Sie sind auf Seite 1von 16

Problem Set 11 Solutions

137A Spring 2015


I. Siddiqi
Due 4/24/2015 at 5pm in the HW box in 251 LeConte.
Problem #1
(Based on Griffiths, 2nd ed, 4.33) An electron is at rest in an oscillating magnetic field
B = B0 cos(t)
z
where B0 and are constants.
The electron has a magnetic dipole moment that is proportional to its spin: = S for a constant of
proportionality , the gyromagnetic ratio. As in problem 6 of Problem Set 10, the Hamiltonian of this

= B = B S.
system is H
(a) Construct the Hamiltonian matrix for this system in the basis of Sz eigenstates. Note that it depends
explicitly on t.
above. Then we
Solution: We can get the Hamiltonian by substituting our expression for B into H
~

getH = S B = B0 cos(t)S z = B0 2 cos(t)z . This is just a scalar multiple of the Pauli


matrix z , so we get:


1 0
H = ~2 B0 cos(t)
0 1
(b) The electron starts out (at t = 0) in the spin-up state with respect to the x-axis (that is: (0) is the
eigenstate of Sx with eigenvalue +~/2). Write this state as a spinor (that is, a 2-component vector) in
the basis of Sz eigenstates.


0 1
~
Solution: We diagonalize the 2 2 matrix Sx = 2
to get the eigenspinor with eigenvalue
1 0


1
.
~/2, which is 12
1
(c) Determine (t) at any subsequent time. (Note: the Hamiltonian in this problem is time-dependent! That
means that we cannot use separation of variables to reduce the time-dependent SE to the time-independent
SE. You must solve the time-dependent SE directly.)
Solution: As the note says, we cannot use separation of variables to get rid of the time-dependence,
since the potential depends explicitly on t. So we will use the time-dependent Schrodinger equation:

= H
 t



~
i
1
1
B0 cos(t)
=
2
0
t
~
2
i~

0
1

Separating the two components:

iB0 cos(t)
1 =
1
t
2
and

iB0 cos(t)
2 =
2
t
2
1



1
2



These differential equations can be solved directly. I will show 1 explicitly, and 2 is the same apart
from a minus sign.
iB0 cos(t)

1 =
1
t
2
iB0 cos(t)
d1
=
dt
1
2
Z
Z
iB0
d1
=
cos(t)dt
1
2
iB0
ln |1 | =
sin(t) + C
2
iB0 sin(t)/(2)
1 = Ce

1 = eiB0 sin(t)/(2) / 2

where in the last line I have used 1 (0) = 1/ 2. Likewise, we get 2 = eiB0 sin(t)/(2) / 2. Combining
the two results, we find:
 iB sin(t)/(2) 
e 0
(t) = 12
eiB0 sin(t)/(2)
(d) If you measure Sx at time t, what values could you obtain? What is the probability of each?
Solution: As always, when we measure some operator, the possible results are its eigenvalues. So we
could get ~/2. To find the corresponding probabilities, we need to write our state as a sum of the
eigenstates, at which point we can get the probabilities from the norm squared of the coefficients. So
lets write (t) from the previous part as a linear combination of the two Sx eigenstates:

 iB sin(t)/(2) 
1
e 0

(t) =
eiB0 sin(t)/(2)
2
 


1
1
1
1
+ b
= a
1
1
2
2
so that
a + b = eiB0 sin(t)/(2)
a b = eiB0 sin(t)/(2)
To make the solution clearer, write this as a + b = eix , a b = eix . Adding the two equations gives
2a = eix + eix , so that a = cos(x), while subtracting the equations gives 2b = eix eix = 2i sin(x) so
that b = i sin(x). Substituting this result back into the earlier expression gives
(t) = cos(B0 sin(t)/(2))| i + i sin(B0 sin(t)/(2))| i




1
1
is the Sx eigenstate with eigenvalue ~/2 and | i 12
is the Sx
where | i 12
1
1
eigenstate with eigenvalue ~/2.
The probabilities are given by the norm squared of the coefficients, so we can measure:
~/2
~/2

with probability cos2 (B0 sin(t)/(2))


with probability sin2 (B0 sin(t)/(2))

Note that the probabilities sum to 1, as they must.


(e) What is the minimum field (B0 ) required to force a complete flip in Sx ?
Solution: We started in the state | i, so a complete flip means that we must end up in the state
| i at some time t. This is only possible if the coefficient of | i in (t) has norm 1, so that
1 = sin2 (B0 sin(t)/(2)). This can only happen if the argument of sine is an odd multiple of /2.
Depending on the values of B0 , , and , this may or may not be achievable at any time t. To find
the minimum B0 for which it is possible, just look at the smallest necessary value of the argument, /2.
The biggest that sin(t) can be is 1, so we want to solve B0 /(2) /2, or equivalently B0 .
Problem #2
Consider two particles, each with spin 21 . Written in terms of the Sz eigenstates for each particle, the singlet
state with total spin 0 is
1
|i = (|i1 |i2 |i1 |i2 )
2
where for each particle |i is the s = 12 , mz =
(a) Find the
find that

1
2

state and |i is the s = 12 , mz = 12 state.

of Sx and Sy (for s = 21 ) in terms of the Sz eigenstates. For instance, you should


eigenstates
s = 1 , mx = 1 = 1 (|i |i).
2
2
2

Solution: The easiest way to do this problem is to write the matrices for Sx and Sy in the basis of Sz
eigenstates, then diagonalize those matrices. We get Sx = ~2 x and Sy = ~2 y . But multiplying a matrix
by some constant doesnt change its eigenvectors, so actually we just need to find the eigenvectors of




0 1
0 i
x =
and y =
1 0
i 0
This is now a standard calculation, one which I have also spelled out in more detail in previous weeks
solutions. Thus I will just present the answer:


mx = 1 =
2


my = 1 =
2

1
2

1
2

1
1

1
i

+ |i



m x = 1 =
2

1
2

+ i |i



my = 1 =
2

1
2

1 (|i
2

1 (|i
2

1
1
i
1

1 (|i
2

|i

1 (i |i
2

+ |i


=

Note that these answers are not unique - a normalized eigenvector multiplied by a scalar with norm 1,
like 1 or i, is still a normalized eigenvector. Also note that I have not written out the s = 21 . This is
implied since we are talking entirely about particles that have spin 12 .
(b) Use your answers to part (a) to write the Sz eigenstates as linear combinations of (I) the Sx eigenstates
and (II) the Sy eigenstates. (Note: these are two distinct calculations!)
Solution: For the Sx states, we see that adding them together will cancel the |i component, will
subtracting them will cancel the |i component. This lets us write the SZ eigenstates in terms of the Sx
eigenstates as:
|i =

1
2




mx = 1 + mx = 1 , |i =
2
2

1
2




mx = 1 mx = 1
2
2



Similarly, we can cancel the |i components from the Sy eigenstates if we multiply the my = 21 state
by i and then add them, and we can cancle the |i components if we multiply the my = 21 state by
3

i and then add. This gives:


|i =

1
2





my = 1 i my = 1 , |i =
2
2

1
2




i my = 12 + my = 21

(c) Substitute the expressions from part (b) into |i to rewrite the singlet state (I) in terms of Sx eigenstates
and (II) in terms of Sy eigenstates.
Solution: Lets go through the algebra, first for the Sx eigenstates and then for the Sy eigenstates. To
make
the algebra look a little

neater and to help with the comparison for part (d), I will denote the
mx = 1 state by |x i, the mx = 1 state by |x i, and likewise for the Sy eigenstates.
2
2

1
|i = |i1 |i2 |i1 |i2
2


1
1
1
1
1
(|x i1 + |x i1 ) (|x i2 |x i2 ) (|x i1 |x i1 ) (|x i2 + |x i2 )
=
2
2
2
2
2


1
|x i1 |x i2 |x i1 |x i2 + |x i1 |x i2 |x i1 |x i2

=
|x i1 |x i2 + |x i1 |x i2 |x i1 |x i2 |x i1 |x i2
2 2


1
= 2 |x i1 |x i2 2 |x i1 |x i2
2 2

1
= |x i1 |x i2 |x i1 |x i2
2

1
|i = |i1 |i2 |i1 |i2
2


1
1
1
1
1
(|y i1 i |y i1 ) (i |y i2 + |y i2 ) (i |y i1 + |y i1 ) (|y i2 i |y i2 )
=
2
2
2
2
2


1
i |y i1 |y i2 + |y i1 |y i2 |y i1 |y i2 i |y i1 |y i2

=
i |y i1 |y i2 |y i1 |y i2 + |y i1 |y i2 i |y i1 |y i2
2 2


1
= 2 |y i1 |y i2 2 |y i1 |y i2
2 2

1
= |y i1 |y i2 |y i1 |y i2
2
(d) Compare the expressions for the singlet state in the bases of Sx , Sy , and Sz eigenstates. You should find
that in every case the form of the singlet state is the same, with the possible exception of a physically
irrelevant overall phase. Explain this fact using a symmetry argument.
Solution: In the three bases, the singlet state looks like

1
|i = |z i1 |z i2 |z i1 |z i2
2

1
|i = |x i1 |x i2 |x i1 |x i2
2

1
|i = |y i1 |y i2 |y i1 |y i2
2
These are all the same except for the overall minus sign in the x case, which is irrelevant because an
overall phase doesnt affect ||2 and therefore doesnt affect probabilities.

It makes sense that the singlet state would look the same in each case. The singlet state has total spin l,
which means it has only one m value, m = 0. But this is true if we look at mx or my or mz (or mn for
any other direction n). In other words, the state must look the same along any axis, since the component
of angular momentum is 0 along each of the three axes simultaneously! (Note: this doesnt violate the
uncertainty principle. The product of uncertainty in, say, Lx and Ly is related to the expectation value
of Lz , which is 0 if mz = 0.) If the physical state of the particle doesnt pick out any direction in space
as being special, then we should be able to just relabel the x direction or the y direction as z without
changing anything. But relabeling doesnt change how the state looks in terms of the single-particle
eigenstates, so the state had better look the same in the x basis as in the z basis so that when we relabel
x z, it looks correct, the same way it did in the z basis in our original labeling scheme.
Problem #3
(Based on Griffiths, 2nd ed, 4.35) Quarks are elementary particles with spin 21 . Particles composed of three
quarks, such as protons and neutrons, are called baryons. Particles composed of two quarks, such as the
pion, are called mesons.
(a) What are the possible values of the total spin quantum number, s, for mesons?
Solution: The idea for this problem is to use the selection rule for total angular momentum: when
writing a state |j1 , m1 i |j2 , m2 i as a sum of total angular momentum eigenstates |j1 , j2 ; j, mi, the possible
values for j are |j1 j2 |, |j1 j2 | + 1, , j1 + j2 .
In this case, we have two quarks, each with spin 21 , so s1 = s2 = 12 and thus the total spin can be
s = 0 or 1 . The corresponding states are the spin singlet and three spin triplet states, as presented in
lecture, the textbook, and discussion section.
(b) What are the possible values of the total spin for baryons?
Solution: We now have three quarks, each with spin 12 . Putting the first two together, we can get total
spin s = 0 or 1 as in part (a), but we still have to add the third quark. If the first two have total spin 0,
then adding an additional quark with spin 12 gives a total spin between |j1 j2 | = |0 1/2| = 1/2 and
j1 + j2 = 1/2. So in that case s = 1/2.
On the other hand, if the first two quarks together have total spin 1, adding the third quark could give
total spin |j1 j2 | = |1 1/2| = 1/2 or j1 + j2 = 1 + 1/2 = 3/2. (Note that s = 1 is not allowed because
corresponding m values would then be integers, which is not possible here.)
We can conclude that the total spin can be s =

1
2

or

3
2

Problem #4
B&J 6.23
(Note: for this problem and the following one, you will need to refer to a table of Clebsch-Gordan coefficients.
One is given on page 320 of B&J. You can find an alternative presentation on bCourses, in the Discussion
Section Notes category. It is called Clebsch-Gordan Table.)
Solution: In this problem, were basically doing the algebra behind problem 3, part (b). We are explicitly
putting together the three particles of spin 12 , and we will explicitly find the eigenstates of S 2 and Sz where
S is the total spin operator for the three-particle system, S = S1 + S2 + S3 . One option would be to use the
procedure from discussion section, which would give us 8 8 matrices for Sz and S 2 , but its a bit easier to
use the Clebsch-Gordan tables instead.
Our goal is to find the eigenstates of total S 2 and total Sz as linear combinations of the product basis states
|i , |i , |i. There are eight such product basis states, so we should expect to get eight states in

the total angular momentum basis as well. As the problem statement suggests, these will be organized into
quadruplet states (four states for the four different m values if the total spin is 3/2) and doublet states (two
states for the two different m values if the total spin is 1/2). In fact, to get eight total states, we will get
one quadruplet and two independent doublets. In other words, we expect to get four states with total spin
3/2 and another four with total spin 1/2.
The procedure will be as follows: (1) we will write the possible states of the first two particles as the triplet
or singlet states; (2) forgetting about the fact that the spin-1 triplet states and spin-0 singlet state actually
represent a combination of two particles, we will take
the product
|s = 1, m = 1i, |s = 1, m = 0i,

of the
|s = 1, m = 1i, and |s = 0, m = 0i states with the s3 = 12 , m3 = 21 and s3 = 12 , m3 = 12 states and use
the Clebsch-Gordan table to write these products in the total spin basis; (3) we will invert these relationships
to find the total spin states as linear combinations of the (2 particle)(1 particle) product basis states; and
finally (4) we will substitute back in the expressions for the s = 0 and 1 states of the first two particles in
terms of the individual spins of the two particles. This is a fairly long procedure, but in practice it is not
quite as hard as it sounds.
The main work in the problem is in step (2). The first step in writing out each product is to apply the
selection rules: m = m1 + m2 and j = |j1 j2 |, , j1 + j2 . We also use the extra rule that |m| j. This
gives:



|j1 = 1, m1 = 1i j2 = 12 , m2 = 12


|1, 1i 12 , 12


|1, 0i 12 , 12


|1, 0i 12 , 12


|1, 1i 21 , 12


|1, 1i 21 , 12


|0, 0i 12 , 12


|0, 0i 1 , 1
2



= a1 j = 23 , m = 32




= a2 32 , 21 + b2 21 , 12




= a3 32 , 21 + b3 21 , 12




+ b4 12 , 12
= a4 32 , 1
2




+ b5 12 , 12
= a5 32 , 1
2


= a6 32 , 3
2


= a7 12 , 21


= a8 1 , 1
2

For the cases where there is only one term on the right, the corresponding coefficient must be 1 because all of
the individual states are normalized. For the remaining cases, we have to use the Clebsch-Gordon coefficient
table to find the coefficients. This gives:




|1, 1i 12 , 12 = 32 , 32
q


q2 1 1
1 3 1

|1, 1i 12 , 21 =
,
3 2 2 +
3 2, 2
q


q1 1 1
2 3 1

|1, 0i 12 , 12 =
3 2, 2
3 2, 2
q
q






2 3 1
|1, 0i 12 , 21 =
+ 13 12 , 12
3 2, 2
q


q2 1 1
1 3 1
|1, 1i 12 , 12 =
3 2, 2
3 2, 2
1 1
3 3
|1, 1i 2 , 2 = 2 , 2




|0, 0i 12 , 12 = 12 , 12




|0, 0i 12 , 21 = 12 , 12
Thats it for step (2). Now onto step (3): we need to invert this relationship to isolate the total spin
eigenstates. This isnt too bad though: four of them are already done, and the remaining four are actually

separated into two pairs, one for total m = 1/2 and the other for total
m
at the m = 1/2
= 1/2. Lets look
pair, which are in the second and third lines above. To eliminate 12 , 12 , we should add 2 times the third


line to the second line, which gives 33 32 , 12 . We should thus divide by 3 to get


3 1 q1
q

,
=
|1, 1i 1 , 1 + 2 |1, 0i 1 , 1
2 2

Following a similar procedure, we get


3 3
,
2 2
3 1
,
2 2
3 1
,
2
2
3 3
,
2
2

2 2

the other states with total s = 3/2:




= |1, 1i 12 , 21
q
1 1 q2
1 1
1
, +

=
|1,
1i
3
2
2
3 |1, 0i 2 , 2
q
1 1 q1
1 1
2
, +

=
|1,
0i
3
2
2
3 |1, 1i 2 , 2
1 1
= |1, 1i 2 , 2

The situation with the total spin 12 states seems a bit more complicated. In particular, it seems like we are
getting contradictory results: on the one hand, by eliminating s = 32 states by taking linear combinations,
we get
q
1 1
1 1 q1
1 1
2
,
,

=
|1,
1i
2 2
3
2
2
3 |1, 0i 2 , 2
q
1 1
1 1 q2
1 1
1
,


=
2
2
3 |1, 0i 2 , 2
3 |1, 1i 2 , 2
But on the other hand we also have

1 1

,
= |0, 0i 21 , 12
2 2
1 1


,
= |0, 0i 21 , 12
2
2




There seem to be two completely different expressions for the 21 , 12 and 12 , 12 states! In fact, this is
correct: although we have written these states the same way, they are not the same state: there are two
independent (in fact, orthogonal) states for each of the pairs s = 21 , m = 12 and s = 21 , m = 12 . In each of
these sectors, the most general state is some linear combination of the two independent states. For this
problem, we wont worry about the preferred way of picking those linear combinations, although I will
make a brief comment about an alternative choice after finishing the main part of the problem.
Anyway, weve now finished step (3).
3 3
,
2 2
3 1
,
2 2
3 1
,
2
2
3 3
,
2
2
1 1
,
2 2 a
1 1
,
2
2 a
1 1
,
1 2 12 b
,
2

2 b

Our result is:




= |1, 1i 12 , 12
q
1 1 q2
1 1
1


=
3 |1, 1i 2 , 2 +
3 |1, 0i 2 , 2
q
q
1 1
1 1
2
1


=
3 |1, 0i 2 , 2 +
3 |1, 1i 2 , 2


= |1, 1i 21 , 12
q
1 1 q1
1 1
2
,

=
|1,
1i
3
2
2
3 |1, 0i 2 , 2
q
q
1 1
1 1
1
2


=
3 |1, 0i 2 , 2
3 |1, 1i 2 , 2
1 1
= |0, 0i 2 , 2


= |0, 0i 1 , 1
2

where the subscripts a and b distinguish between the two different doublets. Note that as expected we have
7

eight total states in our new basis.


Finally, we are ready for step (4): we substitute in the expressions for the combined states of two particles,
namely



|1, 1i = 12 , 12 12 , 12





|1, 0i = 12 12 , 12 12 , 12 + 12 , 12 12 , 12



|1, 1i = 12 , 12 12 , 12





|0, 0i = 1 1 , 1 1 , 1 1 , 1 1 , 1
2 2

2 2





Then, writing 12 , 12 as |i and 12 , 12 as |i, we get:
3 3
,
= |i
2 2
q
q h
3 1
i
1
2 1
,
=
|i
+
|i
+
|i
|i
2 2
3
3
2
q
q
i
h
3 1

2 1
,
=
|i + |i |i + 13 |i
2
2
3
2
3 3
,
= |i
2
2
q
q h
1 1
i
2
1 1
,
|i
+
|i
|i

|i
2 2 a =
3
3
2
q h
q
i
1 1

1 1
,
|i + |i |i 23 |i
2
2 a =
3
2
h
1 1
i
,
1
|i

|i
|i
=
2 2 b
2
h
i
1 1

,
1
|i |i |i
2
2 b =
2
Or, simplifying the expressions:
3 3
,
2 2
3 1
,
2 2
3 1
,
2
2
3 3
,
2
2
1 1
,
2 2 a
1 1
,
2
2 a
1 1
,
2 2 b
1 1
,
2
2 b

= |i
q 

1
=
3 |i + |i + |i
q 

1
=
|i
+
|i
+
|i
3
= |i
q
q
q
2
1
1
=
|i

|i

3
6
6 |i
q
q
q
1
1
2
=
6 |i +
6 |i
3 |i
=
=

1
2
1
2

|i
|i

1
2
1
2

|i
|i

We have found all of the spin functions, corresponding to the total spin quantum numbers s =
Note that for the spin-half case, other linear combinations are possible.

3
2

and s = 12 .

In fact, we can find linear combinations that look a little nicer than these ones! One particularly nice looking

set of total spin 1/2 states is:


1 1
,

2 2 1

1 1
,
2

2 1

1 1
,

2 2 2

1 1
,
2

2 2

1 1


,
i 1 1
2 2 a + 2 2, 2 b




= 12 21 , 12 a + i2 12 , 21 b




= 12 21 , 12 a i2 21 , 12 b




= 12 21 , 12 a i2 12 , 21 b
=

1
2

1
3

|i + ei

2
3

|i + ei

4
3

|i

|i + ei

2
3

|i + ei

4
3

|i

1
3

|i + ei

4
3

|i + ei

2
3

|i

|i + ei

4
3

|i + ei

2
3

|i

1
1
I have chosen
which
1 m1 = 2 state goes with which m 1= 2 state by requiring that the total Sx operator

1 1


turn 2 , 2 1 into 2 , 2 1 , etc, just as for a single spin 2 particle the Sx operator flips the up state to the
down state and vice-versa.

Problem #5
(Based on Griffiths, 2nd ed, 4.55) In this problem we explore addition of angular momentum and combined
wave functions of position and spin. Consider an electron in a hydrogen atom in the combined spin and
position state

p
p
1/3Y10 | i + 2/3Y11 | i
R21
where | i |s = 21 , ms = 21 i and | i | 12 , 21 i. Also, in the following problems let J be the total angular
momentum, J = L + S. (The expression for R21 can be found on page 361 of B&J. What is the value of Z?)
(a) If you measured the orbital angular momentum squared (L2 ), what values might you get, and what is the
probability of each?
Solution: The state of the system is described by a position part and a spin part. The basis states are
all the pairs of the individual basis states: (position basis state, spin basis state). Then the electron we
consider is in a linear combination of two of these product basis states: R21 Y10 | i and R21 Y11 | i.
Since the L2 operator acts only on the Ylm part of each basis function and leaves the rest invariant, each
of these basis states is an eigenstate of L2 . For example,
L2 R21 Y10 | i = R21 (L2 Y10 )| i = R21 (2~2 Y10 )| i = 2~2 R21 Y10 | i
Since both basis states have quantum number l = 1, they are both eigenstates of L2 with eigenvalue
2~2 . Thus in a combination of these states, you must measure 2~2 , with probability 1 .
(b) Same for the z component of orbital angular momentum (Lz ).
Solution: The logic is just as above, only Lz returns ~m for each of the two basis states. The first has
m = 0, while the second has m = 1. We get the corresponding probabilities from the norm squared of
the coefficients as usual. This gives:
0, probability = 1/3
~, probability = 2/3
(c) Same for the spin angular momentum squared (S 2 ).
Solution: Again, the idea is the same. We have the two product basis states as above, and both are
eigenstates of S 2 with eigenvalue ~2 s(s + 1) = 34 ~2 . So we measure 34 ~2 with probability 1 .
(d) Same for the z component of spin angular momentum (Sz ).

Solution: Just like Lz : each of the two basis states is an eigenstate of Sz , and the eigenstate of each is
~ms . For the first state, ms = 21 , while for the second, ms = 12 . Thus we measure
~
2,
~2 ,

probability = 1/3
probability = 2/3

(e) Same for the total angular momentum squared (J 2 ).


Solution: Now we get to the tricky part. Our basis states are currently expressed using the quantum
numbers n, l, m, s, and ms : Rnl Ylm ms . But to say anything useful about J, we need to switch to the
total angular momentum basis, which has quantum numbers n, j, mj , l, and s. We do this by means of
Clebsch-Gordon coefficients. Lets proceed carefully, one state at a time.
First we look at the basis state R21 Y10 | i. We dont care for the time being about the radial part since
that has nothing to do with J 2 . Focusing on the angular momentum, we see that the quantum numbers
are l = 1, ml = 0, s = 1/2, ms = 1/2. I will denote this state as |l, ml i|s, ms i = |1, 0i| 12 , 12 i. We want
to rewrite this state as a sum of total angular momentum states, which I will denote by |j, mi. These
states still have the implicit quantum numbers l and s, but I will omit them to make the identity of the
state more readily apparent.
The first step is to use our selection rules to determine which values of j and m are possible. The two
rules of interest are:
lsj l+s
m = ml + ms
In the case of our problem, these imply that j = 1/2 or 3/2 and mj = 1/2 (note that we do not allow
j = 1 because that there is no state with j = 1 and m = 1/2). Thus we can write our original state as




E 1 1 E
1 1

3
1


+ b ,
= a j = , m =
1, 0 ,
2 2
2
2
2 2
for some coefficients a and b. These are precisely the Clebsch-Gordon coefficients, and they can be looked
up in a table, either the one on page 320 of B&J or the one provided on bCourses. Using the table, we
conclude that


E 1 1 E r 2 3 1 
1 1 1


0
,
=
,
,

1,
2 2
3 2 2
3 2 2
We can do the exact same thing with the second basis state. The only difference is that m2 = 1/2.
We get:

 r

E 1 1 E
1 3 1
2 1 1


+
,
= ,
1, 1 ,
2 2
3 2 2
3 2 2

10

Now we should substitute these results back into our original expression to complete our change of basis:

E 1 1 E p E 1 1 E

p
p
p




1/3Y10 | i + 2/3Y11 | i = R21
1/3 1, 0 ,
R21
+ 2/3 1, 1 ,
2 2
2 2






p
p
p
3 1
1 1


= R21
1/3
2/3 ,
1/3 ,
2 2
2 2





p
p
p
3 1
1 1
1/3 ,
+ 2/3 ,
+ 2/3
2 2
2 2



!
1 1 1
2 2 3 1
= R21
,
,

3 2 2
3 2 2
So finally we have written our state as a sum of two basis states in the total angular momentum basis.
Each of the two states is an eigenstate of J 2 , and the eigenvalues are given by ~2 j(j + 1). The first state
has j = 3/2, while the second has j = 1/2. The probabilities are again given by the norm squared of
the coefficients. Thus we measure:
15 2
4 ~ ,
3 2
4~ ,

probability = 8/9
probability = 1/9

(f) Same for the z component of total angular momentum (Jz ).


Solution: This works just like Lz and Sz above, but we use the state as written in our new total
angular momentum basis, as we did for J 2 . But this time, both mj values are 1/2, so we measure
~/2 with probability 1 .
(g) If you measured the position of the particle, what is the probability density for finding it at r, , ?
Solution: When we measure position, we need to be careful in how we deal with the spin. We somehow
want to say that it is irrelevant, so lets think carefully about the interpretation. To do so, we will first
think about rules of probability. I want to know if some event A happens. But for any other event C, it
must be true that P (A) = P (A&C) + P (A&( C)), where means not. Going back to the problem
at hand, the probability of the particle being in a certain location (A) is the probability of it being in
that location and being spin up (A&C) plus the probability of it being in that location and being spin
down (A&( C)). (Note that I do not need to worry about the order of these measurements - see the
next part of the problem.)
The probability density (which satisfies the same rule described above for the probability) for finding
the particle at r, , is:
P(r, , ) = P(r, , ; ) + P(r, , ; )
p
p
= |R21 (r) 1/3Y10 (, )|2 + |R21 (r) 2/3Y11 (, )|2


= |R21 (r)|2 (1/3)|Y10 (, )|2 + (2/3)|Y11 (, )|2
Leaving the answer in this form is perfectly fine - that is a correct answer. But we could also plug in
the expressions for R21 , Y10 , and Y11 , in which case we find:
1
r2 er/a0
96a50

[Note that I have not included r2 sin() from the integration measure. Since the problem asks about
probability density for being at a particular point in space, it makes the most sense to define it as a per
volume density, and the r2 sin() is part of dV (or if you prefer, dr).]

11

(h) Suppose I want to measure both the z component of the spin and the distance from the origin. Explain
why I can measure both simultaneously.
Solution: We can measure two observables simultaneously if the operators commute. Since the spin
operator and the position operator act on completely different types of functions, they must be unaware
of each other; that is, [S, r] = 0. So we can measure them simultaneously.
(i) If I make the measurement from the previous part, what is the probability density for finding the particle
with spin up and at radius r?
Solution: This is actually very similar to the question of finding P(r, , ), where we summed over
possible values for the quantum numbers we didnt care about (namely ms ). To help make the steps
intuitively clear in the earlier part, I phrased this as some event (spin up) happening or not, but the
important aspect was really just that spin up and spin down exhausted all possibilities and were mutually
exclusive. So I could rewrite the earlier principle as
X
P (A&Ci )
P (A) =
All disjoint
outcomes, Ci
In the current case, we dont care about and , but those are continuous, so the sum should become an
integral. In particular, we will integrate over all value of and using the usual 3D spherical integration
measure, r2 sin()dd. (The fact that we include the factor of r2 even though we are integrating only
over and might seem strange, but it is correct. The r2 in the integration measure is actually
associated with d and d, not with dr; in the intuitive derivation of the spherical coordinate volume
element, dV = r2 sin()dr d d, we consider a small rectangular prism of at the end of a wedge extending
from the origin, and the side lengths are dr in the r direction, r d in the direction, and r sin()d
in the direction. Alternatively, you can think of finding the probability density at r as finding the
probability to be in a thin shell at radius r with thickness dr, then dividing by dr to get density, in
which case the r2 will be present. See Problem set 12, problem 4.) So we have:
Z2 Z
P(r; ) =

P(r, , ; )r2 sin()dd

=0 =0

Z2 Z
=

p
|R21 (r) 1/3Y10 (, )|2 r2 sin()dd

=0 =0

= (1/3)|R21 (r)|2 r2

Z2 Z

|Y10 (, )|2 sin()dd

=0 =0
2 2

= (1/3)|R21 (r)| r

where I have used the fact that the Ylm are normalized. Again, this is a valid answer as is, but we could
also plug in the expression for R21 to get
1
r4 er/a0
72a50

(j) If you measured the position of the particle, what is the probability density for finding it at polar angle
?
Solution: Using the same idea as in the previous part, we can find the probability to be at polar angle
with spin up, or at polar angle with spin down. The total probability is just the sum of the two:

12

P() = P(; ) + P(; )

Z Z2
Z Z2

P(r, , ; )r2 sin()drd


=
P(r, , ; )r2 sin()drd +
r=0 =0

r=0 =0

Z Z2

Z Z2
p
p

|R21 (r) 1/3Y10 (, )|2 r2 sin()drd +


|R21 (r) 2/3Y11 (, )|2 r2 sin()drd

r=0 =0

|R21 (r)|2 r2 dr

r=0

Z2

r=0 =0

Z2 

2
1

|Y10 |2 + |Y11 |2 sin()d


3
3

=0

 

  
2 3
1 3
2
2
cos () +
sin () sin()d
3 4
3 8

=0


= 2
=


1
1
2
2
cos () +
sin () sin()
4
4

sin()
2

Problem #6
In lecture we have solved the Schr
odinger equation for the free particle both in Cartesian coordinates and in
spherical coordinates. In this problem we will solve it in cylindrical coordinates, where r = (r cos(), r sin(), z).
Note that the following steps will walk you through the solution. In order to keep your solution synchronized
~2
with the intended solution and therefore consistent with later steps, do not divide through by 2
until
directed to do so.
(a) Write down the time-independent Schr
odinger equation for a free particle with mass in three dimensions. Specialize to the case of cylindrical coordinates by using the corresponding formula for the
Laplacian (you can look this up - you dont need to derive it).
Solution: The general Schr
odinger equation, for a free particle with V = 0, looks like

~2 2
= E
2

In the case of cylindrical coordinates, the Laplacian operator is


1
=
r r
2

r
r


+

1 2
2
+
r2
z 2

We just substitute this in to get our final answer:


"
#


2
~2 1

1 2
2

r
+ 2
+
= E
2 r r
r
r
z 2

13

(b) We will now perform separation of variables in two steps. First, write (r) = F (r, )Z(z). Substitute
this expression into your Schr
odinger equation from part (a) and perform separation of variables to get
two differential equations - one for F (r, ) and one for Z(z). Call the separation constant Ez . In terms
of a fixed value for Ez , find the most general solution for ZEz (z).
Solution: When we substitue in = F Z, F passes through derivatives with respect to z and Z passes
through derivatives with respect to r and . We thus get:
"
#


2
F
Z 2F
2Z
~2 Z
r
+ 2
+ F 2 = E FZ

2 r r
r
r
z
We then divide through by F Z and move the parts with F to the right-hand side to get:
h

2i
F
1 2F
~2 1
~2 2 Z
r
+
2
2
2 r r
r
r
z 2
=
+E
Z
F
Both sides are constant, and so we can give this constant a name. This is the separation constant
which we call Ez . This gives us two new differential equations which are independent apart from sharing
that one constant:
~2 2 Z
= Ez Z
2 z 2
"
#


2
F
1 2F
~2 1
= (E Ez )F
r
+ 2

2 r r
r
r

We can now solve the equation for Z(z). This is just the 1D free particle Schrodinger equation, so the
result is
Z(z) = Aeikz z + Beikz z
q
z
for kz = 2E
~2 .
(c) Next, write F (r, ) = R(r)() and subsitute that expression into your equation for F . Divide the
~2
equation through by 2
, then separate variables again to get two equations, one for R(r) and one for
(). Write the separation constant as m2 . Then solve for m (). What are the allowed values of m?
Solution: Following the suggested steps gives:
"
#


2
~2
R
R 2

r
+ 2
= (E Ez )R
2 r r
r
r


2

R
R 2
2(E Ez )
r
+ 2
+
R = 0
r r
r
r
~2


2
R
2
2(E Ez ) 2

r
+R
+
r
r R = 0
r
r

~2
2
 2(EEz ) 2
2

r r
r R
r R

r +
~2
=
R

14

Both sides are constant, and we call this separation constant m2 to get two equations:
2

2
= m2




R
2(E Ez ) 2
r R = m2 R
r
r
+
r
r
~2
Now we can solve the equation. The solution is again a sum of complex exponentials,
() = Ceim + Deim
Since an angle ( + 2) describes the same physical location as the angle , we must have ( + 2) =
(). This imposes the condition that m is an integer.
(d) Find the subsitutions required to make the differential equation for R look like:
2

dR
d2 R
+
+ (2 2 )R = 0
2
d
d

This is called the Bessel equation. The (normalizable) solutions are the Bessel functions, J , for
= 0, 1, 2, . These are the same functions that we wrote in lecture with a capital J, not to be confused
with the spherical Bessel functions that we wrote with a lower case j.
Solution: The first step is to get 0 on the right-hand side. Then we can proceed by pattern-matching.

We will also need the useful fact that if = kr, then r


=
r = k . This means for instance that
for = kr,







R
R

R
k
r
r
= k
=

r
r
k k

So lets find our substitutions.





R
2(E Ez ) 2
r
r R = m2 R
r
+
r
r
~2


R
2(E Ez ) 2

r
+
r R m2 R = 0
r
r
r
~2


R
2R
2(E Ez ) 2
r
+r 2 +
r R m2 R = 0
r
r
~2
d2 R
dR
r2 2 + r
+ (k 2 r2 m2 )R = 0
dr
dr
z)
z)
where k 2 = 2(EE
. Then lets let = kr = 2(EE
r and = m. Using the fact proved above that
~2
~2
the derivative term is the same in terms of r and in terms of , we then get the desired equation,

2
for = kr =

2(EEz )
~2

d2 R
dR
+
+ (2 2 )R = 0
2
d
d

r and = m.

(e) The most general solution to the free particle time-independent Schr
odinger equation in cylindrical coordinates can be written as
(r, , z) =

X Z
m

cm (, Ez )J ()m ()ZEz (z) dE 0 dEz

Ez =0 E 0 =0

15

where E 0 = E Ez and and will depend on E 0 and m as found in part (d). Note that every value
of E from 0 to is allowed.
Now suppose that instead of being free, the particle lives in an infinite cylindrical well. That is, V (r, , z)
is 0 if r < a and |z| < b and otherwise. (This is the cylindrical coordinate equivalent of the usual 1D
infinite square well.) Explain why this quantizes the energy levels, then find the lowest allowed energy
and the corresponding eigenstate. [Hints: you may use the fact that the smallest positive root of Jn (x) is
greater than the smallest positive root of Jm (x) if n > m. The answer cannot be found analytically, but
both the energy and the eigenvalue can be written in terms of a single unitless constant, . You should
give analytical expressions in terms of and you should separately give the value of that constant. You
can find that value in numerous references online (a Google search immediately turns up many of them),
or you can use the root command on Wolfram Alpha.]
Solution: The potential quantizes the energy levels because must be continuous at the boundary of
the cylinder, and therefore the free particle solution given in the problem statement must go to 0 at
r = a and at z = b. Given that Jm () is only zero at specific values of and that Aeikz z + Beikz z is
only 0 at specific values of kz z, once we plug in r = a and z = b these will fix allowed values of k and
kz , giving us (E Ez ) and Ez respectively.
Our goal is to find the ground state wave function and energy. Since Ez is determined by the boundary
conditions on Z(z) and (E Ez ) is independently determined by the boundary conditions on R(r), we
should separately find the lowest energy eigenstate and eigenvalue in each case.
For Z(z),qthis is just the infinite square well as solved in class back in chapter 4. The ground state is

2 2

z , with corresponding energy Ez 1 = ~8b2 .


Z1 (z) = 1b cos 2b
For R(r), the boundary condition is more complicated. We want to find the smallest allowed value of
(E Ez ) and therefore the smallest allowed value of k. Our condition is that Jm () must be zero at
r = a, so the allowed values of k are given by Jm (ka) = 0. Then to find the smallest allowed k, our goal
becomes: find the smallest value such that for some allowed m, Jm () = 0. Now we need to use the
given fact that a smaller value of m means the corresponding Jm has a smaller first root. Thus of all
the integers m, the Jm with the smallest first root will be J0 , and our desired smallest will be the first
root of J0 . This value can be looked up in a table, and the result is 2.40. The corresponding value
2 2
~2 2
of k is /a, so the corresponding energy is ~2k = 2a
2 and the eigenstate is R(r) = F J0 (r/a) for some
normalization constant F .
Finally we need to put all the pieces together. It would seem that we havent yet found (), but in fact
we determined that for the ground state, m must be 0. (In fact, noticing that the equation actually is
L2z = (~m)2 , we see that the m number represents angular momentum, so it shouldnt be surprising
that the lowest energy state has m = 0.) This gives () = constant. Thus we get:
E=

~2 2
~2 2
+
2a2
8b2

(r, , z) = N J0 (r/a) cos

 
z inside the cylinder, 0 outside
2b

2.40

16

Das könnte Ihnen auch gefallen