Anda di halaman 1dari 22

Physics 214 Solution Set 4 Winter 2017

1. [Jackson, problem 12.3] A particle with mass m and charge e moves in a uniform, static,
~ 0.
electric field E
(a) Solve for the velocity and position of the particle as explicit functions of time,
assuming that the initial velocity ~
v 0 was perpendicular to the electric field.

~ = 0, we have:
Using eqs. (12.1) and (12.2) of Jackson and setting B
d~
p ~, dW ~,
= eE = e~
v·E
dt dt
where W is the total mechanical energy (usually called E, but we have renamed this W in
order to better distinguish it from the electric field) and ~
v is the particle velocity (which is
denoted as u~ by Jackson).
Clearly, the motion takes place in a plane containing the E-field.~ Without loss of
generality, we assume that
E~ = E x̂ ,
and assume that the motion takes place in the x–y plane. By assumption, ~ v·E ~ = 0 at
t = 0, in which case px = 0 at t = 0. Solving the equations,
dpx dpy
= eE , = 0,
dt dt
in follows that
px = eEt , py = p0 ,
where p0 is a constant.
Using p~ = γm~ v and E = γmc2 , it follows that
c2 p
~ c2 p
~
v=
~ =p .
W |~
p|2 c2 + m2 c4
Hence,
c2 eEt c2 p 0
vx = p , vy = p 2 .
(p20 + e2 E 2 t2 )c2 + m2 c4 (p0 + e2 E 2 t2 )c2 + m2 c4

Since ~
v = d~
x/dt, it follows that
tdt dt
Z Z
2 2
x = c eE p 2 , y = c p0 p , (1)
W0 + (ceEt)2 W02 + (ceEt)2

where W02 = p20 c2 + m2 c4 .

1
We shall define the origin of the coordinate system to coincide with t = 0. Then
computing the integrals in eq. (1) yields
q   
1 2 p 0 c −1 ceEt
x(t) = W0 + (ceEt)2 − W0 , y(t) = sinh . (2)
eE eE W0

(b) Eliminate the time to obtain the trajectory of the particle in space. Discuss the
shape of the path for short and long times (define “short” and “long” times).

We can eliminate t from eq. (2),


 
W0 eEy
t= sinh .
ceE p0 c
Inserting this into the equation for x(t) and using the identity cosh2 z − sinh2 z = 1, it
follows that    
W0 eEy
x= cosh −1 ,
eE p0 c
which is the equation for a catenary curve.
To describe the shape of the path for short and long times, we note that W0 /(ceE)
has units of time. This we can define short and long times relative to this quantity. For
t ≪ W0 /(ceE), we have
(ceEt)2
 
ceEt ceEt
q
2 −1
W0 + (ceEt) ≃ W0 +
2 , sinh ≃ .
2W0 W0 W0
Hence the approximate form of eq. (2) is
c2 eEt2 p 0 c2 t
x(t) ≃ , y(t) ≃ .
2W0 W0
Solving for t and inserting the result back into the above equations yields
eEW0 y 2
x≃ .
2p20 c2
Since v0 = c2 p0 /W0 , we can eliminate W0 from the above expression to obtain,
eEy 2
x≃ . (3)
2p0 v0
That is, as short times, the motion is parabolic.1
1
The result of eq. (3) also coincides with the non-relativistic limit (in which case p0 = mv0 ). To verify
this assertion, we can perform a formal expansion in powers of 1/c. In this limit, W0 ≃ mc2 and
W0 mc
t≪ ≃ ,
ceE eE
which is always true in the limit of c → ∞ (which is equivalent to taking the non-relativistic limit).

2
For t ≫ W0 /(ceE), eq. (2) yields:
 
p0 c 2ceEt
x(t) ≃ ct , y(t) ≃ ln .
eE W0
In the latter case, we used:
 √ 
sinh−1 z = ln z + z 2 + 1 ≃ ln 2z , for z ≫ 1 .

Hence, to a good approximation,


 
p0 c 2eEx
y≃ ln ,
eE W0
or equivalently,  
W0 eEy
x≃ exp .
2eE p0 c
That is, at long times the motion is exponential.

2. [Jackson, problem 12.9] The magnetic field of the earth can be represented approximately
by a magnetic dipole of magnetic moment M = 8.1 × 1025 gauss-cm3 . Consider the motion
of energetic electrons in the neighborhood of the earth under the action of this dipole field
(Van Allen electron belts). [Note that M~ points south.]

(a) Show that the equation for a line of magnetic force is r = r0 sin2 θ, where θ is the
usual polar angle (colatitude) measured from the axis of the dipole, and find an expression
for the magnitude of B~ along any line of force as a function of θ.

Let the z-axis point from the origin in the direction of the north pole. Then, the magnetic
dipole moment (which points south) is given by M ~ = −M ẑ, where M ≡ |M ~ |. The vector
potential is given in gaussian units by:
 
~ x̂ ŷ ẑ
~ x) = M × ~ x 1
A(~ = 3 det  0 0 −M 
|~
x| 3 r
r sin θ cos φ r sin θ sin φ r cos θ

M sin θ M sin θ
= (x̂ sin φ − ŷ cos φ) = − φ̂ ,
r 2 r2
where r ≡ |~
x|. Then,
 
r̂ r θ̂ r sin θφ̂
 
~ =∇
~ ×A
~= 1 ∂ ∂ ∂ 
B det  ,
r 2 sin θ  ∂r
 ∂θ ∂φ 

0 0 r sin θ Aφ

3
where Aφ = −M sin θ/r 2 . Evaluating the above determinant yields:

~ = − 2M cos θ r̂ − M sin θ θ̂ .
B (4)
r3 r3

Given the magnetic field at every point in space, B(~ ~ x), one can consider a related
~ (~
vector field, F ~ x), which gives the force on a magnetic test charge qm due to
x) = qm B(~
the magnetic field at the point ~x.2 If we choose our test charge to have qm = 1, then there
is no distinction between the “lines of magnetic force” and the “magnetic field lines.” We
choose to follow this convention in what follows.
The lines of force follow a curve ~
x(x), where the arclength s parameterizes the location
~
along the curve. By definition B(~ x) is tangent to the lines of force. That is,

~ ~

d~
x B x(s)
= , (5)
ds B

where B ≡ |B|.~ To understand the normalization on the right hand side above, we note
that eq. (5) is equivalent to the three equations,

dx Bx dy By dz Bz
= , = , = .
ds B ds B ds B
Squaring each equation and summing the three resulting equations yields

(ds)2 = (dx)2 + (dy)2 + (dz)2 ,

which is the well-known formula for the differential arclength.


It is convenience to work in spherical coordinates. Consider an infinitesimal displace-
ment d~x, where
x = r sin θ cos φ x̂ + r sin θ sin φ ŷ + r cos θ ẑ .
~
By the chain rule,

∂~
x ∂~
x ∂~
x
d~
x= dr + dθ + dφ
dr dθ dφ
= (cos φ sin θ x̂ + sin θ sin φ ŷ + cos θ ẑ) dr + r (cos θ cos φ x̂ + cos θ sin φ ŷ − sin θ ẑ) dθ
+r (− sin θ sin φ x̂ + sin θ cos φ ŷ) dφ
= r̂ dr + θ̂ r dθ + φ̂ r sin θ dφ . (6)
2
Of course, magnetic charges do not exist in classical electromagnetism. But the concept of “lines of
force” were developed before this fact was understood. In the case of the electric field, we do have F~ = q E,
~
so the terminology “lines of force” makes sense. In the case of magnetic fields, it would be better to refer to
the lines of force as the magnetic field lines. Nevertheless, following Jackson, we retain the old terminology
in this problem.

4
The tangent to the curve ~
x(s) then takes the form
d~x dr dθ dφ
= r̂ + θ̂ r + φ̂ r sin θ . (7)
ds ds ds ds
Using eq. (4), it follows that the line of magnetic force is determined by the equation,
d~x ~ (~
B x(s)) 2M M sin θ
= = − 3 cos θ r̂ − θ̂ , (8)
ds B Br Br 3
where r, θ and φ are functions of s. Equating eqs. (7) and (8) yields three differential
equations,
dr 2M cos θ dθ M sin θ dφ
=− , r = − , = 0. (9)
ds Br 3 ds Br 3 ds
Dividing the first two equations above yields,
dr 2r cos θ
= ,
dθ sin θ
which is easily integrated,
dr cos θ
Z Z
=2 dθ ,
r sin θ
Evaluating the integrals and imposing the condition r = r0 at θ = 12 θ, we obtain
 
r
ln = 2 ln sin θ ,
r0
or equivalently
r = r0 sin2 θ , (10)
which we identify as the equation for the line of magnetic force. Note that the third
equation in eq. (9) implies that φ is a constant along the line of magnetic force.
Finally, we evaluate the magnitude of B ~ along the line of force. Since

~ = Br2 + B 2 + B 2 = M 4 cos2 θ + sin2 θ ,


q p
B ≡ |B| θ φ
r3
We simply plug in eq. (10) to obtain B as a function of θ along the line of magnetic force,

M 1 + 3 cos2 θ
B(θ) = 3 , (11)
r0 sin6 θ
after using sin2 θ = 1 − cos2 θ in the numerator above.

(b) A positively charged particle circles around a line of force in the equatorial plane
with a gyration radius a and a mean radius R (where a ≪ R). Show that the particle’s
azimuthal position (east longitude) changes approximately linearly in time according to:
3  a 2
φ(t) = φ0 − ωB (t − t0 ) .
2 R
where ωB is the frequency of gyration at radius R.

5
Assuming that a ≪ R, we can use eq. (12.55) of Jackson to obtain an approximate formula
for the gradient drift velocity,
~
vG a ~ 
~ ⊥B ,
= B × ∇ (12)
ωB a 2B 2
~ is the field at the equator (θ = 1 π). Using eq. (4),
where a is the gyration radius and B 2
this means that

~ M M ~ ∂B 3M
B = −θ̂ 3 = − 3 θ̂ , ∇⊥ B = r̂ = − 4 r̂ , (13)
r r=R R ∂r
r=RR

where R is the mean radius. In computing ∇ ~ ⊥ B, we used the fact that B ≡ |B|
~ = M/r 3
and
~ = r̂ ∂ + φ̂ 1
~ ⊥ = n̂ · ∇


,
∂r r sin θ ∂φ
~ = 0. Inserting the results of eq. (13) into eq. (12), we end up with
where n̂ · B
ωB a2 R6 3ωB a2
   
M 3M
vG =
~ θ̂ × r̂ = − φ̂ . (14)
2 M2 R3 R4 2R
Finally, we can express ~
v G in terms of the angular velocity dφ/dt by

vG = R
~ φ̂ .
dt
Comparing this equation with eq. (14), we conclude that
dφ 3a2
= − 2 ωB .
dt 2R
Solving this differential equation, and imposing the initial condition φ(t0 ) = φ0 , we end up
with
3a2
φ(t) = φ0 − ω (t − t0 ) . (15)
2R2 B

(c) If, in addition to its circular motion of part (b), the particle has a small component
of velocity parallel to the lines of force,
√ show that it undergoes small oscillations in θ around
1
θ = 2 π with frequency Ω = (3/ 2)(a/R)ωB . Find the change in longitude per cycle of
oscillation in latitude.

As discussed in Chapter 12, section 4 of Jackson, the transverse velocity of gyration is


v⊥ = ωB a [cf. discussion below eq. (12.61) of Jackson]. If in addition, we now include
the small component of the velocity parallel to the lines of magnetic force, we may use
eq. (12.72) of Jackson to write:
B(z)
vk2 = v02 − v⊥
2
0 .
B0

6
Here, the subscript 0 refers to the equator z = 0 (or equivalently to θ = 12 π). In particular,
we can write v02 = vk2 0 + v⊥
2
0 so that
 
2 2 2 B(z)
vk = vk 0 + v⊥ 0 1 − . (16)
B0
In part (a), we found that along the lines of magnetic force,

M 1 + 3 cos2 θ
B(θ) = 3 , (17)
r0 sin6 θ
where r0 ≡ r(θ = 12 π). In this problem, we are interested in the behavior of the particle at
the mean radius R, so we take r0 = R. To compute B(z), we expand about z = 0. Since
z = R cos θ, we expand about z = 0 by writing θ = 21 π + ǫ. Then,

z = R cos θ = R cos 21 π + ǫ = −R sin ǫ ≃ −Rǫ .
Hence, ǫ ≃ −z/R and θ ≃ 12 π − z/R. It follows that
π z z π z z
cos θ ≃ cos − = sin , sin θ ≃ sin − = cos .
2 R R 2 R R
Using eq. (17),
p p
1 + 3 sin2 (z/R) 9z 2
 
M M 1 + 3z 2 /R2 M
B(z) ≃ 3 ≃ 3 ≃ 3 1+ .
r0 cos6 (z/R) r0 [1 − z 2 /(2R2 )]6 r0 2R2
Plugging this result into eq. (16) yields
9  ωB a 2 2
vk2 (z) = vk2 0 − z . (18)
2 R
As discussed below eq. (12.72) of Jackson, this equation is equivalent to the conservation
of energy of a one-dimensional non-relativistic mechanics problem with total mechanical
energy,
E(z) = 21 mvk2 + V (z) ,
where
9ωB2 a2
 
V (z) = 1
2
m z2 , (19)
2R2
is the potential energy of a one-dimensional harmonic oscillator. Indeed, eq. (18) is equiv-
alent to the statement that E(z) = E(0), i.e. conservation of energy. If we write the
harmonic oscillator potential in the standard form,
V (z) = 21 mΩ2 z 2 ,
the eq. (19) implies that the effective oscillator frequency Ω is given by
3 ω a
Ω= √ B .
2 R

7
1
That is, the charged particle undergoes small oscillations in θ around θ = 2
π with fre-
quency Ω.
One period T of oscillation is given by

2π 2 2πR
T = = . (20)
Ω 3ωB a

Using the results of part (b) [cf. eq. (15)], the change of longitude is

3a2
∆φ = − ω ∆t . (21)
2R2 B
Choosing ∆t = T then yields the change of longitude per cycle of oscillation in latitude,

2πa
∆φ = − .
R

(d) For an electron of 10 MeV kinetic energy at a mean radius of R = 3 × 107 m, find
ω and a, and so determine how long it takes to drift once around the earth and how long
it takes to execute one cycle of oscillation in latitude. Calculate the same quantities for an
electron of 10 keV at the same radius.

Given M = 8.1 × 1025 gauss-cm3 and R = 3 × 109 cm, the magnetic field at the equator is
M
B= = 3 × 10−3 gauss .
R 3

Using eq. (12.39) of Jackson,


eB ecB
ωB = = . (22)
γmc γmc2
Although the last step above is rather trivial, it is convenient to write ωB in this form. The
numerical value of the quantity ec is given by

ec = (4.8 × 10−10 statcoulombs)(3 × 1010 cm s−1 ) = 14.4 statcoulombs cm s−1 . (23)

It is convenient to eliminate statcoulombs in favor of gauss. That is,

1 gauss = 1 dyne statcoulomb−1 = 1 erg cm−1 statcoulomb−1 .

Using 1 eV = 1.6 × 10−12 ergs, we can write:

1 gauss = (1.6 × 10−12 )−1 eV cm−1 statcoulomb−1 = 6.25 × 1011 eV cm−1 statcoulomb−1 .

Hence, it follows that

1 statcoulomb = 6.25 × 1011 eV cm−1 gauss−1 .

8
Inserting this result into eq. (23) yields

ec = 9 × 1012 eV gauss−1 s−1 .

Therefore, the gyration frequency can be written as


B (gauss)
ωB = 9 × 1012 s−1 . (24)
γmc2 (eV)
For the electron, we have mc2 = 511 keV. If the electron has a kinetic energy of K =
10 MeV, then E = γmc2 = mc2 + K, which yields K = (γ − 1)mc2 . Hence,
K 10 MeV
γ =1+ =1+ = 20.57 .
mc 2 0.511 MeV
It follows from eq. (24) that

−1 3 × 10−3
ωB = 9 × 10 12
s · = 2.57 × 103 s−1 .
(20.57)(5.11 × 10 )
5

Next we use v ≃ v⊥ = ωB a to determine a. Since γ ≫ 1, it follows that v ≃ c, so that


c 3 × 1010 cm s−1
a= = = 117 km .
ωB 2.57 × 103 s−1
To drift once around the earth requires the longitude (or azimuthal angle φ) to change by
2π. Inserting ∆φ = −2π in eq. (21) [the overall sign is not significant here], we obtain
4πR2 4π(3 × 109 cm)2
∆t = = = 107 s .
3a2 ωB 3(1.17 × 107 cm)2 (2.57 × 103 s−1 )
Finally, the time it takes to execute one cycle of oscillation in latitude was obtained in
part (c) [cf. eq. (20)]:
√ √
2 2πR 2 2π(3 × 109 cm)
T = = = 0.3 s .
3ωB a 3(1.17 × 107 cm)2 (2.57 × 103 s−1 )
For an electron with kinetic energy of 10 keV,
K 10 keV
γ =1+ = 1 + = 1.02 . (25)
mc2 511 keV
It follows from eq. (24) that
(9 × 1012 s−1 )(3 × 10−3 )
ωB = = 5.18 × 104 s−1 .
(1.02)(5.11 × 105 )
To determine a, we first compute v using eq. (25):
1 v
p = 1.02 =⇒ = 0.195 .
1 − v 2 /c2 c

9
Hence,
v (0.195)(3 × 1010 cm s−1 )
a= = = 1.13 km .
ωB 5.18 × 104 s−1
Finally, following the previous computation,
4πR2 4π(3 × 109 cm)2
∆t = = = 5.7 × 104 s ,
3a2 ωB 3(1.13 × 105 cm)2 (5.18 × 104 s−1 )
and √ √
2 2πR 2 2π(3 × 109 cm)
T = = = 1.52 s .
3ωB a 3(1.13 × 105 cm)2 (5.18 × 104 s−1 )
Note that in both computations above, we have a ≪ R, which implies that the gradient
of the magnetic field is small over the orbit of the electrons. Hence, the approximations
introduced in Chapter 12, sections 4 and 5 of Jackson are valid for the charged particle
motions examined in this problem.

3. [Jackson, problem 12.11] Consider the precession of the spin of a muon, initially lon-
gitudinally polarized, as the muon moves in a circular orbit in a plane perpendicular to a
uniform magnetic field B.~

(a) Show that the difference Ω of the spin precession frequency and the orbital gyration
frequency is
eBa
Ω= ,
mµ c
independent of the muon’s energy, where a = 21 (g − 2) is the magnetic moment anomaly.
Find the equations of motion for the components of the spin along the mutually perpen-
dicular directions defined by the particle’s velocity, the radius vector from the center of the
circle to the particle, and the magnetic field.

Our starting point is the Thomas equation, which Jackson writes in the following form
[cf. eq. (11.170) of Jackson]:
    
d~s e g 1 ~ g  γ
~ ~ ~ g γ ~ ~
= ~
s× −1+ B− −1 (β · B)β − − β×E ,
dt mc 2 γ 2 γ+1 2 γ+1
(26)
where the time derivative of the velocity vector is given by [cf. eq. (11.168) of Jackson]:
dβ~ e h~ ~ i
= ~ ~ ~
E + β × B − β(β · E) .~ (27)
dt γmc
For a particle moving in a circular orbit in a plane perpendicular to a uniform magnetic
~ we have β
field B, ~·B ~ = 0, where ~ v ≡ cβ ~ is the particle velocity. Hence, eqs. (26) and
(27) reduce to
 
d~
s e g 1 ~, d~
v e ~,
= −1+ ~
s×B = ~
v×B (28)
dt mc 2 γ dt γmc

10
~ = 0). That is, eq. (28) can be
since by assumption there is no electric field present (E
written in the form of precession equations,
d~
s d~
v
=~ ~,
s×ω =~ ~B ,
v×ω
dt dt
where the spin precession frequency ω
~ and the orbital gyration frequency ω
~ B are given by:
   
e g−2 ~, e ~
~ ≡
ω 1+ γ B ~B ≡
ω B.
γmc 2 γmc
The difference of these two frequencies is
 
~ ≡ω e g−2 ~
Ω ~ −ω
~B = B,
mc 2
and the magnitude of this frequency difference is given by
eBa
Ω= , where a = 12 (g − 2) .
mc
To find the equations of motion for the components of the spin vector, we first decompose
this vector into longitudinal and transverse components with respect to the direction of the
~
velocity, β̂ ≡ β/β. That is, ~
s=~ sk + ~
s⊥ , where

sk = (β̂ · ~
~ s)β̂ , s⊥ = ~
~ s−~
sk .

By construction,
s⊥ · β̂ = 0 .
~ (29)
We first work out d~
sk /dt.

d~
sk d   d   dβ̂
= (β̂ · ~
s)β̂ = β̂ s +~
β̂ · ~ s · β̂ . (30)
dt dt dt dt
Jackson gives the following result in his eq. (11.171),
   
d   e g 
~ gβ 1 ~
s =− ~
β̂ · ~ s · − 1 β̂ × B + − E .
dt mc ⊥ 2 2 β
~ = 0, we obtain
Setting E
 
d   eB g − 2
s =−
β̂ · ~ s⊥ · (β̂ × B̂) .
~ (31)
dt mc 2

We also need to work out dβ̂/dt.


!
dβ̂ d ~
β ~
1 dβ ~ dβ
β
= = − 2 . (32)
dt dt β β dt β dt

11
Using

dβ d  ~ ~ 1/2 1  ~ ~ −1/2 d  ~ ~  1 ~ dβ ~ ~

= β·β = β·β β·β = 2β · = β̂ · ,
dt dt 2 dt 2β dt dt
in eq. (32), we conclude that
" !#
dβ̂ 1 ~
dβ ~

= − β̂ β̂ · .
dt β dt dt

From eq. (28), we obtain


~
dβ e ~ ~
= β×B.
dt γmc
~
Hence β̂ · dβ/dt = 0, and we end up with

dβ̂ eB
= β̂ × B̂ . (33)
dt γmc
Inserting eqs. (31) and (33) into eq. (30), we obtain
 
dsk eB g − 2 eB
=− [~
s⊥ · (β̂ × B̂)]β̂ + s · β̂(β̂ × B̂) .
~
dt mc 2 γmc

Since ~
sk ≡ (~
s · β̂)β̂, it immediately follows that

s · β̂(β̂ × B̂) = ~
~ ~.
sk × B

We can further simplify the quantity [~ s⊥ · (β̂ × B̂)]β̂ by using ~s⊥ · β̂ = 0 [cf. eq. (29)] and
β̂ · B̂ = 0. First, consider the triple cross product
h i
s⊥ × β̂ × (β̂ × B̂) = [~
~ s⊥ · (β̂ × B̂)]β̂ − (β̂ × B̂)~s⊥ · β̂ = [~s⊥ · (β̂ × B̂)]β̂ .

However, β̂ × (β̂ × B̂) = β̂(β̂ · B̂) − B̂ = −B̂. Hence,

[~
s⊥ · (β̂ × B̂)]β̂ = −~
s⊥ × B̂ .

Inserting eqs. (35) and (36) into eq. (34) then yields

d~
sk
  
eB g−2 1
= s⊥ + ~
~ s × B̂
dt mc 2 γ k

Using this result, we can evaluate d~


s⊥ /dt.
    
d~
s⊥ d   eB g 1 ~ = eB g − 2 1
= s−~
~ sk = −1+ (~
sk + ~
s⊥ ) × B s⊥ + ~
~ s × B̂ ,
dt dt mc 2 γ mc 2 γ k

12
which simplifies to
  
d~
s⊥ eB g−2 1
= sk + ~
~ s × B̂
dt mc 2 γ ⊥
Finally, we need to further decompose ~ s⊥ into components along the direction of the
magnetic field and along the radius vector ~r which points to the center of the circular path of
~ = 0], d~
the moving spin. In light of eq. (27) [with E v /dt ∝ β̂ × B̂. But for circular motion,
r̂ · β̂ = 0 and the acceleration d~v /dt points radially into the origin, i.e. d~v /dt ∝ −r̂. It
follows that r̂ = B̂ × β̂, and we conclude that the unit vectors {B̂ , β̂ , r̂} form a mutually
orthogonal right-handed triad of vectors. Thus, we can write:
s⊥ ≡ ~
~ sB + ~
sr , where ~
sB ≡ (~
s · B̂)B̂ and ~
sr ≡ (~
s · r̂)r̂ . (34)
Note that  
d~
sB d~
s
= B̂ · B̂ = 0 , (35)
dt dt
~ is time-independent by assumption and
since B

~ · d~
B
s ~ · (~
∝B ~ = 0,
s × B)
dt
in light of eq. (28). Thus, ~
sB is a constant in time, from which it follows that
d~
sr d d~
s
= (~ sB ) = ⊥ .
s⊥ + ~ (36)
dt dt dt
Hence, the equations of motion for the components of the spin vector are:
d~
sB
= 0,
dt
  
d~
sr eB g−2 1
= sk +
~ s × B̂ ,
~
dt mc 2 γ r
d~
sk
  
eB g−2 1
= sr +
~ s × B̂ ,
~
dt mc 2 γ k

after using ~
sB × B̂ = (~
s · B̂)B̂ × B̂ = 0.

(b) For the CERN Muon Storage Ring, the orbit radius is R = 2.5 meters and B =
17 × 103 gauss. What is the momentum of the muon? What is the time dilation factor γ?
How many periods of precession T = 2π/Ω occur per observed laboratory mean lifetime
of the muons? [Relevant data: mµ = 105.66 MeV, τ0 = 2.2 × 10−6 s, a ≃ α/(2π) where
α ≃ 1/137.]

For circular motion,


d~
v v2
a=
~ = − r̂ . (37)
dt R
13
Since the circular motion is in a plane that is perpendicular to the magnetic field B, ~ it
~
follows that B, ~
v and r̂ are mutually orthogonal vectors. Moreover, eqs. (12.38) and (12.39)
of Jackson yield
d~
v e ~.
= ~
v×B (38)
dt γmc
Thus, if B~ points in the z-direction, then ~
v = −v θ̂ and the circular motion is clockwise in
the x–y plane. Combining eqs. (37) and (38), it follows that
eBR
γmv = , (39)
c
which we recognize as the relativistic momentum of the muon, pµ . Using eq. (12.42) of
Jackson, we can rewrite eq. (39) as3

pµ (MeV/c) = 3 × 10−4 BR (gauss-cm) .

Hence,
pµ = (3 × 10−4 )(1.7 × 104 )(250) MeV/c = 1.275 × 103 MeV/c .
The γ-factor is
1/2
(p2 c2 + m2 c4 )1/2 p2

E
γ= = = +1 .
mc2 mc2 m2 c2
The muon rest energy is mc2 = 105.66 MeV. Hence,
1/2
(1.275 × 103 )2

γ = 1+ = 12.11 .
(105.66)2
The number of periods of precession, T = 2π/Ω, occurring per observed mean muon
lifetime, γτ0 = γ(2.2 × 10−6 s), is given by4
γτ0 γτ0 Ω γτ0 eBa γ 2 τ0 va
= = = ,
T 2π 2πmc 2πR
where eq. (39) was used to arrive at the final result above. Since γ ≫ 1, we can approximate
v ≃ c. In addition, we take
α 1
a = 12 (g − 2) ≃ , where α ≃ ,
2π 137
as predicted at lowest non-trivial order in quantum electrodynamics. Hence,
γτ0 γ 2 τ0 cα (12.11)2 (2.2 × 10−6 s)(3 × 1010 cm s−1 )
≃ = = 7.156 .
T 4π 2 R 4π 2 (250 cm)(137)
3
The factor of 3 × 10−4 arises as follows. In gaussian units, e = 4.8 × 10−10 esu and 1 MeV= 1.6 ×
10 ergs. Hence, the conversion factor between ergs and MeV is 4.8 × 10−10 /1.6 × 10−6 = 3 × 10−4 .
−6
4
Note that in the laboratory frame, the observed muon lifetime is given by γτ0 , where τ0 is the muon
lifetime in the muon rest frame.

14
(c) Express the difference frequency Ω in units of orbital rotation frequency and com-
pute how many precessional periods (at the difference frequency) occur per rotation for a
300 MeV muon, a 300 MeV electron, a 5 GeV electron (this last typical of the e+ e− storage
ring at Cornell).
NOTE: The energy values above correspond to the total relativistic energies.

For a 300 MeV muon,


E 300
γ= = = 2.839 ,
mc 2 105.66
and
eBa γωB α
Ω= = γωB a ≃ = 3.3 × 10−3 ωB .
mc 2π
One revolution occurs in time t = 2πR/v. In this time, the number of periods of precession,
T = 2π/Ω, is given by   
t 2πR Ω ΩR
= = .
T v 2π v
We can rewrite the above result using eq. (39), which yields

R γmc 1
= = .
v eB ωB

Hence, for a 300 MeV muon, we have


t Ω γα
= ≃ = 3.3 × 10−3 .
T ωB 2π

For a 300 MeV electron, we use me c2 = 511 keV to obtain


300
γ= = 587 .
0.511
Hence,
t Ω γα
= ≃ = 0.682 .
T ωB 2π
Finally, for a 5 GeV electron, we have
5000
γ= = 9.785 × 103 .
0.511
It follows that
t Ω γα
= ≃ = 11.37 .
T ωB 2π

15
4. [Jackson, problem 14.4] Using the Liénard-Wiechert fields, discuss the time-averaged
power radiated per unit solid angle in nonrelativistic motion of a particle with charge e,
moving:
(a) along the z axis with instantaneous position z(t) = a cos ω0 (t) ,
(b) in a circle of radius R in the x–y plane with constant angular frequency ω0 .
Sketch the angular distribution of the radiation of the radiation and determine the total
power radiated in each case.

(a) Case 1: Non-relativistic motion of a particle with charge e moving along the z-axis with
instantaneous position z(t) = a cos ω0 (t) .
We make use of eq. (14.20) of Jackson, which is relevant for non-relativistic motion,
! 2
dP 2
e ~

= n̂ × n̂ × , (40)
dΩ 4πc dt

where
~=~
v 1 d~
x
β = .
c c dt
In this case, we have
x(t) = ẑ a cos ω0 t ,
~
which yields
dβ~ aω 2
= −ẑ 0 cos ω0 t .
dt c
Working out the absolute square of the triple product in eq. (40),
! ! !
dβ~ 2 dβ~ dβ~ 2 dβ
~ 2 ~ 2

n̂ × n̂ × = n̂ n̂ · − = − n̂ · (41)

dt dt dt dt dt

a2 ω04 2
 2
 a2 ω04
= cos ω 0 t 1 − (n̂ · ẑ) = 2 cos2 ω0 t sin2 θ .
c2 c
In obtaining the final result above, we chose to work in a coordinate system in which the
origin corresponds to the instantaneous position of the charged particle, and the unit vector
n̂ has polar angle θ and azimuthal angle φ with respect to the z-axis,

n̂ = x̂ sin θ cos φ + ŷ sin θ sin φ + ẑ cos θ . (42)

The time-averaged power is easily obtained by noting that5

hcos2 ω0 ti = 1
2
.
5
To compute the time-average of cos2 ω0 t, note that the time averages satisfy hcos2 ω0 ti = hsin2 ω0 ti,
and cos2 ω0 t + sin2 ω0 t = 1.

16
Hence, it follows that
e2 a20 ω04

dP
= sin2 θ . (43)
dΩ 8πc 3

In Figure 1, the angular distribution of the radiated power is exhibited as a polar plot.

Figure 1: A polar plot of the angular distribution of the power radiated by a charged particle
moving non-relativistically along the z axis with instantaneous position z(t) = a cos ω0 (t). The
angular distribution is given by eq. (43) and is proportional to sin2 θ. This plot was created with
Maple 15 software.

Integrating over the solid angle yields the total radiated power,
e2 a2 ω04
hP i = .
3c3

(b) Case 2: Non-relativistic motion of a particle with charge e moving in a circle of radius R
in the x–y plane with constant angular frequency ω0 .

For circular motion in the x–y plane, the trajectory of the particle is given by
x(t) = R(x̂ cos ω0 t + ŷ sin ω0 t) .
~
Then, we easily compute
dβ~ 1 d2 ~
x ω02
= = − x(t) .
~
dt c dt2 c
We again choose to work in a coordinate system in which the origin corresponds to the
instantaneous position of the charged particle, and the unit vector n̂ given by eq. (42) has
polar angle θ and azimuthal angle φ with respect to the z-axis. Consequently,
~
dβ ω2R
n̂ · = − 0 (cos ω0 t sin θ cos φ + sin ω0 t sin θ sin φ) .
dt c
17
Evaluating the absolute square of the triple cross product as in part (a) [cf. eq. (41)], we
obtain:
!
~ 2 ω 4 R2 

= 0 2 1 − sin2 θ(cos φ cos ω0 t + sin φ sin ω0 t)2

n̂ × n̂ ×

dt c
ω04 R2 
1 − sin2 θ cos2 (ω0 t − φ) .

=
c 2

Using eq. (40), it follows that


dP e2 ω04 R2 
1 − sin2 θ cos2 (ω0 t − φ) .

=
dΩ 4πc 3

The time-averaged power is easily obtained by noting that hcos2 (ω0 t − φ)i = 21 . Employing
the trigonometric identity, 1 − 12 sin2 θ = 12 (1 + cos2 θ) , it follows that
e2 ω04 R2
 
dP
1 + cos2 θ .

= (44)
dΩ 8πc 3

In Figure 2, the angular distribution of the radiated power is exhibited as a polar plot.

Figure 2: A polar plot of the angular distribution of the power radiated by a charged particle
moving non-relativistically in a circle of radius R in the x–y plane with constant angular fre-
quency ω0 . The angular distribution is given by eq. (44) and is proportional to 1 + cos2 θ. This
plot was created with Maple 15 software.

Integrating over solid angles yields the total radiated power,


2e2 ω04 R2
hP i = .
3c3

5. [Jackson, problem 14.5] A nonrelativistic particle of charge ze, mass m and kinetic
energy E makes a head-on collision with a fixed central force field of finite range. The
interaction is repulsive and described by a potential V (r), which becomes greater than E
at close distances.

18
(a) Show that the total energy radiated is given by
r Z ∞ 2
4 z 2 e2 m dV dr
∆W = p ,
3mc 2 3 2 rmin dr
V (rmin ) − V (r)
where rmin is the closest distance of approach in the collision.

Consider a particle with kinetic energy E at t = −∞ that is initially an infinite distance


away and is headed in a radial direction toward the origin. Because the potential V (r)
is repulsive and becomes greater than E at close distances, there is a distance of closest
approach, rmin , where the particle’s radial velocity drops to zero. At this point, the particle
“collides” head on with the central force field and is turned around. It now travels back
along its original radial path until it reaches its original starting point (an infinite distance
away) at t = ∞.
If the particle does not radiate, then we can use energy conservation to compute the
instantaneous velocity of the particle at all points along its trajectory. In particular, the
conservation of the sum of the kinetic and potential energy yields

E = 21 m[v(r)]2 + V (r) , for rmin ≤ r < ∞ , (45)

since the potential is assumed to be of finite range which means that limr→∞ V (r) = 0. At
the point of closes approach to the origin, v(rmin ) = 0. Hence, it follows from eq. (45) that

E = V (rmin ) . (46)

Writing v(r) = dr/dt, we can solve for the velocity using eq. (45),
r
dr 2p
v(t) = =± E − V (r) , (47)
dt m
where we employ the minus sign as the particle moves toward the origin and the plus sign as
the particle moves away from the origin. The acceleration can be obtained by differentiating
eq. (47) with respect to t. However, a more direct computation uses Newton’s second law,

~ = m~
F ~ (r) = −r̂ dV ,
a = −∇V
dr
which yields
1 dV
a = −r̂
~ . (48)
m dr
During the period of acceleration, the particle radiates and hence loses energy, ∆W .
Thus, it is not justified to ignore this energy loss in the energy balance equation given in
eq. (45). Nevertheless, if ∆W ≪ E, then it is justified in first approximation to ignore the
radiated energy loss in deriving the acceleration given in eq. (48).6 Thus, we shall assume
6
To properly take the radiation loss into account, we must address the question of radiation reaction,
which is treated in Chapter 16 of Jackson. It turns out that this topic involves numerous subtleties, not
all of which are completely understood. For further details, check out the first few sections of Chapter 16.

19
that one can neglect the energy loss due to radiation and check for consistency at the end
of the computation. In this case, we can use the Larmor formula [cf. eq. (14.22) of Jackson]
for the instantaneous power emitted by a nonrelativistic, accelerated charge,
2
2 z 2 e2 2 2 z 2 e2 dV
P = |~
a| = . (49)
3 c3 3 m2 c3 dr

where eq. (48) has been used for the acceleration. To compute ∆W , we first consider the
energy emitted by the radiation from the initial position of the particle at r = ∞ until the
point of closest approach to the origin, rmin. Then,7
Z rmin r Z ∞
dt m dr
Z

∆W = P r(t) dt = P (r) dr = P (r) . (50)
∞ dr 2 rmin E − V (r)

Since we are neglecting the energy loss in computing the instantaneous acceleration of the
particle, the energy loss of the particle as it moves from the distance of closest approach back
out to infinity again yields eq. (50). Hence the total energy radiated during −∞ < t < ∞
is just twice that of eq. (50),
r Z ∞
m dr
∆W = 2 P (r) ,
2 rmin V (rmin) − V (r)

after employing eq. (46). Finally, we substitute for P using eq. (49), which yields
r Z ∞ 2
4 z 2 e2 m dV dr
∆W = p , (51)
3mc 2 3 2 rmin dr
V (rmin ) − V (r)

To justify this computation, we would have to show that

∆W ≪ 21 mv02 , (52)

where E ≡ 12 mv02 and v0 is the initial velocity at time t = −∞. Since the motion is non-
relativistic, we have v0 ≪ c, and one can check that for for reasonable potentials, eq. (52)
is satisfied [cf. eq. (56)].

(b) If the interaction is a Coulomb potential V (r) = zZe2 /r, show that the total energy
radiated is
8 zmv05
∆W = ,
45 Zc3
where v0 is the velocity of the charge at infinity.

7
Note that we use the minus sign in eq. (47) when the particle moves toward the origin. This minus
sign is then used to reverse the limits of integration in eq. (50).

20
Substituting V (r) = zZe2 /r into eq. (51) yields
r Z ∞
4 z 2 e5 3/2 m dr 1
∆W = (zZ) r . (53)
3 m2 c3 2 rmin r 4 1 1

rmin r
It is more convenient to rewrite this as

r
4 z 2 e5 m dr 1
Z
∆W = (zZ)3/2 .
r
r
3 m2 c3 2 rmin r 7/2
−1
rmin
We now change variables by defining
r
u= − 1.
rmin
Then dr = rmin du and r = (u + 1)rmin. Hence,
Z ∞ Z ∞
dr 1 1 du
= √ .
r
r
rmin r u (u + 1)
7/2 5/2 7/2
−1 r min 0
rmin
2

We make one more
√ change of variables by defining u = x . Then du = 2xdx = 2 u du so
that 2dx = du/ u. Hence,
Z ∞ Z ∞
dr 1 2 dx
= 5/2 . (54)
r
r
rmin r rmin 0 (x + 1)
7/2 2 7/2
−1
rmin
To evaluate this integral, we first consider the well-known result,
dx x
Z
= √ .
(x + a )
2 2 3/2
a x2 + a2
2

The desired integral can be obtained by differentiating twice with respect to a. Alterna-
tively, one can use an integral table to obtain
x3 x5
 
dx 1 x 2 1
Z
= 6 √ − + .
(x2 + a2 )7/2 a x2 + a2 3 (x2 + a2 )3/2 5 (x2 + a2 )5/2
Hence,
Z ∞ ∞
dx x 2 x3 1 x5 =1− 2 + 1 = 8 .

= √ − +
0 (x2 + 1)7/2 x2 + 1 3 (x2 + 1)3/2 5 (x2 + 1)5/2 0 3 5 15
It follows that ∞
dr 1 16
Z
= .
r
r
rmin r 7/2 5/2
15rmin
−1
rmin

21
Plugging this result back into eq. (53), we end up with
r
4 z 2 e5 3/2 m 16
∆W = (zZ) 5/2
. (55)
3mc 2 3 2 15rmin

We can simplify this expression by writing E = 12 mv02 , where v0 is the initial velocity of
the particle at t = −∞. Using eq. (46),

zZe2
1
2
mv02 = V (rmin ) = .
rmin
Solving for rmin and inserting this result back into eq. (55) yields our final result,
r 5/2
64 z 2 e5 mv02 8 zmv05

m
∆W = (zZ)3/2 = .
45 m2 c3 2 2zZe2 45 Zc3

The condition of eq. (52) then implies that


16z  v0 3
≪ 1, (56)
45Z c
which is always satisfied for non-relativistic motion [assuming that z/Z ∼ O(1)].

22

Anda mungkin juga menyukai