You are on page 1of 50

Physics H7C Section Notes

Kevin Grosvenor

Last Edited: April 24, 2011


b
Contents

1 Special Relativity 1
1.1 Relativistic Train 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Relativistic Train 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Colliding Photons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.4 Length Contraction in Arbitrary Direction . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2 The Propagation of Light 9


2.1 Rayleigh Scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2 Apparent Depth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

3 Geometrical Optics 11
3.1 Refraction at a Spherical Interface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3.2 Hecht P.6.8 p.278 (Modified) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.3 Silvered Thin Lens . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

4 Interference 15
4.1 Laser Wavelength Measurement via Metal Ruler . . . . . . . . . . . . . . . . . . . . . . 15
4.2 Continuous Case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

5 Polarization 19
5.1 Hecht P.8.32 p.381 (Augmented) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
5.2 Hecht P.8.41 p.382 (Augmented) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

6 Midterm Review 23
6.1 Michelson Interferometer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
6.2 Pion Photoproduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
6.3 Microscopes and Embryos . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
6.4 Three-Slit Interference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
6.5 Polarizers Away! . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

7 Blackbody Radiation 31
7.1 Planck Distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
7.2 Planck Spectrum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
7.3 Stefan-Boltzmann Integral . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

8 Quantum Operator Expectation Values 37


8.1 Infinite Square Well . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

i
ii CONTENTS

9 Operators and States 41


9.1 Harmonic Oscillator Bra-Kets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

10 Final Review 45
10.1 Wavefunction Shapes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
10.2 Harmonic Oscillator Coherent States . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
10.3 Finite Square Well Bound States . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
10.4 A Theorem about Bound States . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
10.5 Spin-Spin Interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
10.6 Fine and Hyperfine Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
Chapter 1
Special Relativity

1.1 Relativistic Train 1


A relativistic train moves at velocity 0.8c in the +x direction. Observer A is at the back of the
train. Observer B is at the front of the train. The conductor, C, is at the middle of the train. The
stationmaster, S, stands on the platform as the train passes. The conductor, C, and the stationmaster,
S, agree to set their clocks to t = 0 when they are directly next to each other and to define this position
as x = 0. A, B and C all have identical synchronized clocks. The conductor, C, measures the train to
be two lightseconds (ls) long.

(a) At t = 0, C turns on a bulb. What time will A and B read on their clocks when they see the light?

(b) What does S measure for the length of the train?

(c) What time(s) does S observe on his clock when he sees the light reach A and B.

(d) What do your answers to (a) and (c) tell you about simultaneity? Draw two spacetime diagrams
for this process: one for the train reference frame and one for the platform reference frame.

SOLUTION:

(a) A and B are at rest relative to C and thus measure the same length for the train. In their
perspective, they are each 1 ls away from C. From their perspective, it takes the light one second
to go from C to them. Thus, A and B will both read 1 sec.

(b) The train will appear contracted to S by a factor of γ = 5/3. Thus, its length is 2/(5/3) = 6/5 = 1.2
ls for the stationmaster.

(c) From the point of view of S, observer A is moving towards the light ray and observer B is moving
away. Thus, from the point of view of S, the relative speed of the light ray and A is 1.8c whereas
it is 0.2c for the light ray and B. The distance that needs to be covered for each is 0.6 ls (half of
part (b)). Thus, the time as viewed by S for A is (0.6 ls)/(1.8c) = 1/3 = 0.33 sec and the time for
B is (0.6 ls)/(0.2c) = 3 sec.
You might want to think of this in the follow way instead. We can write the trajectory of A as
xA = −0.6 ls + 0.8ct and the trajectory of B as xB = 0.6 ls + 0.8ct. Meanwhile, the light ray going
towards A has the trajectory x`1 = −ct and the one heading for B is x`2 = ct. The time when
xA = x`1 is t = 0.33 sec and the time when xb = x`2 is t = 3 sec.

1
2 CHAPTER 1. SPECIAL RELATIVITY

(d) In the frame of reference of the train (and A, B and C), the light reaches the back and the front
of the train at the same time. But, from the point of view of S, it reaches the back in 0.33 sec and
reaches the front in 3 sec. Thus, events that appear simultaneous in one frame will generically not
appear simultaneous in a different frame.
You might be wondering what S observes on the clocks of A and B when the light reaches each
one (as opposed to what he observes on his own clock at these two events). On the answer to this
question, the two reference frames cannot possibly disagree. That is, A and B see their clocks
read 1 sec, when the light reaches them. S must also see 1 sec on A’s and B’s clock when the
light reaches them. But, we already know that S sees the light reach A long before he sees it
reach B. Hence, we must conclude that, while S observes A’s and B’s clocks ticking at the same
rate, he observes that A’s clock is systematically ahead of B’s clock by a fixed amount, namely
3 − 0.33 = 2.67 sec. Even though A’s and B’s clocks are synchronized in their reference frame,
they are not synchronized in S’s reference frame!
Below (left) is the spacetime diagram from the train perspective and below (right) in the platform
perspective. The vertical direction is time with each mark corresponding to 0.5 sec. The horizontal
direction is the x-direction with each mark corresponding to 0.5 ls. The dotted lines are the
trajectories of each observer and the solid directed line shows the light signal propagating from
C to A and B. In the train perspective, S is moving with velocity −0.8c x̂. Thus, the slope of
the trajectory of S in this frame is 1/(−0.8) = −5/4. Similarly, the slopes of the A, B and C
trajectories in the platform frame is +5/4. Meanwhile, all light signals have slopes of ±1 in both
frames. Note that A and B are 1 ls away from C in the train frame, whereas they are 0.6 ls away
from C in the platform frame. The squiggly dots show the moment when the light signal reaches
A and B. In the train frame, these occur at the same time whereas in the platform frame, these
occur at drastically different times.

1.2 Relativistic Train 2


A train and a tunnel both have proper lengths L. The train moves towards the tunnel at speed v. A
bomb is located at the front of the train and is designed to explode when the front of the train passes
the far end of the tunnel. A deactivation sensor is located at the back of the train. When the back of
the train passes the near end of the tunnel, the sensor tells the bomb to disarm itself. Does the bomb
explode?

SOLUTION:

Yes, the bomb explodes. Let us first consider the train reference frame, in which the answer is obvious.
In this frame, the train has length L and the tunnel has length L/γ < L and is heading towards the
1.2. RELATIVISTIC TRAIN 2 3

train at speed v. Therefore, it is clear that the back end of the tunnel will pass the front of the train
before the front end of the tunnel reaches the back of the train.
In the tunnel frame, the tunnel has length L and the train has length L/γ < L. Therefore, the
back of the train reaches the near end of the tunnel before the front of the train reaches the back of
the tunnel. You might be tempted to say that the bomb is then deactivated before it can explode.
However, you have to keep in mind that the deactivator at the back of the train needs to send a signal
to the bomb at the front of the train saying that it has reached the front end of the tunnel and that
the bomb should therefore disarm itself. At best, that signal can travel at the speed of light. It will
take time for that signal to reach the bomb at the front of the train. If that time is longer than the
time it takes for the front of the train to reach the back of the tunnel, then it will be too late and the
bomb will explode.
Let the front of the tunnel correspond to x = 0 and let t = 0 be when the back of the train passes
the front of the tunnel. Henceforth, the signal sent by the deactivator travels forward at the speed of
light, its worldline described by xs = ct (the s subscript stands for “signal”). At t = 0, the front of the
train is at x = L/γ, since that is the length of the train in the tunnel reference frame. The trajectory
of the front of the train is xb = L γ + vt (the b subscript stands for “bomb”). Which one reaches x = L
(the back of the tunnel) first? Well,
 the time it takes for the signal is ts = L/c whereas for the bomb
ts
takes tb = L − L 1

γ /v = β 1 − γ , where β ≡ v/c. We claim that tb < ts and so the bomb explodes.
To prove this, start with the inequality β < 1, which just says that the train must be moving at
less than the speed of light. Multiply by 2β and add 1 to both sides to get 1 + 2β 2 < 1 + 2β. Now,
subtract 2β + β 2 from both sides to get 1 − 2β + β 2 < 1 − β 2 . p Rewrite the left hand side as (1 − β)2 ,
then take the positive square root of both sides to get 1 − β < 1 − β 2 = γ1 . This final inequality can
be rearranged to read β1 1 − γ1 < 1. But, the left hand side is just tb /ts , and so tb < ts .


Below are spacetime diagrams in both reference frames in the case β = 4/5. Note that we have set
t = t0 = 0 when the front of the train lines up with the front of the tunnel. But, note that we have not
set x = x0 = 0 to be the position of this event. The spatial origins of the frames are different: x0 = 0
for the center of the train and x = 0 for the center of the tunnel. In the train frame, the explosion
happens before the deactivation signal is sent. In the tunnel frame, those two events occur in the
opposite order. However, in both reference frames, the bomb explodes; it certainly cannot be the case
that the train explodes in one frame whereas it does not in the other! Assuming that the explosion
“signal” (i.e. the fires, etc.) travel at the speed of light, the red shaded regions represent the region of
the train that is engulfed in fire, or at least the region that is aware of the fact that the explosion has
occurred.
4 CHAPTER 1. SPECIAL RELATIVITY

1.3 Colliding Photons


[Goldstein, Poole & Safko 7.22 ] A photon of energy E 0 collides at angle θ with a photon of energy E.
Determine the minimum value of E 0 permitting the formation of a pair of particles of mass m.

SOLUTION:

Set c = 1 for the moment. Let the four-momenta of the photons be

p = E(1, 1, 0, 0), p0 = E 0 (1, cos θ, sin θ, 0).

This is written in a coordinate system where x points in the direction of propagation of the photon
with energy E. In these coordinates, which we will call x0 and y 0 , the total 4-momentum is

p0t = p + p0 = (E + E 0 , E + E 0 cos θ, E 0 sin θ, 0).

These coordinates turn out to be inconvenient. Define the x and y coordinates to be such that x is in
the direction of the total 3-momentum (the last three components of the total 4-momentum). These
coordinates will simply be rotated with respect to the x0 and y 0 coordinates. Thankfully, we do not
really need to know what this rotation is, since we don’t care about what p and p0 look like in these
coordinates. What we know is that, if pt is the total 4-momentum in these unprimed coordinates,
and p0t and pt are the 3-momentum parts of p0t and pt , then |p0t | = |pt |, since rotations do not change
magnitudes. Conveniently, in these coordinates, the y and z components of pt vanish, so we can just
neglect them altogether and just write the t and x components:

E + E0 E + E0 1
     
√ 0
pt = = = (E + E ) q 0
(1−cos θ) .
|p0t | E 2 + E 02 + 2EE 0 cos θ 1 − 2EE (E+E 0 )2
| {z }
A

Henceforth, we may just drop the y and z entries since they are identically zero. Now, boost in the
+x direction by some amount β:
     
1 −β 1 1 − βA
γ (E + E 0 ) = γ(E + E 0 ) .
−β 1 A −β + A

To make the boosted frame the center of momentum frame (COM), we need the x component of the
boosted total momentum to vanish, which implies β = A. The corresponding γ factor is
−1/2
2EE 0 (1 − cos θ) E + E0

γ = (1 − β 2 )−1/2 = 1−1+ =p .
(E + E 0 )2 2EE 0 (1 − cos θ)

Thus, the total energy in the COM is

E + E0 p
Ecom = γ(E + E 0 )(1 − βA) = = 2EE 0 (1 − cos θ),
γ

where use was made of β = A and 1 − β 2 = γ −2 .


1.4. LENGTH CONTRACTION IN ARBITRARY DIRECTION 5

Actually, there is a much simpler way to get to this result for Ecom . The combination Aµ Aµ is
Lorentz inveriant for any covariant vector, Aµ . Recall that Aµ Aµ ≡ A20 − A21 − A22 − A23 . For p0t this
gives
(p0t )µ (p0t )µ = (E + E 0 )2 − (E + E 0 cos θ)2 − (E 0 sin θ)2 = 2EE 0 (1 − cos θ).
The total 4-momentum in the COM is pt,com = (Ecom , 0, 0, 0) with invariant

(pt,com )µ (pt,com )µ = Ecom


2
.

The invariant for p0t must be equal to the invariant for pt,com thereby yielding the previously calculated
result for Ecom .
We must have Ecom ≥ 2m in order to form a pair of particles of mass m, thus yielding

2m2
E0 ≥ .
E(1 − cos θ)

Now, we must reintroduce c. In order for the RHS to have units of energy, we must multiply by c4 :

2m2 c4
E0 ≥ .
E(1 − cos θ)

1.4 Length Contraction in Arbitrary Direction


Consider a cube of side-length 1 (in some appropriate units). Let S be the rest frame of the cube in
which the sides of the cube are parallel to the coordinate axes, one corner is at the origin, and the cube
occupies the positive octant (i.e. the coordinates of any point in the cube are non-negative). Let S 0
be a frame whose axes are parallel with those of S, but with respect to which the cube is moving with
velocity βcx̂0 . Let S 00 be a frame that is rotated in the xy-plane relative to S 0 in the counter-clockwise
direction by an angle θ. Determine the volume of the cube in the frames S 0 and S 00 .

[Note: check out Wikipedia for the volume of a Parallelepiped.]

SOLUTION:
−1/2
The volume in S 0 is 1/γ, where γ = 1 − β 2

. Hopefully, this much is familiar from problem set 1.

Incorrect solution for S”: First, let us try to follow the naı̈ve procedure that I had proposed
in section: break up the velocity into its x00 and y 00 components in frame S 00 and contract each
direction accordingly. So, βx00 = β cos θ and βy00 = β sin θ, with corresponding gamma factors, γx00 =
 −1/2  −1/2
1 − βx200 and γy00 = 1 − βy200 . After some algebra, we can solve for these gamma factors in
0
 −1/2
terms of the gamma factor of S with respect to S, namely γ = 1 − β 2 :

1 h cos2 θ i1/2 1 h sin2 θ i1/2


= 2
+ sin2 θ , = 2
+ cos2 θ .
γx00 γ γy00 γ
Therefore, we might expect the volume to decrease by
1 1 1 1/2
= 1 + β 4 γ 2 cos2 θ sin2 θ ,
γx00 γy00 γ
where I have performed a fair amount of algebra to simplify the answer to this form.
Notice that this is not the same as the volume we found in S 0 , which is 1/γ. However, the difference
is of order β 4 , which is at least very small except in the limit β → 1.
6 CHAPTER 1. SPECIAL RELATIVITY

Correct solution: Since S 00 is related to S 0 only by a spatial rotation, and spatial rotations DO NOT
change spatial distances, the volume in S 0 ought to be the same as the volume in S 00 exactly, not only
approximately! Let us rigorously trace the steps to get the volume in S 0 , then it should be clear how
to extend the procedure to S 00 .
Let the spacetime coordinates of the cube corners in S be

p1 = (t, 0, 0, 0), p5 = (t, 1, 1, 0),


p2 = (t, 1, 0, 0), p6 = (t, 1, 0, 1),
p3 = (t, 0, 1, 0), p7 = (t, 0, 1, 1),
p4 = (t, 0, 0, 1), p8 = (t, 1, 1, 1).

Note that the spatial coordinates are t-independent since S is the cube’s rest frame. Also, so that we
can keep track of volume, p2 can be considered the width, p3 the length, and p4 the height, where pi
is the spatial vector corresponding to pi (i.e. the last three components). Therefore, the volume in S
is 1, as expected:
 
(p2 × p3 ) · p4 = (1, 0, 0) × (0, 1, 0) · (0, 0, 1) = (0, 0, 1) · (0, 0, 1) = 1.

We must boost these coordinates into the S 0 frame, which is moving relative to S with a velocity
−βc x̂. The appropriate boost transformation is
 
γ βγ 0 0
βγ γ 0 0
Λ= 0
.
0 1 0
0 0 0 1

The coordinates of the corners in S 0 are p0i = Λpi :

p01 = (γt, βγt, 0, 0), p05 = (γ(t + β), γ(βt + 1), 1, 0),
p02 = (γ(t + β), γ(βt + 1), 0, 0), p06 = (γ(t + β), γ(βt + 1), 0, 1),
p03 = (γt, βγt, 1, 0), p07 = (γt, βγt, 1, 1),
p04 = (γt, βγt, 0, 1), p08 = (γ(t + β), γ(βt + 1), 1, 1).

This form is not too useful. We want to write each 4-vector in terms of t0 , where t0 is set equal to the
zeroth component of each 4-vector separately. For example, for p01 , set t0 = γt, then x01 = βγt = βt0 . As
another example, for p02 , set t0 = γ(t + β), then, after some algebra, one finds x02 = γ(βt + 1) = βt0 + γ1 :

p01 = (t0 , βt0 , 0, 0), p05 = (t0 , βt0 + γ1 , 1, 0),


p02 = (t0 , βt0 + γ1 , 0, 0), p06 = (t0 , βt0 + γ1 , 0, 1),
p03 = (t0 , βt0 , 1, 0), p07 = (t0 , βt0 , 1, 1),
p04 = (t0 , βt0 , 0, 1), p08 = (t0 , βt0 + γ1 , 1, 1).

Now, let us see the positions of the cube corners in S 0 at some fixed time in S 0 , say t0 = 0:

p01 = (0, 0, 0), p05 = ( γ1 , 1, 0),


p02 = ( γ1 , 0, 0), p06 = ( γ1 , 0, 1),
p03 = (0, 1, 0), p07 = (0, 1, 1),
p04 = (0, 0, 1), p08 = ( γ1 , 1, 1).

The volume in S 0 is 1/γ, as expected:

(p02 × p03 ) · p04 = ( γ1 , 0, 0) × (0, 1, 0) · (0, 0, 1) = (0, 0, γ1 ) · (0, 0, 1) = γ1 .


 
1.4. LENGTH CONTRACTION IN ARBITRARY DIRECTION 7

Now, the appropriate rotation matrix that takes S 0 into S 00 is


 
1 0 0 0
0 cos θ sin θ 0
R= 0 − sin θ cos θ 0

0 0 0 1

Then, p00i = Rp0i are the spacetime coordinates of the cube corners in S 00 . Note that R does not change
the zeroth (time) component and so t00 = t0 . You can go ahead and do the algebra to calculate the
volume, but the result is obvious: spatial rotations cannot change lengths and thus the volume will
still be 1/γ.

[Note: I invite you to try and see what happens if you first rotate using R from S to another frame,
S 000 , then boost by −βx00 in the horizontal direction, and then boost by +βy00 in the vertical direction.
My initial calculations are encouraging in the sense that I get approximately the same result as the
“incorrect solution”. But, it’s such a big mess of a calculation that I’m just going to stop and not
write any more. You may want to ruminate on the distinction between the frame one gets with this
set of transformations versus the frame one gets from first boosting from S to S 0 and then rotating
from S 0 to S 00 .]
8 CHAPTER 1. SPECIAL RELATIVITY
Chapter 2
The Propagation of Light

2.1 Rayleigh Scattering


[Hecht P.4.2 (modified and augmented)] A white floodlight beam crosses a large volume containing
a tenuous molecular gas mixture of mostly oxygen and nitrogen. Compare the relative amount of
scattering occurring for the red (∼ 750 nm) component with that of the blue (∼ 450 nm) component.
What does this have to say about the color of sky and the sunset?

SOLUTION:

This problem is about Rayleigh scattering, whose intensity goes like ∼ 1/λ4 . Thus,
red scattering (1/750 nm)4
= = 0.13. (2.1)
blue scattering (1/450 nm)4
When sunlight passes through the atmosphere, the blue component is scattered more than the red.
What we see as the color of the sky is just scattered sunlight, which will be dominated by blue. What
we see in the sunset is mostly direct sunlight, the blue component of which has mostly been scattered
by the time it travels through so much atmosphere, leaving mostly red.

2.2 Apparent Depth


[Hecht P.4.25 ] A coin is resting on the bottom of a tank of water (nw = 1.33) 1.00 m deep. On
top of the water floats a layer of benzene (nb = 1.50), which is 20.0 cm thick. Looking down nearly
perpendicularly, how far beneath the topmost surface does the coin appear?

SOLUTION:

Consider the following diagram

9
10 CHAPTER 2. THE PROPAGATION OF LIGHT

There are a couple of things to note about this diagram: in order to actually show the angles, we have
made them quite large, when, in actual fact, they’re supposed to be very small. Secondly, in order
to show the separate triangles (black, red, and blue), we have displaced them relative to each other,
slightly. The light ray (lines with arrows) is black in the water, red in benzene, and blue in air, but
this is unrelated to their actual color in those media! The incident and transmitted angles relative the
normal at each surface in water, benzene, and air are denoted θw , θb , and θa . By simple geometry,
these also happen to be the angles at the bottom corners of the appropriate triangles (θw for the black
triangle, θb for the red triangles, and θa for the blue triangle). Finally, the apparent depth relative to
the topmost surface (benzene-air interface) is h00 .
The point of looking down nearly perpendicularly is to make the incident and refracted angles
small, and thus to justify the small angle approximation sin θ ≈ tan θ ≈ θ. In this case, Snell’s law
reads ni θi = nt θt , or θθti = nnti . Thus, θθwb = nnwb and θθab = nnab .
Comparing the black triangle with the smaller red triangle, we have the two equations
x x
tan θw = , tan θb = .
h h0
Dividing these two equations and using the small angle approximation yields

h0 tan θw θw nb
= ≈ ≈ .
h tan θb θb nw
Comparing the larger red triangle with the blue triangle, we have the two equations
y y
tan θb = , tan θa = .
h0 + t h00
Dividing these two equations and using the small angle approximation yields

h00 tan θb θb na
= ≈ ≈ .
h0 +t tan θa θa nb
Thus, the apparent depth is
na 0 na h nb i 1 h 1.50 i
h00 = (h + t) = h+t = (1.00 m) + 0.20 m = 0.883 m = 88.3 cm.
nb nb nw 1.50 1.33
Chapter 3
Geometrical Optics

3.1 Refraction at a Spherical Interface


[Hecht P.5.5 ] Making use of the diagram below, Snell’s Law, and small angle approximations for all
the relevant angles, derive the equation nso1 + nsi2 = n2 −n
R .
1

SOLUTION:

θ1 is the supplement of ∠ABG = 180 − α − ϕ. Therefore, θ1 = α + ϕ. Similarly, ϕ is the supplement


of ∠BGE = 180 √− θ2 − β. Therefore,
p ϕ = θ2 + β, or θ2 = ϕ − β.
CD = R − R2 − h2 = R 1 − 1 − (h/R)2 ≈ R 1 − 1 + 21 (h/R)2 ≈ 0, where we have dropped
 

the quadratic term, (h/R)2 , since small angles α, β and ϕ imply small h/R.
The small angle approximations for α, β and ϕ now read
h h h
α ≈ tan α ≈ , β ≈ tan β ≈ , ϕ ≈ sin ϕ = .
so si R
Therefore, the corresponding small angle approximations for θ1 and θ2 read
h h h h
sin θ1 = sin(α + ϕ) ≈ α + ϕ ≈ + , sin θ2 = sin(ϕ − β) ≈ ϕ − β ≈ − .
so R R si
Therefore, Snell’s Law reads
n1 h n1 h n2 h n2 h
n1 sin θ1 = n2 sin θ2 =⇒ + = − ,
so R R si
from which the desired relation follows directly.
From here on, the derivation of the lensmaker formula proceeds as in Hecht resulting in Hecht’s
equation (5.14) with (5.15) as the thin lens limit.

11
12 CHAPTER 3. GEOMETRICAL OPTICS

3.2 Hecht P.6.8 p.278 (Modified)


Compute the magnification that results when the image of a flower 4.0 m from the center of a solid,
clear-plastic sphere with a 0.20-m radius (and a refractive index of 1.4) is cast on a nearby wall. De-
scribe the image in detail.

SOLUTION:

Denote the object and image distances with respect to the vertices with overlines. Those without
overlines will be taken with respect to the principal points. Note that s̄o1 = 3.8 m since it is with
respect to the left vertex and not the center! Let nm = 1 be the index of the medium and let n` = 1.4
be the index of the lens. Let R1 = 0.2 m and R2 = −0.2 m be the radii of the front and back surfaces,
respectively. Let d = 0.4 m be the thickness of the lens.

Method 1: For the first surface,


nm n` n` − nm
+ = .
s̄o1 s̄i1 R1

Everything is known here except s̄i1 whose solution yields s̄i1 = 0.806 m, which is with respect to the
left vertex. The transverse magnification is MT 1 = − nnm` s̄s̄o1i1 = − 1(0.806)
1.4(3.8) = −0.151.
Now, for the second surface:
n` nm nm − n`
+ = .
d − s̄i1 s̄i2 R2
Again, everything is known except for s̄i2 whose solution yields s̄i2 = 0.184 m, which is now with
1.4(0.184)
respect to the right vertex. The transverse magnification is MT 2 = − nnm` s̄s̄i2o2 = 1(0.4−0.806) = 0.633.
The total transverse magnification is MT = MT 1 MT 2 = −0.0959. The image height is about a tenth
of the original flower height and it is up-side-down. Let us assume the paraxial limit and neglect all
aberrations.

Method 2: Let us first calculate the focal length with respect to the principal points. Define
n`m ≡ n` /nm = 1.4 as the relative index of refraction.

1  1 1 (n`m − 1)d 
= (n`m − 1) − + .
f R1 R2 n` R1 R2

Plugging in the numbers gives f = 0.35 m. Next, we calculate the principal points:

(n`m − 1)df (n`m − 1)df


h1 = − = 0.2 m, h2 = − = −0.2 m.
n` R2 n` R1

This means that H1 is 0.2 m to the right of V1 and H2 is 0.2 m to the left of V2 , which puts them both
right at the center. We could just remember the fact that for a perfect sphere, the principal points
converge at the optical center. Thus, the distances with respect to the principal points are actually
with respect to the optical center, which simplifies things! In particular, so = 4 m and we have
1 1 1
+ = =⇒ si = 0.384 m.
so si f

This agrees with Method 1 since si = 0.384 = 0.184 + 0.2 = s̄i2 − h2 . The transverse magnification
also agrees:
f f 0.35
MT = − =− =− = −0.0959.
xo so − f 4 − 0.35
3.3. SILVERED THIN LENS 13

3.3 Silvered Thin Lens


Suppose you have a double convex lens made of glass (n = 1.5) such that the magnitude of the radius
of curvature of both sides is R = 42 m. One of the sides is painted with a layer of silver and thus acts
like a spherical mirror. Assume the thin lens approximation and thus, that the lens and the mirror
are essentially at the exact same position. An object lies to the left of the unsilvered face of the lens
a distance so = 84 m away.

(a) What is the focal length of the lens by itself? What about the mirror by itself?

(b) How many stages are there in this optical system. For each stage, state whether the light is incident
from the left or from the right.

(c) For each stage, calculate the image distance and linear magnification using the image of one stage as
the object for the following stage. Make a diagram for each stage separately. Caution: remember
that using the image of one stage as the object for another does not mean using the image distance
of one stage as the object distance of another.

(d) Where is the final image located and what is the total linear magnification? Describe the image
(i.e. is it real or virtual, upright or inverted, bigger or smaller?)

SOLUTION:

(a) Since the lens is double convex, the center of curvature of the left surface is on the right and
the right surface is on the left. Thus, R1 = R and R2 = −R, by the sign conventions, for light
originating from the object on the left. From the Lensmaker equation,
1 1 1 1 1 1
 
f` = (n − 1) R1 − R2 = (1.5 − 1) R − −R = R =⇒ f` = R = 42 m.

For the mirror, the focal length is just half of the radius of curvature. Since this is a concave
mirror, it is converging and thus has a positive focal length:

fm = R/2 = 21 m.

(b) There are three stages:

(1) The lens. Light goes from left to right through the lens;
(2) The mirror. Light goes from left to right, reflects off of the mirror, and goes right to left;
(3) The lens again. Light goes from right to left through the lens.

(c) Let soj and sij be the object and image distances for stage j = 1, 2, 3 and so so1 ≡ so = 84 m = 2R
is the initial object distance and si3 ≡ si is the final image distance.

(1) The image distance just for the lens alone (stage 1) is
1 1 1 1 1 1
si1 = f` − so1 = R − 2R = 2R =⇒ si1 = 2R = 84 m.

The linear magnification of stage 1 is

M1 = − sso1
i1 84
= − 84 m
m = −1.
14 CHAPTER 3. GEOMETRICAL OPTICS

(2) Since si1 is positive, image 1 lies to the right of the lens. This is the same as object 2. Thus,
object 2 lies to the right of the mirror whereas the light is actually hitting the mirror from the
left! This is a virtual object with negative object distance: so2 = −si1 = −2R = −84 m. So,
1 1 1 2 1 5
si2 = fm − so2 = R − −2R = 2R =⇒ si2 = 52 R = 16.8 m.

The linear magnification of stage 2 is

M2 = − sso2
i2
= − 16.8
−84
m
m = 15 .

(3) Since si2 is positive, image 2 lies to the left of the mirror. This is the same as object 3. Thus,
object 3 lies to the left of the lens whereas the light is actually entering the lens from the right!
This is a virtual object with negative object distance: so3 = −si2 = − 52 R = −16.8 m. So,
1 1 1 1 5 7
si3 = f` − so3 = R − −2R = 2R =⇒ si3 = 72 R = 12 m.

The linear magnification of stage 3 is

M3 = − sso3
i3 12 m
= − −16.8 5
m = 7.

Below, virtual objects are drawn in gray. Some rules are tricky: for example, in stage 2, we
can’t actually draw the light ray that starts from the object parallel to the axis and reflects
off of the mirror heading towards the focal point. This is because the object is virtual and
behind the mirror - there is no actual light there! In fact, the light has to go left to right. So
we have to draw the light ray that looks like its headed towards the object parallel to the axis,
and then this ray reflects towards the focal point.
(d) The final image is 12 m to the left of the lens. The total linear magnification is M = M1 M2 M3 =
−1/7. The image is real, inverted and smaller by a factor of 7.
Chapter 4
Interference

4.1 Laser Wavelength Measurement via Metal Ruler


Devise and explain a method for measuring the wavelength of a laser pointer chiefly using a finely
graded metal ruler (e.g. with 1/64 inch markings or smaller).

SOLUTION:

Consider reflecting the laser off of the ruler at a shallow angle. If there were no notches on the surface
of the ruler, then each point on the ruler where the light hits becomes a source for outwardly spreading
spherical waves. The superposition of these waves produces wavefronts that travel in the direction of
specular reflection. That is, on a far-away screen, we get constructive interference only around the
point of specular reflection, as expected.
Imagine we make wide notices with narrow reflective bands in between. Then, consider the following
diagram showing two adjacent light beams headed towards a far-away screen (e.g. a wall) having
reflected off of two adjacent reflective bands.

The optical path length difference is d(cos α − cos β), which must be set equal to mλ for some integer
m for constructive interference. For a fixed α (angle at which we shine the laser on the ruler), this
gives discrete values for β where bright spots occur (i.e. we get a diffraction pattern).
The claim is that we see the same thing if instead we have narrow non-reflective notches with wider
reflective bands. Can you think of why? Hint: superposition. This goes under the name of Babinet’s
principle, by the way.
Consider the following setup

15
16 CHAPTER 4. INTERFERENCE

This gives us an expression for the wavelength


dh 1 1 i
λ= p −p .
m 1 + (s0 /L)2 1 + (sm /L)2
If you make α very small and use only low orders, then we can assume that sm /L << 1:
d(s2m − s20 )
λ≈ .
2mL2
As an example, when I did this experiment at home, I used the marks on the ruler that were d = 0.5
mm apart and the distance to the wall was L = 105 cm. I found s0 = 10.5 cm and s1 = 11.9 cm.
Assuming small angles, this gives
(5 × 10−4 m)[(11.9 cm)2 − (10.5 cm)2 ]
λ= ≈ 711 nm.
2 × 1 × (105 cm)2
That’s not bad! The wavelength should be around 635 nm. Before we rejoice, however, we should note
that a millimeter difference in any sm makes a huge difference in the final answer. For example, if I
change s1 to 11.8 cm, I get λ = 657 nm! So, unless I can measure sm and L with very high precision,
the uncertainties are likely to swamp the final measurement of λ anyway.

Note: In section, I claimed that if the laser light reflects off of a smooth metalic surface, then we only
get specular reflection. This is true only if the region on the surface that is illuminated is much wider
than the wavelength of the light. Well, our laser beam has a width of a few millimeters, which will
obviously do since its wavelength is on the order of 10−4 mm! As calculated above, the extra optical
path length travelled by one beam relative to another that hits the surface a distance x to the left of it
is x(cos α − cos β). So, the phase shift is φ = 2π λx (cos α − cos β). Let a be the width of the illuminated
region and let x run from −a/2 to a/2 with the “zero phase” corresponding to x = 0. The intensity is
proportional to
Z a/2 x
2 cos2 A
I∝ e2πi λ (cos α−cos β) dx ∝ ,

−a/2 A2
where A = π λa (cos α − cos β).
In the limit a/λ << 1, the intensity becomes a delta function:
a/λ→0
I −−−−→ δ(A).
Thus, the intensity vanishes everywhere except when A = 0, or when cos α = cos β, or α = β, which
is the condition for specular reflection!

4.2 Continuous Case


A single slit has length a and width b. Determine the interference pattern that it produces on a faraway
screen assuming monochromatic, coherent light is normally incident upon it.

SOLUTION:

Consider the following setup of coordinates:


4.2. CONTINUOUS CASE 17

We have separate but parallel coordinates on the slit screen (X and Y ) and the observation screen
(x and y), a large distance L away. I have chosen the very center of the slit to be the origin of
coordinates on the slit screen and the point on the observation screen directly in line with that to
be the corresponding origin on the observation screen. Let the length of the slit be along horizontal
direction, so that − a2 ≤ X ≤ a2 and − 2b ≤ Y ≤ 2b .
Consider a little rectangle located in the slit at the coordinates (X, Y ) and with length dX and
width (or height?) dY . The contribution, dE, of this little bit of area to the electric field at the point
on the observation screen with coordinates (x, y) should be proportional to its area, dX dY :
E dX dY i(kr−ωt)
dE = e ,
r
where E is some constant having to do with the source about which we need not care.
We calculate r using the Pythagorean theorem:
p h
x 2 y 2 2yY X 2
 i1/2
Y 2
− 2xX
  
r= L2 + (x − X)2 + (y − Y )2 = L 1 + L + L L2 − L2 + L + L
h i
1 x 2 1 y 2 yY
− xX
 
≈L 1+ 2 L + 2 L L2 − L2
xX yY
=R− L − L ,

where we have defined


x2 +y 2
R≡L+ 2L .

R is a constant as far as position on the slit screen is concerned (i.e. it does not change with X or
Y ). Therefore, it does not contribute to the diffraction pattern. We don’t care about the differences
in the path lengths, r, when it comes to the amplitude of the electric field contribution of the little bit
of area. This is because tiny differences in amplitude make tiny overall differences to the total electric
field. But, we must keep this tiny differences in the complex exponential because even tiny values of
X and Y can produce dramatic phase shifts. Thus,

dE ≈ E
R ei(kR−ωt) e−ikxX/L e−ikyY /L dX dY.
E i(kR−ωt)
Denote Re by A. We must integrate this:
Z Z a/2 Z b/2
E= dE = A e−ikxX/L dX e−ikyY /L dY
slit −a/2 −b/2
L −ikxX/L a/2 L −ikyY /L b/2
=A e e
−ikx −a/2 −iky

−b/2
2 h −ikax/2L ikax/2L ih −ikby/2L
AL e −e e − eikby/2L i
= 2
k xy −i −i
4AL2  kax   kby 
= 2 sin sin .
k xy 2L 2L
18 CHAPTER 4. INTERFERENCE

This gives us E as a function of the coordinate (x, y) on the observation screen. Let us take the limit
as (x, y) → (0, 0). For this, we can use the small angle approximation, sin θ ≈ θ:

4AL2 kax kby


E(0, 0) = lim · · = abA.
(x,y)→(0,0) k 2 xy 2L 2L

Therefore, we can eliminate A and relate the electric field at an arbitrary point (x, y) on the observation
screen to the electric field at the origin of the observation screen, (0, 0):

4L2 E(0, 0)  kax   kby 


E(x, y) = sin sin .
k 2 abxy 2L 2L
Define the following expressions,
ka πa kb πb
α≡ = , β≡ = .
2L λL 2L λL
Then, we can write

sin(αx) sin(βy)
E(x, y) = E(0, 0) = E(0, 0) sinc(αx) sinc(βy).
αx βy
Finally, the intensity is proportional to the square of this:

I(x, y) = I(0, 0) sinc2 (αx) sinc2 (βy).

You should recognize this as simply the product of the single slit diffraction patterns for slits of width
a in the x-direction and b in the y-direction.

We could use this result almost directly to calculate the diffraction pattern produced by a cross-shaped
aperture, which you considered in a previous problem set. The cross can be thought of as the sum of
two of our single slits considered above, one with its length along the x-direction and one along the
y-direction. However, we would be counting their intersection twice! So, we must subtract the result
of a square slit of sidelength b:
 
E(x, y) = E(0, 0) sinc(αx) sinc(βy) + sinc(βx) sinc(αy) − sinc(βx)sinc(βy) .

Again, the intensity is just proportional to the square of this, which doesn’t really simplify; there’s
no point in multiplying it all out. Below is a contourplot of the intensity as a function of (x, y). The
diagram on the left (right) is for a cross with arms whose length is three (nine) times their width.
Chapter 5
Polarization

5.1 Hecht P.8.32 p.381 (Augmented)


A beam of natural light is incident on an air-glass interface (nit = 1.5) at 40◦ . Compute the degree of
polarization of the reflected light.

SOLUTION:

From Hecht Eqns. 8.26 and 8.27 p.351, the parallel and perpendicular reflectances are

tan2 (θi − θt ) sin2 (θi − θt )


R|| = , R⊥ = .
tan2 (θi + θt ) sin2 (θi + θt )

In this case, θt = sin−1 ni


nt sin θi = 25.37◦ and so

R|| = 0.014, R⊥ = 0.077.

If the total incident intensity is I, then the incident intensity that is polarized in the parallel direction is
I/2, and the same for perpendicular: Ii|| = Ii⊥ = I/2. Then, the reflected intensities in the respective
polarizations is

R|| R⊥
Ir|| = R|| Ii|| = 2 I, Ir⊥ = R⊥ Ii⊥ = 2 I.

The difference between the two is the part that is completely polarized and the remainder is the part
that is completely unpolarized:
R − R  R + R  R − R 
⊥ || ⊥ || ⊥ ||
Ip = I, In = Ir,tot − Ip = I− I = R|| I.
2 2 2

Thus, the degree of polarization is

Ip Ip R⊥ − R||
V = = = = 0.687.
Ip + In Ir,tot R⊥ + R||

The reflected light is 68.7% polarized in the direction perpendicular to the plane of incidence.

19
20 CHAPTER 5. POLARIZATION

5.2 Hecht P.8.41 p.382 (Augmented)


Take two ideal Polaroids (the first with its axis horizontal and the second, vertical) and insert between
them a stack of 10 half-wave plates, the first with its fast axis rotated π/40 rad from the horizontal,
and each subsequent one rotated π/40 rad from the previous one. Determine the ratio of the emerging
to incident irradiance.

SOLUTION:

Method 1: If I0 is the incident intensity, then after the first polarizer, the intensity is I0 /2 and
polarized in the x̂ direction. Its angle relative to the optical axis of the first HWP is −π/40, which
becomes π/40 after the first HWP. Remember that a HWP turns a polarization angle α into −α (or
π − α or any whole number of π’s −α). But, π/40 relative to the optical axis of the first HWP is π/20
relative to the x-axis (i.e. the transmission axis of the first polarizer. That happens to be the angle
of the optical axis of the second HWP relative to the x-axis. Thus, after the first HWP, the light is
polarized along the optical axis of the second HWP. Therefore, the second HWP does not change its
polarization. The third HWP will rotate the polarization to point along the axis of the fourth HWP,
and so on. Therefore, as long as we have an even number of HWP’s, the final polarization will be along
the final HWP. In this case, that would be at an angle of π/4 relative to the x-axis. The HWP’s change
the polarization but not the intensity. Passing it through the final polarization cuts the intensity down
by a factor of cos2 (π/4) = 1/2. Thus, the final intensity is I0 /4.

Method 2: This is quite a bit more involved, but it’s also more algorithmic and would be superior
in dealing with, for example, the case when the HWP’s are not all evenly rotated. Let N = 10 be the
number of half-wave plates. First, we would like to relate the fast and slow axes of the nth half-wave
plate to the original axes of the first Polaroid. We’ll call the optical axis the fast axis. Consider the
diagram below.

It should be clear that


nπ nπ nπ nπ
   
f̂n = cos 4N x̂ + sin
ŷ, 4N ŝn = − sin 4N x̂ + cos 4N ŷ.


In the sequel, we will abbreviate cos 4N by cn and the sine by sn.
Denote the vector xx̂ + yŷ by xy and the vector fn f̂n + sn ŝn by fsnn n . Note that the subscript n
 

on the column vector reminds us that it is written in terms of the {f̂n , ŝn } basis and not the original
{x̂, ŷ} basis. We have
    
fn cn sn fn
= fn f̂n + sn ŝn = fn (cn x̂ + sn ŷ) + sn (−sn x̂ + cn ŷ) = .
sn n −sn cn sn

The inverse relation would read     


x cn −sn x
= .
y sn cn y n
Let’s unpack these equations. The first one says that if I know the components of a vector in the xy-
basis, namely fn for the x component and sn for the y component, then the corresponding components
cn sn

in the fn sn -basis is given by the rotation matrix −sn cn multiplying the vector in the xy-basis. The
5.2. HECHT P.8.41 P.382 (AUGMENTED) 21

second equation is just the inverse: if you know the components in the fn sn -basis, namely x for the fn
component and y for the sn component, then the equation will give you the corresponding components
in the xy-basis.
Let’s just check one example to make sure this makes sense. In the xy-basis, the horizontal vector
pointing to the right is 10 . In the fn sn -basis, just by looking at the diagram, you should see that it

cn

should be −sn n
. Well, this is indeed what we get using the transformation rule:
    
cn sn 1 cn
= .
−sn cn 0 −sn n

Now, the matrix representing a HWP with its optical axis along the x-direction is
 iπ   
e 0 −1 0
Φ≡ = .
0 1 0 1

It is important to note that this matrix is written with respect to the xy-basis. Also, we have made
the choice, as was done in the lab, of saying that the electric field along the optical axis is flipped while
the perpendicular component is unchanged.
What matrix would represent one of our HWP’s  whose optical axis is rotated by nπ/4N relative
to the positive x-axis? Well, let’s call it CA B . This means that if you pass light with polarization
D
given by fsnn (in the xy-basis!), then what comes out is


  
A B fn
.
C D sn

This is in the xy-basis. To convert it to the fn sn -basis, we multiply by the appropriate rotation:
   
cn sn A B fn
.
−sn cn C D sn

However, we could have written the initial vector in the fn sn -basis before it passed through the HWP.
Then, we are already in the fn sn -basis and we just have to write the matrix for the HWP in this basis,
which is just −10 01 , since the optical axis of the HWP points along f̂n . Therefore, the result above
should be the same as    
−1 0 cn sn fn
.
0 1 −sn cn sn
This implies that      
cn sn A B −1 0 cn sn
= .
−sn cn C D 0 1 −sn cn
Therefore, the matrix representing the rotated HWP in the xy-basis is related to its matrix in the
fn sn -basis by a similarity transformation, each step of which is just a rotation:
     
A B cn −sn −1 0 cn sn
= ≡ Rn ΦRn−1 .
C D sn cn 0 1 −sn cn
−1
Note the relation Rn+1 Rn = R1−1 , which just says that a forward rotation by nπ/4N followed by a
backward rotation by (n + 1)π/4N is equivalent to a backward rotated by −π/4N . Hopefully that’s
obvious.
The matrix representing the N successive HWP’s is
−1 −1 −1 −1 −1 −1
(RN ΦRN )(RN −1 ΦRN −1 ) · · · (R1 ΦR1 ) = RN Φ(RN RN −1 )Φ · · · (R2 R1 )ΦR1
= RN ΦR1−1 Φ · · · R1−1 ΦR1−1
= RN (ΦR1−1 )N .
22 CHAPTER 5. POLARIZATION

Now, we calculate:     
−1 0 c s −c −s
ΦR1−1 = = .
0 1 −s c s c
Conveniently, we find

cos2 + sin2
   
cos sin − sin cos 1 0
(ΦR1−1 )2 = = .
sin cos − cos sin sin2 + cos2 0 1

Therefore, as long as N is even, the matrix for the N HWP’s is


   
−1 N cos(π/4) − sin(π/4) 1 1 −1
RN (ΦR1 ) = RN = =√ ,
sin(π/4) cos(π/4) 2 1 1
1

Now, the light incident on this collection of HWP’s has a polarization vector 0 (i.e. horizontal
direction). After the HWP’s, it has a polarization vector given by
    
1 1 −1 1 1 1
√ =√ ,
2 1 1 0 2 1

which is indeed in the direction π/4 = 45◦ from the positive x-axis. As in method 1, the final polarizer
lets cos2 (π/4) = 1/2 of this intensity through. So the first polarizer cuts the intensity down to I0 /2,
the ten HWP’s rotate the polarization by 45◦ , and the final polarizer cuts the intensity down to I0 /4.
Chapter 6
Midterm Review

6.1 Michelson Interferometer


Suppose the ether theory were true (i.e. that the speed of light is only c in the rest frame of the
ether) and that we lived in a Galilean-relativistic world, rather than an Einsteinian-relativistic one.
Suppose that the Michelson interferometer were moving relative to the rest frame of the ether with a
velocity v = (vx , vy , vz ), where the plane of the interferometer is the xy-plane with one arm along the
x direction and one along the y direction. Let ` be the length of both arms. We do not know what
v is, but suppose we assume that each of its components is characterized by a Gaussian probability
distribution centered at zero with a width, v0 << c. This means that we are assuming that it is very
unlikely for us to be moving relative to the ether with a speed anywhere near c or greater; in fact,
our most likely speed is 0. We are also assuming that it is equally likely for us to be moving in any
direction (i.e. the distribution is isotropic). So, let the probability distribution for vx be
2
P (vx ) = √1
π v0
e−(vx /v0 ) ,

with the exact same distribution


R ∞ for vy and vz . Note that we have already ensured that this probability
distribution is normalized: −∞ P (vx ) dvx = 1.
Let N be the number of fringes through which the interferometer cycles in the course of turning
clockwise by 90◦ . Determine the root-mean-squared value of N as a function of `, v0 , and λ, which is
the wavelength of the laser used in the experiment.

You may need to use the following integrals:


Z ∞ Z ∞ √ ∞ √

Z
2 2 π 2 3 π
e−x dx = π, x2 e−x dx = , x4 e−x dx = .
−∞ −∞ 2 −∞ 4

SOLUTION:

Relative to the rest frame of the interferometer, the ether is moving with a velocity, −v = −(vx , vy , vz ).
This is what French calls the “ether wind”. Let arm 1 be the one along the x direction and arm 2 be
the one along the y direction. Let u± ±
1 = (u1 , 0, 0) be the velocity of the light beam in the rest frame
of the interferometer along arm 1 with the + sign corresponding to going “up” the arm (i.e. in the +x
direction) and the − sign corresponding to going “down” the arm (i.e. in the −x direction). Similarly,
define u± ± ± ± ± ±
2 = (0, u2 , 0) for the velocity up and down arm 2. Let v1 = c(ξ1 , ψ1 , ζ1 ), and similarly for
±
v2 , be the corresponding velocities in the ether rest frame. Galilean relativity relates these via
u± ±
1 = v1 − v,

23
24 CHAPTER 6. MIDTERM REVIEW

and similarly for the velocities along arm 2. This equation reads

(u± ± ± ±
1 , 0, 0) = (cξ1 − vx , cψ1 − vy , cζ1 − vz ),

which implies that


vy
ψ1± = c , ζ1± = vz
c .

The speed of the beam in the rest frame of the ether is supposed to be c, which gives the relation
q q q
± ± 2 ± 2 ± 2 ± ± 2 ± 2 v 2 +v 2
c = |v1 | = c (ξ1 ) + (ψ1 ) + (ζ1 ) =⇒ ξ1 = ± 1 − (ψ1 ) − (ζ1 ) = ± 1 − yc2 z .

We have chosen signs appropriately, so that ξ1+ is positive (i.e. up the arm is the positive direction
and down the arm, the negative). Therefore,
q
v 2 +v 2

1 = cξ1
±
− v x = ±c 1 − yc2 z − vx .

The exact same analysis yields


q
2
vx +vz2

2 = cψ2
±
− v y = ±c 1− c2 − vy .

The time that it takes for the beam to go up and down arm 1 is
` ` ` `
t1 = + + − =
q + q .
|u1 | |u1 | 2
v +v 2
v +v 2
2
c 1 − yc2 z − vx c 1 − yc2 z + vx

Define βx ≡ vx /c, and similarly for the other components, define v 2 = vx2 + vy2 + vz2 and β ≡ v/c.
p
Finally, define γ = 1/ 1 − β 2 , as usual. Then, after some algebra, we can simplifty t1 to read
 q 
t1 = γ 2 1 − βy2 − βz2 2`
c .

Let us check the limit as βx,y,z → 0 (i.e. as we approach the ether rest frame). We should expect to
get t1 = 2`
c , which is indeed what we get if we plug βx,y,z = 0 into the formula above.
The exact same analysis for the second arm yields
 p 
t2 = γ 2 1 − βx2 − βz2 2`c .

The difference between these times is


q p 
∆ ≡ t1 − t2 = γ 2 1 − βy2 − βz2 − 1 − βx2 − βz2 2`
c .

Let us now make the small βx,y,z approximation, since we know that the probability for any of these
to be anywhere near or greater than 1 is very small. That is, β0 ≡ v0 /c << 1, and the Gaussian
probability distribution vanishes very quickly above this value. We find
h i 2`
∆ ≈ (1 + β 2 ) 1 1 2 1 
 − 2 βy − 
2
2 βz − 1
1 2 1 
 − 2 βx − 
2
2 βz
2 2 `
c ≈ (βx − βy ) c .

1
Note that we first approximated γ 2 = 1−β 2
2 ≈ 1 + β , and then dropped the β
2
anyway because it is
multiplied by terms quadratic in βx and βy .
Rotating the interferometer by 90◦ clockwise is equivalent to βx → −βy and βy → βx . The time
difference between the two arms in this case is

∆0 ≡ t01 − t02 ≈ (βy2 − βx2 ) c` .


6.2. PION PHOTOPRODUCTION 25

The phase of a beam increases by 2πct


λ after travelling for time t. Each 2π worth of phase is one fringe.
Thus, the number of fringes through which the inteference pattern cycles during a 90◦ rotation of the
interferometer is
N = λc (∆ − ∆0 ) ≈ (βx2 − βy2 ) 2`
λ.
It shouldn’t surprise you that if you take the average of N , you will get 0, because the effect of a
velocity v is cancelled by the effect of the exact opposite velocity, since the probability distribution is
isotropic. The RMS value is
ZZZ ∞
2
Nrms = N =2 dβx dβy dβz P (βx )P (βy )P (βz ) N 2
−∞
 2` 2  1 3 Z Z ∞ Z ∞
−(βx /β0 )2 −(βy /β0 )2 2
= √ dβx dβy e e 2 2 2
(βx − βy ) dβz e−(βz /β0 )
λ π β0 −∞ −∞
 2` 2  1 Z ∞ 2
Z ∞
2

= √ β 4 e−(βx /β0 ) dβx + βy4 e−(βy /β0 ) dβy
λ π β0 −∞ x −∞
Z ∞ Z ∞ 
2 2 −(βx /β0 )2 2 −(βy /β0 )2
− β e dβx βy e dβy .
πβ02 −∞ x −∞

Define A ≡ βx /β0 and B ≡ βy /β0 , then


 2` 2  β 5 Z ∞ Z ∞ 
4 −A2 2
2
Nrms = √ 0
A e dA + B 4 e−B dB
λ π β0 −∞ −∞
2β06 ∞ 2 −A2
Z Z ∞ 
2 −B 2
− A e dA B e dB
πβ02 −∞ −∞
 2` 2  β 4  3√ √ 
2β 4
√ √ 
= √0 π + 3 π
− 0 · ·
π π
λ   π 4 4 π
 2 2
 2` 2
= β04 .
λ
Therefore, the RMS is
2`β02 2(v0 /c)2
Nrms = = .
λ λ/`
Note: This is the same result as French Eqn. (2-14) p.56 if we suppose that the interferometer is
moving relative to the ether only in the x direction with a speed v = v0 .

6.2 Pion Photoproduction


Consider the pion photoproduction reaction γ + p → p + π 0 , where the rest energy is 938 MeV/c2 for
the proton and 135 MeV/c2 for the neutral pion.
(a) If the initial proton is at rest in the laboratory, find the laboratory threshold gamma-ray energy
for this reaction to go.
(b) Consider a head-on collision between a proton and a photon from the isotropic 3K cosmic blackbody
radiation, which has an average photon energy of 10−3 eV. Find the minimum proton energy that
will allow this pion photoproduction reaction to go.
(c) Speculate briefly on the implications of your result for part (b) for the energy spectrum of cosmic-
ray protons.

SOLUTION:
26 CHAPTER 6. MIDTERM REVIEW

E 2
(a) The initial value of E 2 − (pc)2 in the lab frame is (Eγ + mp c2 )2 − cγ c = 2Eγ mp c2 + (mp c2 )2 . At
threshold, in the center of mass frame, the resulting proton and pion are at rest. Thus, the final
value of E 2 − (pc)2 is (mp + mπ )2 c4 = (mp c2 )2 + 2mp mπ c4 + (mπ c2 )2 . Since these are invariant,
set these two equal and solve for Eγ :

Eγ = mπ c2 1 + 2m mπ

p
= 145 MeV.

(b) Take the energy of the photon in the lab frame to be Eγ = 10−3 eV. Again, equating the initial
and final values of E 2 − (pc)2 yields
Eγ 2
(γmp c2 + Eγ )2 − c − γmp βc c2 = (mp + mπ )2 c4 .

Solving for β gives


s
mπ c2
 
1+β mπ
γ(1 + β) = = 1+ = 1.447 × 1011 .
1−β Eγ 2mp

Since γ(1 + β) is so large, β must be very close to 1. Let us solve approximately for γ by setting
β = 1 in the equation above. This gives γ = 7.235 × 1010 and so the proton energy is

Ep = γmp c2 = 7.235 × 1010 × 0.938 MeV = 6.787 × 1010 GeV .

(c) This implies that the energy spectrum of cosmic ray protons with E > 6.79 × 1010 GeV will be
depleted to some degree due to interaction with the cosmic blackbody radiation.

6.3 Microscopes and Embryos


1
The egg of a very tiny creature has a radius of R = 10 mm and an index of refraction negg = 32 . A tiny
embryo sits in its center. This egg sits on the surface of a microscope slide submerged in a layer of
3
water 10 mm thick. Above this sits a compound microscope made with identical converging objective
and eyepiece thin lenses with focal length f = 2 cm. The objective lens sits a distance of 2.65 cm
above the surface of the water and the distance between the objective and eyepiece is 10 cm. Locate
the final image of the embryo and calculate the angular magnification of the microscope by itself. Use
10 cm as your near point.

SOLUTION:

At the egg-water interface, we have


negg nwater nwater − negg
+ = .
R si1 −R
6.4. THREE-SLIT INTERFERENCE 27

Plugging in the numbers gives


si1 = −R = −0.1 mm,
which places it exactly where the embryo started out in the first place! You can show that this is true
no matter the medium in which the egg is immersed nor the index of refraction of the egg itself!
At the water-air interface, we have
nwater nair nair − nwater
+ = = 0.
2R si2 ∞

Plugging in the numbers gives


3R
si2 = − = −0.15 mm.
2
Note that this provides a different method for calculating apparent depth.
The distance between this intermediate image and the objective lens is so3 = (2.65 + 0.015) cm =
2.665 cm. The lens makes an image at

1 1 1
= − =⇒ si3 = 8.01504 cm.
si3 f so3

The distance between this intermediate image and the eyepiece lens is so4 = (10 − 8.01504) cm =
1.98496 cm. The lens makes an image at

1 1 1
= − =⇒ si4 = −264 cm.
si4 f so4

That is, the final image is produced 2.64 m behind the eyepiece (i.e. it is a virtual image).
The linear magnification of the objective by itself is
si3
mob = − ≈ −3.
so3

The angular magnification of the eyepiece by itself is

dnp dnp
Mep = ≈ = 5.
so4 f

Therefore, the overall angular magnification of the microscope is

Mm = mob Mep = −15.

6.4 Three-Slit Interference


[Hecht P.9.18 ] Imagine that you have an opaque screen with three horizontal very narrow parallel slits
in it. The second slit is a center-to-center ditance a beneath the first, and the third is a distance 5a/2
beneath the first.

(a) Write a complex exponential expression in terms of δ = ka sin θ for the amplitude of the electric
field at some point P at an elevation θ on a distant screen.

(b) Prove that I(θ) = I(0) 2I(0)


cos δ + cos 3δ 5δ

3 + 9 2 + cos 2 .

SOLUTION:
28 CHAPTER 6. MIDTERM REVIEW

(a) Let y be the distance of P from the point on the detection screen directly facing the midpoint
between the top and bottom slits. We take the latter point to be the point from which θ is
measured.
Let ri be the straight-line distance from the center of slit i to the point P . Let L be the distance
between the slit and detection screens. Then, making the approximation a/L << 1,

5a 2 1/2 1 y−5a/4 2 y2 5ay


r1 = L2 + y −
     
4 ≈L 1+ 2 L ≈L+ 2L − 4L .

2
y
Define R = L+ 2L , which is actually common to all ri and thus does not contribute to interference.
Note that sin θ ≈ θ ≈ tan θ = y/L. Thus, we may write ay/L ≈ a sin θ = δ/k. The same
calculation can be done for r2 and r3 . In terms of δ,
5δ δ 5δ
r1 = R − 4k , r2 = R − 4k , r3 = R + 4k .

The electric field at P due to ray 1 is rE1 ei(kr1 −ωt) ≈ R E i(kR−ωt) −i(5δ/4)
e e . Note that we have
neglected the small difference in r1 relative to R that is involved in the amplitude, but NOT when
E i(kR−ωt)
it is involved in the phase (i.e. in the complex exponential). Define A ≡ R e , which is
common to the electric field contributions of the beams. The total electric field is

E(θ) = A e−i5δ/4 + e−iδ/4 + ei5δ/4 .




Plugging in θ = 0 gives δ = 0 and E(0) = 3A. Therefore, we may write

E(θ) = 13 E(0) e−i5δ/4 + e−iδ/4 + ei5δ/4 .




(b) Since the irradiance is proportional to |E|2 , we have

I(θ) = 19 I(0) e−i5δ/4 + e−iδ/4 + ei5δ/4 ei5δ/4 + eiδ/4 + e−i5δ/4


 

= 91 I(0) 1 + e−iδ + e−i5δ/2 + eiδ + 1 + e−i3δ/2 + ei5δ/2 + ei3δ/2 + 1




= 91 I(0) 3 + 2 cos δ + 2 cos 3δ 5δ



2 + 2 cos 2

= I(0) 2I(0)
cos δ + cos 3δ 5δ

3 + 9 2 + cos 2 .

6.5 Polarizers Away!


Two identical HN-46 polarizers, at rest relative to each other, with area 10 cm2 (shaped as disks,
perhaps) and mass 1 g, are parallel to each other, separated by 1 cm and floating around in space far
away from any gravitational fields. Their polarization axes make an angle 30◦ relative to each other.
A beam of natural light, with cross-sectional area larger than the polarizers’ and with irradiance 488.6
W/cm2 , is normally incident on the first polarizer with the other polarizer directly behind. How much
time will it take for the two polarizers to make contact with each other?

SOLUTION:

Let the polarizers be HN-100x, so x = 0.46 in this case. Let A = 10 cm2 be the area of the polarizers,
m = 1 g their mass, d = 1 cm their separation, θ = 30◦ their relative angle and I = 488.6 W/cm2 the
irradiance of the light beam. Let us trace what happens to the beam. 100% of the light is incident
on the first polarizer. Of this, 8% is reflected away, 46% is absorbed and 46% is transmitted over
to the second polarizer. At this point, 8% of the remaining 46% is reflected back towards the first
polarizer, and, were θ to equal 0 (i.e. the polarization axes were to be parallel), the remaining 92% of
the 46% would all be transmitted. But, since θ 6= 0, instead, sin2 θ = 41 is absorbed and cos2 θ = 34 is
6.5. POLARIZERS AWAY! 29

transmitted. The light that was reflected off of the second polarizer can partially reflect and transmit
through the first polarizer, but it won’t be absorbed because it is now polarized along the direction of
the first polarizer. We can draw this schematically as follows

Each successive absorption at polarizer 2 gets a factor of 2x sin2 θ, which has been abbreviated to 2xs2
in the diagram, each successive transmission at polarizer 2 gets a factor of 2x cos2 θ, each successive
reflection (on either polarizer) picks up a factor of y ≡ 1 − 2x, and each successive transmission back
through polarizer 1 gets a factor of 2x.
However, recall that reflections count twice as much as absorptions as far as radiation pressure is
concerned. For the first polarizer, the initial reflection and absorption count as pressure to the right
while the successive reflected waves going from 1 to 2 contribute leftward pressure. The total pressure
to the right in units of I/c is
2xy 2
P1 = 2y + x − 2xy 2 − 2xy 4 − · · · = x + 2y − ,
1 − y2
1
where we have used the geometric series formula 1 + y 2 + y 4 + · · · = 1−y 2 , for |y| < 1. Plugging in
y = 1 − 2x and simplifying gives
3 − 6x + 2x2
P1 = = 0.614.
2(1 − x)
As is to be expected, this is only slightly smaller than 0.62 = 0.46 + 2(0.08), which is the result if we
neglect all subsequent reflections between the two polarizers.
The pressure on polarizer 2 just due to absorption is
x sin2 θ
P2abs = (2x2 sin2 θ)(1 + y 2 + y 4 + · · · ) = = 0.1065.
2(1 − x)
The pressure on polarizer 2 just due to reflection is
1 − 2x
P2ref = 2xy(1 + y 2 + y 4 + · · · ) = = 0.0741.
2(1 − x)
Therefore, the total pressure on polarizer 2 is
P2 = 0.181.
First, let’s get some units right:
W J/s Nm N dyne
2
= 2
= 2
= 102 = 107 ,
cm cm cm s cm · s cm · s
where dyne = g · cm/s2 .
The force on polarizer j is Fj = Pj Ic A:

4.886 × 109 dyne/cm · s


F1 = 0.614 × × 10 cm2 = 1 dyne,
3 × 1010 cm/s
0.181
F2 = F1 = 0.295 dyne.
0.614
30 CHAPTER 6. MIDTERM REVIEW

The accelerations, aj , are the same, with dynes replaced by cm/s2 , since m = 1 g.
Set z = 0 and z = d to be the initial positions of the polarizers. Their positions as functions of
time are simply given by kinematics:

z1 (t) = 12 a1 t2 , z2 (t) = d + 21 a2 t2 .

The time of coincidence (z1 = z2 ) is


r s
2d 2 × 1 cm
t= = 2 = 1.68 sec.
a1 − a2 (1 − 0.295) cm/s
Chapter 7
Blackbody Radiation

One lesson we learn from the photoelectric effect is that light may be thought of as being built out
of particles called photons, even though it behaves like a wave in most familiar situations. Somehow,
very many photons conspire to produce wave-like behavior. This concept of photons is what starts
us down the road towards blackbody radiation, although historically the ideas of Planck that follow
preceded, and in fact inspired, Einstein’s explanation of the photoelectric effect.
One thoroughly embarrassing problem that remained before Planck came on the scene is called the
ultraviolet catastrophe. Consider a thermally insulated box of sidelength L containing electromagnetic
fields. The boundary conditions at the inner edges of the box imply that the electric and magnetic
fields have to vanish there. Thus, the fields must form standing waves. In one dimension, these look like
the fundamental and harmonic vibrations on a string that is fixed at both ends. In three dimensions,
we just multiply three of these standing waves (they can all have different wavelengths), one in the x
direction, one in the y direction, and one in the z direction. That gives us three degrees of freedom,
namely the wavelengths of the standing wave in each direction. We can actually denote the different
modes by integers: the fundamental mode fits half of its wavelength within L, the first harmonic fits a
full wavelength within L, the second harmonic fits 3/2 of a wavelength in L, etc. Thus, the nth mode
fits n/2 of a wavelength in L, or, in other words λn = 2L/n is the wavelength of the nth mode. Note
that the fundamental mode corresponds to n = 1, and so on. Thus, a single standing wave mode in
the cube can be distinguished by three positive integers (nx , ny , nz ), one for each direction. However,
since this is an electromagnetic wave, each standing wave component (one in each direction) has two
orthogonal choices for the direction of the electric field (i.e. the polarization). This gives us a total of
six degrees of freedom for each mode (nx , ny , nz ). The equipartition theorem states that each degree
of freedom will contribute an average of kT /2 worth of energy, where k is Boltzmann’s constant and T
is the temperature of the system. That means that each standing wave mode contributes 3kT worth of
energy on average. But, there are infinitely many possible modes: the three integers can get arbitrarily
large, or, in other words, the wavelengths can get arbitrarily small. That would make the total energy
infinite! Schroeder describes just how embarrassing this conclusion is: if it were correct, you would
expect to be blasted with an infinite amount of radiation every time you open the oven door to check
the cookies!
The classical assumption is that each mode, denoted by the three integers, can have any non-
negative energy, E. From H7B, we know that the probability for a mode to have energy E is propor-
tional to the Boltzmann factor: P (E) ∝ e−E/kT . Calculating the average energy per mode as we did
for an ideal gas in H7B for such a continuous spectrum produces the equipartition theorem.
Planck’s neat idea was that electromagnetic energy is not continuously distributed, but is quantized
in integer units of hν, where ν is the frequency of radiation and h is Planck’s constant. He proposed
that light was absorbed and emitted by matter in quanta called photons. So, a single mode with
frequency ν can have an energy of 0, or hν, or 2hν, etc. But it cannot have an energy between these

31
32 CHAPTER 7. BLACKBODY RADIATION

values, like hν/2, since that would correspond to half a photon!

7.1 Planck Distribution


I know of two main paths to Planck’s solution for the UV catastrophe. One makes use of the classical
idea of enumerating modes of electromagnetic fields within a cube and then jumping to the quantum
concept of distributing photons among these modes. The second is to adopt a quantum mechanical
picture right from the start and enumerate the possible states of photons within some arbitrary enclosed
volume.
There are pros and cons to both methods. The first method makes it easy to compare Planck’s fix
to the fully classical approach and clearly highlights exactly where the classical picture breaks down.
On the other hand, I have always been dissatisfied with its mixture of classical and quantum concepts:
on the one hand describing modes of classical electromagnetic fields and on the other hand populating
these modes with quanta (photons) while never really clarifying how the former arise from the latter.
The second method starts from the photon immediately, without recourse to classical fields. It also
possesses the virtue of not requiring a specific shape like a cube for the enclosure. On the other hand,
fully describing the photon states will be a bit more challenging. Since the first method is described
in Serway and most other statistical mechanics textbooks, I have elected to take the second path.
First, we must determine which variables we must consider in order to fully describe the state of
photons inside some arbitrary enclosure. There are four variables: (1) the number of photons, (2)
their helicities, (3) their positions, and (4) their momenta. Since we are in three dimensions, the last
two actually constitute six variables altogether: x, y, z, px , py , and pz . The combination of these
last six variables taken as an abstract six-dimensional “space” is called “phase space”. By the way,
helicity describes the intrinsic angular momentum of photons and can come in two flavors: left and
right handed.
Suppose we fix the helicity, ε, the position, x, and the momentum, p, of a one-photon state. When
the system sits at temperature T , what is the average number of photons in this particular state? That
is, how heavily occupied is this particular one-photon state (OPS)? Hopefully, you remember from 7B
that the probability, Ps , of a state, s, with energy, Es , to be occupied is proportional to the Boltzmann
factor: Ps ∝ e−Es /kT . We must require that the sum of the probabilities of occupation over all of the
states be 1. Thus,
1 X
Ps = e−Es /kT where Z= e−Es /kT , (7.1)
Z s

is called the partition function.


In the case at hand, the energy of the OPS is pc, where p = |p|. The OPS may be occupied by any
number, n, of photons, with total energy npc. Thus, the partition function for a particular OPS with
helicity, position and momentum, (ε, x, p) is

X
Z(ε, x, p) = e−npc/kT (7.2a)
n=0
1
= . (7.2b)
1 − e−pc/kT
The average number of photons in this OPS is
∞ ∞
X photon #probability X ne−npc/kT 1 X −npc/kT
hnp i = = = ne , (7.3)
states
in state of state n=0
Zp Zp n=0

where, I will refrain from writing (ε, x, p) over and over again. It happens that Z and n actually only
depend on p, so we will write Zp and hnp i for brevity.
7.2. PLANCK SPECTRUM 33

Now, consider differentiating Eqn. (7.2a) with respect to temperature, T :


∞  ∞
dZp X npc  −npc/kT pc X −npc/kT
= e = ne . (7.4)
dT n=0
kT 2 kT 2 n=0

On the other hand, differentiating Eqn. (7.2b) yields

dZp 1 pc pc 2 −pc/kT
=− (−e−pc/kT ) 2 = Z e . (7.5)
dT (1 − e−pc/kT )2 kT kT 2 p

Setting Eqn. (7.4) and Eqn. (7.5) equal yields



X
ne−npc/kT = Zp2 e−pc/kT . (7.6)
n=0

Plugging this into Eqn. (7.3) yields

1 2 −pc/kT e−pc/kT 1
hnp i = Zp e = −pc/kT
= pc/kT . (7.7)
Zp 1−e e −1

This is called the Planck distribution. It measures how many photons will occupy a particular OPS
with momentum magnitude p when the system sits at temperature T .

7.2 Planck Spectrum


We would like to calculate the average electromagnetic energy (actually, energy density) inside the
enclosure at temperature T . To do so, we will multiply the average number, hnp i, of photons in the
specific OPS, (ε, x, p), with the energy, pc, of this OPS, and then “sum” over all possible OPS’s.
Conceptually, this is easy enough. But, how exactly do we enumerate the possible OPS’s? Helicity
is easy: left and right handedness does not affect energy, and so this just gives a factor of 2 to the
enumeration of OPS’s. What about phase space? Well, these are continuous variables taking any
possible value. Aren’t the states infinitely dense? The answer is no and the reason is the uncertainty
principle.
The uncertainty principle says that there is an intrinsic lower limit to how well one can determine or
define the position and momentum of a particle. This means that points in phase space that are so close
together as to be within the uncertainty principle of each other must be considered as representing the
exact same state. In other words, there is a minimum volume of phase space and all points within this
volume of some phase space point are considered to be indistinguishable from each other and represent
one and the same state. It turns out that the correct quantization is that there is one quantum state
for every h3 worth of phase space volume. Therefore, to “sum” over all states in phase space simply
means to integrate over phase space assuming that there is only one state per h3 phase space volume.
Thus, we can now calculate the average energy:
X
U ≡ hEi = hnp i pc
OPS
X 1 Z Z
pc
= d x d3 p pc/kT
3
h3 e −1
hel.
Z ∞
1 pc
= 2 × 3 × V × 4π dp p2 pc/kT . (7.8)
h 0 e −1
Since the integrand is independent of helicity and spatial coordinates, the helicity sum gives the factor
of 2 and the spatial integral gives the factor of the volume, V , of the enclosure. Also, since the integrand
34 CHAPTER 7. BLACKBODY RADIATION

is independent of the angular variables (i.e. direction) of the momentum, that integral just gives a
factor of 4π. Note that we converted the momentum integral into spherical momentum coordinates.
We can divide by the volume to get the energy density. We would also like to change the integration
variable to the wavelength of the OPS, rather than the momentum, where the two are related via de
Broglie’s equation p = h/λ. When p = 0, we have λ = ∞, and vice versa. Thus, the integral over λ
goes from ∞ to 0. Also, dp = − λh2 dλ. Thus,
Z 0  h 3 Z ∞
U 8πc h 1 8πhc
= 3 − 2
dλ =  dλ. (7.9)
V h ∞ λ λ ehc/λkT − 1 0 λ5 (ehc/λkT − 1

What if we punch a small hole through the wall of the enclosure? Since there is electromagnetic energy
density inside, it will escape. The energy density multiplied by the rate at which the energy is moving
(speed of light), should give the intensity of the radiation. However, most of the photons are not
heading towards the hole and thus do not escape. It turns out that the intensity radiated through the
hold is only one-fourth of the energy density multiplied by the speed of light:

2πhc2
Z
1U
I= c= dλ. (7.10)
4V 0 λ5 (ehc/λkT − 1)

I believe that the factor of one-fourth is described in your textbook. This immediately gives us the
Planck spectrum:
dI 2πhc2
= 5 hc/λkT . (7.11)
dλ λ (e − 1)

Setting d2 I/dλ2 = 0, we find that dI/dλ peaks at

hc
λpeak ≈ . (7.12)
4.97 kT

This is Wien’s displacement law. This is essentially how a radiation thermometer works. You point
this machine at something and it tells you the temperature of that object without direct contact, which
is extremely important for things that are far away (like the Sun) and/or things that are too hot or
cold for contact thermometers to work properly. The machine analyzes the spectrum of radiation that
it receives, finds the wavelength at which the spectrum peaks and uses Wien’s displacement law to
determine the temperature, T .
Let us go back to Eqn. (7.8) for U/V and change variables to u = pc/kT :

8π(kT )4 u3 du
Z
U
= . (7.13)
V (hc)3 0 eu − 1

It is possible to calculate the integral above. This is done at the end of this chapter. The result is
π 4 /15. Thus,
U 8π 5 (kT )4
= . (7.14)
V 15(hc)3

Therefore, the total irradiance emitted by a blackbody at temperature T is

cU 2π 5 k 4 T 4 2π 5 k 4
I= = = σT 4 where σ= . (7.15)
4V 15h3 c2 15h3 c2

This is Stefan’s law and the constant σ is the Stefan-Boltzmann constant.


7.3. STEFAN-BOLTZMANN INTEGRAL 35

7.3 Stefan-Boltzmann Integral


We would like to calculate the integral

u3 du
Z
J≡ .
0 eu − 1

First rewrite the integrand by multiplying above and below by e−u . Then, for u > 0, we have e−u < 1
and so we may use the appropriate geometric series as seen below:
∞ ∞
u3 u3 e−u 3 −u
X
−nu
X
= = u e e = u3
e−nu .
eu − 1 1 − e−u n=0 n=1

We would like to integrate each term separately. Before doing so, a simple calculation gives
Z ∞
1
e−nu du = .
0 n
Then, taking three derivatives with respect to n pulls down three factors of −u, and thus
Z ∞ Z ∞
d3 −nu
− 3 e du = u3 e−nu du.
dn 0 0

On the other hand,



d3 d3 1 d2 1
Z
d 2 6
− e−nu du = − = =− = 4.
dn3 0
3
dn n dn2 n2 dn n3 n
Therefore, the integral of each term in the expansion is
Z ∞
6
u3 e−nu du = 4 ,
0 n
and the total integral is
∞ Z ∞ ∞
X X 6
J= u3 e−nu du = ≡ 6 ζ(4), (7.16)
n=1 0 n=1
n4
P∞
where ζ(s) ≡ n=1 n1s is the Riemann zeta function.
Hecht Eqn. 7.49 p.306 gives the Fourier series for the square wave:
(
1 for mλ < x < (m + 12 )λ, 4 X 1  2πnx 
f (x) = 1
= sin .
−1 for (m + 2 )λ < x < mλ, π n λ
odd n

where m can take on any integer value. Note we have used the relation k = 2π/λ to change Hecht
Eqn. 7.49 to the above form.
In a moment, we will be integrating both expressions for f (x) multiple times. We just need to be
careful about integration constants. We want to set the integration constants for the integral of each
term in the Fourier series to zero. This simplifies things because the integral itself remains periodic
with the same period and remains centered around the x-axis. This uniquely fixes the constants of
integration for the first definition of f (x).
For example, in the region −λ/2 < x < 0, we have f (x) = −1. You might therefore think that
the integral is simply F 1 (x) = −x in this region. Similarly, you might think that F 1 (x) = x in the
region 0 < x < λ/2. This completes one period and the whole thing is repeated over and over again.
But that would make F 1 (x) always ≥ 0. Namely, it would not be centered around the x-axis and the
second integral, F 2 (x), would deviate further and further away from the x-axis as |x| → ∞. Therefore,
36 CHAPTER 7. BLACKBODY RADIATION

to make F 1 (x) centered around the x-axis, we must set F 1 (x) = −x − λ4 in the region −λ/2 < x < 0
and F 1 (x) = x − λ4 in the region 0 < x < λ/2, with all other regions just translated copies of these.
The second integral is a littler easier because F 2 (x) = 21 x2 − λ4 x + C in the region 0 < x < λ/2,
where C is some constant, but we also need F 2 (0) = 0, which implies C = 0. Hence, F 2 (x) = 21 x(x− λ2 )
in this region.
Perhaps the pattern is clear now: for odd s, we require F s (0) = −F s (λ/2), while for even s, we
3
require F s (0) = F s (λ/2) = 0. The third integral must therefore be F 3 (x) = 16 x3 − λ8 x2 + 2λ6 ·3 in the
region 0 < x < λ/2. Setting this equal to the term-by-term integral of the Fourier series yields

1 3 λ 2 λ3 4 X 1  λ 3  2πnx  λ3 X 1  2πnx 
x − x + 6 = F 3 (x) = cos = 4 4
cos .
6 8 2 ·3 π n 2πn λ 2π n λ
odd n odd n

Evaluating both sides at x = 0 yields

λ3 λ3 X 1 X 1 π4
= =⇒ = .
26 · 3 2π 4 n4 n4 25 · 3
odd n odd n

Now, we can separate the sum in the zeta function into one over odds and evens. The sum over the
evens is
X 1 1 1 1 1 1h1 1 1 1 i 1
4
= 4
+ 4
+ 4
+ 4
+ · · · = 4 14
+ 4
+ 4
+ 4
+ · · · = 4 ζ(4).
even n
n 2 4 6 8 2 2 3 4 2

Hence,
X 1 X 1 π4 1
ζ(4) = + = + ζ(4).
n4 even n n4 25 · 3 24
odd n

Solving for ζ(4) yields


π4 π4
ζ(4) = 2
= .
2·3 ·5 90
Plugging this into Eqn. (7.16) gives us the desired integral,
Z ∞ 3
u du π4
J≡ u
= 6 ζ(4) = . (7.17)
0 e −1 15
Chapter 8
Quantum Operator Expectation Values

8.1 Infinite Square Well


Calculate the expectation values of x, x2 , p and p2 in the nth bound state of the infinite square well
of width L. Statistically speaking, the uncertainty in position is its standard deviation, defined by
rD
2 E
∆x = x − hxi ,

and similarly for momentum. Calculate ∆x and ∆p and show that they satisfy the uncertainty rela-
tion. How does this alter your estimate for the force exerted on the confining walls by a particle in
this bound state from problem set 8 question 2?

SOLUTION:

Let |ni denote the appropriate abstract state in the abstract Hilbert space for this potential. Let |xi
represent the eigenstates of the position operator, which are orthogonal and complete. In other words,
Z ∞
hy|xi = δ(x − y), |xi hx| = 1,
−∞

where δ(x − y) is the Dirac delta function and 1 is the identity operator. The position and momentum
operators act on these states via


x̂ |xi = |xi x, p̂ |xi = −i~ |xi ∂x .

Notice that |xi can be moved to the left of the derivative because the state itself does not change as a
function of position.
The wavefunction of the state |ni expressed in position space is ψn (x) = hx|ni. We already know
what this wavefunction looks like:
r
2  nπx 
ψn (x) = sin .
L L

37
38 CHAPTER 8. QUANTUM OPERATOR EXPECTATION VALUES

Thus, the expectation value of x is


Z ∞
hxi = hn|x̂|ni = hn| x̂ |xi hx| dx |ni
−∞
Z ∞
= hn|x̂|xihx|ni dx
−∞
Z ∞
= hn|xixhx|ni dx
−∞
Z ∞
= ψn∗ (x) x ψn (x) dx

2 L  nπx 
Z  nπx 
= sin x sin dx
L 0 L L
2  L 2 nπ
Z
= ξ sin2 ξ dξ.
L nπ 0

In the first line, we inserted the identity operatore, 1, in the form given by the completeness relation
of the position eigenstates. Note that we changed the integration variable to ξ ≡ nπx/L.
To calculate the integral, write sin2 ξ as 12 1 − cos 2ξ . The integral of ξ cos 2ξ can be calculated


by integration by parts and one finds that it vanishes. Therefore, we get hxi = L/2, which you might
have guessed anyway: the average position of the particle is the center of the well.
Similarly, the expectation value of x2 is
2  L 3 nπ 2
Z
ξ sin2 ξ dξ.

2
x =
L nπ 0

The integral can be calculated the same way, but requries two integration by parts. Since I messed
this up initially, I will write down the steps.
Z nπ

2 2L2 1 − cos 2ξ
x = ξ2 dξ
(nπ)3 0 2
2L2 h 1 nπ 2 1 nπ 2
Z Z i
= 3
3ξ dξ − ξ · 2 cos 2ξ dξ
(nπ) 6 0 4 0
2 h nπ nπ Z nπ
2L 1 3 1 2   1
i
= ξ − ξ sin
 2ξ + 2ξ sin 2ξ dξ
(nπ)3 6 0 4 4 0

 0
2 h 3
1 nπ i

Z
2L (nπ) 1
= − ξ cos 2ξ +  cos 2ξ dξ

(nπ)3 6 4 0 4 0

L2 L2
= − .
3 2(nπ)2
The expectation value of p is
Z ∞
hpi = hn|p̂|ni = hn| p̂ |xi hx| dx |ni
−∞
Z ∞
= hn|p̂|xihx|ni dx
−∞
Z ∞
d
= −i~ ψn (x) ψn (x) dx
−∞ dx
2 nπ L  nπx 
Z  nπx 
= −i~ sin cos dx
L L 0 L L
Z nπ
2i~
=− sin ξ cos ξ dξ.
L 0
8.1. INFINITE SQUARE WELL 39

The integral can be calculated in numerous ways (e.g. using 2 sin ξ cos ξ = sin 2ξ). One finds that the
integral vanishes and so hpi = 0. This is not surprising since the particle moves left and right with
equal probability by symmetry.
We can read off the expectation value of p2 from the kinetic energy, which is the total energy in
this case since the potential is zero in the well:

n 2 π 2 ~2
2 n 2 π 2 ~2
En ≡ hHi = =⇒ p = 2m hHi = .
2mL2 L2
Before we calculate the uncertainties, let us simplify its expression:
D E

2 2 2 2
(∆x)2 = x2 − 2 hxi x + hxi = x2 − 2 hxi + hxi = x2 − hxi .

Note, of course, that hhxii = hxi.


Now, we can calculate the uncertainties in position and momentum:
s s r
L2 L2  L 2 L 6 n 2 π 2 ~2 nπ~
∆x = − 2
− = √ 1− 2
, ∆p = −0= .
3 2(nπ) 2 2 3 (nπ) L2 L

The product of the two satisfies the uncertainty relation:


s r
nπ 6 ~ (nπ)2 − 6 ~ ~
∆x∆p = √ 1− 2
= > .
3 (nπ) 2 3 2 2

I was concerned that this might not satisfy the uncertainty principle for the ground state (n = 1), but
those fears were unwarranted; it turns out that the prefactor of ~/2 is about 1.14 when n = 1. Phew!
It works out after all.
The wavelength of the nth energy eigenstate is λ = 2L/n. Just check this for a couple low values
of n; it should be pretty obvious if you draw the states. De Broglie’s relation gives the momentum
magnitude:
h 2π~ nπ~
|p| = = = .
λ 2L/n L
Note that this doesn’t take into account the direction of the momentum, which is the thing that made
hpi vanish. So, hopefully you don’t get confused as to why |p| =6 0 while hpi = 0. The speed of the
particle is
|p| nπ~
|v| = = .
m mL
Thus, the time between collisions against one of the walls is the time it takes for the particle to go
back and forth once across the well:
2L 2L 2mL2
∆t = = = .
|v| nπ~/mL nπ~

Finally, our estimate for the force is


∆p n 2 π 2 ~2
=
F = ,
∆t 2mL3
which is smaller than F = −dEn /dL by a factor of 2. That is a huge improvement from the previous
estimate which was smaller by a factor of 16π 2 ≈ 160.
40 CHAPTER 8. QUANTUM OPERATOR EXPECTATION VALUES
Chapter 9
Operators and States

The bra-ket notation may seem like simply a rewriting of stuff we already know from the language
of wavefunctions, albeit in more compact form. For example, with the ket |ni representing the same
quantum harmonic oscillator state as the wavefunction ψn (x), the expectation value of the position
operator in this state may be written as
Z ∞
hn|x̂|ni = ψn∗ (x) x ψn (x) dx.
−∞

In the language of wavefunctions, the “state” of a system is represented by the wavefunction. An


operator will almost always be a function of position and momentum, and all you have to do is replace

momentum is −i~ ∂x . To take the expectation value of operators, you sandwich it between the complex
conjugate of the wavefunction on the left and the wavefunction on the right and integrate over position.
There is a somewhat physical interpretation underlying this perspective, which is that ψ ∗ ψ = |ψ|2 is
supposed to be the probability density function.
In the bra-ket language, the “state” of a system is represented by a vector in an abstract complex
vector space, V, (a Hilbert space, to be precise). So, we must be able to add two vectors and the result
is supposed to represent the superposition of the two states that are represented by each vector. We
must be able to multiply a vector by an arbitrary complex number. We must also have an inner product
defined between vectors so that we actually mean something when we say one state is orthogonal to
another. An inner product is nothing more than a map from two copies of the vector space to the
non-negative real numbers: it takes two vectors and spits out a non-negative real number. If it spits
out 0, then the two vectors are orthogonal. Often, the inner product is denoted by (·, ·), where each ·
is slot in which one inserts a vector in the vector space. So, (·, ·) : V × V → R≥0 .
In the case of the harmonic oscillator, the most convenient choice of basis for the vector space is
the basis of energy eigenstates denoted by |ni for n = 0, 1, . . .. Note that this is an abstract vector
representing the nth eigenstate of the harmonic oscillator. It is NOT the wavefunction representing
the same state. The wavefunction and the ket (or vector) are two different representations.
In this picture, an operator, Ô, is nothing more than a map from the vector space to itself:
Ô : V → V. Having chosen a basis, one is free to represent that map by its matrix relative to the basis,
just as in linear algebra. In principle, this is no different than stating that once you have chosen the
standard Cartesian basis for R2 (Euclidean plane), then a rotation of the plane by an angle φ in the
counter-clockwise direction is represented by the matrix
   
(x̂, Rx̂) (x̂, Rŷ) cos θ − sin θ
= ,
(ŷ, Rx̂) (ŷ, Rŷ) sin θ cos θ
where R is the rotation operator, e.g. Rx̂ = cos θ x̂ + sin θ ŷ. The inner product is the standard dot
product in two dimensions, e.g. (x̂, x̂) = x̂ · x̂ = 1, (x̂, ŷ) = x̂ · ŷ = 0, etc.

41
42 CHAPTER 9. OPERATORS AND STATES

Actually, there is some difference. For one thing, the vector (Hilbert) spaces are often infinite
dimensional. Sometimes, they are not even discrete, as in the case of position and momentum eigen-
states. But, let’s not worry too much about this. Anyway, it’s not difficult to think about infinite
dimensional matrices, as long as there is a discrete basis. For example, we could tack on as many
extra dimensions as we want to R2 . The matrix representing the rotation described above would be
an infinite-dimensional matrix that has the same top-left 2 × 2 components as the one above, then 1’s
along the remaining diagonals and 0’s everywhere else.
Using the analogy of matrices to represent operators, you can think of kets as column vectors. An
operator turns a ket into another ket just like a square matrix turns a column vector into another
column vector. We can dream up a different vector space, called the dual vector space, V ∗ . Note that
the star here does not, a priori, have anything to do with complex conjugation. In the most abstract
sense, the elements of V ∗ are defined to be linear maps from V into R. In our particular case, we
restrict to maps into R≥0 .
To make this slightly more concrete, consider the example of the harmonic oscillator. A basis for
V is furnished by the kets {|0i , |1i , |2i , . . .}. Elements of V ∗ are linear maps V → R≥0 . One such
map is the map that sends |0i to 1 and everything else to 0. Another is the map that sends |1i to
1 and everything else to 0. In fact, these types of maps furnish a basis for V ∗ since the map that
sends |0i to a, |1i to b and so on is just a times the first map mentioned plus b times the second map
mentioned, etc. We denote the first map by the bra, h0|, the second map by h1|, etc. There is an
obvious isomorphism between V and V ∗ , which sends |0i to h0|, etc. If in addition to this, an arbitrary
complex constant multiplying a ket becomes complex conjugated under this isomorphism, then this
map is called the adjoint. In other words, an arbitrary linear combination, a |0i + b |1i + · · · is mapped
via the adjoint map into a∗ h0| + b∗ h1| + · · · . Then, we can define a bilinear map (·, ·) : V ∗ × V → R≥0
as follows: (hm| , |ni) is equal to the number to which hm| maps |ni, which happens to be 1 if m = n
and 0 if m 6= n. To save ink, we write hm|ni ≡ (hm| , |ni).
In terms of the inner product in V, all that the adjoint of an operator means is that it must satisfy
the relation (|mi , Ô† |ni) = (Ô |mi , |ni).
Big deal! We’ve managed to double the number of ways of describing or notating the states of our
system. Well, it turns out that thinking of abstract vectors and operators instead of wavefunctions
and differential operators is often more fruitful than it might seem, as exemplified by the following
question.

9.1 Harmonic Oscillator Bra-Kets


Consider the one-dimensional harmonic oscillator with frequency ω. Define
1
â = √ (mωx̂ + ip̂).
2m~ω
Use the results of problem set 9 question 4 to show that
√ √
â |ni = n |n − 1i , ↠|ni = n + 1 |n + 1i .

Let |ψi be a superposition of the ground and first excited states. Calculate the time-dependent ex-
pectation value of the position operator in this state, which is the average position as a function of time.

Note: This small â is related to the big  in the problem set via â = Â/ ~.

SOLUTION:

We have the commutation relation


[Ĥ, â] = −~ωâ.
9.1. HARMONIC OSCILLATOR BRA-KETS 43

Since [Ĥ, â] ≡ Ĥâ − âĤ, we may write



Ĥâ = â Ĥ − ~ω 1̂ ,

where 1̂ is just the identity operator: â = â1̂.


Act both sides of this equation on |ni and bracket out â |ni:

Ĥ(â |ni) = â Ĥ − ~ω 1̂ |ni = â ~ω n + 12 − ~ω |ni = ~ω (n − 1) + 12 (â |ni).


     

This shows that â |ni is an eigenstate of the Hamiltonian, which means that it must be proportional to
one of the |mi’s. The eigenvalue, or energy, happens to coincide with the energy of |n − 1i. Therefore,
â |ni = c(n) |n − 1i, where c is just a number, which may depend on n.
Now, take the inner product of â |ni with itself:

(â |ni , â |ni) = hn|↠â|ni = hn| ~ω


1
Ĥ − 12 |ni = hn| ~ω
1
~ω n + 12 − 12 |ni = nhn|ni = n.


But, this is also equal to


ˆ 1, cn −
(cn − ˆ 1) = |c|2 hn − 1|n − 1i = |c|2 .

Since we don’t distinguish between two


√ states that are related to each other be a complex phase, we
may assume c to be real. Thus, c = n, which√ is the result we wanted.
The same procedure will yield ↠|ni = n + 1 |n + 1i, but you may have to use the commutation
relation [â, ↠] = 1.

To get a sense of how much simpler using abstract states and operators is, let us first try to calculate
hx̂i using the wavefunctions. The wavefunction of our superposition state is
 mω 1/4 2
h  2mω 1/2 i
ψ(x, t) = e−mωx /2~
Ae−iωt/2 + B xe−3iωt/2 .
π~ ~
Here, A and B are arbitrary constants, which we will take to be real, for convenience. Then,
Z ∞
hx̂i (t) = ψ ∗ (x, t) x ψ(x, t) dx
−∞
 mω 1/2  Z ∞ 2
  2mω 1/2 Z ∞
2
= A2 xe −mωx
 /~
dx − AB (eiωt + e−iωt ) x2 e−mωx /~ dx
π~ Z −∞
 ~ −∞
 ∞ ((
2 2mω
 (
2
3 −mωx
( ( /~
+B ((( x e
( ( dx
(((~( −∞
√  ~ 3/2 Z ∞
2 2  mω  2
= √ AB cos ωt ξ 2 e−ξ dξ
π ~ mω −∞
 2~ 1/2 √
2 π
= √ AB cos ωt
π mω 2
 2~ 1/2
= AB cos ωt.

Note that an odd function in x multiplied by the Gaussian is an odd function whose integral from −∞
to +∞ vanishes. Now, first of all, I had to look up what the wavefunctions to begin with. Secondly, I
2
had to know the integral of ξ 2 e−ξ . Ugh!

Now, let us use kets and operators instead. We have

|ψi = Ae−iωt/2 |0i + Be−3iωt/2 |1i .


44 CHAPTER 9. OPERATORS AND STATES

We want to calculate

hx̂i = hψ|x̂|ψi = A2 h0|x̂|0i + AB(eiωt h1|x̂|0i + e−iωt h0|x̂|1i) + B 2 h1|x̂|1i

But, x̂ can be written in terms of the raising and lowering operators:


 ~ 1/2
x̂ = (â + ↠).
2mω
So a piece of x̂ lowers |ni and a piece raises |ni. Therefore, in order for hm|x̂|ni to be nonzero, we
must have either m = n − 1 or m = n + 1. Also,
√ √
hn − 1|â|ni = n, hn + 1|↠|ni = n + 1.

Hence, we find
 ~ 1/2  2~ 1/2
hx̂i = AB (eiωt + e−iωt ) = AB cos ωt.
2mω mω
Not once did I actually need to know what the exact wavefunctions were. Nor did I ever have to
calculate a single integral!
Chapter 10
Final Review

10.1 Wavefunction Shapes


Below is picture of an infinite potential well with a non-flat bottom. Explain your answers to the
following questions.
(a) For some arbitrary allowed energy, E, rank positions A, B and C by the classical kinetic energy
of the particle at these positions from largest to smallest.
(b) Repeat for de Broglie wavelength.
(c) Repeat for the amount of time a classical particle spends traversing an interval of width δx at each
position.
(d) Repeat for the spacings between the zeros of the wavefunction in the regions near each point.
Assume that the energy level is sufficiently high that the wavefunction oscillates many times
between the two walls.
(e) Repeat for the amplitude of the wavefunction in the region near each point.
(f) Sketch plausible wavefunctions for the n = 4 and n = 8 energy levels.

10.2 Harmonic Oscillator Coherent States


1
Eigenvectors of the lowering operator, â = √2m~ω (mωx̂ + ip̂), are called Coherent states. So, these
satisfy â |αi = α |αi for some α ∈ C.



(a) Calculate hxi, x2 , hpi and p2 in the state |αi.
(b) Calculate ∆x and ∆p and show that the uncertainty is saturated.

45
46 CHAPTER 10. FINAL REVIEW

(c) Expand |αi in terms of the basis of energy eigenfunctions, |ni for n = 0, 1, 2, . . .. Determine the
coefficients of the expansion as functions
P∞of α. Remember to properly normalize the state. You
n
may need the Taylor expansion, ex = n=0 xn! .

10.3 Finite Square Well Bound States


Let V (x) = −V0 for |x| ≤ L/2 and V (x) = 0 otherwise. By symmetry, the bound states are either
even or odd functions. Let E < 0 be the energy of a bound state. Define the constants
p p
L 2m|E| L 2m(V0 − |E|)
η≡ , ξ≡ .
2~ 2~
(a) Show that η 2 + ξ 2 = mL2 V0 /2~2 .

(b) Show that η = ξ tan ξ for even bound states and ξ = −ξ cot ξ for odd bound states.
(c) Draw a plot of η vs. ξ. Count the number of bound states as a function of V0 , L and m.

10.4 A Theorem about Bound States


(1)
Let V (1) (x) and V (2) (x) be two potentials such that V (1) (x) ≤ V (2) (x) for all x. Suppose E1 <
(1) (2) (2)
E2 < · · · and E1 < E2 < · · · are the bound state energies of the two potentials. Note: they need
not have the same number of bound states; they don’t even have to have any, necessarily. A theorem
(1) (2)
claims that En ≤ En .
Use this theorem to givep a lower and upper bound to the number of bound states for the potential
V (x) = 21 mω 2 x2 for |x| ≤ 8 ~/mω and V (x) = 32~ω otherwise.

10.5 Spin-Spin Interaction


Two particles with spins s1 = 32 and s2 = 12 interact with each other via a spin-spin interaction whose
Hamiltonian is Ĥ = α~ Ŝ1 · Ŝ2 , where α is some constant. At t = 0, the system is in the following
eigenstate of Ŝ12 , Ŝ1z , Ŝ22 , Ŝ2z : 3 1 1 1
, ; , .
2 2 2 2

Find
3 3 the state
of the system at times t > 0. What is the probability of finding the system in the state
, ; 1 , − 1 as a function of t?
2 2 2 2

10.6 Fine and Hyperfine Structure


(a) What is the zeroth order (Bohr model) value for the n = 2 energy for Hydrogen? Counting the
possible spin states of the proton, what is the degeneracy of this state (i.e. how many different
states have this same energy)?

(b) Describe qualitatively the effect of fine structure on this energy level. What does it do to the
degeneracy? Give an estimate (order of magnitude) of how much the energy levels change. Label
the states as n“`”j , where “`” = s, p for ` = 0, 1 respectively and give the degeneracy of each.
(c) Do the same thing for hyperfine structure.

You might also like