Solved Exercises in Electromagnetism - Mastnak
Solved Exercises in Electromagnetism - Mastnak
Elijan J. Mastnak
Winter semester 2020-2021
Last update: October 10, 2022
Faculty of Mathematics and Physics, University of Ljubljana
Contents
1 First Exercise Set 4
1.1 Electric Field of a Charged Disk . . . . . . . . . . . . . . . . . . . . 4
1.2 Charged Plate with a Slit . . . . . . . . . . . . . . . . . . . . . . . . 5
The electric field of a spatial charge distribution with volume charge density ρ(r) is
˚
1 ρ(r̃) d3 r̃ r − r̃
E(r) = 2 ,
4πϵ0 V |r − r̃| |r − r̃|
dE z z σr dϕ dr
=√ =⇒ dE = 2 2 )3/2 4πϵ
.
dE1 2
z +r 2 (z + r 0
We find the total electric field E(z) by integrating over the contributions dE:
ˆ ˆ 2π ˆ a ˆ a
σzr σz r dr
E(z) = dE = dϕ dr = .
0 0 4πϵ0 (z 2 + r2 )3/2 2ϵ0 0 (z 2 + r2 )3/2
4
1.2. Charged Plate with a Slit
a2
For z ≪ a (very close to the disk), we have 1 + z2
→ ∞ and 1
2 → 0, leaving
1+ a2
z
σ
E(z) → (z ≪ a),
2ϵ0
which is the electric field of an infinite charged plane.
2
For z ≫ a (very far from the disk), we have az 2 ≪ 1 and use the Taylor approximation
(1 + x)p ≈ 1 + px for x ≪ 1 to get
" −1/2 #
a2 a2 σa2
σ σ
E(z) = 1− 1+ 2 ≈ 1− 1− 2 = .
2ϵ0 z 2ϵ0 2z 4ϵ0 z 2
Finally, we multiply above and below by π to match the above result to the expression
for a point charge:
πa2 σ σS q
E(z) = 2
= 2
= (z ≫ a),
4πϵ0 z 4πϵ0 z 4πϵ0 z 2
where q = σS is the disk’s charge.
Because of mirror-image symmetry, both the x and y components of the electric field
cancel, leaving only the z component dEz . We then relate dEz and dE1 using similar
triangles
dEz z σz dr
=√ =⇒ dEz = .
dE1 2
z +r 2 2πϵ0 (z 2 + r2 )
5
1.2. Charged Plate with a Slit
We find the total electric field along the z axis by integrating over the contributions
dEz of all the ribbons. Because of mirror symmetry, we need to calculate only the
contribution of e.g. the right plane and multiple the result by two.
ˆ ˆ ∞ ˆ
σz ∞ dr
σz dr
E(z) = dEz = 2 2 2
=
a/2 2πϵ0 (z + r ) πϵ0 a/2 z 2 + r2
∞
σz 1 r σ hπ a i
= arctan = − arctan .
πϵ0 z z a/2 πϵ0 2 2z
σ π π 2z 2σ z
E(z) ≈ − − = .
πϵ0 2 2 a πϵ0 a
6
2.1. Theory: The Poisson Equation and the Fourier Transform
where U (k) is the amplitude of the plane wave with wave vector k. To find U (k) we
take the inner product of both sides of the above equation with the basis function
e−ik̃·r , which gives
˚ ˚ ˚
U (r)e−ik̃·r d3 r = d3 k d3 rU (k)ei(k−k̃)·r .
The integral over r on the right-hand side is in fact a delta function, because the
orthogonal plane waves cancel out over all space except at the origin, where they
constructively interfere to infinity. Recognizing the delta function simplifies the
equation to
˚ ˚
−ik̃·r 3 3
U (r)e d r = (2π) U (k)δ(k − k̃) d3 k = (2π)3 U (k̃).
The delta function suppresses the integral everywhere except at (k − k̃), which leads
to the expression for the amplitude U (k̃) of a wave vector k̃:
˚
1
U (k̃) = U (r)e−ik·r d3 r.
(2π)3
1
In general any equation of the form ∇2 f (r) = g(r) is called a Poisson equation.
7
2.2. Poisson Equation for a Point Particle
which is a Fourier transform of the function U (k)ik. In other words, the gradient op-
erator ∇ transforms to multiplication by ik under the Fourier transform. Analogously,
the Laplacian ∇2 transform into multiplication by (ik)2 = −k 2 .
Finally, we consider the behavior of the delta function under the Fourier transform.
Let δ(k) denote the amplitude in the expansion of δ(r), analogous to the relationship
between U (k) and U (r). Using the inverse Fourier transform and the integral
properties of the delta function produces
˚
1 1 −ik·0 1
δ(k) = 3
δ(r)e−ik·r d3 r = 3
e = .
(2π) (2π) (2π)3
8
2.2. Poisson Equation for a Point Particle
The plan is to transform into k space, solve for U (k), then transform back to U (r).
First, we take the Fourier transform of both sides—use the Fourier transform identities
∇2 → −k 2 and δ(r) → (2π) 1
3 from the theory section.
q 1 q
−k 2 U (k) = − =⇒ U (k) = .
ϵ0 (2π)3 (2π)3 ϵ0 k 2
Next, we integrate first over θ, (to avoid ei cos θ·∞ from the upper k limit) and recognize
the sine function in the difference of exponential functions:
ˆ ∞ 1 ˆ ∞ ikr
q ei cos θkr q e − e−ikr
U (r) = dk = dk
(2π)2 ϵ0 0 ikr θ=−1 (2π)2 ϵ0 0 ikr
ˆ ∞ ˆ ∞
q 2 sin(kr) 2q
= dk = sinc(κr) dk
(2π)2 ϵ0 0 kr (2π)2 ϵ0 0
9
2.3. Theoretical Interlude: Electric Field of a Charge Distribution
This result is quite powerful—by solving the Poisson equation for a delta function
and then expanding an arbitrary ρ(r) in terms of the delta function, we now have
the solution to the Poisson equation for any ρ(r). We find the corresponding electric
field with ˚ 3
1 d r̃ρ(r̃) r − r̃
E = −∇U = ,
4πϵ0 |r − r̃|2 |r − r̃|
which agrees with the equation quoted in the previous exercise set.
q e−αr αr 2
U (r) = 1+ , α= .
4πϵ0 r 2 rB
ρ(r)
∇2 U (r) = − .
ϵ0
We then calculate the Laplacian of our U (r) and work in spherical coordinates, since
the potential is spherically symmetric (depends only on r). As a review, when acting
on a function that depends only on r, ∇2 in spherical coordinates reads
1 ∂ ∂
∇2 = 2 r2 .
r ∂r ∂r
qα3 −αr
ρ(r) = − e .
8π
10
2.4. Electric Field of a Hydrogen Atom
Note that the charge density is negative, which corresponds to the negatively charged
electron cloud. Inserting the definition of α = r2B gives
q − r2r
ρ(r) = − 3 e
B .
πrB
− 2r − r 2
Another interpretation: e rB is equivalent to e rB , which is the square of the
hydrogen atom’s ground state wave function. The square of the wave function is
probability, and multiplying the probability by rq3 gives a charge density.
B
Note that—incorrectly—the proton’s charge does not contribute to our expression for
ρ(r). This is because the proton occurs at the origin, which corresponds to a charge
density singularity at the origin. To avoid dealing with this singularity, we simply
ignored it when evaluating the Laplacian ∇2
We can resolve this problem by separately considering the special case
q
lim U (r) = .
r→0 4πϵ0 r
which just says the that potential should approach the ∼ 1/r proton potential at the
origin. We would then have to solve the Poisson equation for this potential. However,
we already now the solution—the potential is the potential for a point charge and
corresponds to a charge density
ρ(r) = qδ(r).
The correct total result for the hydrogen atom is the sum of the electron cloud result
and the charge density of the proton nucleus:
qα3 −αr
ρ(r) = qδ(r) − e .
8π
Lesson: Be careful when working with the Poisson equation if U (r) has singularities!
11
3.1. Conducting Ribbon in a Parallel-Plate Capacitor
ρ(r) ∂2 ∂2 ∂2
∇2 U (r) = − where ∇2 = + + .
ϵ0 ∂x2 ∂y 2 ∂z 2
Often ρ(r) = 0 in places we’re solving for the electric potential. In this case ∇2 U (r) =
0. This equation is called a Laplace equation.
∇2 U (x, y) = 0.
Charge can occur only along the ribbon or on the capacitor plates.
Next, we determine boundary conditions for U (x, y) so the equation has a unique
solution; the problem’s boundaries are the ribbon and edges of the capacitor plates.
On the bottom plate, U (x, 0) = 0. On the upper plate, U (x, a) = 0; both potentials
are zero because the plates are grounded, while along the ribbon we have U (0, y) = U0 .
We need one more boundary—infinity. At infinity, we require only that U (x → ∞, y)
is bounded, i.e. that U does not diverge at ∞.
First, we attempt solving the problem with separation of variables and write U (x, y) =
X(x)Y (y)—this approach tends to work well with symmetric problems. Substituting
12
3.1. Conducting Ribbon in a Parallel-Plate Capacitor
′′ ′′ X ′′ Y ′′
X Y + XY = 0 =⇒ =− .
X Y
The separation of variables is successful—we succeeded in isolating x-dependent and
y-dependent terms on different sides of the equation.
We then set the equation equal to a separation constant κ2 and get two equations,
X ′′ − κ2 X = 0 and Y ′′ + κ2 Y = 0.
Both equations have simple solutions! The equations for X and Y are solved by
exponential and sinusoidal functions, respectively.
We then substitute the expressions for X and Y back into the ansatz U = XY :
0 ≡ U (x, 0) = 1 · (0 + D) =⇒ D = 0.
where we’ve joined the product of two constants into a single constant F = BC.
We have two options: either F = 0 or sin(κa) = 0. The option F = 0 gives the
trivial solution U (x, y) = 0. The non-trivial solution comes from
sin(κa) = 0 =⇒ κa = nπ, n = 1, 2, 3, . . . .
13
3.1. Conducting Ribbon in a Parallel-Plate Capacitor
On the right hand side, with the sum, we switch the sum and integral to get
∞ ˆ a nπy mπy ∞ ˆ a mπy
X X Fm a
Fn sin sin dy = Fn δmn sin2 dy = .
0 a a 0 a 2
n=1 n=1
Because of the orthogonality of the sine functions, the integral is zero for m =
̸ n,
and only the case m = n gives a non-zero result. Equating the two sides gives
the desired expression for Fm :
U0 a Fm a 2U0
1 − (−1)m = 1 − (−1)m .
=⇒ Fm =
mπ 2 mπ
Some limit cases: for x ≫ a, the exponent terms very small, and we can neglect all
πx
terms in the series except the first term e− a with n = 1. The result is
4U0 πx πy
U (x, y) = exp − sin .
π a a
A separate case for which we can find a nice analytic solution is at the center of the
capacitor at y = a2 . The solution reads
∞
a 2U0 X 1 − (−1)n nπx nπ
U (x, 2) = exp − sin .
π n a 2
n=1
Instead of finding U (x, a2 ), we’ll find the electric field E(x, a2 ). Because of reflection
symmetry across the line y = a2 , the electric field cannot have a y component—E
only has an x component. We’ll find Ex (x) from the potential:
∞
∂ 2U0 X nπx nπ
U (x, a2 ) = − 1 − (−1)n exp −
Ex (x) = − sin .
∂x a a 2
n=1
14
3.2. A Halved Conducting Cylinder
n even
0
n
nπ
1 − (−1) sin = 2 n = 1, 5, 9, . . .
2
−2 n = 3, 7, 11, . . . .
15
3.2. A Halved Conducting Cylinder
Following the usual separation procedure, we set both sides equal to the separation
constant m2 . The equations for Φ and R read
Now, our cylindrical problem is periodic in ϕ with period 2π—this just means the
cylinder repeats after one revolution. Periodicity in ϕ is possible only if m takes on
integer values, so we can immediately index the solutions for Φ with
We only use positive integers because the odd/even symmetry of sin and cos means
negative integers give the same result as positive one—solving for negative m would
be redundant. We reject m = 0 because this solution leads to Φ′′ = 0, meaning Φ is
a linear function. But a linear function can’t be periodic in ϕ, so we reject m = 0.
The second equation for R is solved with powers of r. The result is
Rm (r) = Cm rm + Dm r−m .
Note: We ended the problem at this point (ran out of time) and continued in the
fourth exercise set.
16
4.1. A Halved Conducting Cylinder (continued)
for the potential inside the cylinder. To find a solution specific to our problem, we
apply boundary conditions. We already applied the periodic boundary condition
U (r, ϕ) = U (r, ϕ + 2π), which required m be integer-valued.
A second boundary condition requires the capacitor halves have a potential difference
U0 between them. It is best to write this potential difference in the symmetric form
(
U0
ϕ ∈ (0, π)
U (a, ϕ) = 2 U0
− 2 ϕ ∈ (π, 2π).
There is another condition—that U does not diverge at r = 0. This condition implies
the Dm coefficients are zero, because the Dm r−m term diverges at r = 0.
Observation: the second boundary condition is an odd function of ϕ. This implies
that only odd (sine) terms can appear in the final solution. This allows us to set the
Am coefficients equal to zero to eliminate the cosine terms. We are left with
∞
X
U (r, ϕ) = Fm rm sin(mϕ),
m=1
To solve this, we take the scalar product of the equation with sin(nϕ):
ˆ 2π ˆ 2π X
∞
U (a, ϕ) sin(nϕ) dϕ = Fm am sin(mϕ) sin(nϕ) dϕ.
0 0 m=1
17
4.1. A Halved Conducting Cylinder (continued)
Next, some limiting cases. It will be easier to work in terms of electric field instead of
potential. We will find the electric field in the two planes parallel and perpendicular
to the slit between the capacitor halves.
First, in the perpendicular (vertical) plane. The field points from high to low potential,
so from the top half of the capacitor to the bottom half. In this plane we can work
with just one coordinate r, which represents the vertical distance from the cylinder’s
center. Note that ϕ = π2 . The component Er we’re after is
∞
U0 X r n−1 1 1 − (−1)n
∂
Er = − U (r, ϕ) =− n sin(nϕ)
∂r ϕ= π2 π a a n ϕ= π2
n=1
∞
U0 X r n−1 1 − (−1)n
nπ
=− sin .
πa a n 2
n=1
n even
1 − (−1) n nπ 0
sin = 2 n = 1, 5, . . .
n 2
−2 n = 3, 7, . . . .
18
4.2. Conducting Sphere in a Uniform Electric Field
As before, the sum simplifies considerably to a geometric series. The result turns out
to be
2U0 1
Eϕ = − .
πa 1 − r 2
a
Note that Eϕ diverges at r = a. This is a consequence of the very small slit spacing
between the capacitor halves at r = a; schematically have E = Ud0 → ∞ as d → 0.
∇2 U (r, θ) = 0.
We then separate variables via U = R(r)Θ(θ), which leads to the general solution
∞ h
X i
U (r, θ) = Al rl + Bl r−(l+1) Pl (cos θ),
l=0
This is the potential of a uniform electric field (at infinity, the potential from the
sphere is negligible). This potential is chosen so that
∂U
− = E0 ,
∂z
i.e. so that the potential at infinity recovers the uniform electric field E0 .
We’ll start with the second boundary condition at r → ∞. Applying the condition to
the general solution gives
∞
X
Al rl Pl (cos θ) = −E0 r cos θ.
l=0
Note that the r−(l+1) terms vanish as r → ∞. The entire series equals a single
term proportional to cos θ and the equality holds if only the l = 1 term in the series
19
4.2. Conducting Sphere in a Uniform Electric Field
is non-zero, which generates a corresponding cos θ term from P1 (cos θ) = cos θ ie.
P1 (x) = x. The l = 1 term is
Next, the second boundary condition: U (a, θ) = 0. Substituting the condition into
the intermediate solution gives
∞
X
E0 a cos θ = Bl a−(l+1) Pl (cos θ).
l=1
Again, the entire series equals only a single term. Again, this will be the l = 1 term
corresponding to P1 (cos θ) = cos θ. For l ̸= 1 we have Bl = 0. The l = 1 term gives
E0 a3
U (r, θ) = −E0 r cos θ + cos θ.
r2
The first term, −E0 r cos θ, is the potential of the uniform external field E0 . The
second term comes from the sphere. In fact, this second term has the same form as
the potential of an electric dipole!
Limiting cases are discussed in the next exercise set.
20
5.1. Conducting Sphere in a Uniform Electric Field (continued)
∂U
E⊥ = − = E0 cos θ + 2E0 cos θ =⇒ σ = 3ϵ0 E0 cos θ.
∂r r=a
To find dS, we find the area of a small band of width da around the sphere’s surface.
The band’s area is 2πr da = 2π(a sin θ)(a dθ). The dipole moment pez is then
ˆ π ˆ 1
2 3
pez = (3ϵ0 E0 a cos θ) · 2π(a sin θ)(a dθ) = 6a πϵ0 E0 cos2 θ d[cos θ]
0 −1
1
3 1
= 6a πϵ0 E0 cos3 θ = 4πϵ0 E0 a3 .
3 −1
21
5.2. Electric Dipole in a Conducting Spherical Shell
We begin with the simpler second boundary condition, (the boundary r → 0 eliminates
the rl -dependent term). Inserted into the general solution, the second condition reads
∞
X pe cos θ −2
U (r → 0, θ) = Bl r−(l+1) Pl (cos θ) = r .
4πϵ0
l=0
Note that the entire series sums to only a single term; for this to work, only the l = 1
term in the series can be non-zero, leaving
pe cos θ −2 pe
B1 r−2 cos θ = r =⇒ B1 = and Bl̸=1 = 0.
4πϵ0 4πϵ0
Note the use of P1 (cos θ) = cos θ. The intermediate solution at this stage is
∞
X pe −2
U (r, θ) = Al rl Pl (cos θ) + r cos θ.
4πϵ0
l=0
As before, only the l = 1 term can be non-zero to satisfy the equality. The result is
pe −2 pe
A1 a cos θ = − a cos θ =⇒ A1 = − and Al̸=1 = 0.
4πϵ0 4πϵ0 a3
22
5.3. Point Charge Above a Conducting Plane
23
5.3. Point Charge Above a Conducting Plane
Considering both points, the potential at an arbitrary position r from the origin is
q 1 q 1
U (r) = − ,
4πϵ0 |r − d| 4πϵ0 |r + d|
where the vector d points perpendicularly up from the plane toward the positive
charge. Introducing an angle θ between d and r, we have
p
|r ± d| = r2 + d2 ± 2rd cos θ.
The expression for U (r) for the two charges is then simply
q 1 1
U2 (r) = √ −√ .
4πϵ0 r2 + d2 ± 2rd cos θ r2 + d2 ± 2rd cos θ
For the original configuration of a single positive charge a distance d above the plane,
the potential above the plane agrees with U2 , while the potential below the plane,
where there is in reality no charge, is zero. The correct expression for the single
charge +q is then (
U2 (r) above the plane
U (r) =
0 below the plane.
Next, we’re interested in the surface charge density σ(ρ) on the plane where ρ is the
radial distance in the plane from the origin. As usual, we start with Gauss’s law for
a small surface element of the plane:
dq dq
−E⊥ dS = =⇒ σ ≡ = −σE⊥ .
ϵ0 dS
To find E⊥ , we differentiate U with respect to the vertical coordinate z. First, we
introduce z into the expression |r ± d|
p p
|r ± d| = r2 + d2 ± 2rd cos θ = ρ2 + z 2 + d2 ± 2dz.
We then have
" #
q 1 1
U (ρ, z) = p −p .
4πϵ0 ρ2 + z 2 + d2 − 2dz ρ2 + z 2 + d2 + 2dz
With surface charge density σ known, we then ask what is the total charge on the
plane. Integrating the plane over rings with area dS = 2πρ dρ, we have
¨ ˆ ∞
qd
qplane = σ dS = − (2πρ dρ).
0 2π(ρ + d2 )3/2
2
24
5.3. Point Charge Above a Conducting Plane
Summary of what we did: recognize that the field above the plane from the positive
charge looks like half the field of an electric dipole. Since we know the solution
for a dipole, instead of solving the charge-plane system, we solve the (imaginary)
two-charge system, which gives the same field above the plane anyway. We then reuse
the upper half of the dipole solution for the single-charge plane system, and set the
field below the plane equal to zero. The basic idea is: the field above the plane is the
same for both the positive-charge plane system and for a dipole system, so we can
use either approach to solve for the field above the plane.
Next, we ask what is the electrostatic force on the point charge above the plane?
First, some theory: Theory: Electrostatic Force
The total electrostatic force F acting on the charges enclosed in a region of space V
permeated with an electric field E is
‹
1 2
F = ϵ0 E(E · n̂) − E n̂ dS,
∂V 2
where n̂ is the normal to the surface ∂V enclosing the charges. Like with Gauss’s
law, a good choice of the boundary surface, usually taking advantage of the problem’s
symmetries, tends to simplify the problem. Alternatively, if the electric field vanishes
at infinity, we choose a surface that closes at infinity.
Back to our problem: we choose an infinite surface whose base runs along the plane,
then turns upward and closes at infinity to enclose the upper half of space above the
plane. The field in this case is the same E⊥ calculated above:
q d
E⊥ = .
2πϵ0 (ρ + d2 )3/2
2
For the part of the surface running parallel to the plane, the normal to the surface n̂
points perpendicularly into the plane, parallel to the electric field. The force equation
for the bottom half of the surface reads
¨ ¨
1 ϵ0
Fe = ϵ0 E 2 n̂ − E 2 n̂ dS = n̂ 2
E⊥ dS.
bottom 2 2 bottom
In fact, the contribution from the upper half of the surface is zero—the upper half
extends to infinity, where the electric field vanishes. We only need to integrate over
the bottom of the surface, running parallel to the plane. Writing dS = 2πρ dρ and
substituting in the expression for E⊥ , the force reads
ˆ ˆ ∞
ϵ0 ∞ q 2 d2 q 2 d2 ρ
Fe = n̂ 2 2 2 3
2πρ dρ = n̂ dρ
2
2 0 4π ϵ0 (ρ + d ) 4πϵ0 0 (ρ + d2 )3
2
ˆ ∞ ∞
q 2 d2 du q 2 d2 1 q2
= n̂ 3
= − n̂ = n̂.
8πϵ0 d2 u 16πϵ0 u2 d2 16πϵ0 d2
25
5.3. Point Charge Above a Conducting Plane
The force points in the direction of n̂—downward into the plane. A final note: if we
write
q2
Fe = n̂,
4πϵ0 (2d)2
the force takes the form of the electric force between a positive and negative charge
separated by a distance 2d—the same situation we used in the method of images
above.
26
6.1. Force on a Conducting Spherical Shell
27
6.2. Point Charge Between Two Conducting Plates
In spherical coordinates, the unit normal n̂ to the sphere’s surface is n̂ = (sin θ cos ϕ, sin θ sin ϕ, cos θ).
The surface element dS at the surface r = a is (just like in the previous exercise sets)
dS = a2 dϕ sin θ dθ. The force on the sphere’s upper half is then
ˆ ˆ
sin θ cos ϕ
9ϵ0 E02 π/2 2π
Fe = cos2 θ sin θ sin ϕ (a2 sin θ dθ dϕ).
2 θ=0 ϕ=0 cos θ
Both the x and y components will be zero—integrating cos ϕ and sin ϕ over a full
period 2π give zero, while the ϕ contribution to the z component is 2π. We make
this explicit with
ˆ
0
9ϵ0 E02 π/2
Fe = cos2 θ 0 (a2 sin θ dθ).
2 θ=0 2π cos θ
The non-zero z component Fz is
ˆ ˆ π/2
9ϵ0 E02 π/2 3 2 2 2
Fz = 2π cos θ(a sin θ dθ) = 9πϵ0 E0 a cos3 θ(sin θ dθ)
2 θ=0 θ=0
ˆ 1
9πϵ0 a2 E02
= 9πϵ0 E02 a2 cos3 θ d[cos θ] = .
0 4
The vector force can be written simply as
9πϵ0 a2 E02
F= ẑ.
4
In other words, the force on the upper half points upward in the positive z direction.
28
6.2. Point Charge Between Two Conducting Plates
where p is the electric dipole moment and Q is the quadrupole moment tensor.
Note: we could think of the charge q is a scalar monopole moment, creating a
logical progression from scalar monopole moment to vector dipole moment to tensor
quadrupole moment.
We find the dipole moment with
˚
p= d3 r̃ρ(r̃)r̃.
Note that both definitions produces a symmetric tensor. Also important: the tensor’s
trace—the sum of the diagonal elements is zero:
X X 2
qn 3x2n − rn2 + 3zn2 − rn2 + 3zn2 − rn2 = qn 3rn − 3rn2 = 0.
tr Q =
n n
The other diagonal terms Qyy and Qzz will analogously sum to zero.
All off-diagonal terms with a z component are also zero, since the charges lie in a
plane with z = 0. We thus have Qxz = Qzx = Qyz = Qzy = 0. We have just two
29
6.2. Point Charge Between Two Conducting Plates
terms left calculate: Qxy and Qyx . By the tensor’s symmetry, the two are equal, so
we really only have one term:
N
X
qn 3xn yn − 0 · rn2 = 3qa2 + 3(−q)(−a2 ) + 3qa2 + 3(−q)(−a2 )
Qxy =
n=1
= 12qa2 .
12qa2 0
0
Q = 12qa2 0 0 .
0 0 0
30
7.1. Theory: Magnetic Vector Potential and the Biot-Savart Law
∇ · B = 0.
This equation rules out the possibility of magnetic monopoles and allows B to be
written as the curl of a vector potential A as
B = ∇ × A.
∇ × B = µ0 j.
∇2 A = −µ0 j,
where j is the current density vector. This equation is a vector analog of the Poisson
equation ∇2 U = − ϵρ0 from electrostatics. Similarly to how the electrostatic potential
U at a point r in region of space with charge density ρ is found with
˚
1 ρ(r̃) d3 r̃
U= ,
4πϵ0 |r − r̃|
where t̂ is the unit normal vector tangent to the conductor, I is the current through
the conductor at the point r̃ and d˜l is a small distance element along the conductor’s
31
7.2. Magnetic Field of a Circular Current Loop
32
7.3. Magnetic Field of a Rotating Charged Disk
where the last lines uses r ≫ a (recall we’re interested in the solution far from the
conducting loop). We now have, again using a ≪ r =⇒ ar ≪ 1,
−1/2
1 1 2a 1 1 2a
= 1− sin θ cos ϕ̃ ≈ 1+ sin θ cos ϕ̃ .
|r − r̃| r r r 2 r
ˆ ˆ
2π − sin ϕ̃
µ0 I t̂ dl µ0 I a 1 2a
A(r) = = dϕ̃ cos ϕ̃ 1 + sin θ cos ϕ̃ .
4π |r − r̃| 4π r 0 2 r
0
The integrals conveniently simplify, since we are integrating sinusoidal terms over an
entire period. Only the integral of cos2 ϕ̃ in the y component is nonzero. We end up
with Ax = Az = 0 and
ˆ
µ0 I a 2π a sin θ µ0 I a2
Ay (r) = cos2 ϕ̃ dϕ̃ = sin θ,
4π r 0 r 4 r2
or, in vector form,
µ0 I a 2 µ0 I a2
A(r) = sin θ ŷ = ẑ × r̂.
4 r2 4 r2
The second expression is preferable: the first, in terms of ŷ, holds only with x, y plane
rotated so ϕ = 0 and r lies in the x, z plane. The second, which uses sin θ ŷ = ẑ × r̂,
holds for any orientation of the x, y plane.
Finally, using Iπa2 = |m| (where m is the loop’s magnetic dipole moment, which
points in the direction of the loop’s normal), the expression for A simplifies to
µ0 m × r̂ µ0 m × r
A(r) = 2
= .
4πr 4πr3
This is the same form of magnetic vector potential as for a magnetic dipole. In other
words, a circular current loop behaves as magnetic dipole at long distances.
33
7.3. Magnetic Field of a Rotating Charged Disk
Using the expressions for our vector quantities, the cross product j(r̃) × (r − r̃) is
− sin ϕ̃ −r̃ cos ϕ̃ z cos ϕ̃
j(r̃) × (r − r̃) = j cos ϕ̃ −r̃ sin ϕ̃ = j z sin ϕ̃ .
0 z r̃
Substituting j(r̃) × (r − r̃) and |r − r̃| into the Biot-Savart law gives
˚
3 z cos ϕ̃
µ0 j d r̃
B(z) = 3/2 z sin ϕ̃ .
4π 2
r̃ + z 2
r̃
Next, we write j d3 r̃ = j dS̃ d˜l = j dS̃(r̃ dϕ̃) = dI r̃ dϕ̃. Note that the product j dS̃ is
the current element dI in the surface element dS̃. The current dI at the radius r̃ on
the disk rotating with period t0 = 2π ω is
dq dq (σ2πr̃ dr̃)
dI = = ω= ω = σωr̃ dr̃.
t0 2π 2π
Substituting j d3 r̃ = dI r̃ dϕ̃ = σωr̃2 dr̃ dϕ̃ into the Biot-Savart law gives
ˆ a ˆ 2π
2 z cos ϕ̃
µ0 σωr̃
B(z) = dr̃ dϕ̃ 3/2 z sin ϕ̃ .
4π 0 0 2
r̃ + z 2
r̃
The first and second components of B contain integrals cos and sin terms over a full
period—the result is zero. After integrating over ϕ̃, the magnetic field simplifies to
ˆ a
2 0
µ0 σωr̃
B(z) =
4π 0
dr̃ 3/2 0 .
r̃ + z 2
2
2πr̃
Only the z component of B is non-zero; it is
ˆ
µ0 a σωr̃3
Bz (z) = dr̃ 3/2 .
2 0 r̃2 + z 2
In terms of the new variable u = r̃2 + z 2 , the integral evaluates to
ˆ 2 2
µ0 σω z +a u − z 2 µ0 σω h 1/2 2 −1/2
iz 2 +a2
Bz (z) = du = 2u + 2z u
2 2 z2 u3/2 4 a2
2
p
µ0 σω z
= z 2 + a2 − z + √ −z
2 z + a2
2
µ0 σω 2z 2 + a2
= √ − 2z .
2 z 2 + a2
34
7.3. Magnetic Field of a Rotating Charged Disk
The magnetic field along the z axis is thus B = (0, 0, Bz ) with Bz as above.
Next, we consider the limit case z ≫ a, far from the rotating disk. Expanding the
square root to fourth order in the small quantity az , multiplying out and simplifying
like terms gives
µ0 σω 2z 2 + a2 a2 a2 3 a4
µ0 σω
Bz = − 2z ≈
2z + 1− 2 + − 2z
8 z4
q
2 z 1+ a
2 2 z 2z
z2
a2
a2 1 a4 3 a4 3 a6 µ0 σω 1 a4 3 a6
µ0 σω
= 2z + − − + + − 2z = + .
2 z z 2 z3 4 z3 8 z5 2 4 z3 8 z5
a6
Neglecting the highest-order z5
term gives the simple result
µ0 σω a4
Bz ≈ , z ≫ a.
8 z3
In other words, far from the disk, the magnetic field falls off as z −3 , just like the field
of a magnetic dipole.
Next, we will try to write the magnetic field in the form Bz ∝ |m| z3
where m is the
disk’s magnetic dipole moment. Integrating over concentric rings with area S carrying
current dI, the disk’s magnetic dipole moment is
ˆ ˆ a ˆ a
2 π
|m| = S dI = (πr̃ ) · (σωr̃ dr̃) = πσω r̃3 dr̃ = σωa4 .
0 0 4
Comparing this expression for |m| to the similar expression for Bz leads to
µ0 σω a4 µ0 |m|
Bz ≈ 3
= ,
8 z 2πz 3
|m|
which is in the desired form Bz ∝ z3
. The general form for the magnetic field of a
magnetic dipole is
µ0 3(m · r)r − mr2
B(r) = .
4π r5
0 |m|
In fact, this expression is equivalent to our result Bz (z) = µ2πz 3 . Since m and r
both point along the z axis, their dot product is m · r = |m|r. Along the z axis,
r = (0, 0, z) and the general expression for the dipole magnetic field simplifies to
35
8.1. Magnetic Force in a Coaxial Cable
The magnetic force F on the matter contained in the region of space enclosed in the
region V and permeated by the magnetic field B is
‹
1 1
F= B(B · n̂) − B2 n̂ dS,
µ0 ∂V 2
We choose an integration surface tightly hugging the half-sheath and work in cylindrical
coordinates r, ϕ. Outside the sheath, the magnetic field is zero and there is no
36
8.2. Tension in a Toroidal Inductor
contribution to Fm . Inside the sheath, the magnetic field is tangent to the semicircle,
so B · n̂ = 0, and the force integral reads
¨
1 1
0 − B 2 n̂ dS.
Fm =
µ0 2
µ0 I
The magnetic field at along the sheath (where r = a) is constant and equal to B = 2πa
and can be moved outside the integral. The normal vector in cylindrical coordinates
reads
n̂ = (− cos ϕ, − sin ϕ, 0),
while the surface element is dS = al dϕ. The force integral reads
ˆ π − cos ϕ
1 µ0 I 2 − sin ϕ la dϕ.
Fm = −
2 4π 2 a2 0
0
The x component with cos ϕ integrates to zero over ϕ ∈ [0, π]. Only the y component
is nonzero, and the vector force reads
1 µ0 I 2 µ0 I 2 l
Fm = + (2al) ŷ = ŷ.
2 4π 2 a2 4π 2 a
Fm µ0 I 2 l F µ0 I 2
F1 = = =⇒ = .
2 8π 2 a l 8π 2 a
Consider a semicircular half of a single inductor coil. As in the previous problem, the
magnetic force with magnitude Fm acts upwards on the top of the semicircle, while
two “surface tension” forces F1 = F2m act downward at the semicircle’s two ends. As
usual, we find the magnetic force Fm using
‹
1 1 2
Fm = B(B · n̂) − B n̂ dS.
µ0 ∂V 2
37
8.3. Resistance of a Thin Conducting Plate
For the integration surface (awkward to describe, best to see a picture) we choose a
surface that looks like a coin cut in half, basically a thin three-dimensional extension
of a semicircle enclosing the coil’s semicircular upper half. We split the surface into
four parts: the left and right semicircular faces, the circular ribbon along the surface’s
outer radius, and the thin rectangular plane along the surface’s bottom.
The circular upper ribbon occurs just outside the inductor coil (where B = 0) and
thus does not contribute the magnetic force. The left and right semicircular faces
have equal magnitude and opposite-sign contributions, since the normal to the surface
changes sign for each face.
Only the rectangular plane has a nonzero contribution to Fm . Under the assumption
r1 ≫ r2 , the magnetic field along the plane simplifies to
ˆ r1 +r2
µ0 N I µ0 N I
B= dr̃ ≈ .
r1 2πr̃ 2πr1
The magnetic field B points along the toroid’s longitudinal axis (into the plane of a
cross-sectional coil) and is perpendicular to the normal n̂ to the planar integration
surface, so B̂ · n̂ = 0. The magnetic force simplifies to
¨
µ0 N I 2
1 1 2 1
Fm = 0 − B n̂ dS = − n̂ dS
µ0 2 2µ0 2πr1
µ0 N I 2 µ0 N I 2 r2
1 2πr1
=− n̂ r2 =− n̂.
2µ0 2πr1 N 2π r1
The unit vector n̂ points downward, so − n̂ and thus Fm point upward and pull the
coil apart. The magnitude of the tension on the coil is then
Fm µ0 N I 2 r2
F1 = = .
2 4π r1
j = σE,
where j and E are the current density and electric field in the conductor. We write j
in terms of the continuity equation
∂ρ
∇·j+ = 0.
∂t
38
8.3. Resistance of a Thin Conducting Plate
∇ · j = σ∇ · E = −σ∇2 U = ∇ · j = 0 =⇒ ∇2 U = 0.
We end up with a Laplace equation for U in the conductor. The plan is to find U (r),
then E(r), then j via j = σE, then I and finally resistance with R = UI .
The general solution of the Laplace equation in cylindrical coordinates (including the
m = 0 term) is
∞
X
Am cos(mϕ) + Bm sin(mϕ) · Cm rm + Dm r−m
U (r, ϕ) =
m=1
+ (aϕ + b)(c ln r + d).
To find a unique solution, we need boundary conditions for our particular problem.
At the first electrode at ϕ = 0, the electric potential is constant; we’ll set U (r, 0) = 0
for convenience. At the second electrode, the potential is U (r, π) = −U0 to make a
potential difference U0 between electrodes (we choose −U0 so the current runs in the
direction of increasing ϕ).
Along the annulus’s semicircular boundaries the electric field must be tangent to the
surface to satisfy ∇ · j = 0. Along these surfaces the radial component of both j and
E is zero; Er = 0 gives the boundary condition
∂U
Er = − = 0.
∂r r1 ,r2
U (r, 0) ≡= 0 + b̃ =⇒ b̃ = 0.
39
8.3. Resistance of a Thin Conducting Plate
40
9.1. Mutual Inductance
Φ1 = L11 I1 ,
where the quantities L12 and L21 are the loops’ mutual inductances. Without
derivation, we state that L12 = L21 for reasons of symmetry.
If the parallel wires carry a current I2 (in opposite directions, since they form a closed
loop), the corresponding magnetic field is
µ0 I 2 1 1 µ0 I 2 d
B= + = .
2π y d−y 2π y(d − y)
´
We find magnetic flux with Φ = B · dS; the dot product drops because B is parallel
to the frame’s cross section dS. The square frame’s surface element is dS = 2y dy
41
9.1. Mutual Inductance
(the frame’s width is 2y), and the magnetic flux through the square frame is
¨ ˆ d/2
µ0 I2 d 2y dy 2µ0 I2 d d/2
Φ1 = B dS = 2 =− ln(d − y) 0
0 2π y(d − y) π
2µ0 I2 d d 2µ0 d ln 2
= ln d − ln = I2 .
π 2 π
Note the use of symmetry—we integrate only from 0 to d/2 and multiply by two.
The mutual inductance—the proportionality between Φ and I2 is
Φ1 2 ln 2
L12 = = µ0 d.
I2 π
Note that L12 depends only on system’s geometry.
Induced Current in the Parallel Wires
Recall the square frame carries an alternating current
so we expect the induced current I2 (t) in the parallel wires to alternate with the
same frequency ω and a general form
Φ2 = L21 I1 = L12 I1 ,
where L12 was found the first part of the problem. We then find U2 with
where L22 is the parallel wire’s self-inductance. Substituting U12 into the earlier
expression U12 = −L12 I˙1 gives
42
9.2. The Cabrera Experiment and Magnetic Monopoles
If we assume both I1 and I2 are sinusoidal with amplitudes I10 and I20 , the above
reduces to
I20 L12
= .
I10 L22
What about the minus sign?
To find the current amplitude ratio, we just need to find the parallel wires self-
inductance L22 using Φ2 = L22 I2
If we send a hypothetical current current I2 through the wires, the corresponding
magnetic field is, as before,
µ0 I2 1 1
B= + .
2π y d−y
The field and surface are parallel, so B · dS = B dS, and the magnetic flux is
¨ ˆ
µ0 I2 ld d−a 1
1
Φ2 = B dS = + dy,
2π a y d−y
Applying d ≫ a we have
µ0 I2 l d
Φ2 = ln .
π a
The self-inductance—the proportionality constant between Φ2 and I2 —is thus
µ0 l d
L22 = ln .
π a
To get a better feel for the numbers involved, if we assume l/d = d/a = 10, we have
I20
≈ 0.06.
I10
43
9.2. The Cabrera Experiment and Magnetic Monopoles
where g is the “magnetic charge”, with units A m. Our plan is to find the magnetic
flux through the loop, use this to find voltage induced in the loop, use the induced
voltage to find current.
Let d(t) be the monopole’s perpendicular distance from the loop’s center, and let ρ
denote radial distance from the current loop’s center (in the plane of the loop).
The magnetic field magnitude a distance r from the monopole is
µ0 g µ0 g 1
B= = .
4π r2 4π d2 + ρ2
We find the magnetic field component B⊥ perpendicular to the current loop with
similar triangles:
d µ0 gd 1
B⊥ = p B= .
2
d +ρ 2 4π (ρ + d2 )3/2
2
We assume the monopole moves with constant speed v and passes through the loop’s
center at t = 0. On the left of the loop (and thus for negative time), the monopole’s
perpendicular distance from the loop is
d(t) = −vt.
44
9.2. The Cabrera Experiment and Magnetic Monopoles
where jm is magnetic current density and accounts for the possible existence of
magnetic monopoles. We then use this modified Maxwell equation to re-derive the
law of induction in the presence of magnetic monopoles. We integrate the equation
over the loop’s surface to get
¨ ¨ ¨
∂
(∇ × E) dS = − B · dS − µ0 jm dS.
S ∂t S S
where Im is magnetic current through the loop’s cross section and Ui is the induced
voltage in the loop. The magnetic current through the loop is nonzero only at the
singular instant when the magnetic monopole passes through, which we model with
the delta function:
Im = gδ(t),
where δ(t) has units s−1 .
If we neglect resistive and capacitive effects, the loop’s circuit equation reads Ui = LI.
˙
Substituting in Ui = −Φ̇ − µ0 Im gives
˙
Ui = Ui = −Φ̇ − µ0 Im = −Φ̇ − µ0 gδ(t) = LI.
where H(t) is the Heaviside step function. Since I(t → −∞) = Φ(t → −∞) = 0
(when the monopole is infinitely far from the loop), the equation reduces to
Remember that Φ changes from µ20 g to − µ20 g as the monopole passes through the loop
at t = 0. The discontinuity in Φ is exactly balanced by the Heaviside step function
activating with magnitude µ0 g at t = 0, and the effect is that LI(t) and thus the
current through the loop is as a continuous, measurable quantity.
If we use the Dirac quantization of magnetic charge, which reads
µ0 ge0 2h
= h =⇒ µ0 g = ,
2 e0
we get a numerical result for the quantity µ0 g, and a theoretically expected value
value of LI(t).
45
10.1. Skin Effect in a Ribbon-Like Conductor
∇ × B ≈ µ0 j.
∂
−∇2 E = − µ0 σE.
∂t
Since the ribbon is attached to an alternating voltage source, the electric field E = E(t)
reads
∂
E(t) = E0 eiωt =⇒ E(t) = iωE(t).
∂t
Finally, we substitute the ∂E
∂t into the earlier equation −∇2 E = − ∂t
∂
µ0 σE to get
∇2 E − iωµ0 σE ≡ ∇2 E − k 2 E = 0,
46
10.1. Skin Effect in a Ribbon-Like Conductor
Next, we define coordinate system whose x axis aligns with the ribbon’s width a and
whose y axis aligns with the ribbon’s height. The conductor’s length corresponds to
the z axis. We choose the origin so that x ∈ [−a/2, a/2], meaning the ribbon’s center
occurs at x = 0.
With respect to this coordinate system, we can then simplify the Laplacian ∇2 E.
Since the ribbon’s width a is much smaller than the height and length, the Laplacian’s
derivatives with respect to y and z are negligible, i.e.
2
∂2 ∂2 ∂2E
2 ∂
∇ E= 2
+ 2 + 2 E≈ .
∂x ∂y ∂z ∂x2
The alternating voltage is applied along the conductor’s length—along the z axis—so
the electric field reads E = Ez (x) ẑ. Plugging all of these simplifications into the
amplitude equation gives
∂2
∇2 E − k 2 E ≈ − k 2 Ez (x) ẑ = Ez′′ (x) − k 2 Ez (x) = 0.
Ez (x) ẑ
∂x2
The solutions to the equation
Ez′′ − k 2 Ez = 0
can be written either as exponents or hyperbolic functions; we will use hyperbolic
functions, which are best suited to the problem’s reflection symmetry about the y
axis. The general solution is
However, the problem’s reflection symmetry means Ez will have only the even
component cosh, and the solution simplifies to
Next, we draw some qualitative sketches of E(x) with k as a parameter. The resulting
curves show that as k (and thus frequency ω) increases, E(x) becomes concentrated
near the ribbon’s outer surfaces x = ±a/2. This is a qualitative demonstration of the
skin effect, where electric field and current become concentrated along a conductor’s
surface at high frequencies.
Potential Difference, Current and Impedance
Next, with Ez (x) known, the potential difference across the ribbon’s is simply
Uz (x) = Ez (x)l,
47
10.1. Skin Effect in a Ribbon-Like Conductor
2σE0 b ka
= tanh .
k 2
Next, we introduce the dimensionless quantity κ = ka 2 and substitute in U0 = E0 l to
get
σU0 ba 2 ka σU0 ba tanh κ U0 tanh κ
I= tanh = = ,
l ka 2 l κ R0 κ
where, in the last equality, we have substituted in the conductor’s static resistance
l
R0 = .
σS
In terms of R0 , the conductor’s impedance is then
U0 κ
Z≡ = R0 .
I tanh κ
48
10.2. Theory: Conservation of Electromagnetic Energy
where we have used the divergence theorem to convert the Poynting vector term to a
surface integral.
• The w term corresponds to the changing electromagnetic field energy within
the region V
• The S term encodes energy flow (power) through the surface
• The j · E term corresponds to Ohmic energy losses within the region.
49
10.3. Power in a Coaxial and Cylindrical Conductor
Meanwhile, we find electric field with Gauss’s law, using a cylinder enclosing the
inner conductor:
Q
Q = ϵ0 E · 2πrl =⇒ E = .
2πϵ0 lr
The induced charge Q and potential difference between the inner and outer conductor
are related by
ˆ b
Q b
U= E dr = ln .
a 2πϵ0 l a
In terms of U , the electric field is thus
Q U
E= = .
2πϵ0 lr r ln ab
The Poynting vector points in the direction E × B, along the conductor’s longitudinal
axis. In our case we have E ⊥ B, so the Poynting vector magnitude S is
!
1 1 µ0 I U UI
S= EB = b
= .
µ0 µ0 2πr r ln a 2π ln ab r2
50
10.4. Cylindrical Conductor with a Slit
The power through the entire conductor’s lateral surface occurs at r = a, which
produces the familiar result
U Ia2
P = = U I.
a2
Next, we consider the conductor’s energy balance
˚ ‹ ˚ ˚
∂
− w dV = S · dS + j · E dV = P + j · E dV.
∂t V ∂V V V
∂t = 0, and thus
Because the situation is stationary we have ∂w
˚
P+ j · E dV ≡ P + Ploss = 0 =⇒ Ploss = −P = −U I.
V
Note that Ohmic losses amount to Ploss = −U I, which is the familiar expression from
e.g. high school physics. The energy balance reads
˚
P+ j · E dV = P + Ploss = U I − U I = 0,
51
10.4. Cylindrical Conductor with a Slit
The charge accumulating on the slit is q(t) = It, and, using the model of a parallel-
plate capacitor, the corresponding electric field is
σ(t) 1 q(t) I
E(t) = = = t.
ϵ0 ϵ0 S ϵ0 S
Next, we find the magnetic field in the slit. The time-varying changing electric field
E(t) creates displacement current, and the Maxwell equation for ∇ × B reads
∂E ∂E
∇ × B = µ0 j + ϵ0 µ0 = ϵ0 µ0 .
∂t ∂t
Note that µ0 j = 0 inside the slit, since there’s no free current. We integrate the
equation over the conductor’s cross-sectional surface S to get
¨ ˛ ¨
∂ ∂E
∇ × B · dS = B · dl = ϵ0 µ0 E · dS = ϵ0 µ0 · (πr2 ).
S ∂t S ∂t
We evaluate the line integral over the surface’s boundary and substitute in electric
field to get
˛
∂E(t) µ0 I µ0 I
B · dl = B · (2πr) = ϵ0 µ0 =⇒ B(r) = r= ,
∂t 2S 2πr
where S = πr2 is the cross-sectional surface area. The magnetic field points tangent
to the cylindrical lateral, just like the magnetic field of a conductor, except that free
current is replaced by displacement current.
Since E points along the conductor’s longitudinal axis and B is tangent to the lateral
surface, S = E × B points along the radial direction. Because B ⊥ S, the power
through the lateral surface is simply
¨ ¨ ¨
1 1 I µ0 I
P = S dS = EB dS = t r dS.
Slat µ0 Slat µ0 Slat ϵ0 S 2S
I 2a I 2d
P = t(2πad) = t,
2ϵ0 S 2 ϵ0 S
where WEM = WE + WB is the sum of magnetic and electric field energy. Note that
B = B(r) and thus magnetic energy WB is independent of time. It follows that
∂WEM ∂WE
= + 0.
∂t ∂t
52
10.4. Cylindrical Conductor with a Slit
ϵ0 I 2 I 2d 2
1 2 2
W E = wE V = ϵ0 E · (Sd) = Sdt = t .
2 2 ϵ20 S 2 2ϵ0 S
∂WEM ∂WE ∂ I 2d 2 I 2d
= = t = t = P.
∂t ∂t ∂t 2ϵ0 S ϵ0 S
53
11.1. A Radially Polarized Sphere
ρb = −∇ · P,
while electric field and total charge density ρ are related by Gauss’s law
ρ = ϵ0 ∇ · E.
We find volume charge density with ρb = −∇ · P and the known expression for P:
where we’ve used ∇ · r = 3. Note the bound volume charge density is negative,
since the electric dipoles in the sphere have their positive poles radially outward and
negative pole radially inward, so negative charge is concentrated toward the sphere’s
center.
To relate surface and volume charge density, we integrate ρb = −∇ · P over the
sphere’s volume and apply the divergence theorem
˚ ˚ ¨
out
ρb dV = qb = − ∇ · P dV = − P · dS = −P · S in ,
V V S
where the last step writes S = n̂S and notes that polarization is zero outside the
sphere. The bound surface charge density is then
qb
qb = P · n̂S =⇒ σB = = P · n̂.
S
At the surface r = a, charge density σb evaluates to
σb = P · n̂ = (kr) · n̂ = kr r=a
= ka.
Note that surface charge density is positive, since the electric dipoles have their
positive pole oriented radially outward.
We find total bound charge from ρb and σb by integrating both volume and surface
charge densities:
˚ ¨
4
qb = ρb dV + σb dS = −3k πa3 + ka4πa2 = 0.
V S 3
We should expect total charge to be zero, since all dipoles within the sphere should
cancel out.
54
11.2. A Halved Polarized Sphere
ρb = −∇ · P = 0,
We proceed with the Poisson equation, which simplifies to the Laplace equation
within the sphere were ρb = 0.
ρb
∇2 U (r) = − = 0.
ϵ0
We account for the non-zero surface charge density with boundary conditions.
55
11.2. A Halved Polarized Sphere
Since the surface charge distribution σb depends only on cos θ, our solution can
include only cos θ-dependent terms, which occurs for l = 1. The general solution
simplifies to
U (r) = (A1 r + B1 r−2 ) cos θ.
To avoid divergence at r → 0 and r → ∞, we separate U (r) according to
(
Uin = A1 r cos θ r<a
U (r) ≡ −2
Uout = B1 r cos θ r > a.
We find A1 and B1 with boundary conditions.
First boundary condition: the electric potential must be continuous at the boundary
(a discontinuous potential would imply infinite charge, which is nonphysical). The
continuity condition at r = a requires
B1
A1 a = =⇒ B1 = A1 a3 .
a2
For the second boundary condition we require electric field is perpendicular to the
surface. We then apply Gauss’s law near the sphere’s surface to get
⊥ ⊥ ⊥ ⊥
qb = ϵ0 S(Eout − Ein ) =⇒ σb = ϵ0 (Eout − Ein ).
∂Uin/out
We know σb = P cos θ and use Ein/out
⊥ =− ∂r r=a
to get
⊥ ⊥ 2B1
σb = P cos θ = ϵ0 (Eout − Ein ) = ϵ0 + A1 cos θ.
a3
We substitute the first boundary condition B1 = A1 a3 into the second to get
P cos θ = ϵ0 (2A1 + A1 ) cos θ =⇒ P = 3ϵ0 A1
and thus
P P a3
A1 = and B1 = .
3ϵ0 3ϵ0
The electric potential is then
P
Uin = 3ϵ r cos θ r<a
0
U (r) ≡
P a3 1
Uout =
cos θ r > a.
3ϵ0 r2
We now discuss the solution. The 3 in the denominators is called the depolarization
factor. Next, we note that r cos θ = z—since Uin depends only on z, the field inside
the sphere is homogeneous:
∂Uin P P
Ez = − =− =⇒ E = − ,
∂z 3ϵ0 3ϵ0
56
11.3. Theory: Dielectric and Displacement Field
where we note that P points in the z direction. The field is negative because with our
choice of polarization, positive charges are in the positive z direction and negative
charges in the negative z direction.
The field outside the sphere is field of an electric dipole; the cosine term is analogous
to the dipole dot product term pe · r. A dipole field is expected, since the sphere is
polarized like an electric dipole.
Halved Sphere; Electric Field in the Slit
We now consider the electric field in the slit when the sphere is cut in half in a plane
perpendicular to the polarization.
Because of the slit through the middle, bound charges accumulates on the cut surfaces
as well as the outer surface. Since polarization points “upwards” in the positive z
direction, the bound charges on cut surface of the upper hemisphere are negative,
while the bound charges on the cut surface of the lower sphere are positive. The
relationship σb = P · n̂ is preserved.
To find electric field because of the slit, we model the slit as a parallel-plate capacitor,
for which the electric field reads
σ̃b (P · n̂) P
Ẽ = n̂ = n̂ = ,
ϵ0 ϵ0 ϵ0
where tilde corresponds to the flat surface and n̂ is the normal to the flat surface.
We must also consider the electric field contribution from the bound charge on the
hemispherical surfaces. Reusing the already derived electric field from the bound
charges on the spherical surfaces, which read
P
Esphere = − ,
3ϵ0
the total field in the slit is
P P 2P
E = Ẽ + Esphere = − = .
ϵ0 3ϵ0 3ϵ0
D = ϵ0 E + P.
57
11.4. Parallel-Plate Capacitor with an Anisotropic Dielectric
The D field arises from free charge and is useful when analyzing dielectrics.
For small fields, we make the approximation D ∝ E where D and E are linearly
dependent, which results in the approximate linear relationship
D = ϵ0 χE,
where the line integral runs over a rectangular path thinly hugging a capacitor plate.
Because the path thinly hugs the plate, the integral registers only the component of
E parallel to the plates:
(∥) (∥) (∥) (∥)
Eout · l − Ein · l = 0 =⇒ Eout = Ein .
58
11.4. Parallel-Plate Capacitor with an Anisotropic Dielectric
U = Ed,
qf = Dy S.
59
11.4. Parallel-Plate Capacitor with an Anisotropic Dielectric
where C0 = ϵ0 S
d is the capacitance of an empty capacitor.
Note that for ϕ = 0, corresponding to an isotropic dielectric in which the dielectric
tensor’s second principle axes does align with normal to the capacitor plates, the
capacitor’s capacitance reduces to C = C0 ϵ2 .
60
12.1. Point Dipole in a Spherical Dielectric Cavity
∇2 U (r) = 0.
In spherical coordinates, which are best suited to the problem’s spherical geometry,
the general solution is
∞
X
U (r, θ) = (Al rl + Bl r−(l+1) )Pl (cos θ),
l=0
For the potential to converge at r → ∞, the A1 r term must vanish at large r (outside
the cavity). The general solution further simplifies to
(
Udipole + A1 r cos θ r < a
U (r) =
B1 r−2 cos θ r>a
p cos θ
e
2
+ A1 r cos θ r < a
= 4πϵ0 r
B1 cos θ
r > a.
r2
61
12.1. Point Dipole in a Spherical Dielectric Cavity
E∥ tangent to the boundary surface must be equal on both sides of the boundary.
The tangential component of electric field is found with
1 ∂U
E∥ = − ,
r ∂θ r=a
∥ ∥
and the boundary condition Ein = Eout then implies
pe B1
− 2
sin θ − A1 a sin θ = 2 sin θ,
4πϵ0 a a
which is the same result we arrived at from requiring continuity of electric potential
U at the boundary.
Finally, we apply the boundary condition on the D field, which applies to the
components D⊥ perpendicular to the boundary surface and reads
⊥ ⊥
Din − Dout = σf ,
where σf is the surface density of free charges along the boundary. We find D⊥ with
∂U
D⊥ = −ϵϵ0 ,.
∂r r=a
In our case, which involves a dielectric with only bound charges, we have σf , and the
boundary condition on D reduces to
⊥ ⊥ ∂Uin ∂Uout
Din = Dout ⇐⇒ ϵ0 ϵin = ϵ0 ϵout .
∂r r=a ∂r r=a
The dielectric constant inside the empty cavity is ϵin = 1, and we denote ϵout ≡ ϵ
in the dielectric. We then substitute Uin and Uout into the boundary condition and
simplify to get
pe B1
2 3
− A1 = 2ϵ 3 .
4πϵ0 a a
We then add the earlier boundary condition requiring continuity of U , which read
pe B1
3
+ A1 = 3 ,
4πϵ0 a a
to the just-derived boundary condition on D⊥ to get
pe B1 3 pe
3 = 3 (1 + 2ϵ) =⇒ B1 = .
4πϵ0 a3 a 1 + 2ϵ 4πϵ0
With B1 known, we can find A1 according to
B1 pe pe 3 2(1 − ϵ) pe
A1 = 3 − 3
= 3
−1 = .
a 4πϵ0 a 4πϵ0 a 1 + 2ϵ 1 + 2ϵ 4πϵ0 a3
With the coefficients A1 and B1 known, the solution for U (r, θ) is then
1 2(ϵ − 1) r
pe 2−
r<a
U (r, θ) = cos θ r 1 + 2ϵ a3
4πϵ0 3 1
r > a.
1 + 2ϵ r2
62
12.2. Dielectric Constant of Cold Plasma
Inside the cavity where r < a, the first 1/r2 term is the potential of the electric dipole
pe , while the term containing r cos θ ≡ z corresponds to a homogeneous field inside
the cavity.
Outside the cavity where r > a, we have another 1/r2 term—again like a dipole, but
with an additional factor 1+2ϵ
3
, which would reduce to the usual 1 for a dielectric
constant ϵ = 1. This is the so-called effective dipole term in the problem instructions.
Outside the cavity we thus have the effective dipole moment
3
p′e = pe .
1 + 2ϵ
To find the surface density of bound charges, we have two options. We could could
use Gauss’s law in the form
⊥ ⊥
∂U
σb = ϵ0 Eout − Ein , E⊥ = − .
∂r r=a
Alternatively, we could use the relationship between surface charge density and
polarization, which reads
σb = P · n̂.
Note that the surface’s normal vector n̂ points from the dielectric material’s surface
into the cavity, while the polarization P points out of the cavity. The product P · n̂
thus evaluates to
P · n̂ ≡ −P⊥ ,
where P⊥ is the component of polarization normal to the cavity’s surface. Assuming
a linear relationship between D and E, we then find polarization with
ϵ0 ϵE ≈ D = ϵ0 E + P =⇒ P = ϵ0 (ϵ − 1)E.
63
12.2. Dielectric Constant of Cold Plasma
Assuming the positive ion cores have charge q and electrons have charge −q, the force
on a representative free electron of mass m is
r = r0 ei(kz−ωt) .
Substituting the expressions for r and E into Newton’s law mr̈ = −qE gives
The displacement r0 of the negative electron from the positive ion core creates a
dipole moment with amplitude
q2
pe0 = −qr0 = − E0 .
mω 2
In terms of pe0 , the polarization of the plasma is then
nq 2
P0 = npe0 = − E0 ,
mω 2
where n is the number density of electric dipoles in the material.
We then use the relationship P, E and permittivity, which reads
P0 = ϵ0 (ϵ − 1)E0 ,
P0 nq 2
ϵ=1+ =1− .
ϵ0 E0 mϵ0 ω 2
nq 2
Next, we note that mϵ0 has units of frequency, and define the plasma frequency
nq 2
ωp2 = ,
mϵ0
which describes the frequency at which displaced electrons oscillated about the positive
ion cores. In terms of plasma frequency, the plasma’s dielectric constant is
ωp2
ϵ=1− .
ω2
Note that for electromagnetic wave frequencies ω > ωp we have ϵ > 0 and for ω < ωp
we have ϵ < 0. However, electromagnetic waves cannot propagate through a material
64
12.2. Dielectric Constant of Cold Plasma
with dielectric constant ϵ < 0, which we will discuss more shortly. Dispersion
Relation for EM Waves in Plasma
Next, we will find the dispersion relation for the electromagnetic waves in the plasma.
In material with zero electric current density (i.e. j = 0) the Maxwell equations read
−k 2 E − µ0 ϵ0 ϵ(−ω 2 )E = 0 =⇒ E(µ0 ϵ0 ϵω 2 − k 2 ) = 0.
We now consider the limiting cases for the dispersion relation in plasma, i.e.
q
ω = ωp2 + c20 k 2 .
65
12.2. Dielectric Constant of Cold Plasma
As a side note, we remark that in this limit of small k, where the dispersion relation
obeys ω ∼ k 2 . A quadratic dispersion relation means the photons making up the
electric field behave like particles with mass, which leads to the concept of effective
mass, like in solid state physics.
Further Discussion
Next, we more thoroughly consider the regime of ϵ < 0. Assuming ϵ < 0, we
rearranging the equation
ω2 c20 ω√ ω √
= to get k= ϵ = i ϵR ,
k2 ϵ c0 c0
where k is imaginary because ϵ < 0. Substituting this expression for k into the
oscillator ansatz for the electric field E gives
√
− cω ϵR z −iωt
E = E0 ei(kz−ωt) = E0 e 0 e .
Note that vphase > c0 —the phase velocity is greater than the speed of light in vacuum.
The group velocity—the quantity relevant to the universal speed limit of c0 —is
∂ω 2c2 k c0
vgroup = = q 0 =r .
∂k 2 ωp2 + c20 k 2 ωp2
1 + c2 k 2
0
66
13.1. Theory: Electromagnetic Wave Propagation in Waveguides
1 ∂2E
∇2 E − = 0,
c20 ∂t2
E(r, t) = E(ρ)ei(kz−ωt) ,
where ρ is the radial position within the waveguide’s planar cross section, i.e. the
(x, y) plane perpendicular to the direction of EM propagation along ẑ, k is the wave
vector corresponding to wave propagation in the ẑ direction, and ω is the wave’s
frequency. With respect to this ansatz, the z and t derivatives produce
∂ ∂2 ∂ ∂2
= ik =⇒ = −k 2 and = −iω =⇒ = −ω 2 .
∂z ∂z 2 ∂t ∂t2
Using the expression for ∂z ,
∂
we then write the Laplacian in the separated form
∂2
∇2 = + ∇2⊥ = −k 2 + ∇2⊥ ,
∂z 2
2 2
where ∇2⊥ = ∂x∂
2 + ∂y 2 corresponds to differentiation in the (x, y) plane, i.e. the
∂
We can include the ei(kz−ωt) , whose only variables are z and t, because the above
wave equation doesn’t contain ∂z∂
or ∂t
∂
derivatives (but only ∂x
∂
and ∂y
∂
derivatives
because of ∇⊥ ).
2
Since E ∈ R3 has three components, the above wave equation in vector form really
has 3 equations for each of E’s three components.
Similarly, we could derive an analogous vector wave equation for magnetic field H:
2
2 ω 2
∇⊥ + −k H(r, t) = 0.
c20
67
13.1. Theory: Electromagnetic Wave Propagation in Waveguides
This vector equation also contain three equations for each of H’s three components.
As a result, the electromagnetic waves’ propagation in the vacuum is nominally
described by 3 + 3 = 6 equations. However, it turns out that we can express multiple
H in terms of E, which simplifies the equations involved. Our goal in the coming
page will be to express Ex , Ey , Hx and Hy in terms of only Ez and Hz .
Relationship Between E and H
We begin with the Maxwell equations
∂H ∂E
∇ × E = −µ0 and ∇ × H = ϵ0 .
∂t ∂t
In component form, the equation for ∇ × E reads
∂
∂E
∂y − ikEy
z
∂x Ex Hx
∂ ∂E
× Ey = ikEx − ∂xz = iµ0 ω Hy ,
∂y
∂ ∂Ey ∂Ex
∂z Ez ∂x − ∂y Hz
The first and fifth equations both contain Hx and Ey . They read
∂Ez ∂Hz
− ikEy = iµ0 ωHx and ikHx − = −iϵ0 ωEy .
∂y ∂x
Our next step is to eliminate Hx : we multiply the first equation by k, the second by
µ0 ω, and subtract the two equations to get
∂Ez 2 ∂Hz
k − ik Ey = µ0 ω − iϵ0 ωEy .
∂y ∂x
ω2
∂Ez ∂Hz 2 2
2
k − µ0 ω = iEy k − ω µ0 ϵ0 = iEy k − 2 ,
∂y ∂x c
k ∂E ∂Hz
∂y − µ0 ω ∂x
z
Ey = i ω2
.
c2
− k2
We could perform an analogous procedure the remaining equations to solve for the
remaining components Ex , Hx and Hy in terms of Ez and Hz . Without derivation,
68
13.2. A Parallel-Plate Waveguide
Hx = i ω2
c2
− k2
k ∂E ∂Hz
∂x + ωµ0 ∂y
z
Ex = i ω2
c2
− k2
k ∂H ∂Ez
∂y + ωϵ0 ∂x
z
Hy = i ω2
.
c2
− k2
Note that these equations apply only in Cartesian coordinates for a pipe-like waveguide
where waves propagate along the ẑ direction.
Next, we consider boundary conditions at the waveguide’s walls. We begin with the
Maxwell equations
∂H
∇ × E = −µ0 and ∇ · H = 0,
∂t
and encircle a small portion of the waveguide boundary wall with a hypothetical
closed loop. We then send to loop to zero thickness while applying Stokes’ law, which
results in
∇ × E → 0 =⇒ E∥ = 0 and H⊥ → 0.
In other words, the component of electric field parallel to the waveguide boundary
and the component of magnetic field normal to the waveguide boundary are zero.
Finally, we find Ez and Hz from the waveguide wave equations
2
w Ez
∇2⊥ + − k 2
= 0.
c2 Hz
69
13.2. A Parallel-Plate Waveguide
Ex = i ω2
Ey = i ω2
c2
− k2 c2
− k2
k ∂H ∂Ez
∂x − ωϵ0 ∂y
z
k ∂H ∂Ez
∂y + ωϵ0 ∂x
z
Hx = i ω2
Hy = i ω2
,
c2
− k2 c2
− k2
which express Ex , Ey , Hx and Hy in terms of Ez and Hz . Using Hz = 0, which holds
in general for TM waves, and ∂y ∂
= 0, which holds for our translationally-invariant
waveguide geometry, the field components simplify to Ey = Hx = 0 and
ik ∂Ez iωϵ0 ∂Ez w2
Ex = and Hy = , κ2 = + k2 .
κ2 ∂x κ2 ∂x c2
Note that H = (0, Hy , 0) has only an y component, while E = (Ex , 0, Ez ) has both x
and z components.
Next, we find Ez using the wave equation
2
∂ 2 Ez
2 2 ∂ 2
= −κ2 Ez ,
∇⊥ + κ Ez = + κ Ez = 0 =⇒
∂x2 ∂x2
2
where we have again used κ2 = wc2 + k 2 for conciseness. The last equality holds
because ∂y
∂
= 0. The equation is solved with the oscillatory ansatz
Next, we consider boundary conditions along the waveguide plates. The general
condition E∥ = 0 applies to the parallel components to waveguide surface, which in
our geometry are Ez and Ey . The result is Ez = Ey = 0—note that we knew the
result Ey from earlier. Evaluated at the boundaries surfaces x = 0 and x = a, the
condition reads
E∥ = Ez (0) = Ez (a) = 0.
70
13.2. A Parallel-Plate Waveguide
We substitute the boundary condition Ez (0) = 0 into the solution ansatz to get
Dispersion Relation
Next, we aim to find the dispersion relation between frequency ω and wave vector k.
2
Combining the definition κ2 = wc2 + k 2 with the just-derived result κ = nπ
a shows the
dispersion relation is
r
2 ω2 2
nπ 2 nπ 2
κ ≡ 2 −k = =⇒ ω = c0 k 2 + .
c a a
At large k, this approaches the free space relation ω = c0 k
Note that frequencies below c0 nπ
a are unattainable, since k cannot decrease below
2
71
13.2. A Parallel-Plate Waveguide
Ex ik ∂E
∂x
z
k
Z= = ∂E
= .
Hy iωϵ0 ∂x z ωϵ0
The coefficient c01ϵ0 has a special meaning—it is the impedance of free space Z0 , which
is more visible in the form
√ r
1 ϵ0 µ0 µ0
= = ≡ Z0 .
c0 ϵ0 ϵ0 ϵ0
We can interpret κ as a wave vector in the x direction, while k is the wave vector
along the direction of propagation z.
TE Propagation Mode
In TE mode, we have Ez = 0 and we find Hz ̸= 0. Once again, we start with the
field component equations
i ∂Ez ∂Hz i ∂Ez ∂Hz
Ex = k + ωµ0 Ey = k − µ0 ω
κ ∂x ∂y κ ∂y ∂x
i ∂Hz ∂Ez i ∂Hz ∂Ez
Hx = k − ωϵ0 Hy = k + ωϵ0 ,
κ ∂x ∂y κ ∂y ∂x
72
13.2. A Parallel-Plate Waveguide
Note that E = (0, Ey , 0) has only an y component, while H = (Hx , 0, Hz ) has both x
and z components.
Next, as for Ez , we find Hz using the wave equation
2
∂ 2 Hz
2 2 ∂ 2
= −κ2 Hz .
∇⊥ + κ Hz = + κ Hz = 0 =⇒
∂x2 ∂x2
The boundary conditions in TE mode are different than in TM mode. For TE mode,
we use the boundary condition H⊥ = 0. For our coordinate system and parallel-plate
waveguide, the component of H perpendicular to the surface is Hx , which is
ik ∂Hz
Hx = .
κ ∂x
73
13.2. A Parallel-Plate Waveguide
Ey iµ0 ω ∂H
∂x
z
µ0 ω
Z= = ∂H
= .
Hx ik ∂x z k
∂2 ∂2
∇2⊥ = + .
∂x2 ∂y 2
We solve the equation with separation of variables: Ez (x, y) = X(x)Y (y). We then
substitute the ansatz for Ez into the wave equation to get
X ′′ Y ′′
X ′′ Y + XY ′′ + κ2 XY = 0 =⇒ + = −κ2 .
X Y
X ′′ Y ′′
Since the X terms depend only on x and the Y terms only on y, but X and Y
must be constant for the equation to hold. We thus define
X ′′ Y ′′
= −κ2x =⇒ X ′′ = −κ2x X and = −κ2y =⇒ Y ′′ = −κ2y Y.
X Y
Both equations are solved with sinusoidal solutions of the form
74
13.2. A Parallel-Plate Waveguide
The boundary conditions now involve both a and b; the results are
nπ mπ
κx = and κy = , n, m = 1, 2, 3, . . . ,
a b
where κ2 = κ2x + κ2y . The dispersion relation now reads
r nπ 2 mπ 2
ω = c0 k2 + + .
a b
In TM mode, both indices run over n, m = 1, 2, 3, . . ..
Meanwhile, in TE mode, which has cosine solutions, because of the presence of two
indices n and m one of either n or m can be zero.
75
14.1. Cylindrical Waveguide
Ez (r, ϕ) = R(r)Φ(ϕ).
Φ′′ + m2 Φ = 0
and
r2 R′′ + rR′ + (κ2 r2 − m2 )R = 0.
The equation for Φ has a sinusoidal solution, which we write in the form
Φ(ϕ) ∝ sin(mϕ + ϕm ).
Note that we’ve written the solution with a phase shift ϕm instead of as a linear
combination of sine and cosine terms.
The radial equation is a Bessel equation and is solved by the Bessel functions Jm and
the Neumann functions Nm . The general solution is a linear combination of the form
76
14.1. Cylindrical Waveguide
The Neumann functions apply in situations with divergence near the origin (e.g. a
coaxial cable), and the Bessel functions for convergence near the origin. Our problem
is free of divergence near the origin, so we will use only the Bessel functions. In this
case, the general solution for Ez = ΦR is
∞
X
Ez (r, ϕ) = Am Jm (κm r) sin(mϕ + ϕm ),
m=0
E∥ ∂
= 0,
Jm (κa) = 0 =⇒ κa = ξm,n ,
where m indexes the Bessel functions and n indexes the zeros of a given Bessel
function.
n J0 (x) J1 (x) J2 (x) J3 (x) J4 (x) J5 (x)
1 2.4048 3.8317 5.1356 6.3802 7.5883 8.7715
2 5.5201 7.0156 8.4172 9.7610 11.0647 12.3386
3 8.6537 10.1735 11.6198 13.0152 14.3725 15.7002
4 11.7915 13.3237 14.7960 16.2235 17.6160 18.9801
5 14.9309 16.4706 17.9598 19.4094 20.8269 22.2178
Table 1: The first 5 zeros ξmn of the first 6 Bessel functions Jm (x).
77
14.1. Cylindrical Waveguide
We then find the waveguide’s bandwidth from the difference between the lowest two
frequency modes, which occur for the smallest two zeros ξmn when k = 0, where the
dispersion relation simplifies to
c0
ω(k = 0) = ξmn .
a
The bandwidth is then the difference between the two smallest possible frequencies
ω, which occur for the two smallest values of ξmn . These are:
c0 c0 c0
∆ω = (ξ11 − ξ10 ) ≈ (3.8317 − 2.4048) = 1.47 · .
a a a
TE Mode
In TE mode we have Ez = 0 and solve for Hz . As in TM, we use separation of
variables with the ansatz
Hz (r, ϕ) = R(r)Φ(ϕ).
Following an analogous separation of variables procedure as in TM mode produces
the general solution
∞
X
Hz (r, ϕ) = Am Jm (κr) sin(mϕ + ϕm ).
m=0
H⊥ ∂
= 0,
n J0′ (x) J1′ (x) J2′ (x) J3′ (x) J4′ (x) J5′ (x)
1 3.8317 1.8412 3.0542 4.2012 5.3175 6.4156
2 7.0156 5.3314 6.7061 8.0152 9.2824 10.5199
3 10.1735 8.5363 9.9695 11.3459 12.6819 13.9872
4 13.3237 11.7060 13.1704 14.5858 15.9641 17.3128
5 16.4706 14.8636 16.3475 17.7887 19.1960 20.5755
Table 2: The first 5 zeros ξmn of the first 6 Bessel function derivatives Jm
′ (x).
78
14.2. Overview: Waveguide with a Quarter-Circle Cross Section
The correct general solution for Hz is written with two indices in the form
∞ X
∞
′ r
X
Hz (r, ϕ) = Amn Jm ξmn sin(mϕ + ϕmn ),
a
n=1 m=0
where m indexes the Bessel functions and n indexes the zeros of the Bessel function
derivatives ξmn
′ .
We find the bandwidth analogously to TM mode, we just using the new zeros ξmn
′
Substituting these ansatzes into the waveguide wave equation leads to the familiar
general solutions
X ∞
Ez (r, ϕ)
= Am Jm (κr) sin(mϕ + ϕm ).
Hz (r, ϕ)
m=0
m = 2, 4, 6, · · · .
79
14.3. Theory: TEM Waves in Waveguides
Finding the radial component of the general solution is analogous to the procedure
for a circular cross section and involves finding zeros of the Bessel functions.
Boundary Conditions in TE Mode
In TE mode, the relevant boundary condition involves the perpendicular magnetic
field H⊥ . Along the circular surface, the perpendicular magnetic field is Hr ∝ ∂H
∂r .
z
Along the vertical flat surface at ϕ = π/2, the boundary condition reads
∂Hz π
H⊥ = , r = 0.
∂ϕ 2
This condition leads to the requirement m = 0, 2, 4, . . .. In TE mode, the case m = 0
is valid because the angular cosine solutions are constant and thus non-trivial, at
m = 0. Similarly, the radial Bessel functions are nontrivial for m = 0, and we get a
valid solution. In TE mode for a quarter-circle cross section, the valid values of m
are thus m = 0, 2, 4, . . ..
E(r) = E(ρ)ei(kz−ωt) ,
∇ × E = ik × E and ∇ × H = ik × H.
80
14.3. Theory: TEM Waves in Waveguides
where the second equation holds because the exponent term is a scalar quantity. Next,
we make the calculation
To show ∇ × E(r) = 0, we just need to show that the term ∇ × E(ρ)ei(kz−ωt) equals
zero. Noting that ρ = (x, y, 0) i.e. the ρz = 0, and also that Ez ≡ 0 in TEM mode,
we have
∂ ∂
∂x Ex (ρ) ∂x Ex (ρ)
∂ ∂
∇ × E(ρ)ei(kz−ωt) = ∂y × Ey (ρ) ei(kz−ωt) = ∂y × Ey (ρ) ei(kz−ωt)
∂ Ez (ρ) 0
0
∂z
∂Ey (ρ) ∂Ex (ρ) i(kz−ωt)
= − e ẑ = ∇ × E(r) z ẑ.
∂x ∂y
Applying Maxwell’s equations shows that
∂Hz
∇ × E(r) z = µ0 ≡ 0,
∂t
since Hz = 0 in TEM mode. The result is the desired equality
∇ × E = ik × E,
Next, we take the cross product of the above equations with k to get
k × (k × E) = µ0 ωk × H = −ϵ0 µ0 ω 2 E.
k × (k × E) = k(k · E) − k 2 E = 0 − k 2 E,
ω2
2 2 2
−k E = −ϵ0 µ0 ω E =⇒ E k − 2 = 0 =⇒ ω = c0 k.
c0
81
14.4. TEM Waves in a Coaxial Waveguide
In other words, we have proven that TEM mode has the dispersion relation ω = c0 k
Last Step
We begin with the waveguide wave equation, which reads
2
2 ω 2 E
∇⊥ + −k = 0.
c20 H
∇2⊥ E∥ = 0.
82
14.4. TEM Waves in a Coaxial Waveguide
Note that this is the same dispersion relation as for cold plasma, even though the
plasma problem was solved in unbounded space, while the coaxial waveguide has
bounded geometry.
Impedance
We find the waveguide’s impedance with the relationship Z = UI where U and I are
the potential difference between the inner and outer sheath and the current through
the conductor, respectively, corresponding to TEM waves in the waveguide. In TEM
mode, the EM field is static, which simplifies finding U and I.
We assume the coaxial cable carries a hypothetical current I. Using Ampere’s law,
the current through the inner conductor leads to a tangential magnetic field Hϕ of
the form
I
I = Hϕ · 2πr =⇒ Hϕ = ,
2πr
where Hϕ is the ϕ component of H (tangential to the waveguide’s circular cross
section).
Next, assume a potential difference U between the inner and outer conductors. We
find the associated radial electric field Er using Gauss’s law:
Q
Q = ϵ0 ϵ · 2πrl · Er =⇒ Er = ,
2πϵϵ0 lr
where q is the charge on the inner conductor. Note that because of the coaxial
conductor’s concentric geometry, E has only a radial component and H has only a
tangential component. With Er known, we find the associated potential difference
with ˆ b ˆ b
q dr q b
U= Er dr = =⇒ U = ln .
a wπϵϵ 0 l a r 2πϵϵ0 l a
We can relate U and Er with
b
U = Er r ln .
a
Next, we substitute U and I into the impedance equation to get
U 1 Er b
Z= = ln .
I 2π Hϕ a
We can relate Er and Hϕ with the earlier theoretical TEM result
k
H= × E.
ωµ0
Since E = (Er , 0, 0) and H = (0, Hϕ , 0) are perpendicular, the above vector equation
simplifies to √
k 1 ϵϵ0 µ0 ϵ
r
Hϕ = Er = Er = 2 Er = Er ,
ωµ0 cµ0 µ0 Z0
83
14.4. TEM Waves in a Coaxial Waveguide
where Z0 = √ϵ10 µ0 is the impedance of free space. We substitute this result for Hϕ in
terms of Er into the impedance equation to get
Z0 b 1 Z b
Z= √ ln = q 0 ln .
2π ϵ a 2π ωp2 a
1− ω2
84