 mhos/meter (5.8 X 107 for copper) and we can set e, & | (ex-
ample, page 439). The ratio Q is thus very small for the usual conductors. At
optical frequencies o is a function of frequency, as explained in the Example
on page 424.
Then
o _ Str f, J\T ;
= re (1 Z) (11-71)
1/2 1/4
ie = (ertir) 1+ - exp {—/j are tan (k;/k,)\. (11-75)
ie @
Our expression for the arc tan function is correct, since k, is a positive quantity
(Appendix E). In a vacuum, k, = 1/%o, ki = 0.
The real part k, of the wave number is 1/A = 27/), where ) is the wave-
length in the medium; the imaginary part k; is the reciprocal of the distance 6
over which the amplitude is attenuated by a factor of e. The quantity 6 = 1/k;,
is called the attenuation distance.
Again from Appendix E, the phase velocity is
u= , (11-76)
+ Note that k? is used here for the square of the complex number k, and not for the
square of its modulus.
11.4 PROPAGATION OF PLANE ELECTROMAGNETIC WAVES 473
where - :
k;
6 = arc tan k. (11-79)
is the phase of E with respect to H.
Also,
E= Eyexp {jt — k,z) — k,z} i, (11-80)
H = Hyexp {jot — k,z — 6) — kz} j, (11-81)
with
(11-82)
pe I
1 94
(11-83)
a ereee
5 HH (1+)
55(;
oieHee He )\exp
0 + 5 HH6 exp (a kzle
iZ) = See
4 Eo (Tee)
+ & Poe 2 k iZ).
(11-84)
Now, from Eqs. 10-123 and 10-124 for linear and isotropic media,
Integrating over a volume 7 and using the divergence theorem on the left-hand
side, and then changing signs,
fe x H)-da
again gives the total outward flow of electromagnetic energy per unit time
through the surface S, as in Section 11.1.1.
As a rule, both E and H are expressed as exponential functions of the
time and of the space coordinates. However, since we require the product of
two such quantities, we must revert to the cosine functions (Appendix D).
For example, if E is of the form
Ey exp j(wt — kz)
one must write instead
Ey exp (—k,z) cos (wt — k,z).
Re (i x H*), (11-90)
where H* is the complex conjugate of Ht, and where the operator Re means
“Real part of.”
Let us return to the values of k, and k; for conducting media as given in Eqs.
11-73 and 11-74. In good conductors Q is much smaller than unity and
Abe
{ 1 1/2 ie | | ic)2 1/2 11-99
a oe (11-102)
or when the conduction current density cE is at least 50 times larger than
the displacement current density 0D/dt.
Good conductors will be defined as those for which the above condition is
satisfied. One must remember here that o is a function of w at optical fre-
quencies. According to this definition, copper is a ‘“‘good’’ conductor up to
frequencies of about 2 X 10'* hertz, or to the ultraviolet.
For good conductors, Eq. 11-68 simplifies to
k2 = — joy, (11-103)
and
1/2
*T-TT
afqey
syJdaq
ee
eee
dANLIOY
ee
urys
u1ySeee
ee
sdojonpuoy)
AyIANONpuOD
© Ayyiqeewlag zi1t9
@yideq
09 Z1I9H ZOYOITY
[| ZoYyRsIPy
€ ZHOYRSID
Jojonpuod /oyw) (1939 all /s9}0UL1)
PUODaS
—(¢,, (sI9}OUNTJUIO)
© (SIsJOUNTTFTW)
$= (SIo}oUNTTTIUN)
= (suosorur
cIMnUTnyy PSE LOLS 00'T £800 Et Lic £800 yl
sseig g°¢9) ‘ND Z'PE (UZ 69'T X LOT 00'T 9710 €9'T 86 9710 0€C
unruo1y) =Beé= X LOI 00'T 180°0 OT OG 180°0 ot
teddo5 08'S X 101 00'T 990°0 S30 eg 990'0 Cal
PIOD OS'P X LOT 00'T ¢L0'0 L60 8E°C SL0°0 vl
amydein OT lx 00°T I 6S SOC €0S 6ST 0°67
SHeuse
uor OT TX 7 X 20T T10°0 vo se0 T10°0 020
Tejouny
SL) ‘INZ ‘ID
¢ ‘ND81 (94 910 X LOI Z X vOl 6200°0 Le0'0 c60°0 6c00°0 €S0°0
[93{9IN a a Me, rb T X 2Ol vt0'0 810 vv vlo0 97°0
BIS J3}8M OS~ 00'T 7 X 2Ol € X 201 L X s0OI 76 AUS 4
J9ATIS ST9 X LOL 00'T 790°0 €8°0 £0'C 790'0 cl
& an OL8°0
X LOT 00'T ILO ECE Tp’s ILT0 cle
OUT, 98'T X 101 00°T LITO IST OL'E LLCO VT
:dounos pojdepy
wo. IY} uvdiaaup ajnjijsuy
fo saisdyg YOoqpuny ‘TITH-MBIDO])
MON OA ‘(£961
« IV @ = 7000 "E[S2}
4} 1Vv stui Aouanbay
O sI ynoGe
‘Zz puk vas Ja}eM
SI JOUB ,,poos,, 1OJONPUOD
*) & ‘(OL
re
11.5 PROPAGATION OF PLANE ELECTROMAGNETIC WAVES 479
5eB Wwe 1
=—— = 0S = (11-113)
Lelie o 50
D) 0
and the energy is essentially all in the magnetic form. This results from the
large conductivity o, which causes E/J; to be small. The electric field intensity
is weak, but the current density, and hence H, are relatively large.
From Eggs. 11-90 and 11-106, the average value of the Poynting vector is
1
sa 5 Re (ESCH): (11-114)
(11-115)
2 1/2
§=1 = Goon xis)
ll 6.6 < 10~ meter, (11-117)
= 66 microns. (11-118)
This is about ten times smaller than the velocity of sound in copper,
which is 3.6 X 103 meters/second. Also,
| (* x 10° & 49 X 107-7\12
= 3.7 X 10-*ohm, (11-121)
H\ 58 10%
as against 1207, or 377 in free space (Eq. 11-16).
LG oka LS A ar ee
“lean Fon) st a ts
within the sheet. Then the average power lost by the wave in the
sheet is
1
5 Fell + e42/*)\q ~ Ey (1— 5) aw Eva (11-124)
and the resistance in the direction of current flow is a/(cb Az). Then
the Joule losses must dissipate
We shall use the general results of Section 11.2, but first we must find the
conductivity o of an ionized gas.
We shall assume that the gas pressure is low, so that collisions and energy
losses can be neglected. Under those conditions the conduction process in
ionized gases is very different from that in metals, because conduction elec-
trons in metals suffer large numbers of collisions with the crystal lattice.
We also neglect thermal agitation. This is equivalent to setting the
temperature T equal to zero.
Let us consider a plane electromagnetic wave traveling in the positive
direction along the z-axis with the E and H vectors respectively parallel to the
x- and the y-axes. An ion of charge Q, mass m, velocity u, situated at x, y, z,
is subjected to a Lorentz force as in Sections 6.1 and 7.3:
—= Yai dz
Gas —. Bat
IPCs dx
BRS
ORV) aeFay 12
(11-127)
Then
iuieg 4
8)
= i
dz
a nd
79
a 28
dt? m (E z ji) ele
E 1 dz
10 tie) ae
where we have used the free-space relationship E/B = c (Eq. 11-17). As we
shall see later on in Section 11.6.3, the ratio E/B is even larger in ionized
gases so that we are overestimating the magnetic force. Similarly,
482 PROPAGATION OF ELECTROMAGNETIC WAVES I
CV .
=p = 0, (11-130)
az E dx
Ae Oar (11-131)
Ghee
[ie Otaies m
<9}
E8 £08 wt. (11-133)
We must not use the exponential notation here, because we shall have to
solve Eq. 11-131, which involves the product of the two variables E and
dx/dt (see Appendix D).
Thus
x= ve E) cos wt. (11-134)
wm
We have neglected the constants of integration, which are related to the mean
position and velocity of the charge during one cycle. Also,
= Fe F sinut. (11-135)
Substituting in Eq. 11-131,
GF = SE in Dut (11-136)
Fe8 PB 5 2, (11-137)
noes eS sin wt. (11-138)
Before proceeding further, let us find the conditions under which the
inequality expressed in Eq. 11-132 is satisfied. We expect that the ion velocity
will remain low, either if the field is not intense, or if the frequency is so high
that the time available for acceleration is small. Let us calculate
eae 2 ke
c
(\dt ) aT
~ Ae22c?
(11-139)
11.6, PROPAGATION OF PLANE ELECTROMAGNETIC WAVES 483
We shall see later on in Section 11.6.3 that this is correct only at high fre-
quencies. Then
1 fdz
=Al = i)
:) S R630 Speedie
?? =f 4!
(11-142)
The velocity dz/dt is proportional to the average flux of energy Say, and
inversely proportional to the square of the frequency. This seems to indicate
that it is possible for (dz/dt)max to exceed the velocity of light c. This is
because we have neglected relativistic effects in our calculation. The ratio
(dz/dt)max/¢ is usually much smalller than unity in practice.
oh = Jy. (11-143)
Now
and
FS ; 2
g=—-t> (11-146)
w i
Since the masses ofthe ions are larger than the electron mass by several orders
of magnitude, whereas their charges are at most a few times larger than that
of the electron, we can retain only the term corresponding to the electrons and
l . OMe
Z! =" =i xpi nel (11-149)
11.6 PROPAGATION OF PLANE ELECTROMAGNETIC WAVES 485
where
le GE (11-150)
Figure 11-9. The displacement current density 0D/dt and the convection
current density J; in an ionized gas are plotted here as functions of the
circular frequency w. The total current density is inductive below wp,
zero at w,, and capacitive above wp.
486
I = jweS(0/s) cxR22
NeGs (uy)
is) (11-152)
1
= (jaeit.Jol’s
v, (11-153)
where V is the applied voltage. The capacitor therefore behaves as
if were connected in parallel with an inductance L’s/S as in Fig-
ure 11-10.
If an alternating voltage is applied to the circuit, the current
through C and that through sL’/S both remain finite but become
equal and opposite in sign at the resonant frequency.
is called the plasma angular frequency. This quantity depends solely on the
properties of the gas considered. It corresponds to a frequency
y
' € (Zp)
Zo agi Pee — — =. ae i)
|
Zo ae (Zo),
Zo + dz + (20 + dz),
or to
488 PROPAGATION OF ELECTROMAGNETIC WAVES I
Zp + dz + §(Z) + Edi
NaS he (11-158)
d¢
Nene
dz
Assuming now that the displacement ¢ is everywhere small and that it varies
smoothly with z, we can set
d¢
eeeae (11-159)
and then
Nefie= N.( ay
=) (11-160)
Then the net charge density is that due to the ions, minus that due to the
electrons, or
where Q., the magnitude of the electronic charge, is positive, and where N;
is the number of ions, either positive or negative, of a given type. The net
charge density p; is positive if d¢/dz is positive, because the electrons are
then spread out.
From the Maxwell equation 10-105, this net charge density gives rise to
an electric field in the direction of the z-axis and
dE, N.Q.dt
reer (11-162)
E, = NQe gs (11-163)
€0
ad Nod NO:peg
lear (11-165)
This is the equation for an undamped simple harmonic vibration of
angular frequency
which is the value of w, given in Eq. 11-156. Each electron can therefore
execute a simple harmonic vibration about its initial position.
We have therefore calculated the plasma angular frequency w, in two
different ways. We first considered the motion of the electrons under the
action of an applied alternating electric field and found that for one particular
value of w, namely w,, the relation between the applied electric field E and the
resulting electron motion was such that the total current was equal to zero.
We then considered the motion of the electrons under the action of the elec-
tric field created by their own nonuniform displacement and found that they
could oscillate about their initial position with the same angular frequency w».
We can also calculate w, in the following manner. If we assume that the
electron density can oscillate at some angular frequency w,, the resulting elec-
tric field intensity E will oscillate at the same frequency. Then we can use
Eq. 11-147 for the conductivity o of the ionized gas. Now Eq. 10-15 for the
free charge density py assumes only the conservation of charge and Ohm’s law
J; = cE. Substituting the value of o and setting e, = 1,
p= psespys
NQ* i (11-167)
WyINe€o
as previously.
(1 xe\"
k=+ ie ; (11-169)
= rt
=)"
é (11-170)
For a wave propagating in the positive direction of the z-axis and with its E
vector parallel to the x-axis,
E = E,exp jot — kz)i (11-171)
where k is positive. From the general relationship for E/H, Eq. 11-53,
/2 2 Wh
(eee (2) i (“*) Wen at ie ete
wu wo) | ap
When w > w,, the wave number &k is real and the E and H vectors
are in phase, the ratio E/H being larger than in free space by a factor of
1/{1 — (wp/w)?} 1/2. As w approaches w», H tends to zero as expected, since
the total current density J; + (0D/dt) tends to zero. The Poynting vector
S also tends to zero.
The index of refraction is
ee ; 212
Hie (2yNhs (11-175)
mers
At high frequencies where w > w,, the phase velocity u is therefore greater
than the velocity of light, the wave number k is real, and there is no attenu-
ation. Figure 11-12 shows both n and 1/n as functions of w/w,. Since the phase
velocity increases with increasing ion density, waves tend to bend away from
regions of high ion density.
For high frequencies where w >> w;, the transmission is hardly affected
by the presence of ionized gas.
Note that the phase velocity is the velocity at which a given phase is
propagated. This is not the velocity at which a signal can be transmitted.
The reason for this is that a signal can be transmitted only if the wave is
modulated in some way, for example by varying its amplitude at an acoustic
491
For w < w,, the wave number k of Eq. 11-170 is imaginary. The vectors E
and H are then out of phase by 7/2 radian.
* Léon Brillouin, Wave Propagation and Group Velocity (Academic Press, New York,
1960).
492 PROPAGATION OF ELECTROMAGNETIC WAVES I
The ratio |E/H|is smaller than in free space for w < (w,/21/”) and larger at
higher frequencies, from Eq. 11-172.
The average Poynting vector 8,, = (1/2)Re(E X H™*) is zero, and there
is no energy transmission.
The index of refraction and the phase velocity are also imaginary, and
E = Eyexp (jot — k’z), (11-176)
H = Hy exp (jot — k’2), (11-177)
where k’ = jk is a real number. There is no wave, since the phases of E and
of H are independent of z, and the amplitude decreases exponentially with z.
An electromagnetic wave is therefore either transmitted without attenu-
ation, or not transmitted at all through an ionized gas, depending on the ratio
w/w,. High frequencies are transmitted, whereas low frequencies are not. Our
theory, however, is much simplified. In particular, we have neglected energy
losses, which is justifiable only at low gas pressures.
about 10!° to 10!?/meter® from the lowest to the highest layer. Over
this range of altitudes the number of molecules per cubic meter varies
from about 102 to 10!°. The percent ionization therefore increases
very rapidly with altitude, but it always remains low. For N, =
10!!/meter’, the plasma frequency f, ~ 3 megahertz.
Frequencies lower than f, are not transmitted. At radio fre-
quencies, waves are bent back toward the Earth, since the electron
density increases with increasing height, and since the phase velocity
increases with increasing electron density.
The assumption that there are no collisions between the elec-
trons and the gas atoms or molecules is not satisfactory in the
lowest regions of the ionosphere, where the pressure is highest, at
frequencies of the order of 1 megahertz or lower.
In the presence of the Earth’s magnetic field the ionized gas be-
comes doubly refracting, with the result that there are two distinct
phase velocities, depending on whether the E vector of the wave is
parallel or perpendicular to the B of the Earth.
11.7 SUMMARY
WE + equwk = 0, (T-3)
V7 + euo’H = 0. (11-6)
These are the wave equations for E and H in free space. They show that the
phase velocity of such waves is
l
oS ae (11-7)
(opto)?
Plane electromagnetic waves in free space are transverse. Their E and H
vectors are orthogonal and oriented so that E X H points in the direction of
propagation. The magnitudes of E and H are related to give equal electric
and magnetic energy densities:
H €0
The quantity
s=-EXH (11-23)
is called the Poynting vector. It is expressed in watts/meter’ and, when its
normal component is integrated over a closed surface, it gives the electro-
494 PROPAGATION OF ELECTROMAGNETIC WAVES I
magnetic power flowing out of the surface. In the case of a plane wave, §$ is
equal to the energy density multiplied by the phase velocity.
In any homogeneous, isotropic, linear, and stationary medium, a plane
electromagnetic wave has the following characteristics:
(11-70)
is the magnitude of the ratio of the displacement current density to the con-
duction current density. The wave number k is complex (Eq. 11-75), and E
leads H (Eq. 11-78). The average value of the Poynting vector is then given by
1
Ss 50° (11-102)
Then
k=
we
5? (11-104)
where
beth
= - (=) s)\ (11-105)
is the attenuation distance. The attenuation is so rapid that the wave is hardly
discernible (Figure 11-7). The E vector leads the H vector by 7/4 radian, and
PROBLEMS 495
E 2
a os (““) esn/4h (11-106)
o
N02
Ces om (11-147)
where WN, is the number of electrons per cubic meter, Q, is the charge of the
electron, and m, is its mass. The total current density is less than in free space:
oD ‘ ON
Ji = op t= jon {1— (2) be, (11-155)
where
2\ 1/2
Wp = (72) (11-156)
PROBLEMS
*11-1. Show that the wave equations for E and B is free space are invariant under a
Lorentz transformation, but not under a Galilean transformation.
Dep = Cs te
2 8719
or
eB _ (ie a
m c
where the mass m is the mass measured in the laboratory system, and where
dz/dt is the longitudinal velocity of the particle.
*(d) You can now use the result you obtained under (b) above to show that,
if there is resonance at ¢ = 0, then resonance will be maintained indefinitely
afterward. The frequency of the wave, measured at the source, is assumed to
be constant.
This result is most interesting. As the velocity of the particle increases,
its mass increases and its cyclotron frequency e®/m decreases. However, the
axial velocity also increases, and the resulting Doppler effect gives a frequency
shift which exactly compensates for the increase in mass.
498 PROPAGATION OF ELECTROMAGNETIC WAVES I
(e) It is quite difficult to find the energy & as a function of z, but you can do
part of the calculation.
First, you can show—if you have not already done so under (b)—that
(S) + V8) = 0,
where
wo Eo 36 — &
VG) a2 each
for a particle which is initially at rest (69 = mc”). Since & > &, V is always
negative, and & increases indefinitely.
It is useful to sketch a curve of V as a function of & to understand quali-
tatively the behavior of d&/dt.
(f) We have found d&/dt in (e) above, and dz/dt in (c); you can show,
finally, that
& — & 1 3 mez"
&o i 21/38 ce
11-9. A plane elliptically polarized wave results from the superposition of two
plane-polarized waves whose E vectors are oriented in perpendicular direc-
tions and out of phase.
For example, E can have the two components
+ C. S. Roberts and S. J. Buchsbaum, Phys. Rey. 135, A383 (1964). See also H. R. Jory
and A. W. Trivelpiece, Jour. Appl. Phys. 39, 3053 (1968).
PROBLEMS 499
Show that the E and H vectors for an elliptically polarized wave are
orthogonal if g = 0 or if the wave impedance E/H is real.
11-10. Show that S,, for a plane wave is equal to the average energy density
multiplied by the phase velocity, in any homogeneous, isotropic, linear, and
stationary medium.
11-11. In the general case of a uniform plane wave in a nonconductor,
E = (Eost + Eqyj + Evck) exp j(wt — krx — kyy — k.z),
where the coefficients Foz, Eo,, Eo:, kz, ky, kz can be complex, but are inde-
pendent of the coordinates x, y, z and of the time ¢.
The vector H is given by a similar expression with the same values of
Kaw Kayeukese
(a) Show that
B+ B+i=h
where k is the wave number corresponding to the medium and the frequency
under consideration.
(b) Discuss the relative orientations of E, H, and k.
11-12. Discuss the propagation of a plane electromagnetic wave in a medium
where (a) Q = 1, (b) Q? > 1.
11-14. It was shown in the example on page 424 that the charge density p in a
conductor decreases exponentially with a relaxation time of €/c:
p = poexp (—at/e)
Show that © must be at most about 3 if p/po is to be less than 1% within
one fourth of a period.
11-15. Using the values of E and of H that we have found for plane electromagnetic
waves in good conductors, verify that all of Maxwell’s equations apply.
11-16. A plane electromagnetic wave propagates in a good conductor in the positive
direction of the z-axis.
Calculate the total power lost per square meter by Joule heating between
z = 0 and z— », and show that this is equal to S,, at z = 0.
11-17. (a) Show that the real part n, of the index of refraction n, the attenuation distance
6 and the Q ofa good conductor are related by the following equations:
03? = 2€ é
po®
Q _ &br
- 6? 2h?
Qn = €H,
Oho
11-18. Compare the ratios |E/H| for an electromagnetic wave in air and in sea
water at 20 kilohertz and at 20 megahertz.
T = Ty exp Vv( - S 5 Al
where
mea wpe
1/2
11-22. Show that there is zero energy dissipation in a medium when E and J; are
out of phase by 7/2 radians.
11-23. (a) A mass m slides on a frictionless horizontal surface under the action of
a spring of stiffness (stretching force/elongation) & fixed to a rigid support,
and of a force F = Fy cos wt in line with the spring.
Show that this mechanical system is mathematically equivalent to a
series-resonant electric circuit.
Discuss the amplitudes of the displacement and of the velocity as func-
tions of w.
(b) A mass m slides on a horizontal frictionless surface under the action of a
force F = Fy cos wt, which is applied to it through a spring of stiffness k.
Show that this mechanical system is mathematically equivalent to a
parallel-resonant circuit.
Discuss the amplitudes and the velocities of the two ends of the spring
as functions of w.
11-25. Two plane electromagnetic waves of equal amplitude propagate in the iono-
sphere where the electron density is N. electrons/meter*. One wave has a
circular frequency w; and a corresponding wavelength \,; the other has a
slightly different circular frequency w, and a wavelength d..
(a) At a given time ¢ there exist values of z for which the two waves are
in phase and values of z for which they are opposite in phase. What is the
distance between the maxima?
(b) What is their velocity?
This velocity is called the group velocity Uy.
(c) Show that, in the limit,
ee gq dk
dw»
(d) Show that
Uplt, = Cc’,
11-26. Pulsars, or pulsating radio stars, emit sharp bursts of radio energy about 5
to 50 milliseconds wide at intervals of about one second. For any given pulsar
the repetition frequency is stable within about one part in 10°. The amplitude
and shape of the pulses vary widely, but each pulsar has its own characteristic
mean pulse profile.
Within a few months after the discovery of pulsars, distance estimates
were obtained in the following manner. It was observed that the arrival time
of a pulse depends on the frequency of observation, the arrival time being
later at lower frequencies. This delay is attributed to dispersion in the inter-
stellar medium which is ionized hydrogen with an electror. “ensity V, ~ 10°
per cubic meter.
(a) Show that, if w? > w,, a plot of the time delay Ar as a function of
LS peed) Tell
lO G Ay
is a straight line whose slope is a measure of the distance to the pulsar.
(b) In the case of pulsar CP 0328, arrival times measured at 151, 408, and 610
megahertz gave the results shown below.
i At
Megahertz Seconds
151
4.18
408
0.367
610
Show that, according to these measurements, the distance to CP 0328 is
268 parsec. (The parsec is 3.086 X 10!° meters. It is the distance from which
the radius of the Earth’s orbit, 1.495 < 10!! meters, would subtend an angle
of one second).
The fact that such plots give straight lines passing through the origin
indicates that the assumption w? > w? is correct. The delay therefore occurs
over large distances in a low-density plasma such as that of interstellar space,
and not inside the pulsar itself, which is quite small.
11-27. The solar wind is a high-conductivity plasma which is emitted radially from
the surface of the Sun. Let us calculate the flux of electromagnetic energy in
the solar wind at the orbit of the Earth.
In the plane of the Earth’s orbit, the magnetic field of the Sun is approxi-
mately radial, pointing outwards in certain regions and inwards in others.
This field is ‘‘frozen” in the high-conductivity plasma. Since the Sun rotates
(with a period of 27 days), and the plasma has a radial velocity, the lines of B
are in fact Archimedes spirals and, at the Earth, they form an angle of about
45° with the Sun-Earth direction. This is the so-called garden-hose effect.
At the orbit of the Earth the solar wind has a density of about 107 proton
masses/meter® and a velocity of about 4 X 10° meters/second, while the
magnetic field of the Sun is about 5 < 10-9 tesla.
PROBLEMS 503
(a) First show that, in an electrically neutral (p: = 0) and nonmagnetic fluid
of conductivity o and velocity v, Maxwell’s equations become
V-E=0, REO,
OB
Me ae VX B= wlolE +0 xB) +o Hh
(c) Show that the Poynting vector for the solar wind is
5
S= 2 Uv, & 6 microwatts/meter?.
This is about 4 < 107° times the average value of the Poynting vector of the
solar radiation, which is about 1.4 kilowatts/meter?.
The Poynting vector of the solar wind is normal to the local B and it
points at an angle 45° away from the Sun-Earth direction.
(d) You can show that the kinetic, magnetic, and electric energy densities are
related as follows:
Ux > Uu > Uz.
Hints
dp _ hae
in dv se
_ Oe PO)
ae ev: E.
dt
CHAPTER l2
PROPAGATION OF
ELECTROMAGNETIC WAVES II
Reflection and Refraction
Interface
where 1 is the phase velocity of the wave in medium 1. (See Appendix E.)
The time ¢ = 0 and the origin r = 0 can be chosen arbitrarily. This equation
defines a plane wave for all values of ¢ and for all values of r, that is, a wave
that extends throughout all time and all space. However, it is used only in
medium 1.
Both the reflected and the refracted waves from a plane interface are then
also plane and plane-polarized, since the laws of reflection and of refraction
for any given incident ray must be the same at all points on the interface.
Then the reflected and the transmitted waves are of the form
where u is the phase velocity of the wave in medium 2. Note that we have
made no assumption whatever as to the amplitudes, phases, frequencies, or
directions of the reflected and transmitted waves. The amplitudes Eo, and Eo,
can be complex to take phase differences into account.
‘We shall choose the origin at some convenient poi
(= = =) -rp = 0. (12-9)
Uy U2
Hence the vector between parentheses must also be normal to the interface,
so that n;, m,, m are coplanar and all four vectors n;, n,, n,, nm are in the
plane of incidence. Moreover, the tangential components of n;/im and of n;/u
must be equal, and
sin@; _ sin 6,
7A a (12-10)
nterface
wy
%%
%,.
v
© 3.
9,a)
Thus, choosing the origin in the interface and axes as in Figure 12-3,
We shall turn now to the third condition mentioned in the preceding section.
We must find the relations between the quantities Eo:, Eo, Eo. that will ensure
continuity of the tangential components of FE and of H at the interface.
We recall from Section 11-2 that the E and H vectors in a plane electro-
magnetic wave are always perpendicular to the direction of propagation and
to each other. The E vector of the incident wave can thus be oriented in any
direction perpendicular to the vector nj.
It will be convenient to divide the discussion into two parts. We shall
consider successively incident waves polarized with their EF vectors normal,
and then parallel to the plane of incidence. Any incident wave can be sepa-
rated into two such components.
The E and the H vectors of the incident wave are oriented as in Figure 12-4.
If the media are isotropic, as we assumed at the beginning of this chapter, the
509
E vectors of both the reflected and the transmitted waves will also be perpen-
dicular to the plane of incidence, as in the figure.
Considering the electric and magnetic field intensities of the incident wave
to be known, we have four unknowns: Eo,, Eo:, Ho, Ho. We also have four
equations that are provided (a) by the continuity of the tangential components
of E and H at the interface and (b) by the relations between E and H for plane
waves in medium | and in medium 2 as given in Eq. 11-53. It will therefore
suffice to
at any given time and at any given point on the interface. Similarly, the
continuity of the tangential component of H requires that
Hy; cos 6; — Ho, cos 0; = Hy, cos 6; (12-18)
or, from Eq. 11-53,
Thus, recalling that k = n/Xo, where n is the index of refraction and Xp is the
free-space wavelength divided by 2r,
510 ROMAGNETIC WAVES II
ue cos 6; — we COs 4;
Part gene ee Rd al O88 (12-20)
n Nye
=~ 608.0; -—— cos 6;
Mri Lr2
2 ak cos 6;
= ——_*t _,
hy 1/0)
(12-21)
— cos 6; + — cos @,;
In this case the E vectors of all three waves must be in the plane of incidence
as in Figure 12-5. We have chosen the orientations of Ey, and of Eo; so that
Figures 12-4 and 12-5 become identical at normal incidence, except for a
rotation of 90° around the normal to the interface.
We now have
Hoi — Hor = Hor (12-22)
or
Myo Ay
— — cos6; + — cos8;
E T2é
a
= — cos 6; + —' cos 6;
Mro Mrt
Ny
2 — cos 6;
(=) =—_ _. (12-26)
ae ® Cos 6, + we COS6;
nterface
mel
as in Eq. 12-1
(#). S oe (12-29)
Eo; =)cos 6; + cos
2
for a wave polarized with its E vector normal to the plane of incidence.
512
Interface Interface
0) ]
Ce
xe
aeBe
\by ie
\
\
n,
ly (a)
Air
Glass
Glass
if | , Air
(b)
513
Figure 12-6. (a) The relative phases, at the interface, of the electric field intensities
in the reflected and transmitted waves for n. > n, and for n. < m with Ey;
normal to the plane of incidence. In the first case the reflected wave is 7 radians
out of phase with the incident wave. The transmitted wave is in phase in both
cases. (b) “‘Crests”’ of E corresponding to (a) at some particular time. They are
spaced one wavelength apart and travel in the directions shown. Note the phase
shift of 7 on reflection from a glass surface. Note also the interference pattern
resulting from the superposition of the incident and reflected waves. Constructive
interference occurs wherever two crests or two troughs meet; destructive inter-
ference occurs where a crest meets a trough. (c) Graphs of E;, E,, E, at normal
incidence on an air-glass interface. Where the wave emerges from glass into air,
as on the right, EZ; is larger than E;. Conservation of energy still applies, however.
See Section 12.3.2.
It will be observed that (E,/Ep:)w is always real and positive. This means
that at the interface the transmitted wave is always in phase with the incident
wave. The ratio (£p,/Eo:)w can, however, be either positive or negative, depend-
ing on the value of m/m, for if m/no > 1, then 6, > 6; and cos 6; > cos 4;
whereas if n/n. < 1, then 6, < 6; and cos 6; < cos 6. The reflected wave is
thus either in phase with the incident wave at the interface if m, > ny or is
radians out of phase if m < ny. Figure 12-6 illustrates the E vectors for both
types of reflection; Figure 12-7 shows the ratios of Eqs. 12-28 and 12-29 for
n/n, = 1/1.5. This corresponds, for example, to a light wave incident in air
on a glass with an index of refraction of 1.5.
For an incident wave polarized with its E vector parallel to the plane of
incidence,
514
= 110
ny
—cos 6; + (2) cos 6;
(=) ee eee eee (12-30)
a cos 6; + (“) cos 6;
ny
2{—)cos@;
=) (“:) .
i
—)
" = ——o“—-—:
cos 6; + (“')COS 6;
Seal
12-31
Ng
The second ratio is always positive. This indicates that the relative phases
of Eo, and Eo; are as in Figure 12-5, which we used in arriving at this result;
that is, the incident and transmitted electric field intensities are in phase at
the interface.
On the other hand, the ratio for Ey, can be either positive or negative,
which indicates that Ey, can point either as in Figure 12-5 or in the opposite
direction. The tangential components of Eo; and of Eo, can thus be either in
phase or w radians out of phase. The Ey, component is in phase with Ep; at
the interface if
Eo, é
—0.4 Eo; ; z||
Figure 12-8. The ratios (E,,/E,;)p and (E,,/E,;)p as functions
of the angle of incidence 0; for n,/n, = 1/1.5. This corre-
sponds to light incident in air on a glass with n = 1.5. The
wave is polarized with its E vector parallel to the plane of
incidence.
or if
sin 6, cos @; — sin 6; cos 6; > O,. (12-33)
sin 20, — sin 26; > 0, (12-34)
sin (0; — 6;) cos (@, + 6;) > 0. (12-35)
This inequality will be satisfied either if
6, > 0; and b,x 5 (12-36)
or if
Oe < 0; and oe p> * (12-37)
The phase of the reflected wave in this case does not therefore depend only on
the ratio n2/n,; it depends on both 6; and 6;. The ratio Ey,/E), can be either
positive or negative, both for np > m, and for np. < nm. Figure 12-8 shows the
ratios of Eqs. 12-30 and 12-31, again for m/n, = 1/1.5.
Z nterface nterface
ly ly
Figure 12-9. When the incident wave is polarized with its E vector parallel to the
plane of incidence, there is no reflected wave at 6; + 6, = 1/2. The angle of
incidence 6; is then called the Brewster angle. The position of the missing reflected
ray is at 90° to the transmitted ray. For any pair of media, the sum of the two
angles 6;2 is 90°.
6;+6, = a (12-39)
there is a reflected wave only when the incident wave is polarized with its E
vector normal to the plane of incidence. This is rather remarkable because
it involves the passage of a wave through a discontinuity in the medium of
propagation without the production of a reflected wave. The conditions of
continuity at the interface are then satisfied by two waves only—the incident
and the transmitted waves—instead of the usual three. This is illustrated in
Figure 12-9. This angle of incidence is called the Brewster angle. It is also
called the polarizing angle, since an unpolarized wave incident on an interface
at this angle is reflected as a polarized wave with its E vector normal to the
plane of incidence.*
* The Brewster angle is often explained incorrectly as follows. For this particular angle
of incidence, the missing reflected ray is at 90° to the transmitted ray. It is argued that the
electrons excited in medium 2 do not radiate in their direction of oscillation (Section
14.1.4) and hence cannot give rise to a reflected ray in medium 1 in this case. This explana-
tion is incorrect, since the Brewster angle is observed even when medium 2 is a vacuum.
12.3. THE INTERFACE BETWEEN TWO NONCONDUCTORS 517
2 \uo
1 fe\” 2
Stay =5 (— Eo. m (12-43)
2 \Ho
518
1
8; av ba 7
ee
€r2 2 EX COS 6,
; 3
Siay'n - (©) E2, cos 6; (12-45)
__ Ny Ep: Cos 6; 3
~ ny, E2, cos 0; (12-46)
4 (2)cos 6; cos 0;
Ty = 4,
— (12-48)
Ny
Rp = i cos 6; + (2
*)e080)
leads to the apparently absurd result that sin @, is greater than unity. The
critical angle of incidence, for which sin @, = 1 and @, = 90°, is
sin 0. = 2. (12-52)
ny
It is observed experimentally that, when 6; > 0,., the wave originating in
medium | and incident on the interface is totally reflected back into medium 1
as in Figure 12-13. This phenomenon, which is called total reflection, does not
depend on the orientation of the E vector in the incident wave. For light
propagating in glass with an index of refraction of 1.6, the critical angle of
incidence is 38.7°.
It turns out, as we shall see in Section 12.4.1, that Snell’s law (Eq. 12-13),
the laws of reflection and refraction (Section 12.1), and Fresnel’s equations
(12-20, 12-21, and 12-25, 12-26) are all applicable to total reflection if we
disregard the fact that sin @, > 1 and if we set
cos @,= — (1 — sin? 6)”, (12-53)
s {1 (“) sin? 6.) (12-54)
oe a ny 2 oe iby ;
=Sei
—2cs{sin®o
BP me (*)
6.) pl ae (12-55)
2 COS 6;
Eo
pace, (12-60)
(=),: Pea cos 6; — j {int6; — (2) }
ny
ol
We first notice that the amplitude of the reflected wave is equal to that
of the incident wave, so that the coefficient of reflection R is equal to unity.
The energy is totally reflected and there is zero net flux of energy through the
interface.
The phase jump upon reflection varies from 0° at the critical angle of
incidence (sin 6; = m2/m) to 180° at glancing incidence, through positive
angles. This is shown in Figure 12-14.
I Oe
=
l. tn 2.0
As regards the transmitted wave, it is obvious that Ep; is not zero, despite
the fact that the net flux of energy across the interface is zero. Medium 2 can
be considered to act like a pure inductance fed by a source of alternating
current. The average power input to the inductance is zero, the power flow
being alternately one way and then the other, but there is nevertheless a
current through the inductance.
Figure 12-15 shows how the electric field intensity of the transmitted
wave varies both in amplitude and in phase with the angle of incidence.
The transmitted wave is quite remarkable. According to Eq. 12-58,
it travels unattenuated (as in Figure 12-16) parallel to the interface, with a
wavelength
Mo | AM
msin6; sind;
(12-62)
A; being the wavelength in medium | above the interface. The wavelength \,
is exactly the distance along the x-axis between two neighboring equiphase
points in the incident wave. This was to be expected, since the continuity
conditions must be satisfied at all points on the interface.
This result, namely that the wave travels unattenuated parallel to the
interface, is most surprising if we think of an incident wave of finite cross-
523
j : >
t: H§ { : Air
Figure 12-16. “‘Crests’’ of E for the incident, reflected, and transmitted waves are
represented here schematically for the case of total reflection. They are spaced one
wavelength apart. The transmitted wave travels unattenuated below the interface,
and its amplitude decreases exponentially with depth in medium 2. The data
used for the figure are the following: nm, = 3.0, m. = 1.0, 6; = 75°. (See Problem
12-12.)
section. Does the transmitted wave on the other side of the interface extend
beyond the illuminated region? Our discussion cannot provide us with an
answer, since it is based on the assumption that the incident wave is infinite
in extent. Physically, what happens is this: a given incident ray, instead of
being reflected abruptly at the interface, penetrates into medium 2, where it
is bent back into medium 1. It is this phenomenon which gives rise to the
“transmitted” wave.*
The transmitted wave is damped exponentially in the direction perpen-
dicular to the interface in such a way that its amplitude decreases by a factor
of e over a distance
Ao
= 2 1/2" (12-63)
{(2) sin’é; — 1}
2
* See A. von Hippel, Dielectrics and Waves (Wiley, New York, 1954), p. 54, for a brief
account of work by F. Goos and H. Hanchen on this subject.
524
lon T
Nn
— =
4 | i (eee i | 1
40 50 60 70 80 90°
i
0.85 4
0.7
0.6
S05
q
@ 0.4
*See, for example, J. Strong, Concepts of Classical Optics (W. H. Freeman and Com-
pany, San Francisco, 1958), Appendix J by G. F. Hull, Jr., p. 516.
12.4 TOTAL REFLECTION AT AN INTERFACE 525
coefficient of reflection R is again equal to unity, but the phase jump upon
reflection is different. Total reflection of a wave polarized in some arbitrary
direction therefore gives an elliptically polarized reflected wave.
F
1 / 2nR sin RAD _ y _ cos 6,,).
Bic 7
(12-64)
~ Oy 0 TR
pee
Reflecting
coating ~7}
Electron
beam
20 ie =
Total reflection
Critical angle
Fa
(SE
ETS
Ve
LS
Ta
LESS
Fag
!
Brewster angle
2 3 4 SO8: SAU
n)/ny
Figure 12-20. The critical angle and the Brewster angle as functions of
the ratio m/n, of the indexes of refraction on either side of the interface.
The wave is incident in medium 1. It is polarized with its E vector
parallel to the plane of incidence for the Brewster angle curve.
527
We have set the derivative with respect to y equal to zero because, by hypoth-
esis, the field does not vary with the y-coordinate (Figure 12-3).
We can use Eq. 12-14 for the incident wave:
E; = Ey; exp j{wt — k(x sin 6; — zcos6;)}. (12-66)
where E,,, kiz, kiz are unknown constants. We have used the same value of w
as for the incident wave since, as in Section 12.1, all three waves have the
same frequency. We have also used only x and z terms in the exponent since
the derivative with respect to y must, again, be zero.
We set
E, = (Borat + EowyJ + Eouk) exp j(wt a KoxX = ky,2), (12-68)
where Eoiz, Eoiy, Eotzs Kor, koz are also unknown constants. Again we have the
same w, and we have no y term in the exponent. We let the unknowns be
complex in order that the various components of E, and of E, can be out of
phase with each other and in order that the dependence on x and on z can be
more general than with a simpler expression, such as that for the incident
wave.
All the unknowns are independent both of the coordinates and of the
time, the only dependence on x, z, ¢ being expressed by the exponential
functions.
We represent the H vectors by similar expressions, in which the letter E
is replaced by the letter H and, for simplicity, we choose the origin of coor-
dinates at the interface, as in Figure 12-3.
We shall determine all these unknowns by using the boundary conditions
at the interface, the wave equation, and Maxwell’s equations.
12.4.1.1 The Wave Numbers ki, and k,, for the Reflected Wave. To satisfy
the boundary conditions at the interface, we proceed just as we did in Section
12.1. We first require that the exponents be equal at the interface z = 0.
Then, from Eqs. 12-66 and 12-67,
ee ee ()> (12-70)
and
ee
ae cos 6; 75
(12-72)
We choose the positive sign in order to have a wave propagating in the proper
direction.
The reflected wave is thus of the form
E, = Eo, exp j{wt — k,(x sin 6; + z cos 6,). (12-73)
This is a plane wave reflected from the interface at an angle equal to the angle
of incidence.
12.4.1.2 The Wave Numbers k», and ky, for the Transmitted Wave. Equating
similarly the exponents for E; and for E,, again at the interface z = 0, we find
the wave number k», of the transmitted wave:
ni
x,0 sin 8. (12-74)
e
Kos = ky sin d; =
To find k»,, we again use the wave equation. Using the expression for E,,
12.4 TOTAL REFLECTION 529
2 1/2
exp j{ot— 2xsinoi} +2 te sin?
6; — lp z}-
Xo Xo Ng
We have replaced the + sign before the z term by a + sign since the electric
field intensity must not become infinite as z > — «©. Then ky, can be written
with a plus sign:
2 1/2
kan=+524 (BY sint
oy—1b (12-79)
If @, were real, we would expect to have
Ny
ko, = — — COS 4. (12-80)
Xo
Comparing these two equations, we set*
2 1/2
COS 0; a {1= (™) sin? 6 : (12-81)
Nog
* Tt is probably useful to recall here that, if A is some positive real number, then
(—A)¥2 = jA'2,
and
ZAP = —j(— A)¥2,
These equations do not apply if A is negative. If you have any doubt concerning such
operations you should plot A, — A, and their square roots in the complex plane. It is rather
easy to be misled; for example,
nterface_
We can dispose of Hp, immediately because the relation between Ho, and
E,, is that which applies to plane waves in dielectrics (Eq. 11-61).
We can assume, as in Section 12.2.1, that the vectors Eo, and Ep, for the
reflected and transmitted waves are normal to the plane of incidence as in
Figure 12-21. Then
E; = Ep; exp j(wt = kyn;r) j, (12-83)
We have written the exponents in the first two equations in a more compact
form than previously. The wave numbers k2, and ks, are now known (Eqs.
12-74 and 12-79).
For the incident and reflected waves, H is in the plane of incidence, just
as in Figure 12-4, and it has both x- and z-components:
H; ll= Ho(cos 6,t + sin 0;k) exp j(wt — kin;-r), (12-86)
H, = Ho(— cos 6,i + sin 6;k) exp j(wt — kyn,-r). (12-87)
For the transmitted wave, however, we must use a more general expres-
sion, since we know nothing as yet concerning its magnetic field intensity.
We therefore set
0 = Hoty, (12-90)
We are now left with four unknowns, E),, Eo:, Hotz, Ho.z, and only three
equations. We therefore turn to Maxwell’s equations. We choose one of the
simpler ones and apply it to the transmitted wave. Since
V-H, = 0, (12-94)
then
KozHoiz + kezAoi. = 0. (12-95)
Solving and substituting the values of ky, and of k»,, we find Eqs. 12-59 and
12-60, and
2
H 1/9 2JM COS 8; (sin?6; — (2) le
( v2) Ltt (*) al (12-96)
Ey: Lo eta ( ney. vil Ny i
N
COSIU any) ‘arg 0; A
(#)
Aon
2 (2)
€9\1
any sin 26;
aon (12-97)
COSI7 aay, eat ts = (™) i
To obtain E,, the value of Ey, must be substituted into Eq. 12-84. Simi-
larly, to obtain E,, H,,, H.,, the values of Ey,, Hoiz, Ho.z must be multiplied by
the exponential function of Section 12.4.1.2 for the transmitted wave.
Since both denominators on the right are identical, whereas one numer-
ator is real and the other imaginary, the two components of I are 7/2 radians
out of phase, and, at any given point, it has neither a fixed amplitude nor a
fixed direction but rotates in the plane of incidence. The transmitted wave is
thus rather complex.
i Pak
(Sav = 5Re (KE. X Ht) = 5Rel0 0 |, (12-98)
He OE;
X exp {ieSe 0
sil 2
and where the stars identify complex conjugates (—/ substituted for /).
Then the x-component of Say is
:Re (EH) = (2)1/2 ny cos 6;. sinQ] 26; mes Ny (2) 2 seo 1}1/2 _
Mo _ (mY Ro (\ia
ny
(12-101)
and the z-component is zero.
The energy flow in the transmitted wave was discussed near the end of
Section 12.4.
Note that we require a positive sign before the square root, whereas we had a
negative sign in the corresponding equation 12-54 for total reflection.
The second term between the braces is complex, because of 1, but neg-
ligible since n, >> m (Section 11.5). Then
4 ke Oe Z Zz
exp {i(wt X,x sin 0; 4- :)te ‘I.
The fact that the first ratio is negative means that the E vector of the reflected
wave is in the direction opposite to that shown in Figure 12-4, which was used
for the calculation. As to the second ratio, it is necessarily quite small since
Eo. = Eo: — Eo; and Eo, & Ep;. Also, the E in the conductor can be expected
to be small. The E vectors are shown in Figure 12-22.
Reflection from the surface of a dielectric with mn, >> m would also give
Ey,/ Eo; — 1 and a weak transmitted wave.
It is interesting to note that there is a small loss of intensity on reflection
from a good conductor, Ey, being somewhat smaller than Eo;. You will re-
member that, with total reflection, there is no loss of intensity and R is equal
to unity (Eq. 12-59).
534
Interface
Figure 12-22. The incident, reflected, and transmitted waves at the interface
between a dielectric and a good conductor. The incident wave is in the dielectric
and is polarized with its E vector normal to the plane of incidence. The H vector
of the transmitted wave is parallel to the interface, as shown, but it lags the E
vector by 7/4 radian.
The coefficient of reflection R is approximately equal to unity; the electric
fields of the incident and reflected waves nearly cancel on the surface, and a weak,
highly attenuated wave penetrates perpendicularly into the conductor.
Interface
Figure 12-23. Reflection and refraction at the interface between a dielectric and a
good conductor. The incident wave is in the dielectric and is polarized with its E
vector parallel to the plane of incidence. The E vector of the transmitted wave is
as shown, but it leads the H vector by 7/4 radian. The coefficient of reflection R
is approximately unity, the tangential components of the E vectors of the incident
and reflected waves nearly cancel on the interface, and a weak attenuated wave
penetrates perpendicularly into the conductor.
12.5 INTERFACE BETWEEN A DIELECTRIC AND A GOOD CONDUCTOR 535
=)La
(= tonal (2)
ais (12-108)
The tangential components of Ey; and of Ey, nearly cancel at the surface of the
conductor, as expected, and the transmitted wave is again a weak, highly
attenuated plane wave which penetrates perpendicularly into the conductor.
H, = Hy: exp {i
(wt+ ;)+ 5p (12-111)
with
€0 1/2
iva (<) Ev: (12-112)
Ho
BOT,
< “Eo Ho
Nn ng
Interface
vi
\,/ Figure 12-24. Reflection at normal
incidence from the surface of a good
conductor. The electric fields of the
incident and reflected waves cancel
at the interface, or nearly so. The
magnetic field intensities add,
however, with the result that the
amplitude of the Ho, vector at the
y interface is 2Ho;.
537
Ey Eotr EXP iu
Li(wt = ) - 3} (12-113)
€) \!/2 ; Zz
Ay (2) Four EXP J (or+- Z) (12-116)
Ho Xo
or
= Ex
a-)% ?(E,— Ev) (12-119)
538
eae
ee exp | Eo
at aliek
+A) p> (12-120
Eu
ae _ 2eer
2 exp {i _ 444
el TD , (12-121)
1 €0 1/2 >
3; ay = Sl3 iin ESOi> (12-125)
=
1
Stay = 5 Re (E; X Hh), (12-126)
1/2
Stay = 23/2 (<2) eS. 4, (12-127)
Just below the first interface, at z = OM ona aoe eSIIICE
(weo/o)!/? & QV? and © is small (Section 11.5 and the example on
page 479). Then, inside the sheet, the Poynting vector is further
reduced by a factor of e~2/°, On the far side of the sheet, from Eqs.
12-121 and 12-110,
1 1/2
Sit avy — 2 () Ee, (12-128)
16 78
o
e-raldg, (12-129)
For the wave which is reflected four times, we add the term
exp j——
= oe epee TE (12-136)
* sinh +a +a
6
and
Siitvay ss >(=)"
1 /
ae
i
ie
eee | em . (12-137)
Ls(<2)
2
aA
a + es
4 WE,
o
5,5, (12-139)
sinh? ricos? hel + cosh? 7 sin? R
5
1 29107 ys (12-140)
which is larger than when multiple reflections are neglected. This is
of course true only for the particular case we have chosen.
with axes as in Figure 12-27. We have not taken the trouble to show that E, is
of the form shown above; the demonstration is identical to that which led to
Eq. 12-73 for total reflection. We have set all three circular frequencies equal,
as previously.
12.5.1.1 The Wave Numbers ko, and ko, for Refraction into a Good Conductor.
We can again equate the exponential functions for E; and E, on the interface,
just as we did in Sections 12.1 and 12.4.1.2. Then
ny
(es [ea in Oe = x, sin 6;, (12-144)
as for total reflection.
We can find k,, from the wave equation for good conductors,
PE:
Ox:
,aE,
02?
= joo. (12-145)
The derivative with respect to y is zero since the wave is assumed to be inde-
pendent of the y-coordinate. Then
ky = =i = - (12-147)
0
Thus
kee 1/2
hap = oe he (1- iH) ; (12-148)
2
== + va
hi (2)
a are
sin a (12-1499)
I = — ky cos. (12-152)
Comparing Eqs. 12-149 (with a negative sign before k2) and 12-152, we see
that we were justified in writing cos @, as in Eq. 12-104.
€ 1/2
(=) n,(Eo; + #,,) sin 0; = H,,., (12-156)
0
We have now discussed at some length the reflection and the refraction of an
electromagnetic wave at the interface between two dielectrics, and then at the
interface between a dielectric and a good conductor. We also wish to investi-
gate the same phenomena for the case of a dielectric and an ionized gas.
Before going on with this, however, we shall study another phenomenon
that is related to reflection from conductors, namely radiation pressure.
Let us consider an electromagnetic wave incident on a good conductor.
We limit ourselves to normal incidence. We have seen above that the electric
field intensity transmitted into the metal is small compared to that of the
incident wave, but that it is not zero. We have also seen that the FE and H
vectors of the transmitted wave inside the conductor are orthogonal, just as
in the incident wave.
According to Ohm’s law, the electric field intensity E, in the conductor
gives a current density cE,, where o is the conductivity of the medium. This
current, which is oriented like E,, is perpendicular to H;,. It turns out, as we
shall see presently, that under these conditions the conduction electrons are
pushed by the magnetic force Qu X B in the direction of propagation of the
wave. The electrons in turn push on the conductor in the process of colliding
with the atoms in their path, giving rise to radiation pressure.
The situation would be very different if the conduction electrons were
entirely free to move through the metal: the conductivity would then be
544
A nterface
Figure 12-28. The conduction
current cE, and the magnetic
induction pol which give rise to
radiation pressure in an element
ae of volume of a reflector. The
ice sides of the element of volume
Ca are parallel to the E and H vec-
dz tors. The element has a thickness
ee dz and is ab square meters in
Z ly area.
imaginary, as in the low density gases that we studied in Section 11.6.1, and
the radiation pressure would be zero.
Let us study the radiation pressure of a wave incident normally on a
good conductor. We consider an element of volume that is parallel to the
interface, of thickness dz as in Figure 12-28, and ab meter? in area. It carries
a current cEb dz, and is submitted to a magnetic force cEy)Hab dz in the
negative direction of the z-axis.
Note that the relative permeability u, of the conductor does not enter
into the calculation. It has been shown that the magnetic force exerted on an
electron within a magnetic material is proportional to u)H and not to uH, for
slow electrons like the ones we are considering.*
Then the pressure exerted by the wave on the element of thickness dz is
dP = oE oH dz. (12-159)
E, — 2Hnmd
“=p Eo;pn exp f
iu(wt + 5:)4 23)” (12-160)
2 p95
E, = aK, Eye? cos (or+ ;+ i) (12-162)
Si av
Py = 2 wri (12-170)
where 8S; ay is the average of the absolute value of the Poynting vector for the
incident wave (Section 11.3).
We can ascribe this pressure to a change in momentum of 2 8,,/c per
unit time and per unit area in the incident wave, the factor 2 being required
because the wave is reflected with a momentum equal to its initial momentum,
but of opposite sign.
546 PROPAGATION OF ELECTROMAGNETIC WAVES II
we ‘= {1_ (*) } .
2)1/
(12-175)
1/
{1— 80,5 x _ (12-176)
from Eqs. 11-173 and 11-174, where c is the velocity of light, u is the phase
velocity of the wave, w, is the plasma angular frequency of the wave, N, is the
number of free electrons per cubic meter, and f is the frequency of the wave.
If the ionized gas had a definite boundary and a uniform value of N,
throughout its volume, reflection and refraction at its surface would be simple
to describe: the ionized gas would simply act as a dielectric with n < 1. As
a rule, neither assumption is valid and the wave 1s reflected in much the same
way that a light wave is reflected in a mirage.
It is possible to calculate the path of a ray by performing a numerical
integration of the ray equation stated in Problem 12-2. However, one can
deduce the main features of the reflection by simply using Snell’s law.
We select coordinates as in Figure 12-29 and assume that the index of
refraction n varies slowly with z, but not with the other two coordinates
x and y. To be more specific, we assume that n varies by a negligible amount
over one wavelength. If N, gradually increases with z, a given ray gradually
bends down as in Figure 12-29 to an angle 6, at a point where the index of
refraction is 7.
We can calculate @ in the following way. When refraction occurs at the
interface between any two media m and mm, the quantity n sin @ is conserved
in going from one side of the interface to the other. This is Snell’s law, Eq.
12-13. If the index of refraction varies continuously, the medium can be
imagined to be stratified in infinitely thin layers, and n sin @ is similarly con-
served all along the ray. Thus
If the ray penetrates into an ionized region where the ion density increases
with z, the index of refraction n decreases with / and the derivative dn/d] is
negative, so that the angle @ increases with distance as in Figure 12-29. After
some distance, if n becomes sufficiently large, @ becomes equal to 90°. At that
point the tangent of @ becomes infinite, but dn/d/ becomes zero. After this,
tan 6 becomes negative, whereas the derivative dn/d/ becomes positive, and @
keeps increasing until the ray escapes from the ionized region at an angle equal
to the angle of incidence 6;.
At the top of the trajectory,
sin 6 I 1, (12-180)
Ngoc = Sin 6;. (12-181)
This is the index of refraction required for reflection when the angle of inci-
dence is 6;.
12.8 SUMMARY 549
sec 6; = —, (12-182)
Wp
where w is the angular frequency of the wave and w, is the plasma angular
frequency at the value of z where 6 = 90°. For normal incidence, 6; = 0,
sec 6; = 1, and reflection occurs at a z sec wp» = w, that is, where n is zero
and the phase velocity is infinite. At oblique incidence, however, sec 6; > 1,
and reflection occurs where wy) < w, or at a lower value of z, if we assume that
the electron density N., and hence w,, increase with z.
We have found above the value of dé/d/. This is the reciprocal of the
radius of curvature R, and thus
1 _ 1dn
R = - dl tan @. (12-183)
1 1 dn
(ae eile (12-184)
12.8 SUMMARY
The laws of reflection, Snell’s law, and Fresnel’s equations result from the
continuity of the tangential components of EF and of H at the interface between
two media.
The Jaws of reflection are the following: (a) the angle of reflection is
equal to the angle of incidence, (b) the normals to the wavefronts of the
incident and reflected waves lie in a plane that also contains the normal to the
interface and that is called the plane of incidence.
According to Snell’s law,
SE
sin@;
te
Ne
(12-13)
where 6, and 6; are the angles of incidence and of refraction respectively and
where 7, is the index of refraction of the first medium and np is that of the
second medium.
Fresnel’s equations relate the amplitudes and phases of the reflected and
550 PROPAGATION OF ELECTROMAGNETIC WAVES II
E 2 eis cos 6;
(?) = Se (12-21)
OT = 60s Ga —— COsU:
Mri Lr2
ey eee » 603 8,
Ew = Br2 Mri 5 (12-25)
Evi] p fz Ay
— cos 6; + — cos #;
br2 bri
E 2Dee cos 6;
(*) Be ER
oea (12-26)
1 = 63 6; — cos @,
ae Mri
my
= COUUsze (12-40)
9
fe (WOoM2
Z iy (12-102)
The Fresnel equations again apply if we set
* deren Ani m\2 . , \ 12
cos #; = +(1 — sin? 6,2 = a = (egypt aa » (12-104)
PROBLEMS
12-1. In a plane sinusoidal wave, the components of EF and H are of the form
W = Wyexp/(wt — k-n),
= W exp j(wt — krx — kyy — k.z),
where W, is a constant that can be complex. See Problem 11-11.
552 PROPAGATION OF ELECTROMAGNETIC WAVES II
The vector k, which is called the wave number (not to be confused with
the unit vector along the z-axis), can also be complex. Then we can write that
i an, — jbny,
where mn, and ny are unit vectors, with k, = am: — jbno,, and similarly for
k, and k,.
(a) Show that surfaces of constant phase are normal to n,; and that surfaces
of constant amplitude are normal to no.
(b) Show that a is related to the wavelength, and 4 to the attenuation.
(c) Show that, in all operations involving V, this operator can be replaced
by —jk.
(d) Rewrite Maxwell’s equations utilizing this fact.
(e) It does not follow that E, H, k are orthogonal when p, = 0 and ¢,, y, are con-
stants. To show this, use Maxwell’s equations, with W expressed as a cosine
function, and then find the conditions for orthogonality.
(a) Show that the angle @ between a ray and the y-axis is given by:
G0 me ee
ds
where the distance s is measured along the ray.
(b) You can now verify the ray equation
d
7 ot) = Vn,
where t is a unit vector tangent to the ray at a point where the index of refrac-
tion is n.{
. Show that
tis 2uriNs COS 6;
seed) = .
Ao; N Mrolly COS 0; i Mrill2 COs 6,
12-6. Show that there is no reflection from a p/ate of material illuminated at the
Brewster angle when the E vector of the incident wave lies in the plane of
incidence.
PROBLEMS 553
12-7. (a) Show that, for a wave incident in air on a nonconducting magnetic
medium, (£,/Ep;)p is zero for
tan? 6; = slereH
Err — 1
and hence that the Brewster angle exists only if €, > p,.
(b) Show that, similarly, (£o,/Eo;)y is zero for
where
A* = 1 — a(a + 2) tan? 6,,
(c) Draw wave fronts for all three waves in the neighborhood of the interface,
showing the phase shifts found above.
Draw parallel lines spaced 27 apart to represent the wave fronts.
(d) Check the continuity of E across the interface.
(e) Compare your results with Figure 12-16.
12-14. In the case of the total reflection of a wave polarized with its E vector normal
to the plane of incidence, we have found in Eqs. 12-96 and 12-97 that the x-
and z-components of the H vector in the transmitted wave are out of phase
by 7/2 radians.
In what direction does the H vector rotate with respect to the direction
of propagation of the transmitted wave: does it rotate like the wheels of a
vehicle moving on a road or does it rotate in the opposite direction?
(a) Calculate the fraction F as a function of = assuming that all the light is
transmitted for angles of incidence smaller than the critical angle and that
the scintillator is surrounded by a nonreflecting substance.
(b) Draw a graph of F for values of 1)/n, ranging from 0.1 to 1.0.
12-16. Construct sin (a + jb) and cos (a + jb), where a and b are both real, in the
complex plane.
Over what ranges of values can these two functions vary ?
6
Eor
Eo a
. ~ 1 — = cos 6;,
a COS
12-18. (a) Calculate (Eo,/Eo:)p for reflection from the surface of a good conductor.
(b) Show that the reflected wave leads the incident wave by the angle
PROBLEMS -
555
2 6
Qe nye aly
Bri Xo
12-20. Draw two figures similar to the left-hand one in Figure 12-6c, showing E and
H for an electromagnetic wave incident on a good conductor.
You will of course have to exaggerate the values of Ey, and of \ in the
conductor.
Be sure to show the phases correctly. Show x, y, z axes on both figures
so as to relate one to the other.
12-21. We have seen in the second example on page 535 that the standing wave
pattern obtained by reflecting an electromagnetic wave at normal incidence
on a good conductor gives nodes of E which are half way between the nodes
of H.
Show that the electromagnetic energy density is constant throughout the
standing wave pattern.
12-22. The surface impedance of a conductor is defined as the ratio E;/H;, at the
surface, where E; and H; are the tangential components of EF and H.
It was shown in the Problem 10-8 that H; is numerically equal to the
current per unit width in the conductor.
(a) Show that the surface impedance of a good conductor is
. (ou?
Ol vd) ($4) a OE
a+a/
od
The quantity 1/c6 is called the surface resistivity.
Both the surface impedance and the surface resistivity are expressed in
ohms/square. They give the impedance, or the resistance, between opposite
sides of asquare surface of any size. For example, the surface resistivity of copper
is 8.25 x 10°? ohm/square at 1 gigahertz.
(b) Show that the power dissipated per square meter in the conductor is
1
206
yp
to
12-24. (a) Compare the forces due to the gravitational attraction and to the radia-
tion pressure of the Sun on a spherical particle of radius r and of specific
gravity 5.
The Sun has a mass of 2.0 * 10% kilograms, and it radiates 3.8 « 10%
watts in the form of electromagnetic radiation. The gravitational constant is
6.7 X 10-" newton-meter?/kilogram?’.
In computing the radiation pressure, assume that the particle absorbs the
radiation. See Problem 12-23.
(b) Calculate the value of r for which the two forces are equal.
12-25. A radio wave is incident at an angle 0; on the ionosphere. One may assume
that the electron density depends only on the altitude.
(a) Show that, at any given point in the ionosphere, the radius of curvature
of a ray is given by
nm
R=
~ (dn/dz) sin 0;
(b) Assuming that n decreases linearly with altitude, where does the ray bend
most sharply, and what is the value of R at that point?
(c) Show that a ray would describe a circular arc of radius R if the index of
refraction were given by
Rsin 6;
n=
z+ Rsin 6;
(d) Draw a sketch of such a ray showing 6; and the center of curvature.
Is n equal to sin 6; at the top of the trajectory?
Hint
12-2. (b) Select your y axis along Wn, and your x axis in the plane of incidence.
CHAPTER l3
PROPAGATION OF
ELECTROMAGNETIC WAVES III
Guided Waves
H = (Aut + Aoyj + Aok) exp j(wt — kz) = Hy exp j(wt — k,z), (13-2)
where the coefficients Ey,, Eo,, Eo:, Hor, «++ are as yet unspecified functions of
x and of y only. The dependence on ¢ and on z appears in the exponential
function, which is characteristic of a wave propagating in the positive direc-
tion of the z-axis.
The wave number for the guided wave,
1
i x,’ (13-3)
is not necessarily equal to the wave number for a plane wave (Section 11.3)
and is real if there is zero attenuation.
It is interesting to compare our present procedure with that of the
preceding chapter. In discussing reflection and refraction, we used exponential
functions to describe the dependence of the reflected and transmitted waves
on all three coordinates and on the time, as for example in Eqs. 12-14 to 12-16.
Here we use an exponential function only for z and rt, whereas the dependence
on x and on y is left unspecified. Let us write out Maxwell’s equations for this
field. Since p; = 0, it follows that V-E = 0. Then
Olea Oleg. = oh
= + By Jk Eve. (13-4)
fo ap
OAs 0H, ee
em (13-5)
From V X E = —0B/0t,
OEte 4 ;
—jouAoes, (13-6)
i + jk, Eo, =
[ ORs 3
Tk, Eon Ia cae = Jou, (13-7)
OE oy, OE ox = A
Ox oy a Jop Hoe, (13-8)
; OM: :
Jk, Hox ae ce cs —JweEoy, (13-10)
0H, OA: ;
Te — ah = JweE;. (13-11)
We can now show that the four transverse components Ep,, Eo,, Hox, Hoy
can be deduced from the two longitudinal components £p, and Ho,. From
Eqs. 13-7 and 13-9,
Foz = i
—jou
i ih (= Facey
k, OE oz wae)
(X, # A). (13-12)
MG Sait
We have made the additional assumption that the radian length 4 =
1/w(eu)'/? for a plane wave is different from the radian length of the guided
wave, or that A, ~ X. The case where A, = A will be discussed separately in
Section 13.1.2.
Similarly,
jou a) ae
_ (13-13)
( wu dy ar (A, # A),
ce Lee
2 x2
x ‘ois
x
OE oz k, at)
— — Joe
(XA,# X). ( )
aera (epee oy
le
He
It 1s obvious, by inspection of the above four equations, that the wave is
completely determined once Ey, and Ho, are known.
The wave equation 11-54,
OE
Vi = oe = —k°E, (13-16)
O7Ey, , 07 Ez see He ’
Ax? + ay? — Ey = —k’En., (13-17)
or
O°Ey, , OE _ (5 x 3) E 13-18
ax? + dy? re x2 x2 Oz» ( = )
The ratio Ey:/Hp; is called the wave impedance and is a real positive
number.
Our demonstration is again valid only for X, ¥ X, but the result is really
correct for any value of X,, as will also be shown in Section 13.1.2.
as
Re
is Zero, the parentheses on the right in Eqs. 13-12 to 13-15 must also be zero:
k Ey , OHo. _
fy Bil ga 0, (13-27)
kena Ep oii
ay ae 0, (13-28)
Oby | wk Oly
Ove bueroxe a ¥ arse)
OFo. , K OH. _
Aiea tockaye a 0. (13-30)
The last two equations are equivalent to the first two, since, by hypothesis,
A, = Aand k/wu = we/k = (c/n)? in this case.
These equations can be satisfied by setting Ey, and Ho, both equal to zero,
in which case we have a Transverse Electric and Magnetic (TEM) wave.
The TEM wave has some interesting characteristics. To begin with,
X, = X; hence the phase velocity u = wA is the same as that of a plane wave in
the medium of propagation. This velocity is 1/(eu)!/2 and is independent of
the frequency, insofar as e and p are themselves independent of the frequency.
If the wave propagates in.a vacuum, its velocity is c, whatever the
geometry of the wave guide, and whatever the frequency. Such a line is said
to be distortionless because the various frequency components of a complex
waveform are all transmitted at the same velocity.*
The TEM wave also has the following remarkable property. We have
seen in Sections 6.5 and 8.2 that
aA vy,
E= ———— (13-31)
where the first term is associated with changes in the magnetic field, and the
second with accumulations of charge. Now the currents are longitudinal, as
we Shall see in Section 13.1.3, and A must therefore also be longitudinal, as
well as 0A/dt. Then A = Ak, and
-1i-Fj-(F a
ax" ays Oz + Ot :
(13-32)
(13-33)
and
E= ——i-—j. (13-34)
(a) (b)
Figure 13-2. (a) Shielded-pair line. The signal is applied between the two
wires; the outer cylindrical shield is grounded. (b) Parallel-wire line.
E= al (Fe
Mt i+ eA eej (« ee‘).
ay )exp (13-37)
= Ey expj @ — ;) (13-38)
with
Ey = _ Vy,
rea _ Vo,
ay ap (13-39)
Let us rewrite Eqs. 13-4 to 13-11 for Eo, = 0, Ao, = 0. These eight
equations are simply Maxwell’s equations, as applied to a wave propagating
in the positive direction along the z-axis. We now have only six distinct
equations:
Oboe , Oy
ae -+ ay e—— 0, (13-40 )
Oz Oo,
—— = 13-41
Ox oy 0, ( )
m EZ
he (*) ee (13-43)
m 1/2
OE
a
OEys
ee
_ 0, (13-44)
ony — Ee = 0. (13-45)
Ox Oy
We first note that, for a plane wave propagating along the z-axis, as in
Chapter 11, the field vectors depend solely on z and on f¢, and the derivatives
with respect to x and to y are zero. Then Eqs. 13-40, 13-41, 13-44, 13-45
become identities, and Eqs. 13-42 and 13-43 are equivalent to Eq. 11-61. Such
a plane wave is the simplest form of TEM wave.
For any TEM wave, we can see, from the third and fourth of the above
equations, that E and H are mutually perpendicular, just as for the transverse
components of TE and TM waves. The wave impedance EF,’ Ho; can be
calculated as in Section 13.1.1, choosing coordinate axes so that &), = Ay. =
0. Thus
Eo: as Eo: ie 7
Ho, % A, id (‘) ose
Sw = 5ries
Ree X HY) = 5heres
(5) ER, (13-49)
—s ({\"
u : rms k /t, 13-50
( a= )
— vEbrask, (13-51)
This vector is directed in the positive direction of the z-axis, which is the
direction of propagation of the wave, and its magnitude is equal to the phase
velocity u = 1/(eu)? multiplied by the average energy density, just as for a
plane wave (Section 11.2).
To find E and H, one first chooses the static field Ey that corresponds to
the mode of propagation desired. For example, with the coaxial line of
Figure 13-4, Eo is radial. Then Eq. 13-37 gives E and Eqs. 13-42 and 13-43
give H. We shall work out the field of the coaxial line in Section 13.2.
le OG
= Hk x Vie (13-55)
Ao ae
where k is the unit vector in the direction of the z-axis. The three vectors Ko,
k, V Ho, are shown in Figure 13-3. It will be observed that V Hp, is necessarily
566
tangent to the conducting wall and, therefore, that the rate of change of Ho.
in the direction normal to the surface must be zero.
in cylindrical coordinates, p; being the unit vector in the radial direction, and
C being a constant.
Since we have a TEM wave, the radian length X, of the guided wave is the
same as that of an infinite plane wave, or 1/w(euo)!/*, and the velocity of
propagation wA = 1/(euo)!/? is the same at all frequencies.*
The line voltage U, which is the potential of the inner conductor with
respect to the outer conductor, is
Po p Zz
V= i Edp = Clin * exp j (at ~Z) (13-59)
pi Pi nN
The vectors E and H are oriented as in Figure 13-4 and the Poynting vector
E X H points in the direction of propagation.
* Assuming that the line has infinite conductivity and that e- is independent of the
frequency. In practice, the velocity of propagation does depend somewhat on the frequency.
See Section 13.1.2.
568 PROPAGATION OF ELECTROMAGNETIC WAVES III
The line current J flowing along the surface of the inner conductor can be
calculated from the circuital law, Eq. 7-77:
1/2
I= i}
Hp; de = 2p; (<) <exp j (ot-3), (13-61)
Bee
See! | w eex ==
€r= 1, br = 1) 2) (13-62)
An equal current flows in the opposite direction along the inner surface of the
outer conductor. Note that C is equal to 60 times the peak value of Jif e, = 1,
Te
The average transmitted power can be calculated by integrating the
average Poynting vector over the annular area between the two conductors:
Po
Wr = HeSav 2mp dp, (13-63)
where
Z Ho 2p?
Thus
Wr =
Ces
= In— watts © =i =i) (13-65)
120 Pi
60 In ohms G. = go 1),
a
The wave impedance E/H is 377 ohms, as we saw earlier in Section 13.1.2.
We make the same simplifying assumptions as for the coaxial line of the
previous section. We also assume that the dielectric is air.
We have shown in Section 13.1.2 that TEM waves cannot propagate
inside a hollow tube. Thus the radian length 4, of the guided TE or TM waves
is different from 4%, which is Xo, or 1/c(epu)'/? in this case.
We select axes as in Figure 13-5, with the wave propagating in the
positive direction along the z-axis.
Oz
= C(f cos Cy — Gsin Cy) = 0 aty=Oandy=b. (13-73)
oy
To satisfy the boundary condition at y = 0, F must be zero. The condition
C = 0 must be rejected because it implies that A, = Ao, which is not compati-
ble with a TE wave. At y = 3,
sin Cb. = 0, (13-74)
and thus
c=" (13-75)
13.3. THE HOLLOW RECTANGULAR WAVE GUIDE 571
where is an integer that cannot be zero, for then C would also be zero. Then
a? | 1
BB WO choad} (13-77)
It will be noticed that A, can have only certain discrete characteristic
values corresponding ton = 1, 2,3,---.
Since the left-hand side of the above equation is negative, A, > Ao. This
means that the wavelength in the guide is /onger than that of a plane wave of
the same frequency propagating in free space. Then the phase velocity wu is
larger than c. We shall return to this point later on.
Now that we have found Ap. and X,, we can calculate the two remaining
unknowns £p, and Ao, from Eqs. 13-12 and 13-15, setting \ = Xp since e, = 1,
Hr = 1, by hypothesis:
ge eee
AT b
y, (13-78)
a Dees
Ay = nae sin | :
(13-79)
or, setting
(Bigg ne
nr
(13-80)
then
: nn : Z
E, = Egor sin | — y })exp j @ -Z) (13-81)
b X,
Hes_ Enon sg (3
nt; cos (0 v)exp j(
{ wt -z)
Kyi (13-83 )
where the values for X, are given by Eq. 13-77. The coefficient Eo, is the
maximum value of E inside the guide. We have already found that Ey, = 0,
Hor = 0 (Eq. 13-66).
Let us consider the value of E,. It can be understood qualitatively as
follows. In any plane wave front such as that shown in Figure 13-6, E is
independent of the x-coordinate; since the wave along the z-axis results from
the superposition of such plane waves by multiple reflections, its E must also
be independent of the x-coordinate. The same cannot be said about the y
dependence, however. If we consider a single elementary plane wave progress-
572
(a) (b)
ing at an angle along the guide, the amplitude and the phase of E are constant
over a wave front such as that shown in the figure. The superposition of such
waves by multiple reflections gives an interference pattern, with the result that
the amplitude varies with the y coordinate.
Let us now consider the meaning of the quantity n. For n = 1, Eo, varies
from 0 at y = Oto a maximum of Eo, at y = b/2 and to zero again at yp= b
as in Figure 13-7a. Forn = 2, Ey, is zero in the middle ofthe guide at yp= 6/2,
and is of opposite sign on either side, as in Figure 13-7b. The different values
of n thus correspond to different modes of propagation inside the guide.
Now let us return to Eq. 13-77 for A,. It can be rewritten as
A l
i = Mo 2 ve (13-84)
= (35) }
1
= EE, fora = 1. (13-85)
2b) J
This equation shows that 4, is real if \) < 26. If X, is real, the exponential
functions of Eqs. 13-81 to 13-83 describe an wnattenuated wave. Therefore, if
the above inequality is satisfied, that is, if the frequency is high enough, and
if the walls are perfectly conducting, a wave can propagate inside the hollow
wave guide without attenuation.
If the above inequality is not satisfied, that is, if the frequency is too low,
then A, is imaginary, and the exponential functions show that the field is
13.3 THE HOLLOW RECTANGULAR WAVE GUIDE 573
attenuated exponentially with z. Then the phase does not vary with z, there is
no wave, and there is zero energy flow into the guide once the field is es-
tablished.
The attenuation is rapid. For example, if the frequency is too low by a
factor of 2, \) = 4b. Then, from Eq. 13-84,
1 pone
hes 3'5/ a (13-86)
We must choose the negative sign, for otherwise the amplitude would
increase exponentially with z. Then
Fine 0. (13-89)
E, = 0, (13-90)
H, = 0, (13-91)
A, =_ Foor. (tv
Sin sin (b )exp/i
| wt = a
wy (13-922)
H, _= id
TEwe cos (
ay
h )Expy ; (1 ba x,
fee LY
4 13-93
(13-93)
sehaS
sh cos (h )exp jeen)
Ccot= x, (z+ I,
4) (13-94 )
574
= (iy shia
2) 1/2 > No, (13-95)
{1- Gi) }
and the phase velocity is
u P eet ae
fi2 é is ag (13-96)
2b
It will be observed that £, and H, are in phase, but that H, has the same
oN
phase at z + z as have the two other vectors at z.
Figure 13-8 shows lines of E and ofH for a TE wave with n = 1 propa-
gating in a hollow rectangular wave guide.
13.3. THE HOLLOW RECTANGULAR WAVE GUIDE 575
(13-97)
If the wavelength were n times smaller, the same lines AB and CD would
be n wavelengths apart, and then we would have
COSa =
= 5, :
(13-98)
There are obviously only certain discrete values of the angle a that permit
destructive interference to occur at the walls of the guide.
We have therefore found a geometrical interpretation for the ratio
\o/(2b/n). From Eq. 13-84,
Nye
do ol sin a
(13-99)
This is in agreement with our requirement that the signal velocity never
exceed c (Section 5.10).
13.3. THE HOLLOW RECTANGULAR WAVE GUIDE 577
BA Sheed:
Say = 5Re 12a Tig 5Re(—E.HYj + E,Hyk). (13-105)
OGAs eH:
Substituting the values of E,, H,, H., we find that the first term in the paren-
theses is imaginary, whereas the second term is real.
The energy therefore flows only in the direction of the z-axis and
Say == ae
Evoz sin” 2 |
(7 y)k.is (13-106)
= Epozab i { raat ey
etl 1/2
13-109
Acto ic 2b J ee )
1 ‘ < WE €() 2
= eghdon
if4 €09£00 Sill” (;
b y JadLy = —8 ab Eppe.
0) (13-110)
To find the average magnetic energy content per unit length, we proceed
578 PROPAGATION OF ELECTROMAGNETIC WAVES III
similarly for both the y- and z-components of H and add the results, since
HoH, ane, (13-111)
Again the result is (€/8)ab Ejor, and the average electric and magnetic
energies per unit length are equal. This is reasonable, since the plane electro-
magnetic waves that produce the field configuration by reflection at the side
walls involve equal electric and magnetic energy densities. It is not obvious,
however, because the interference effects tend to confuse the picture.
The total average electromagnetic energy content per unit length in the
guide is therefore epab Ejo,/4. Upon dividing the total average transmitted
power by this quantity, we find that
13.3.4 Attenuation
We have assumed until now that the walls were perfectly conducting; let us
now consider real wave guides of finite conductivity.
In the process of guiding electromagnetic waves, conductors dissipate
part of the wave energy in the form of Joule losses. This is because the waves
induce electric currents in the guide. A rigorous calculation of the field for a
guide of finite conductivity is difficult, but fortunately unnecessary.
The procedure used for calculating the Joule losses is the following. We
have performed a calculation on the assumption that the guide is perfectly
conducting. This led to a field in which there is a tangential H at the surface of
the guide. Since the tangential H must be continuous across any interface, we
know the value of H inside the conductor. Then, using Maxwell’s equations,
we can find the corresponding tangential E inside, which is not zero unless
the guide material is a perfect conductor. This small tangential E is then
considered to be a perturbation of the ideal field obtained with perfect
conductors. The method is entirely satisfactory because this E is so small that
it hardly disturbs the wave. We thus have a tangential E, a tangential H, and
a Poynting vector that is normal to the conducting surface and directed into
the metal.
That both EFand H vectors must exist inside the conducting walls can also
13.3. THE HOLLOW RECTANGULAR WAVE GUIDE 579
kes = ar
— Wi
(13-116)
i
The real part k,, of k, can be taken to be the k, obtained on the assumption
of perfectly conducting walls.
It might be expected at first sight that the attenuation could be calculated
from the reflection losses. It will be recalled from Section 12.5 that an electro-
magnetic wave reflected from a good conductor is slightly weaker than the
incident wave. This method of calculation is incorrect because, as we shall
see, there are also energy losses in the guide faces parallel to the yz- plane.
Let us calculate k,;. The tangential H produces an electromagnetic wave
that penetrates perpendicularly into the wall. Inside the conducting wall,
as in Eq. 11-106.
We assume that the dielectric inside the guide is dry air, and is therefore
lossless. We also assume that n = 1 in order that the field can be described,
as a first approximation, by Eqs. 13-88 to 13-94.
580 PROPAGATION OF ELECTROMAGNETIC WAVES III
ig ete
wiob
eae
Naat
5)
2.
(ise Olas CS)
Then E, is not equal to zero for y = 0, as in Eq. 13-88, but is rather
1/2
E, = (=)
o
ee
lod
ay (otba XA
2 1)
4
(y = 0). (13-119)
Note that, as the conductivity « approaches infinity, E, approaches zero. The
average Poynting vector (1/2) Re (E X H*) is directed into the guide wall and
is equal to
TE, Ox 2 1
( 5 ) og! Qerpi) 8/2 (= Uy
This is the average energy flowing into the wall at y = 0, per square meter
and per second. It is interesting to note that this energy flux is the same at all
points on the face y = 0. The power lost to the wall per meter of length is a
times larger and, for the two faces parallel to the xz-plane,
tT Epox \2 2a
iva Geta error eee
This is the average power lost by reflection.
At the face x = 0, H has y- and z-components as in Eqs. 13-92 and 13-94.
Associated with H,, we have an electric field intensity
(He) 1 cos? =
b | eeu) by
Integrating the sum of these two quantities from y = Oto y = b and multiply-
ing by 2, we obtain the average power lost per meter in the two walls parallel
to the yz-plane:
We
___Ebe
bo Quy
_[, , (2b)
|} aie (2) | (13-122)
or, using Eq. 13-85,
1 Eo e 2D\*
Wye = bot Quo 7 (~) : (13-123)
581
kid”
law,
he
al
ae
0 N i= i= = \ a ! |
02 0.4 0.6 08 1.00
My Atenagiob
Figure 13-10. Dependence of k,;b*/? on \o/d. for a/b = 0.5
and for copper.
See Cae 2b
= painanar {5 + (5) | a)
To find the attenuation constant k,; for the wave, we now use Eqs.
13-116, 13-125, and 13-109:
ie ae (13-126)
1G)
2ai fren?
RET (13-127)
Oba
For an infinite guide height a, only the second term remains in the
numerator and
ee ae)
No \?
"BOA"? fy Gey fe
(a «). (13-128)
i 2b
This term comes from W,,, and the losses then occur only on the guide faces
paraliel to the xz-plane. It is shown in Problem 13-22 that this value of k,; can
be accounted for entirely by the reflection losses.
In practice, the ratio 2a/b is close to unity, whereas (\o/2)? is of the
order of 1/2. The losses on the faces parallel to xz-plane are thus of the same
order of magnitude as those on the other pair of faces, but smaller by a factor
of about 2.
Figure 13-10 shows k,;b?/? as a function of the ratio \o/A, = do/2b for
582 PROPAGATION OF ELECTROMAGNETIC WAVES III
2a/b equal to unity. According to this curve, the optimum value of \»/26 is
about 0.4, but the minimum is very broad; actual values of \o/26 are larger so
as to achieve strong attenuation for the n = 2 mode.
The attenuation is of the order of a tenth of a decibel per meter at
frequencies of a few gigahertz, increasing as f?/? when the ratios a/b and
do/2b are kept constant (Problem 13-21).
Table 13-1 shows the characteristics of a few types of hollow rectangular
wave guides.
13.4 SUMMARY
x2 x?
mi
Jwe OE). k, 0H.
Ay0 = ———
i i |
(—ay — ree
2 —),) 5
(13-14)
13.4 SUMMARY 583
figs —Jwe (= k, on
Ox we Oy Us2l>)
The vectors E and H can therefore be calculated once Ep, and Hy, are known.
If Ey. = Owe have a TE wave, and if Hy. = Owe haveaTM wave. If both
Eo, and Ho, are zero, then we have a TEM wave and Agar iNe
For TE and TM waves we calculate Ep. and Ho. using the wave equations
and the boundary conditions imposed by the guide. The wave equations
for Ey, and Hp, are similar; that for Ep, is the following:
O*Ens | O° Ey, [1 1
axe a ay? (5 = zs)Ev. (13-18)
In TEM waves,
OV 0A
oe oT. (13-31)
and
LS ae oe : ae
E e Poe ay i)exp j (ot ‘) (13-37)
AB
ao (pape i" for TE waves,
(‘) (13-23)
€ Na
(13-25)
1/2
= (*) for TEM waves. (13-46)
€
To illustrate how one can deal with guided waves, we studied two
common types, namely, the TEM wave in a coaxial line and the TE wave with
= | in a rectangular wave guide. In a coaxial line,
(nen p
ees (ot= 5)
A
re (13-58)
1/2
H = (<) ay (or= 5)01, (13-60)
Ho p x
Wr = 120
— in ©?
ero,
watts (cr = 1, ur = 1). (13-65)
ett 3B
Mo = GI200n)" fay"
2a No 2
(13-127)
i} 2b J
It corresponds to losses in all four faces.
585
PROBLEMS
13-1. The transverse part of the vector E in a guided wave is written as a vector E,,.
What can you say about the orientation of this vector in space, as a func-
tion of the four variables x, y, z, r, in the case of TE waves? What if E,/Eoz
is not real?
13-2. Sketch a rather large cross-sectional view of a coaxial line in a plane contain-
ing the axis.
(a) Show lines of E and of H at a given instant over at least one wavelength.
The lines should be most closely spaced where the field is strongest. Indicate
the directions of the fields by means of arrow heads. Show the direction of
propagation.
(b) Add arrows at various points to represent Poynting vectors, using longer
arrows where the power flow is larger. Assume that the length of the arrow
represents the magnitude of the Poynting vector at its midpoint.
(c) How does the pattern change with time?
(d) Sketch a cross-sectional view of the coaxial line in a plane perpendicular
to the axis and show lines of E and of H at a particular instant.
Relate this plane to the figure you drew under (a) above.
(e) Explain how this pattern changes with time.
(f) Add plus and minus signs to both figures to show the surface charges.
The spacing between the signs should indicate qualitatively the relative mag-
nitude of the surface charge density.
(g) How does the charge pattern change with time at a given z?
(h) Now add arrows of various lengths to your first figure to represent surface
current densities.
(i) How does the current pattern change with time at a given z?
13-3. Show that, in the case of an idealized coaxial line of infinite conductivity, the
current is given by the linear charge density multiplied by the velocity of
propagation.
where L’ and C’ are, respectively, the inductance and capacitance per unit
length.
Show that this is correct in the case of the coaxial line, by calculating L’
and C’, and then comparing with the value of the characteristic impedance
given in Section 13.2.
. (a) If the maximum allowed field strength in a coaxial line is E,,, show that
the maximum allowed voltage is
Va, = pug Bol Po/pi
586
(b)
Figure 13-11.
(b) Show that, for a given value of p., V,, is greatest when p,/p; is equal to e.
(c) Show that the characteristic impedance is then 60 ohms if the line is air-
insulated.
(d) Show that, under those conditions, the maximum allowable current is
PoEm/163 ampere.
13-6. (a) Setting E,, to be the maximum allowed field strength in a coaxial line,
use the result of Problem 13-5 to show that the maximum allowed power is
ES °° 120 ee
(p/p)?es
Ex, 1 (p0/pi)
z.=(G)"
where L’ and C’ are, respectively, the inductance and the capacitance per unit
length of the line.
In designing such lines it is therefore essential to predict the values of L’
and C’. If the geometry is such that L’ and C’ cannot be calculated analytically,
as, for example, in Figure 13-11a, it is sometimes useful to perform the follow-
ing measurements on a resistance-sheet analog.*
(a) We can find the value of C’ by cutting out a sheet of resistive material
in the shape of the cross section of the dielectric as in Figure 13-11b, and
measuring the resistance between electrodes 4 and B.
*D. J. Epstein, Comment on a Theorem in the Field of Steady Current Flow, Proc.
IEEE, 56, 198 (1968).
587
Figure 13-12.
Riles
So
ae @ Ry
See also Problem 10-14.t
. Figure 13-12 shows a cross section of a microstrip line. The lower electrode
is grounded and the wave is TEM.
The microstrip has the advantage of being much less costly than either
the coaxial line or the rectangular wave guide. It is also particularly convenient
for conveying high-frequency signals within printed and integrated circuits.
Its main disadvantage lies in the fact that its field is not limited to the region
immediately below the upper electrode. Microstrip lines can therefore interact
with other elements in a circuit, unless they are either spaced or shielded
properly. Shielding can be achieved by adding a second grounded plane above
the strip, but this, of course, increases the complexity and cost of the line.
(a) Sketch lines of E and of H.
Use arrows to show the directions of E and H at a given time. Show the
direction of propagation.
(b) In practice, the width 4 ofthe strip is much larger than its distance / to the
grounded plane, and edge effects are small.
Show that the instantaneous value / of the transmitted power is equal
to the Poynting vector integrated over the cross section bh.
(c) Show that the characteristic impedance U/J is
p\V2h
(Z) b
(d) Show that one arrives at the same result if one defines the characteristic
impedance as in Problems 13-4 or 13-7.
(e) Show that the addition of a second grounded plane placed symmetrically
with the first reduces the characteristic impedance by a factor of two.
588 PROPAGATION OF ELECTROMAGNETIC WAVES III
13-9. Figure 13-8 shows schematically the E and H fields inside a rectangular wave
guide carrying a TE (n = 1) wave.
Draw sketches as in Problem 13-2 showing Poynting vectors E < H, sur-
face charge densities, and surface current densities.
To represent properly these last two quantities you will have to sketch
two or three neighboring faces. Be sure to relate these faces to those shown
in Figure 13-8.
13-10. If a wave guide is not properly connected to its load, part of the incident
wave is reflected back towards the source and a standing wave is established
along the line. Under such conditions the power fed to the load can become
negligible.
It is therefore useful to be able to move a small probe along a longitu-
dinal slot to sample the field inside the guide. The quantity that is usually
measured is the Voltage Standing Wave Ratio (VSWR). This is the ratio of the
maximum to the minimum time-averaged voltage picked up by the probe as it
slides along the slot. Under ideal conditions of power transfer, there is no
reflected wave and the VSWR is equal to unity.
The probe can be either a small loop that is coupled to the magnetic field,
or a short length of wire that responds to the electric field. In both cases the
projection into the guide is approximately one millimeter or less.
(a) In the case of a rectangular hollow wave guide and a TE mode, as in
Figure 13-8, where should the slot be cut? The slot should, of course, disturb
the wave as little as possible. Sketch a perspective view of the guide, showing
both the EF vector and the slot.
(b) If the probe is a small loop, how should it be oriented with respect to the
guide?
(c) How would you use the movable probe to measure ),?
13-11. Ordinary wave guides are brass or copper tubes with wall thicknesses of the
order of one or two millimeters. However, some microwave structures are
quite complex and would be too expensive to fabricate either from tubing or
from solid stock. Several techniques have therefore been developed for such
cases.
One technique, called e/ectroforming, utilizes a former, in the shape if the
internal volume required, which is electroplated to a thickness of a few milli-
meters. The former is then removed, leaving the required structure.
In some cases the former is a plastic and is left in place. The resulting
structure is thus dielectric-filled and is smaller (Problem 13-13), for a given
operating frequency, than if it were air-filled.
In this latter case, what minimum thickness of copper would you recom-
mend if the operating frequency is to be 3 gigahertz?
13-12. It is found that an electromagnetic wave propagating in the TE mode with
n = 1 (Section 13.3.1) through the 2.84 inch X 1.34 inch wave guide of Table
13-1 has a X, of 13.8 centimeters.
Calculate its frequency.
13-13. (a) Show that, if a hollow rectangular wave guide is completely filled with a
PROBLEMS 589
1 1 ane
Sheei old g
where m is the number of half cycles of E or of Hin the direction of the x-axis,
and similarly for n.
13-16. (a) Sketch a graph of \o/A, as a function of Ao/A, for a rectangular wave
guide.
In actual practice this curve is meaningful only for \o/A, ~ 0.7. At much
smaller values of \o, higher modes of propagation (n = 2,n = 3, etc.), are pos-
sible, and the field configuration inside the guide becomes uncertain. At much
larger values of \o the attenuation becomes excessive.
(b) Sketch a curve of w as a function of k,.
590 PROPAGATION OF ELECTROMAGNETIC WAVES III
For a given operating point, the phase velocity u, is given by the ratio
w/Kg.
(c) Verify that the signal velocity is given by
Us
a ae ee
~ dk,/do dk
and thus by the slope of this curve.
13-17. Check the power rating of the largest rectangular wave guide shown in
Table 13-1. Assume that the operating frequency is 1.5 times the cut-off
frequency.
The maximum permissible electric field intensity in air is 3 x 108
volts/meter. Allow a factor of safety of 2 to take into account the effects of
irregularities in the inner surface of the guide.
13-19. It is suggested that one could measure the power transmitted by a rectangular
wave guide by observing either the deflection or the energy gain of a beam of
electrons crossing the guide.
Discuss the posibility of such measurements.
13-20. (a) Show that the average force per unit area exerted by the field on the face
x = 0 of the rectangular wave guide shown in Figure 13-8 is given by
9 °
_ €9E oor Xo \? s, 2ry
Koehn ge 1 cops 20s tan
(b) Calculate F, for the first guide listed in Table 13-1 at the maximum power
rating and at a frequency of 3.00 gigahertz.
(c) Is it important to take this force into account in the design of rectangular
wave guides?
(d) Could this force be used to measure the transmitted power?
13-21. Show that the attenuation constant k,,; fora TE(n = 1) wave ina rectangular
wave guide, as given in Eq. 13-127, varies as f*/? when the ratios a/b and \o/2b
are kept constant.
13-26. (a) Check the attenuation of the largest type of rectangular wave guide
listed in Table 13-1. Assume that the guide material is copper and that the
operating frequency is 1.5 times the cut-off frequency.
See Problem 13-25 for transforming nepers/meter to decibels/100 feet.
(b) Calculate the power dissipated in the guide per meter of length at its maxi-
mum power rating of 2800 kilowatts.
(c) Calculate the rate at which the temperature of the guide would increase,
at a maximum power rating, if it were thermally insulated. The walls have a
thickness of 0.080 inch.
(d) How should the guide walls be cooled?
Would it be practical to transmit large amounts of electric energy over
long distances in this way ?
Wave guides are often used to transmit large amounts of power at low
duty cycles. The average power dissipation is then much lower than in the
above example.
13-27. Tabulate the values of the following quantities for the five types of rectang-
ular wave guide of Table 13-1:
(a) b/a;
(b) Vea fants :
(c) 0/26 at both limits of the operating range;
592 PROPAGATION OF ELECTROMAGNETIC WAVES III
(d) the attenuation in nepers/meter (see Problem 13-25) for the n = 2 mode
at both limits of the operating frequency.
13-28. (a) Weare required to calculate the maximum power that can be transmitted
at a frequency of 3.00 gigahertz through a short length of (a) coaxial line, and
(b) rectangular wave guide.
The coaxial line has a diameter 2p, of 2.50 centimeters and the wave guide
has an inside cross section of 3.75 & 7.50 centimeters.
The coaxial line satisfies the condition for maximum power transfer,
namely p./p; = 1.65 (Problem 13-6). The radius py is chosen to ensure
attenuation of higher order modes.*
In both cases the dielectric is air and the current-carrying surfaces are
silver plated. The maximum allowed electric field intensity is 1.5 « 108
volts/meter as in Problem 13-17.
It is assumed that there is no reflected wave.
(b) We are also required to calculate the power lost per meter of length in
both cases. See Problem 13-29a.
(c) Finally, we wish to know the rms voltage and current at the input end of
the coaxial line.
13-29, In calculating the resistance of a coaxial line it is correct to assume that the
current is uniformly distributed throughout a thickness equal to the skin depth
6 on both conductors (see Problem 13-23).
(a) Show that the resistance of a coaxial line per unit length is
1 1 1
on 216 Do + *),
13-30. (a) Calculate the power rating of an air-insulated copper coaxial line whose
outer conductor has an inside diameter 2, of 2.84 inches, when the ratio
Po/pi 1S Selected as in Problem 13-6 for maximum power transfer. The maxi-
mum electric field intensity in air is 3 X 10° volts/meter. Allow a factor of
safety of 2 to take into account irregularities in the surfaces of the conductors.
(b) Use the results of Problem 13-29 to calculate the power dissipation in both
the inner and outer conductors per meter of length at the power calculated
under (a) above, and at a frequency of one gigahertz.
13-31. Let us calculate the attenuation constant k,; for a coaxial line, taking into
account energy dissipation in the dielectric.
* See, for example, A. F. Harvey, Microwave Engineering, Academic Press, New York,
1963, page 20.
PROBLEMS
593
Since the dielectric losses are normally small, we shall simply add their
attenuation constant to that associated with resistive losses in the conductors
(Problem 13-29) to obtain the value of k,: for the line.
(a) A dielectric is said to be lossy when it dissipates energy in the course of
the polarization process, or when it is slightly conducting.
Consider first a parallel-plate capacitor containing a lossy dielectric. The
plates have an area A and are separated by a distance s. An alternating voltage
voltage U is applied across the plates.
You should be able to show that, if there is energy dissipation in the
dielectric, then one can either say that the dielectric has a conductivity o, or
that its relative permittivity is of the form
ae > Jess
ken e or aA (2
Thus, taking into account the losses in both the conductors and the dielec-
tric, and using the result of Problem 13-29, the attenuation constant for a
coaxial line is
= zy il = (po/ pi) €.
ke a= 2.64
2.64 X 1 0-8 ((+ ee
Palio sim MMOs) 5 110) —8 Seyi
GE meters =
where a is the conductivity of the conductors, and where e¢} — je!’ is the rela-
tive permittivity of the dielectric.
13-32. A gas lens utilizes the fact that the index of refraction of a gas is dependent
on its density, and therefore on its temperature. Thus, cool gas blown gently
through a heated tube becomes hotter and lighter near the periphery than near
the axis and acts as a weak converging lens.
After a certain distance the radial temperature gradient becomes negli-
gible and, if further convergence is required, the hot gas must be evacuated
and fresh cool gas injected in its place. Another possibility is to use the same
gas flowing through a succession of alternately hot and cold tubes. Then the
hot tubes act as focusing elements, while the cold tubes act as defocusing
elements, and, under proper conditions, there is a net focusing effect. This
technique is known as alternate gradient focusing and was developed for par-
ticle accelerators.
Such beam wave guides can be used for transmitting modulated light
beams carrying, for example, telephone messages.
The index of refraction is related to the density D and to the temperature
T as follows:
594 PROPAGATION OF ELECTROMAGNETIC WAVES III
p= i DRS
no — 1 ID; tan:
Os” Oz
and neglect the term
Op On
Oz Oz
Since 77 is very close to unity,
d’p _ an
dz*~ dp
Then
ap Ty) OT
det ad (1 1) T? dp
This equation can serve as a starting point for the study of such beam
wave guides. Since the right-hand side is a complicated function of both p
and z, the equation must be solved numerically.
Hints
RADIATION OF
ELECTROMAGNETIC WAVES
= Oe?" (14-1)
+Q
-0
We shall use this simple model, but electric dipole radiation is also produced
by any charge distribution whose dipole moment
p= | eras (14-4)
where r and 6 are as in Figure 14-4 and [f] is the retarded time t — (r/c).
For a nonzero frequency, the exponential term appears to show that V
propagates as a wave at a phase velocity c. This is not quite correct, because
of the complex factor (A/r) + j. Rewriting the complex factor in exponential
form,
For r>>X, arc tan (r/A) & 1/2 and is approximately independent of r. The
phase velocity is then c. However, closer in to the dipole, where r is not much
larger than 4, arc tan (r/A) is not constant, and the effect of this term is to
give a phase velocity that is Jarger than c.
The scalar potential V varies as cos @ and is zero in the equatorial plane,
where the fields of the two charges cancel exactly, just as in electrostatics.
It varies as 1/r? as in the static case, but only as 1/r when r > 4. Also, for
598
r >> X, V varies as 1/X and is thus proportional to the frequency. Figure 14-5
shows a radial plot of V as a function of @ and ¢.
A=ey a
eel I, expiD jw[t]
ij s.: :
(14-7)
A= a ys (ex
jw[ t})(cos
p 6 r; — sin 6 @;) (14-8)
= EEP* (ex
jel#]\(cos
p @r: — sin 6 61) (14-9)
where the polar unit vectors ri, 0, g: are as in Figure 1-19. The vector po-
tential A propagates at a velocity c, even for r< i.
599
Figure 14-5. The scalar potential V and the magnitude of the vector potential A
are shown here as functions of 6 and ¢ about an oscillating electric dipole ori-
ented as shown. The radial distance from the center of the dipole to the spheres
marked V is proportional to the value of V in that particular direction. The scalar
potential V is maximum at the poles; it vanishes at the equator, where the indi-
vidual potentials of the charge — QO and +0 of the dipole cancel. It is positive
in the northern hemisphere, where the field of +@Q is predominant, and negative
in the southern hemisphere. The magnitude of A is similarly represented by the
sphere marked A. The vector potential is independent, both in magnitude and in
direction, of the coordinates @ and ¢.
P(r,0,~)
Thus
l 1 fa 0A,
The two other components of the curl are zero because A has no ¢ component,
and because it is not a function of ¢. Thus
Re (RING
herf (exp jolt)(7 +i) sin 61. (14-11)
As usual, we compare this new result with a previously acquired one,
namely the Biot-Savart law for steady currents, as stated in Eq. 7-8. According
to this equation, the magnetic field intensity due to an element of current /s
at the position of the dipole is
Since both parts of H involve sin 6, they are both maximum in the
equatorial plane and zero along the axis of the dipole. The magnitude ofH as
a function of @ and ¢ is plotted in Figure 14-8.
Since [hs = jwpo,
== 1)sin 6 1.
ee=va(exp jol)( i> (14-13)
For r> i,
Pave me exp jw[t] sin 0 gi (PA). (14-14)
—Po 3
GiregrX? exp jolt]
IN nA Kr? I\ <6
xX \(-25 — jot 1)cos 6r; — 5 +] *)sin 604) (14-17)
Then
= = ri exp jolt]
oe Ae \|
1)| cos 6
x/{(-0+ (-25,- yt
2
+44) + (-4
9 — )}sinoa (14-19)
es Ly Po A : :
is ara exp jw[t] sin 6 0, (r > 2X). (14-22)
Recalling that
mnxX = — GW, aixXa=n, (14-26)
the components of E and of H being given by Eqs. 14-20 and 14-13. Thus
Ps ADD)
_ Po sin’0 14-28
ay 32 negrAt ( )
One striking feature of the Poynting vector is that it involves only the
604
Figure 14-8. Polar diagrams of sin @ (outer surface) and of sin? @ (inner surface)
showing, respectively, the angular distributions of E or H, and of 8,, at a distance
r ><X of an oscillating electric dipole situated at the origin. The radial distance
from the dipole to one of the surfaces is proportional to the magnitude of these
quantities in the corresponding direction. Most of the energy is radiated near the
equatorial plane; none is radiated along the axis.
radiation terms, despite the fact that our calculation is valid even near the
dipole, where r is not much larger than X. In fact, it could have been calculated
correctly by disregarding the terms that are unimportant far away from the
dipole. The energy flow is everywhere purely radial, at least as long as the
dipole length s is small compared to both r and X.
The Poynting vector varies as |/r?. This is required for conservation of
energy since, under steady state conditions, the energy flow through any
given solid angle must be the same for all 7. This 1/r? dependence results from
the fact that the radiation terms for E and for H both vary as 1/r.
Since the energy flow varies as sin? 6, it is zero along the axis of the dipole
and maximum in the equatorial plane, as in Figure 14-8. An electric dipole
does not radiate energy along its axis.
The total radiated power W is obtained by integrating $,,. over the
surface of a sphere of radius r:
14.1. ELECTRIC DIPOLE RADIATION 605
pe Tie = (14-31)
7 _ Hoe (3)
(S\* R,22 (14-33)
2
This is the resistance that would dissipate in the form of heat the same power
that the dipole radiates in the form of an electromagnetic wave, if it carried the
same current. It will be recalled that we have assumed that s < X.
co)
Figure 14-9. The electric lines of force of an oscillating dipole for wt = 0, 7/2,
mw, and 32/2. The dipole is situated in the center and is oriented in the vertical
direction. The decrease in wavelength with distance can be observed on these
figures. The magnetic lines of torce are circles perpendicular to the paper and
centered on the axis of the dipole.
Then
E=VXC (14-39)
ee ae
= rsing do. (14-40)
ic) : ie
Hee eae (14-42)
14.1. ELECTRIC DIPOLE RADIATION 607
Or
0 , 0 :
30 (Cr sin 6) do + oe (Cr sin 6) dr = 0, (14-43)
Sin /Ae ;
x G as i)exp ja( ai ‘)= Constant, (14-45)
‘ x2 1/2 r r
sin? @ é os 1) cos (otas + arc tan z)=i, (14-46)
where K is a real parameter that varies from one line of force to the next. The
calculation is exact as long as r>>s5 and X>>s. All three factors determine
the shape of the line of force as a function of r and 6, whereas the cosine term
gives the radial motion. Figure 14-9 shows four families of lines of force.
For r >A,
and the lines of force travel outward at a velocity wA = c. Closer in, however,
the arc tan term varies with r, with the result that the velocity of the lines of
force is larger than c.
The magnetic lines of force are much simpler: they are circles perpen-
dicular to, and centered on, the axis of the dipole. This can be seen from Eq.
14-11, which shows that H is everywhere azimuthal.
* You may wish to omit this rather detailed discussion of the electric lines of force of
the oscillating electric dipole.
608
Figure 14-10. The parameter KA of Eq. 14-46 is plotted here as a function of the
coordinates r and @. The surface shown is that corresponding to tf = 0; with
increasing time f, the central peaks oscillate together from —» to +, and the
ripples move out radially. The loops are electric lines of force corresponding to
constant values of KX.
with the plane 6 = 7/2. This curve is situated inside its envelope, defined by
the curves
Kia + 1(3) me iy" (sin @ = 1), (14-48)
which are shown by broken lines. These curves approach infinity as r/X > 0
and approach unity when (r/A)? > 1. All values of KX are possible, from —
to + «. The figure also shows a succession of curves for successive values of f.
As the ripples move out, their height decreases rapidly at first, and soon
approaches unity. It is clear that lines of force with KX < 1 can travel out to
infinity. They give the radiation field. It is also clear that if KX > 1, they cannot
go far. Let us consider the case where KX is slightly larger than unity. The
loop formed by the corresponding line of force shrinks until it reaches the top
of the ripple and then disappears. This explains the relatively rapid decrease in
the field intensity in the region where r is of the order of X or less.
Closer to the dipole, which is situated at the origin, some lines of force,
such as that for which KX = 5, do not even get into a ripple. These lines of
force simply pulsate in and out without ever escaping into space. This is the
static field.
609
One
ON
PSS
5
THEY
Figure 14-11. This figure illustrates how the intersection of the KX surface (Figure
14-10) with @ = 7/2 changes with time. The curves are all situated within the
envelope shown; thus the amplitude is infinite at r = 0 and approaches unity for
oN
Let us analyze the motion of the lines of force quantitatively. We can set
sin 6 = | and consider their motion only in the 6 = 7/2 plane. Then
iy Ay ie r .
(5 + 1) cos (otzi + arc tan ‘)= Ki (sind = 1). (14-49)
(14-52)
610 RADIATION OF ELECTROMAGNETIC WAVES
The velocity u of the line of force becomes zero for sin X = 0, or for cos
X = 1. This corresponds to a point where the ripple touches the envelope.
The curve then “‘rolls’’ on the envelope and the line of force momentarily has
zero velocity.
The velocity u also approaches plus or minus infinity near
anes. (14-53)
3
or
Kx
r p\ 12 (14-54)
a (1+ 4)
Now this is precisely the condition that defines the top of a crest, or the bot-
tom of a trough. The infinite velocity simply results from the fact that, as the
ripple decreases in amplitude with increasing r, the line of force closes in with
infinite velocity just before disappearing. This is interesting in that it tells us
where the lines of force disappear.
Note that the top of a crest does not occur at cos X = 1, or sin X¥ = QO,
where the curve touches its envelope, but somewhat closer to the dipole at a
value of X defined by the Eq. 14-112, since the cosine term in Eq. 14-49 is
multiplied by another term that decreases with increasing r, namely
(a/ry? + 1p”.
For large values of 7/X, KX is about unity, and the top of a crest, or the
bottom of a trough, occurs at X~ 0. For small values of r/X, and for large
values of KX, these occur at |sin X| > 0.
For r > i, KX is about unity or less, and u— c. The lines of force then
travel outward at the velocity of light, as expected.
For Ki > 1 and r pA,
(14-55)
Io cos wt
Figure 14-12. Half-wave antenna. The current distribution shown as a broken line
is I, cos (//X) cos wt. This is the standing wave pattern at some particular time
when cos wt = 1.
where “‘Re” means, as usual, “Real part of.’ The right-hand side shows that
the standing wave is the sum of two traveling waves, one in the positive di-
rection, and one in the negative direction, with amplitudes /)/2.
Then, from Eqs. 14-2 and 14-3, using the usual complex notation, we can
express the electric dipole moment p of the element d/ as
(14-59)
where r’ is the distance between the element d/ and the point (r, 6, ¢), as in
Figure 14-12,
r' ~r — Icos 8, (14-60)
and
E=PE as SC
Teer X sin @ Gadd
. lag 28
ee (ix (cos @ — 1)
We have removed the 1/7’ from under the integral sign and set it equal to Lyin
14.2. RADIATION FROM A HALF-WAVE ANTENNA 613
since it is assumed that r >> X. With this condition, the dE’s can all be taken to
be parallel for a given point of observation and can be assumed to have the
same amplitude but different phases, and the integration can be limited to the
phases. Integrating,
oe Ih exp jw[t]
me J8rceodr
+/4
(exp AR (cos 6 — »} X exp ee (cos 6 + | :
X sin é Cy aN + Fee ip es oe. 6,. (14-62)
init
sin \2 (cose lho
I) — i COS @5 cos 6 ) (14-65)
; cos (5cos 6)
i = Dacre Ih exp jw[t] = Tein Gaten 0,, (14-66)
cos G COs a)
= 60.07 fooui \—; 0, volts/meter. (r > A). (14-67)
Figure 14-13. Polar diagrams the functions of cos {(7/2) cos 6} /sin 6 and cos?
{(a/2) cos 6\ /sin? 6 showing, respectively, the angular distributions of E or H and
of S.y for the half-wave antenna at r > 4. The angular distributions are similar
to those for the dipole, except that the half-wave antenna radiates a larger frac-
tion of its power in the region of the equatorial plane.
as in Eq. 14-23. For the half-wave antenna, H is again azimuthal, and the
above equation also applies; thus
COs (5cos 0)
i ay Ih exp jw| ft] sin 8 Yi. (14-71)
Z
l
Say =o) Re (E X H*), (14-72)
. \ eos ae
G cos#)
P cos? 5 cos 8)
== 0.50 2 = See watts/meter?. (14-74)
mICOS: ¢ cos 0)
= 60.0 [ens —__—____ qf (14-76)
0 sin 6
and
20
1 — cosa
Wer¢ 60.0rhin2 | eon da. ~
(14-78)
Then if we write
i i fil 1
al4r — 2a) 4r (;a 27 — ) Sete)
Now the two integrals between the braces are equal, as can be seen from the
curves for 1 — cos a, for a, and for 2x — a, shown in Figure 14-14. Since the
curve for 1 — cos a is symmetrical about a = 7, the area under the curve
(1 — cos a)/a must be equal to that under (1 — cos a)/(27 — a) between the
limits 0 and 27. Thus
W =a Te awats, (14-82)
and
Rea = 73.1 ohms. (14-83)
* See, for example E. Jahnke and F. Emde, Tables of Functions (Dover, New York,
1951), pp. 3 and 6.
617
Figure 14-15. A simple antenna array, represented by the two vertical bars, and
its radiation pattern for r >> A. The two half-wave antennas are spaced \/2 apart
and are excited in phase. The surface shows the magnitude of E plotted radially
as a function of 6 and ». The curves shown are situated on the surface for con-
stant values of y chosen at 10° intervals. The surface is cut into two parts for
clarity. Along the y-axis the two waves add, and the resulting electric field in-
tensity is twice that produced by a single antenna. The same applies to all of the
yz-plane, at least for r >> A. Along the x-axis, however, the two waves arrive in
opposite phases and cancel. For other directions along the xz-plane the waves do
not cancel completely, since the difference in path is smaller than \/2. There is
zero field along the z-axis for each of the antennas, and hence for the array.
T
, cos (Jcos ‘)
ee
Va sin 6
J (- D cos (i?cos i)i :
x ieee Ix + exp marge’ i 61, (14-84)
where the quantity on the second line accounts for the fact that the two waves
arrive out of phase, one having traveled a distance r + (D/2) cos y, and the
other a distance r — (D/2) cos y. Thus
Tv
5 cos ( COS a)
E = 120.0; BIEN
r
ke
sin @
cos (5
2K
cos v)6, (14-85)
618
Figure 14-16. The two half-wave antennas shown as vertical bars are spaced \/2
apart, but the one at x = — D/2 leads the other by zm radians. See legend of Figure
14-15 for an explanation of the surface. The two waves now cancel everywhere
on the yz-plane. All along the x-axis, the two waves arrive in phase to give twice
the field of a single antenna. There is again no radiation in the direction of the
z-axis.
and then
, cos (5cos a)
E = 120.0) pel + cos & sin 6 cos e)6; volts/meter.
Dr
(14-87)
When the antennas are one-half wavelength apart, D/(2X) = 2/2. Then,
in the xy-plane, where 6 = 7/2, E varies as
COS
oaks
Lees
COS ¢ cos )
Nea. cos ¢ sin @ )-
sin 6 2)
The first term is the angular distribution for a single half-wave antenna; it is
zero at 6 = 0 and maximum at 6 = 7/2. The second term comes from the
interference between the two antennas; it is maximum at 6 = O and zero at
6 = 1/2. The product of the two is zero both at 6 = 0 and 6 = w/2. Finally,
in the yz-plane, ¢ = 7/2, and E varies simply as
COs ¢ COs s)
sin 6
T
cos { = cos 6
je 60,0) 1h exp jw[t
xp sole) ic__f
2 )
which is the same as Eq. 14-84 except for the addition of jr in the last ex-
ponent. Recalling that e’" = —1,
620 RADIATION OF ELECTROMAGNETIC WAVES
T
: cos (5cos s)
E = 1200 ie
expel — 4 sin (3 sin 6 cos e)6, volts/meter.
r i 2K
(14-89)
The radiation pattern is shown in Figure 14-16.
is a sinusoidal function of time. This type of radiation would arise, for example,
if currents were excited at the surface of a sphere in such a way as to drive the
electrons alternately to both poles and then to the equator.
We could proceed exactly as for electric dipole radiation and calculate
successively V, 4, E, H. This will be done in Problem 14-11. It is easier, how-
ever, to add the fields of the two component dipoles as follows. For simplicity,
we consider only the radiation field and set r >> X.
We now have two dipoles, one with moment —pye’*! centered at —s/2,
and another with moment + poe’! centered at + 5/2, as in Figure 14-18. The
electric field intensities of the two dipoles add vectorially, according to the
621
P(r,0,@)
principle of superposition (Section 2.2). From Eq. 14-22, the electric field
intensity in the radiation field of an electric dipole situated at the origin is
ta iat i :
oe Are? PAINS
The last two exponential functions can be expanded as power series in s/A for
s << X, and their sum then reduces to (js/X) cos 6. Thus
Figure 14-19. Radiation pattern for a vertical oscillating electric quadrupole at the
origin. The amplitude of E or of H in any given direction is proportional to the
distance between the origin and the outer surface in that direction. The inner
surface is a similar plot of the magnitude of §,,. There is no field along the axis or
along the equator of the quadrupole. The maximum field intensity occurs along
the surface of a cone at 45° to the axis.
dipoles forming the quadrupole radiate energy; there is also zero radiation
along the equator @ = 7/2, where the two dipoles give equal and opposite
fields.
The magnetic field intensity H is found in the same manner from Eq.
14-14 and
€0 1/2
H=(=) Fe: (14-98)
The E and H vectors for electric quadrupole radiation are therefore
oriented in the same manner as those for electric dipole radiation. This was
to be expected, since we have simply superposed the fields of two electric
dipoles. The amplitudes of both E and H are inversely proportional to r and
therefore decrease at the same rate as for dipole radiation. This makes the
14.5 MAGNETIC DIPOLE RADIATION 623
average Poynting vector (1/2) Re (KE X H*) decrease again as 1/r?, which is
necessary for conservation of energy.
There are two main differences between electric dipole and electric
quadrupole radiation: the former field increases with the square of the
frequency, whereas the latter increases as the cube of the frequency; the dipole
field is zero along the polar axis, whereas the quadrupole field is zero both at
the poles and at the equator.
Figure 14-19 shows the radiation pattern for an oscillating electric
quadrupole at the origin.
ee
xX
lea
ae
T= he, (14-100)
but magnetic dipole radiation can also be produced by any current distribution
whose magnetic dipole moment
m = 5LG x J) dr (14-101)
_ wo Mo Xr (A : . 5
A= ae ges (-+) exp jw[t]. (14-102)
You will remember from Sections 6.7 and 10.3 that the Lorentz condition
states that V-A is —euo times dV/dt. Since V = 0 for a magnetic dipole, the
divergence of the above A must be zero. This will be shown in Problem 14-19.
(14-104)
=p
ay exp jw[t] sin @ 0, (r > Xv); (14-105)
14.5 |MAGNETIC DIPOLE RADIATION 625
0A 0A
= wo\ "2 mo K .
= =(2) Aan (exp jw t]) (/Pa 1)sin 6 y, :
(14-107)
Say
== Cuomo Sin? 9 ri,
327 2r2X4 (
14-111 )
=r1,19 one
p24 r; watts /meter’; (14-112)
626 RADIATION OF ELECTROMAGNETIC WAVES
We
ecules
127! (
14-113 )
2
a0)
Tale
5 R UV U
ees > |
NE
(a) (b)
Figure 14-22. Electric and magnetic dipoles used as receiving antennas. The
incident electromagnetic radiation induces an electric field in the wires, and the
resulting voltage U across the resistance R is measured with a suitable electronic
circuit.
We can now form an oscillating magnetic quadrupole with two such loops
parallel to each other on either side of the origin and oscillating in opposite
phases. We shall consider the simplest case, in which the distance between the
loops is equal to the radius a of the loops, as in Figure 14-21, and we shall
consider only the radiation field (r >> 4). The lower dipole is centered at
z = —a/2 and has a moment —mpe’*‘, whereas the upper dipole is centered
at z = +a/2 and has a moment +/me'. This arrangement is purposely
made similar to that of the linear electric quadrupole of Section 14.4.
As for the electric quadrupole, we may add the fields of the two magnetic
dipoles, neglecting the differences in amplitude and in direction, but taking
into account the difference in retardation. This is done by multiplying the
dipole fields by plus and minus exp (ja/A) cos 4, as in Section 14.4. Thus,
1/2
(r >A). (14-119)
LEee) Anake exp jw[t] sin 6 cos 6 6;
Figure 14-22 shows electric and magnetic dipoles used as receiving antennas.
The tangential component of the incident electric field induces currents in the
wires that (a) reradiate energy and (b) produce a voltage 0 across the load
628 RADIATION OF ELECTROMAGNETIC WAVES
U = Es, (14-120)
oA
fpeal Il — f (F + vv) - dl, (14-121)
C c\ ot
where S is any surface bounded by the circuit. The second term on the right
vanishes, since the curl of a gradient is identically zero. Also, the order of the
operations in the first term can be interchanged to give
The induced electromotance in the loop is therefore equal to the rate of change
of the flux linking the loop. It is maximum when the normal to the loop is
parallel to the local B.
The voltage © when R—~ is not necessarily equal to the induced
electromotance, because the circuit may also be excited in the electric dipole
mode. For example, with a symmetrical loop such as that shown in Figure
14-22, if the E vector is parallel to the wires leading to R, charge oscillates
from one end of the circuit to the other, U is not affected by the electric dipole
oscillation, and the above relation is correct. On the other hand, if the E
vector is in the plane of the loop but is perpendicular to the pair of wires, an
extra voltage appears on R that comes from the dipole excitation and adds to
the above induced electromotance.
*R. W. P. King, Handbuch der Physik (Springer-Verlag, Gottingen, 1958), Vol. XVI,
p. 267.
629
Re R,
_ Transmitter Receiver
Figure 14-23. Pair of loop antennas. The one on the left is fed by an oscillator
supplying a voltage U; the other is connected to a load resistance R, and to a
zero impedance ammeter. The current / could also be measured with a high-
impedance voltmeter across Rs. When a current / is drawn from a source such as
an oscillator, the voltage drops at the output by an amount AV. The ratio AU/I
is called the output impedance of the source. It is shown here as the resistance Ra.
R,
Transmitter Receiver
Receiver Transmitter
Field vectors E,,, H,,
Figure 14-24. Pair of loop antennas with the source in a (top), and then in 6
(bottom). The frequency and the impedances are the same in both cases.
part of the same figure. The frequency and the impedances are assumed to be
the same in the two cases.
At any point in space including the loops, and even including the source
in loop a, we first have the mathematical identity
V-(E.
X Hy — Ey X H,)
=H)VXE,— EV X A, — Hav X E,+ Ey-V X Hy. (14-125)
For points outside the source, the current density J, is simply cE, if we
assume that Ohm’s law applies. However, within the source, there is a further
electric field intensity E, (Section 8.5) and, in general,
14.8 THE RECIPROCITY THEOREM 631
In general, the right-hand side is not equal to zero. This relation applies to any
pair of electromagnetic fields at any point in space, even inside the sources.
Let us integrate over all space:
If we now assume that the sources are limited to a finite volume, the surface
of integration on the left-hand side is infinitely remote from them, and we
have a plane wave with E and H orthogonal and transverse:
EXCH — EH 1, (14-133)
where r; is the unit radial vector. This is shown in Figure 14-25. Then
€ 1/2
H, = (2) m1 x E,, (14-134)
Ho
E, XH, — EX
Ho
at points infinitely remote from the sources. We can show that the quantity
within braces is equal to zero merely by expanding it, recalling that ry is
perpendicular to both E, and E,. Therefore the integral on the right-hand side
of Eq. 14-132 must also be zero:
the sources, since E,, and E,, are zero everywhere else. Thus
where the integrals are evaluated over the regions where E,, and E,, are
nonzero.
The meaning of this equation can be illustrated by referring again to our
pair of loop antennas. These are shown in Figure 14-24 for the field a, which
obtains when the source is in a, and for the field b with the source in b.
For these antennas,
where U,a is the voltage supplied by the source in a, and J; j, « is the current in
the same loop a when b is energized. The other integral of Eq. 14-137 can be
expressed similarly, and
Usals IG aa Usola in bs (14-139)
or
ye Oe I, ina
14.9 SUMMARY
E Sieasies
eae j
exp jw[t] sin 6 0, (r > nNX), (14-22 )
H = —-?
AarK2
exp jw[t] sind o (r > 2). (14-14)
A transverse spherical wave therefore radiates away from the dipole, and
the ratio E/H is the same as for a plane wave in free space. Closer in, the wave
is more complex: there is a total of seven terms, and FE has a radial component.
The average Poynting vector is everywhere radial:
634 RADIATION OF ELECTROMAGNETIC WAVES
cpio sin? 6
Sam os Biterne | (14-28)
; COs (5cos 0)
Hie 2nr
ST cexpyoltaesin 6ee (14-71)
i
The radiated power is
W = 73.) Tims watts, (14-82)
where J,,,; is the rms current at the center of the antenna at the point where
it is fed by a source of power. The coefficient 73.1 is the radiation resistance of
the antenna.
Antenna arrays are sets of antennas properly spaced and phased so as to
obtain, by interference, special radiation patterns.
Electric quadrupole radiation is produced by a charge distribution whose
quadrupole moment is a function of the time. The simplest quadrupole
source 1s the linear quadrupole. It is a simple matter to deduce its field from
that of the electric dipole:
the magnetic dipole, this voltage is the electromotance induced by the chang-
ing magnetic flux linking the dipole, plus, in some cases, a voltage due to
electric dipole excitation of the antenna.
The reciprocity theorem states that the current in a detector divided by
the voltage applied at the source remains constant when source and detector
are interchanged, as long as the frequency and the impedances are left
unchanged. This applies to any electromagnetic field and, in particular, to
electric circuits and to antennas.
PROBLEMS
14-1. (a) Calculate the vector potential A at a distance r on the axis of an oscillating
electric dipole of length s.
(b) Show that the result is the same as it would be if the current were localized
at the center of the dipole.
14-2. (a) Show that the phase velocity of the H field of an electric dipole is
u= (1=r zs)@
r
where 7’ = r/X.
(b) Draw a qualitative graph of u as a function of 7’.
14-3. (a) Show that the phase velocities of the r and 6 components ofE for an oscil-
lating electric dipole are, respectively,
1
w= (+55 C,
uw = {he
r’4— 2r”?
where r’ = r/X.
(b) Draw qualitative graphs of u, and of ue as functions of r’.
density to the average mag-
14-4. Calculate the ratio of the average electric energy
netic energy density in the field of an electric dipole for (a) r< i, (b) r = 4,
(Cine:
dipole is radiated
14-5. What fraction of the total power radiated by an electric
between -£45° of the equatorial plane?
r > X is equal to the
14-6. Show that the Poynting vector for the electric dipole at
energy density multiplied by c.
14-7. (a) Show that, in the field of an electric dipole,
6
Exms = 6.71 win,
where W is the radiated power and r > A.
636 RADIATION OF ELECTROMAGNETIC WAVES
14-11. Calculate the electromagnetic potentials V and 4 and the field vectors E and
H for the oscillating linear electric quadrupole directly, without using the field
of the oscillating electric dipole.
PROBLEMS 637
14-12. Show that the Lorentz condition applies to the radiation field of a linear
electric quadrupole for r > X.
14-13. Calculate the Poynting vector for the radiation field of an oscillating linear
electric quadrupole, and show that the total radiated power is
ilefe Se MO
QOxo
294
watts.
14-14. A sealed plastic box contains an antenna that radiates electromagnetic waves.
How could you tell whether the antenna is an electric or a magnetic dipole?
14-15. Show that the electric dipole and magnetic dipole fields illustrate the duality
property of electromagnetic fields (Section 10.8).
14-16. The gain G of an antenna is defined as the ratio of the Poynting vector at the
maximum of the radiation pattern to the Poynting vector averaged over a
spherical surface:
G
Spare
|
a I [7ssinkee
Wie fe
6 aode
The gain of an antenna is a measure of its directivity.
(a) Show that the gain of an electric or magnetic dipole is 1.5.
(b) Show that the gain of a half-wave antenna is 1.64.
14-17. (a) Show that the ratio |E/H| for the field of an electric dipole is much larger
close to the dipole than it is far away.
(b) Show that the inverse is true for the magnetic dipole.
(c) Which type of probe would you use to detect (i) electric dipole radia-
tion, (ii) magnetic dipole radiation, near the source?
14-18. In the field of a magnetic dipole V = 0 and A # O. Is it possible to have a
radiation field such that V # 0 and A = 0?
14-19. (a) Show that V- 4 = 0 for the field of the magnetic dipole.
This is the Lorentz condition, since V is equal to zero.
(b) Show also that
Vase ee
Or
0
for the field of the magnetic dipole.
This is the wave equation for A.
14-20. Two identical magnetic dipoles are perpendicular to each other and have a
common diameter.
(a) Show that the radiation pattern (amplitude as a function of @) is a circle
in the plane perpendicular to the common diameter if one dipole leads the
other by 7/2 radian.
(b) Explain the nature of the resulting field.
(c) How would you connect these antennas to a common source he
Such a pair of crossed coils can be used as an omnidirectional trans-
mitting or receiving antenna.
638 RADIATION OF ELECTROMAGNETIC WAVES
Figure 14-27.
14-21. Mobile radio receivers equipped with the usual vertical whip antennas are
subject to signal fading when they move through the standing wave patterns
that exist near electrically conducting buildings.
Let us see how this effect can be eliminated by using a half-loop antenna
like the one shown in Figure 14-27. The E vector is vertical and the plane of
the loop is normal to the local H.
(a) Show that the sum of the signals at A and B is a measure of E, while their
difference is a measure of H.
(b) Then, from Problem 12-21,
WGKCAN Si Te) <i ICAP = 15)
will be a constant in the standing wave pattern if M/WN is chosen properly.
Show that we must have that
14-23. In the design of loop antennas for satellites, weight and size limitations are of
foremost importance.
Show that, for a given loop diameter and for a given mass of copper, the
ratio wL/R is independent of the number ofturns, L and R being, respectively,
the inductance and resistance of the loop.
14-24. Compare the responses of the electric and of the magnetic dipole antennas
when used as receivers (a) in air and (b) in sea water.
In (b), assume a frequency of 20 kilohertz. Assume that the loop antenna
has a single turn, that its diameter is equal to the length / of the electric dipole,
and that /< i.
14-25. (a) Calculate the ratio 7/0 for the circuit of Figure 14-26, and show that
the reciprocity theorem applies.
(b) Calculate the power expended by the source when the circuit is as shown
and when and / are interchanged.
PROBLEMS 639
Hints
14-9. Think of the images of the charges flowing along the antennas.
s, Novem-
* See, for example, A. Braccesi and L. Forniggini, Astronomy and Astrophysic
ber 1969, page 3064.
APPENDIXES
>>
naa
b "i| A
raped 2
ok
7 | | |
APPENDIX A
Conversion Table
Examples: One meter equals 100 centimeters. One volt equals 108 electromagnetic
units of potential.
SI CGS SYSTEMS
esu emu
z=x+jy, (B-1)
where x and y are real numbers and where j = (— 1)". This quantity can be
represented by a point in the complex plane with x as abscissa and jy as
ordinate, as in Figure B-1. This z must not be confused with the z-coordinate
of Cartesian or cylindrical coordinates.
We can have functions W(z) such as z?, or 1/z, or In z, and so on, and
Zz
(OE yy = ? — y*) 4 2ixy, (B-3)
in which
U=x?— y? and V = 2xy. (B-4)
The function W(z) = U + jV can be represented as a point on another
complex plane with U as abscissa and jV as ordinate. We then speak of the
z-plane and of the W-plane.
We shall consider functions W(z) such that the derivative dW/dz exists
in the region considered. This condition leads to an important pair of equa-
tions. First, let us examine the meaning of the derivative dW/dz by consider-
ing Figure B-2. The point W in the W-plane corresponds to the point z in
the z-plane, according to some specified function W(z). If z changes to
z+ Az, W changes similarly to W + AW, where the increments Az and AW
are complex numbers. The derivative dW/dz is the ratio of these increments
at the limit Az — 0. The value of this derivative can take on different values
for different values of z, but we wish to have a single value of dW/dz for a
given value of z, that is, for a given point in the z-plane, no matter how dz is
chosen.
We consider two particular values of dz: dx and j dy. In the first case,
dz is parallel to the x-axis; in the second case, dz is parallel to the jy-axis.
For both of these particular values of dz, we must have the same value of
jV ; iy
dW /dz. For the first case, the value of dW/dz becomes dW/dx, and
OW SoU en
Ox Ox J Ox ee)
For the second case, dW/dz becomes 0W/j dy, and
dW lou | oV
na cs ce e B-6
LOA CVS OY. oe
These two expressions must be equal:
OUD OV OU ey
ax dx 9 7 oy uy oy ee
Then
oU aV aU OV
ES = ay and ay => mae (B-8)
These are the Cauchy-Riemann equations. The functions U and V are related
to each other through these equations and are called conjugate functions.
A function W(z) is said to be analytic if its four partial derivatives exist
and are continuous throughout the region considered and, moreover, if they
satisfy the Cauchy-Riemann equations.
Consider now a point z and two neighboring points z’ and z” in the z-plane.
These three points correspond to three other points in the W-plane: the point
W and the neighboring points W’ and W’’, as in Figure B-3. Since dW/dz
is unique at the point z, then
W _ W'—Ww rabedl | ghee 2°
ssid=" lim a SaaS lim spills (B-9)
Figure B-3. The three points W, W’, W” correspond, respectively, to z, z’, z’’. The
angles a and § are equal when z’ — z and z’’ — z, as long as dW//dz exists and
is not zero.
THE COMPLEX POTENTIAL 647
* See, for example, E. Durand, Electrostatique (Masson et Cie, Paris, 1964), Vol. 1,
and E. Kober, Dictionary of Conformal Representations (Dover, New York, 1952).
649
ae (B-23)
We have seen above that the stream function is a constant along a line of
force. We shall now see that it is quantitatively just as important as the
potential function. We shall assume, as in Section B.3, that V is the potential
function.
Figure B-5 shows three equipotentials and three lines of force in a por-
tion of an electrostatic field. The vector dn is an element of length along the
line of force normal to the equipotential at the point considered. The vector
ds is an element of length along an equipotential and is oriented with respect
to dn so that it points to the /eff when viewed along dn. For convenience,
we choose our x- and jy-axes as shown, so that the x-axis is parallel to dn
and the jy axis is parallel to ds at the point considered.
According to the Cauchy-Riemann equations,
abe nu OV _ an (B-24)
Os Oy Ox on
Since —0V/dn is the electric field intensity E at the point, the positive direc-
tion for E being along dn, then, along an equipotential,
dU = Eds. (B-25)
650
This relation is valid for any point in the field as long as V is chosen to be
the potential function, and E and ds are oriented as in Figure B-5.
The stream function is thus related to Eds. This quantity Eds is the
flux of E crossing the equipotential in the direction of dn through an element
of area on the equipotential surface that is ds wide and whose height, meas-
ured in the direction perpendicular to the paper, is the unit of length, namely
one meter. Integrating Eq. B-25 along an equipotential between the lines of
force U; and Up,
Us Us
U,— U, = / Ms = / Eds. (B-26)
Ui ds Ui
The line of force for which the stream function is zero is chosen arbitrarily,
just as for the equipotential on which the potential is zero. Thus the charge
density o at the surface of a conductor is
og = gE = — —> (B-27)
where the vector ds points toward the /eft when one looks toward the outside
of the conductor into the field.
The total charge Q per unit length on a cylindrical conductor whose axis
is perpendicular to the paper is obtained by integrating o« ds around the
periphery of the conductor in the direction of increasing ds:
Q= fo ds = fdu. (B-28)
The stream function is thus useful for determining surface charge densities
and total charges on conductors.
THE COMPLEX POTENTIAL 651
: f= = =,
aoe 3
(B-30)
and
je Le ess (B-31)
dz 8
as expected.
We have thus verified our method of calculation by applying it
to a well-known field.
or, writing
Zee (B-33)
eee a (B-35)
l a In ee te
ry; ry
Thus
fic eg: (B-36)
652
eG: (B-37)
r La!
In —
ry
Q inner = — Pe (B-38)
_|aw) | wtlee
=i (B-40)
In —
ry
THE COMPLEX POTENTIAL 653
(B-41)
meeQ
Dare
I i
22 (B-43)
This suggests that
Way i Oren n=
ea
(B-44)
, @ z+a|en
=] Dac In ( a) (B-45)
z—a
és =|
ae 27reo mn
zt+a
re
oe se a |, (B-46)
and
rae 21reo
(= 6); (B-47)
Se 2ré9
OS ana
|Z2—a
ee AgHae2
27€)
(B-48)
The lines of force are given by the equation 6. — 6, = constant.
They are thus arcs of circles passing through the line charges and
centered on the jy-axis, both above and below the x-axis.
The equipotential V is determined by the equation
Jy
3 = acoth ea (B-51)
Since
coth? x — csch?x = 1, (B-52)
then
TT€o :
(B-55)
~ cosh! (D/2R)
It will be observed that C’ depends only on the ratio D/R, and
not on the actual dimensions, just as in the case of the cylindrical
capacitor. When D = 4R, C’ = 21.2 picofarads/meter.
PROBLEMS
B.1 Two cylinders of length / and radius a have their axes parallel and separated
by a distance D in a liquid. A potential difference V is applied between the
cylinders. Calculate the force of attraction for / = 1.00 meter, a = 0.500
centimeter, D = 2.50 centimeters, ¢, = 2.6, and V = 1.00 X 104 volts. Neg-
lect end effects.
B.2 Show that the electric field intensity for a two-dimensional field in the xy-plane
iS
dw|
dz}
where W is the complex potential and z = x + jy.
B.3 Show that the equipotentials defined by Eq. B-49 are the circular cylinders
Induced Electromotance
in Moving Systems*
Example EXPERIMENT 1
E-dl
(V X E), = lim —~—: (C-2)
S—0 S
The voltmeter reading divided by the area of the loop is thus the
component of the curl in the direction of the normal to the loop,
if the loop is small enough. The “‘B-meter”’ can be a cathode ray
tube with the deflection of the electron beam on the tube face cali-
brated in teslas. The stop watch is used to measure the time rate
of change of B. The observers know nothing about the magnetic
field except what they measure with their own instruments.
The laboratory observer measures a uniform magnetic field:
B, = B. He determines its direction and magnitude by observing
the deflection of his electron beam for at least two mutually perpen-
dicular orientations. He also observes that, for each orientation, the
beam deflection is time-independent. Then 0B,/dt = 0. Further-
more, his ‘‘curl-meter” reads zero for all orientations, since there is
no changing flux through the loop, and thus
VY Xi = =e = th (C-3)
When the disk observer points his electron beam in the plane
of the disk, he always records the same deflection, no matter where
he is on the disk. When he points it parallel to the axis of the disk, he
always records zero deflection. He therefore concludes that the mag-
netic induction is uniform and perpendicular to his disk. His value
Bp, is the same as that of the laboratory observer: Bp; = Br = B.
The “‘curl-meter”’ on the disk sees only a constant flux, and
V X Ep. = 0 (C-4)
: OB
V X En = —— = 0. (C-5)
(Bor): = 0, (C-6)
(Boo), = Bo sin wt. (C7)
When he points his beam in the y direction, the x-deflection measures
(Bp»)., and the z-deflection measures (Bp»);:
(Bp»), = 9, (C-8)
(Bp), = Bo cos wt. (C-9)
He finds these same fields no matter where he measures on his disk.
He can describe the field he measures as a uniform field By rotating
with angular velocity w about his x-axis. If he were to compare notes
with the laboratory observer, he would find that By) = Br = B.
What about his ‘“‘curl-meter’’? When the disk observer points
it so that the axis of the loop is in the x-direction, there is no flux
through the loop, and the reading is zero. Then
— WBo»)x
(V X Eps): = Pea ee 0. (C-10)
(F
dt /p2
SS BSo stn. (C-11)
where 5S is the area of the loop. Then, for this orientation,
Example EXPERIMENT 2
Let us now suppose that each observer is given some conducting
wire and told to try to arrange it so as to induce an electromotance
in a closed circuit at rest in his own coordinate system. For the
laboratory observer,
OB,
V x E, a Tar = 0 (C-17)
p E,-dl = 0 (C-18)
for any closed path.
The observer on disk D1 has the same experience, and if D1 is
made of conducting material, there are no eddy currents induced
in it.
Again the situation is different in D2. Let the conductor be a
single loop around the rim of the disk, with a voltmeter connected
in series. The voltmeter can either be on the disk or in a fixed position
in the laboratory and connected to the loop by means of slip rings.
The voltmeter is read both by the observer on the disk and by the
laboratory observer. The former can calculate what the meter will
read by invoking the Faraday law, since he knows dBp:/0r:
VX Ep: = — 7B (C-23)
Now let us calculate what the laboratory observer thinks the
meter on the disk will read according to the Faraday law. He says
that db/dt differs from zero because the circuit is rotating in a time-
independent field, whereas the observer on the disk says that d®/dt
differs from zero because his circuit is in a time-varying field. The
laboratory man calculates di/dt through the rotating loop and gets
i)
OB,
a = (C-26)
C-26
the equation
Va br) Vee a BD) (C-27)
as to the voltmeter reading, but they disagree as to the reason for the
induced electromotance. The laboratory observer says that the mag-
netic field is static but that the magnetic force (uw xX B) on the free
charges in the moving conductor produces the electromotance. The
disk D2 observer, on the other hand, says that the conductor is at
rest but that it is in a time-dependent magnetic field.
Example EXPERIMENT 3
f E-dl = — (C-37)
2
portion of the circuit external to the disk, but they always agree on
the voltmeter reading.
If the whole disk is made of conducting material, the electro-
motance is calculated in exactly the same way. It makes no difference
what integration path we choose from the axle to the rim, as long as
the path is at rest relative to the disk. It is essential that this part of
the path be at rest relative to the conductor, since the charges which
experience the force resulting in the electromotance are, on the
average, at rest with respect to the conductor. The electromotance
is independent of the path in the conductor, since only the radial
components of the path elements contribute to the integral of
(u X B)-dl.
The electromotance may also be calculated for the conducting
disk from the Faraday law. Again we may choose any path in the
moving conductor, as long as it is at rest relative to it. The result
is the same as with the wire discussed above.
Our discussion is strictly valid only for large magnetic fields and
small angular velocities, since centrifugal and Coriolis effects have
been neglected. For practical laboratory purposes, however, this is
not a limitation. The ratio of the magnetic force to the centripetal
force on an electron in a disk rotating with angular velocity w in a
magnetic field B is
Fy, eB
itn (C-38)
The procedure for differentiating cos wf, although elementary, is rather in-
convenient: the cosine function is changed to a sine, the result is multiplied
by w, and the sign is changed. To find the second derivative, the sine is
666 APPENDIX D
changed back to a cosine, and the result is again multiplied by w, but this
time without changing sign, and so on.
Differentiation can be simplified if it is kept in mind that
= Re (jwxe*), (D-5)
= —wXp Sin wf, (D-6)
oa = Re ((ia)me4}, (D-1)
= —w?Xy COS wh. (D-8)
The results are, of course, the same as by the usual method.
We now adopt the following convention: we shall express the quantities x,
dx/dt, d’x/dt?, and so forth, as exponential functions without writing the Re
operator, but with the tacit understanding that only the real part must be used.
Then x = xX cos wt will be written as
X = Xeirt,
7)
and
a = Jwxoe'* = jwx, ae
With this convention, the operator d/dt can be replaced by the factor
jw.
This simplification is so useful that the exponential notation is almost in-
variably used to represent sinusoidally varying quantities, whether they are
mechanical, acoustical, or electrical.
The coefficient before the exponential function can itself be complex.
For example, jxoe”*’ means
A ea aT : b
(a + jb)e' = Va? + b? expj (or+ arc tan :) (D-16)
a+
: ees
jb = Va? + b? exp arc tan ({)
b (D-17)
in the usual manner, on the assumption that a is positive, in order that the
angle chosen lies either in the first or fourth quadrant. Thus
Warning No. 1: If a is not positive, one must be careful to use the proper
angle in the exponent. For example, the argument of (—1 + 2/) is (w — arc
tan 2), not arc tan (—2).
Space-dependent functions can also be represented with the exponential
notation. For example, in an electric field, the vector E is oriented in some
direction in space. It can also be a function of the time; for example
are completed, the resulting expressions are often left in exponential form.
However, if amplitudes and phases are required, the imaginary part is re-
jected, and the result is expressed again as a cosine function.
—— TUT ee
E = X cosat pe
X cos wt i |[2
(b)
D-1b: the applied voltage is equal to the sum of the voltages across
a, B, and 1/y.
This equation can be solved, without using the exponential nota-
tion, as follows. We consider only the steady-state solution and neg-
lect all transient effects obtained by setting the right-hand side equal
to zero. There is, on the left, a sum of three terms involving the
unknown function x, its first, and its second derivatives, with con-
stant coefficients. Since the derivatives of the sine and cosine func-
tions are themselves sine and cosine functions, it is plausible to try a
function of the form
x = Acoswt + Bsin wt. (D-25)
Substituting this expression for x in Eq. D-24, and setting the coeffi-
cient of cos wt on the left-hand side equal to X and that of sin wr
equal to zero, one can determine the coefficients 4 and B. The sum
(A cos wt + B sin wf) is then put into the form C cos (wt + 6), and
the result is as follows:
a a a
jwt
(D-28)
Vy = a”)?+ 8%? exp
exp (are
( arc tan
tan ~ 8”)
=
= Me = as
oo)? ieBa pie (
iL wt ee
arc tan _ oe5)
Gast :
sh ee
APPENDIX E
Waves
then, for any position z in the direction of propagation of the plane wave,
WAVES 671
t— 2 = Constant, (E-3)
cl (E-4)
The quantity u is called the phase velocity of the wave, since it is the velocity
(a) (b)
The quantity X is called the radian length. It is the distance over which
the phase of the wave changes by one radian; this is about \/6, and is there-
fore considerably less than one-quarter wavelength. It turns out that the
quantity that appears in nearly all the calculations is A, and not \. We are
already familiar with the fact that it is preferable to use the circular frequency
w instead of the frequency f. In both cases the intuitively simple quantity,
namely, the wavelength or the frequency, is not the one that is “natural”
from a mathematical standpoint.
With the exponential notation, the wave traveling in the positive direc-
tion along the z-axis is written*
a = ayexp j(wt — kz). (E-10)
Similarly, a wave traveling in the negative direction along the z-axis is
described by
= ao exp j(wt + kz). (E-11)
as can be seen in Figure E-2. When n coincides with the unit vector in the z
direction, n-r = z.
If a plane wave traveling along the z-axis is attenuated, its amplitude
decreases exponentially, and
* We shall use the notation of Eq. E-10. Notice, however, that one could equally well
write
a = a exp {—j(wt — kz)},
since the cosine is an even function. The latter notation is frequently used where it is con-
venient to omit the factor e~’*¢ for brevity. A wave traveling in the positive direction is then
simply written
a= a exp jkz.
673
k, == (E-14)
whereas the quantity k;, which is called the attenuation constant, is such that
the wave is reduced in amplitude by a factor of e in a distance
Sans (E-15)
i= ie (E-16)
F/
eo
= \
2
The element of mass p dz takes on an acceleration a under the action
We have written a partial derivative for 6, since the dé found above was for
a given time ¢. For waves of small amplitude, we can set cos@ = 1 and
6 = dy/dz. Then
Oz
ee
F or
(E-26)
“= (=)". (E-27)
This result is general, and any uwnattenuated wave traveling with a velocity
u along the z-axis is described by the following differential equation:
(E-28)
ge ees
Oa w?
(E29)
0a
——
. = Ss E-30
AC = Gel®, (E-38)
where G and ¢ are real. Also, a must be a real quantity, and we must have
WAVES
677
BD = Ge-ie (E-39)
in order that the first two terms of Eq. E-37 can add up to give a cosine
function.
Similarly, we can set
AD = Fe’ (E-40)
and
BG whes. (E-41)
Finally, a can be written as follows:
where the angles @ and ¢ are constant. Comparing this with Eq. E-5, we find
that the above equation determines a pair of waves traveling in opposite
directions with a common velocity u. The separation constant k is the wave
number, and
ia — ae Dry, (E-43)
Since a is a function only of z and of t, the waves are plane, with wave
fronts parallel to the xy-plane.
It will be observed that the formal solution that we arrived at by sepa-
rating the variables z and ¢ has led us to a very special class of waves; namely,
sine waves. We did not find a general function of {t — (z/u)\. This is quite
disturbing at first sight, since our formal solution is presumably general.
There is no contradiction, however, for the following reason. Since our dif-
ferential equation is linear, that is, since its terms are all of the first degree
in a or its derivatives, the sum of any number of solutions is also a solution.
Any type of periodic wave form encountered in practice can be expressed
as a Fourier series of sines and cosines of the fundamental frequency and
of its harmonics (Section 4.4). Even individual nonperiodic pulses can be
analyzed in a somewhat similar manner by means of Fourier integrals. Any
wave form can thus be synthesized by combining terms of the form shown in
Eq. E-42 with appropriate amplitudes and wave numbers.
rem PEIPIIPPIDIE
LOOTED DOPE
F P1 0 P2 F
be partly reflected and partly transmitted at the knot. The phase velocities are
“(9
respectively
F 1/2
Pl
1/2
(E-44)
Wu = () 5)
p2
(E-45)
a) |
ky tS: \>
Uy 55]
=
Let us assume that a wave travels to the right along string 1. We shall
call this the incident wave and set
Since the ratio yo/yo: is always real and positive, the transmitted wave is
always in phase with the incident wave. On the other hand, the ratio yo,/yo;
can be either positive or negative. The reflected wave is in phase with the
incident wave if p; > p2, and m radians out of phase if p1 < po. If p, = py there
is no discontinuity, no reflected wave, and yo: = oi.
Ah ee 2, (E-53)
0*y Oy
pdz ——
ry 2 «= F cos
6 dé a— bd. ahAt (E-54 )
680 APPENDIX E
or
We can show that this is the differential equation for an attenuated wave
by trying a solution of the form
and the wave number k’ is now obviously complex. The above differential
equation is therefore that of an attenuated wave. This is to be expected, since
it is identical to Eq. E-26, except for the addition of the second term on the
right-hand side. This term corresponds to a damping force which dissipates
energy.
Equation E-55 can also be rewritten as
This equation is similar to Eq. E-30, except that the wave number k’ is now
complex. It represents a pair of attenuated waves traveling in opposite
directions along the z-axis.
It will be observed from Eq. E-57 that the differential equation is equally
well satisfied by +k’ or by —k’, that is, by waves traveling in either direction
along the z-axis.
We can find both k, and k; by recalling Eq. E-18 and substituting
Since, by definition, k, and k; are both positive and real (Section E.1), the
+ signs must all be replaced by + signs, and
The differential equation E-55 that we found above is general, and any
attenuated wave traveling along the z-axis is described by
07"
az?
07a"
oF
hoe (E-65)
We can solve this equation formally by separating the variables, as in Section
E.3. We set
gael ()Z.(2), (E-66)
substitute in Eq. E-65, and divide by T’Z’. Then
1 @Z' gda@T' , hdr’
(E-67)
Fwd a Te dpe Taf
where the left-hand side is a function only of z, whereas the right-hand side
is a function only of ¢. Then
Len F
Say = OK E-6
Li az ee Wea)
BOT) oh al si :
Tier T 7 dt we Cad
GES!
+h = 55 ip a) (E-71)
S dP
We can assume a Sinusoidal wave without losing generality, as was noted in
Section E.3, and set
This result is identical to that which we found in Eq. E-57, the coefficients
g and / of the general differential equation for an attenuated wave (Eq. E-65)
being equal, respectively, to p/F and to b/F in the differential equation for
an attenuated wave on a stretched string (Eq. E-55).
If we again set
ki = ky — jki, (E-74)
“OGY ex
we find that
Ee
where k’ = k, — jk;. The first term represents a plane wave traveling in the
negative direction along the z-axis, whereas the second term represents a
similar wave traveling in the positive direction.
In this general case, the wave velocity w is w/k,, which reduces to (1/g)!/?
when h = 0. The wave is attenuated by a factor of e in a distance 6 = 1/k,,
which approaches infinity as A approaches zero.
These results are the same as those of the preceding section.
The coefficients g and / again determine the wave number as in Eqs. E-75
and E-76. For sinusoidal waves,
a + ka = 0, (E-80)
where k is the wave number. We have omitted the primes that were used
previously to identify the attenuated wave.
If there is no attenuation, / = 0, g = 1/u?, and
1 0a
Ae are (E-81)
for any waveform, u being the phase velocity.
As yet, we have only considered waves in which the quantity which is propa-
gated is a scalar. Vector quantities, such as an electric field intensity KE, for
example, can also propagate as a wave, and then, for a plane wave propagat-
ing in the positive direction of the z-axis,
E = Ey,exp j(wt — kz). (E-82)
where Ey,, Eo,, and Ey. are the components of Ey. These components may
conceivably depend on x and on y, but they do not depend on z, because a
plane sinusoidal wave is characterized by the dependence on ¢ and on z
which is shown in the exponential function.
a= —
a 1 ee,
z
y’, z’)
dr /
(E-87)
as in Eq. 2-19, where
Te (Nae) eC ye ke (E-88)
is the distance between the source point x’, y’, z’ and the field point x, y, z.
If now fis a function, not only of x’, y’, z’ but also of r, we must take
into account the time required for the disturbance to travel from x’, p’, z
to x, y, z at the velocity u, as explained in Section 10.2.1, and thus
a=-_ ey J 2~—
(~,Ley
—_lee“w
‘) (E-89)
This is the solution of Eq. E-84.
The solution of the vector nonhomogeneous wave equation
CE
In many cases the answer is included in the statement of the problem; the list below
contains one half of the remaining answers.
CHAPTER |
CHAPTER 2
Dem) a5.O me VOL metens
(b) 5.1 meters.
V E
IK10,° 15
1OxX10 16
Sal 9 Si
E
V
_ 2poa*
CD) ees 15er?
_ 2p9a'
ee 15eor
= Pe ery
(©) ae €0 3 =)
(e)
0 2 3 4 5
Figure A-2. Solution to Problem 2-21e.
ANSWERS 687
2-24. (4/3)ma%ao.
2G
Waite Arreor’
Vor 03
De
7 967reor®
ieay
(c) Since 3n? — 1 is of the order of unity,
V3
alge
1
V7, < MES:
i00 lopip >2a Det.
2Tr€9
* In (6/a)
(CQ. — C01)"
Ze CCAG + C)
(b) The energy is dissipated as heat in the wires connecting the capacitors.
3e? e
2-34. (a) Tine (b) Srregmc?
2-36. (a) 150 kilovolts. (b) About 4 « 10~4 atmosphere.
2-39. 70 2‘
(a) in G/a) (b) The force is due to the fringing field.
CHAPTER 3
3-1. (a) p 5.7 & 107%? coulomb-meter.
(b) s = 35.9' X 10-2 meter.
3-6.
} E D
|
1000 ~=10x10° 110x107)
\.- £ without dielectric
VA
é VA, —V without dielectric
500. Seal Sx10"* ~~
NS
r (centimeters)
(b) «AR a = [ta + (tao + aCOS wt)]o, — taoa, Where A is the area of the film,
t, its thickness, and o, its charge density; o is the charge density on the
electrodes; and t., + a cos wt is the thickness of the air gap. Note that
AR(dcz,/dt) is the output voltage.
3-17. (b) Assuming a 1 HP motor operating for one hour at 100% efficiency, the
capacitor would have a volume of the order of 10 meters* and would
weigh about 10 tons.
3-24. 44 micrometers.
CHAPTER 4
4-4. (a) m0. 8 (c) Forn Ym, EV /(re — ri) is large. For mn Kr,
E is large at the surface of the small inner sphere.
(CRON <0) 323 (f) 4.15 X 105 volts/meter.
4-8. 2.4 104 volts/meter.
4-24, No.
Electron source
on
|
|
|
|
|
|
|
Me
Figure A-4. Solution to Problem 4-28e.
ANSWERS 689
CHAPTER 5
Sail, ayes
5-18. Dme*
|
|
(a) I
|
|
= .
Explosion Time
=Imyc
(b)
|
|
We Time A
Explosion Figure A-5. Solution to Problem 5-18.
CHAPTER 6
6-3. (i) 2.31 < 10°” newton.
(ii) 7.80 < 10~*% newton.
690
6-8. (d)
155
; B
(a) (b)
(d) Yes:
dr Cc
ae 10! at x = 400 meters.
G ©
ANSWERS
691
(e) No.
(f) We have neglected end effects on the cylinders of charge.
GHAPTER 7
7-4. 0.234 newton.
7-10. (a) Perpendicular to the plane passing through the two axes.
(b) Yes.
21245 i
7-13.
T a
7-19. V = (e/m)B*(b? — a®)*/8b2, where e and m are the charge and mass of the
electron, a is the radius of the cathode, and 4 that of the anode.
OB awe 3pola’z
7-29. (a) “Op a 4a? + z2)5/2
3yola*zp
(0) By = Hat + Z25/2
OB, 3uUola*p
gee Gl aye
2 3(a2 — 42”)?
(@ B= xe
5 tle
+ ml t aa sxeh
(1 +
73a. (C) 1.15 AO tesla,
(e) 4.83 & 10-§ ampere/meter?.
(f) 6.14 * 10 ampere.
CHAPTER 8
8-1. 1.2 millivolts.
mroLdE , trr’aR V
is E=-—>
ee l at I /
where L is the inductance, o is the conductivity of the wire, r is its radius, and
/is its length.
692
8-8. (b)
(c)
8-17. L decreases with increasing frequency until the skin depth becomes small com-
pared to 6 — a.
8-71. eT (Ge) p
8-33. (c) Both are equal to the energy density and proportional to the square of the
magnitude of the field.
The electric force is directed into the field; the magnetic force is directed
away from the field.
The electric force per unit area is normal to the surface of a conductor and
is equal to €)?/2 only if the field is static.
The magnetic pressure is normal to the surface of a conductor and equal
to B®/2uo only if there is zero field inside the conductor. This condition is
achieved exactly with superconductors, and approximately if the frequency
is high.
ANSWERS
693
CHAPTER 9
9-5. “2.0.
9-8. (a) E is reduced.
D is unaffected.
(c) Minimum.
(d) B is increased.
H is unaffected.
(f) Minimum.
9-20. Use a circuit similar to that of Figure 9-15 with A,, = 2.5 centimeters? and
L,, = 8.84 centimeters (each magnet is 4.42 centimeters long).
CHAPTER 10
-yh/sa (2<4); a
2
Vie
Ae al
The other two equations (V-B = OandV X E = —0B/0f) lead to identities.
10-9. (c) B = pod.
(d) The magnetic pressure is poA?/2.
(f) Currents are induced which maintain ® constant.
10-12. (Cc) EV = E; WV = dH.
CHAPTER 11
20 kilohertz 20 megahertz
Sea Water 5.4 170
Copper 1.9 x 104 5.9 < 105
c
2 @ CE PE = Py
1/2
(b) u, =c {1- (2“
(e) um = 3.24 108 meters/second.
u. = 3.23 & 108 meters/second.
u, = 2.78 X 108 meters/second.
(f) 2.73 & 10% meters; 45 waves.
CHAPTER 12
(b) FF
0.5
0.3
0.1
2 Ss ee ee ee
eee
12-20.
CHAPTER 13
13-11. The skin depth is 1.5 microns; the coating should have a minimum thickness
of about 5 microns.
13-15. (a)
Reflecting surface
13-19. Both methods can, in fact, be used. See, for example: Harvey, Microwave
Engineering (New York: Academic Press), p. 144.
13-23. } = A,-~o. This current flows inside the conductor, very close to the surface.
13-27.
Attenuation
No /2b (n = 2)
Jest a in Cee oe La a a
Type |e
a inten Uicain Wee iiemtin eae
CHAPTER14
Mee ay
De
?
fee ose ee ee
|
|
Figure A-14. Solution to
Problem 14-2b.
1 + 3 cos? 6 A?
14-4. (a)
Sift Ue eer
1 + 7 cos? 6
(b) 2 sin? 6
(cae
14-18. No.
In sea water,
Ve 4
Vinag 1 (opow)!!?
For a frequency of the order of 104 (see Section 12.5.1.
APPENDIX B
B-1. 6.44 * 1074 newton.
Index
D]
—_ rs
o ya ¢ .
eae
Y°Cc $9
r. 1 SiGe b
¢
i ile
Q
a.
Vv 7
+ |
B>VxA
—_
(n
O OPas= Uo.t
. F: YXA
,
os Py * Sin 4
vA
R=
Z°
x
Bienens R dk dodz
Z
dQ = dkR + RAPF + 42
dS * Rdo dz
es Spheri cot
X= 15in 9 coo?
Mean
a rans sin?
ali : R°sin@ skaeds
dle dr Rrrde Gr rsinode $i
ASunfiee 2R™ sin@ doa’
OK
(4 > sm @ cosh sine Sgr ws
Examples: One meter equals 100 centimeters. One volt equals 108 electromagnetic
units of potential.
SI CGS SYSTEMS
esu emu