You are on page 1of 59

Solutions to Problems in Merzbacher,

Quantum Mechanics, Third Edition


Homer Reid
November 20, 1999

Chapter 2

Problem 2.1
A one-dimensional initial wave packet with a mean wave number
kx and a Gaussian amplitude is given by


x2
+
ik
x
.
(x, 0) = C exp
x
4(x)2
Calculate the corresponding kx distribution and (x, t), assuming
free particle motion. Plot |(x, t)|2 as a function of x for several
values of t, choosing x small enough to show that the wave packet
spreads in time, while it advances according to the classical laws.
Apply the results to calculate the effect of spreading in some typical
microscopic and macroscopic experiments.

The first step is to compute the Fourier transform of (x, 0) to find the
distribution of the wave packet in momentum space:

(k) = (2)1/2

(x, 0)eikx dx


x2
+
i(k

k)x
dx
= (2)1/2 C
exp
0
4(x)2

(1)

(I have dropped the x subscripts, and I write k0 instead of k).


To proceed we need to complete the square in the exponent:

x2
+ i(k0 k)x
4(x)2


x2
2
2
2
2

i(k

k)x

(k

k)
(x)
+
(k

k)
(x)
0
0
0
4(x)2
2

x
i(k0 k)x (k0 k)2 (x)2
=
2(x)

2
1
x 2i(k0 k)(x)2 (k0 k)2 (x)2
(2)
=
2
4(x)

Now we plug (??) into (??) to find:

(k) = (2)1/2 C exp[(k0 k)2 (x)2 ]


exp


1
2 2
]
[x

2i(k

k)(x)
dx
0
4(x)2

In the integral weR can make the shift x x 2i(k0 k)(x)2 and use the

standard formula exp(ax2 )dx = (/a)1/2 . The result is


(k) =



2Cx exp (k0 k)2 (x)2

To put this into direct correspondence with the form of the wave packet in
configuration space, we can write



(k0 k)2
(k) = 2Cx exp
4(k)2
where k = 1/(2x). This is the minimum possible k width attainable for
a wave packet with x width x, which is why the Gaussian wave packet is
sometimes referred to as a minimum uncertainty wave packet.
The next step is to compute (x, t) for t > 0. Since we are talking about
a free particle, we know that the momentum eigenfunctions are also energy
eigenfunctions, which makes their time evolution particularly simple to write
down. In the above work we have expressed the initial wave packet (x, 0)
as a linear combination of momentum eigenfunctions, i.e. as a sum of terms
exp(ikx), with the kth term weighted in the sum by the factor (k). The wave
packet at a later time t > 0 will be given by the same linear combination, but
now with the kth term multiplied by a phase factor exp[i(k)t] describing its
time evolution. In symbols we have
Z
(x, t) =
(k)ei[kx(k)t] dk.

For a free particle the frequency and wave number are connected through
(k) =
2

h
k2
.
2m

Using our earlier expression for (k), we find

(x, t) =

2Cx



h
k2
t dk.
exp (k0 k)2 (x)2 + ikx i
2m

(3)

Again we complete the square in the exponent:




h
k2
i
h
(k0 k)2 (x)2 + ikx i
t = [(x)2 +
t]k 2 [2k0 (x)2 + ix]k + k02 (x)2
2m
2m
 2 2

= k k +



2
2
= 2 k 2 2 k + 4 4

4
4
2

= 2 [k 2 ]2 + 2
2
4
where we have defined some shorthand:
2 = (x)2 +

i
h
t
2m

= 2k0 (x)2 + ix

= k02 (x)2 .

Using this in (??) we find:

(x, t) = 2Cx exp

42

Z


2
exp (k 2 ) dk.
2

The integral evaluates to 1/2 /. We have

(x, t) =
=

 
 2

1

2Cx
exp

42
"
#1/2
#
"

(x)2
(ix + 2k0 (x)2 )2
k02 (x)2
2C
e
exp
i
h
i
h
(x)2 + 2m
t
4[(x)2 + 2m
t]

This is pretty ugly, but it does display the relevant features. The important
point is that term i
ht/2m adds to the initial uncertainty (x)2 , so that the
wave packet spreads out with time.
In the figure, Ive plotted this function for a few values of t, with the following parameters: m=940 Mev (corresponding to a proton or neutron), x=3
A,
k0 =0.8
A1 . This value of k0 corresponds, for a neutron, to a velocity of about
5 104 m/s; and note that, sure enough, the center of the wave packet travels
about 5 nm in 100 fs. The time scale of the spread of this wave packet is 100
fs.

Gaussian wave packet example


350
t=0 s
t=30 fs
t=70 fs
t=100 fs

Wavefunction (arbitrary units)

300

250

200

150

100

50

0
-1e-08

-5e-09

0
Distance (meters)

5e-09

1e-08

Problem 2.2
Express the spreading Gaussian wave function (x, t) obtained
in Problem 1 in the form (x, t) = exp[iS(x, t)/
h]. Identify the
function S(x, t) and show that it satisfies the quantum mechanical
Hamilton-Jacobi equation.
In the last problem we found

(x, t)


 2

x

=
2C
exp

42


 

2
x
+ 2
2C
= exp ln

4


(1)

where weve again used the shorthand we defined earlier:


2 = (x)2 +

i
h
t
2m

= 2k0 (x)2 + ix

= k02 (x)2 .

From (??) we can identify (neglecting an unimportant additive constant):




2
h

ln + 2 .
S(x, t) =
i
4
Things are actually easier if we define = 2 . Then


i
h
1
h

2
2
= (x) +
ln +
t
S(x, t) =
2m
i
2
4
Now computing partial derivatives:
S
t

=
=

S
x

=
=

2S
x2



h

1
2
2
i
2 4 t


2
h
2 1
+ 2

2m 2 4
h

i 2 x
h

2
i
h
2

The quantum-mechanical Hamilton-Jacobi equation for a free particle is



2
1 S
i
h 2S
S
+

=0
t
2m x
2m x2
5

(2)

(3)
(4)

Inserting (??), (??), and (??) into this equation, we find


h
2
h
22
h
22
h
2

+
+
=0
4m 8m 2
8m 2
4m
so, sure enough, the equation is satisfied.

Problem 2.3
Consider a wave function that initially is the superposition of two
well-separated narrow wave packets:
1 (x, 0) + 2 (x, 0)
chosen so that the absolute value of the overlap integral
Z +
(0) =
1 (x)2 (x)dx

is very small. As time evolves, the wave packets move and spread.
Will |(t)| increase in time, as the wave packets overlap? Justify
your answer.
It seems to me that the answer to this problem depends entirely on the
specifics of the particular problem.
One could well imagine a situation in which the overlap integral would not
increase with time. Consider, for example, the neutron wave packets plotted
in the figure from problem 2.1 If one of those wave packets were centered in
Chile and another in China, the overlap integral would be tiny, since the wave
packets only have appreciable value within a few angstroms of their centers.
Furthermore, even if the neutrons are initially moving toward each other, their
wave packets spread out on a time scale of 100 fs, long before their centers
ever come close to each other.
On the other hand, if the two neutron wave packets were each centered, say,
20 angstroms apart, then they would certainly overlap a little before collapsing
entirely.

Problem 2.4
A high resolution neutron interferometer narrows the energy spread
of thermal neutrons of 20 meV kinetic energy to a wavelength dispersion level of /= 109 . Estimate the length of the wave
packets in the direction of motion. Over what length of time will
the wave packets spread appreciably?
6

First lets compute the average momentum of the neutrons.


p0

= [ 2mE ]1/2
[ 2 (940Mev c2 ) (20mev) ]1/2
= 6.1 kev / c

Were given the fractional wavelength dispersion level; what does this tell us
about the momentum dispersion level?
p =
dp =
so

dp
=
p

so the momentum uncertainty is

2
h

2
h
d
2

d
= 109

p = p0 109 = 6.1 106 ev.


This implies a position uncertainty of
x

p
6.6 1016 ev s
6.1 106 ev / c

c 1010 s
30 cm.

This is HUGE! So the point is, if we know with this precision how quickly
our thermal neutrons are moving, we have only the most rough indication of
where in the room they might be.
To estimate the time scale of spreading of the wave packets, we can imagine
that they are Gaussian packets. In this case we start to get appreciable spreading
when

or

t (x)2
2m
t 2m(x)2 /
h
2 9.4 108 ev (0.3 m)2

c2 6.6 1016 ev s
32 ks 10 hours.
7

Solutions to Problems in Merzbacher,


Quantum Mechanics, Third Edition
Homer Reid
March 8, 1999

Chapter 3
Problem 3.1
If the state (r) is a superposition,
(r) = c1 1 (r) + c2 2 (r)
where 1 (r) and 2 (r) are related to one another by time reversal,
show that the probability current density can be expressed without
an interference term involving 1 and 2 .
I found this to be a pretty cool problem! First of all, we have the probability
conservation equation:
d
~ J.
~
=
dt
To show that J~ contains no cross terms, it suffices to show that its divergence
has no cross terms, and to show this it suffices (by probability conservation) to
show that d/dt has no cross terms. We have
=
= [c1 1 + c2 2 ] [c1 1 + c2 2 ]

= |c1 ||1 | + |c2 ||2 | + c1 c2 1 2 + c1 c2 1 2

(1)

Problem 3.2
For a free particle
in one
calculate the variance at time

dimension,
t, (x)2t (x hxit )2 t = x2 t hxi2t without explicit use of the
wave function by applying (3.44) repeatedly. Show that


2 1
(px )2 2
2
2
(x)t = (x)0 +
hxpx + px xi0 hxi0 hpx i t +
t
m 2
m2
and
(px )2t = (px )20 = (px ).2

I find it easiest to use a slightly different notation: w(t) (x)2t . (The w


reminds me of width.) Then


dw
1 2 d2 w
+
(2)
t
w(t) = w(0) + t
+
dt t=0
2
dt2 t=0

We have

i
d h
2
2
x hxi
dt
d
d
2
x 2 hxi hxi
=
(3)
dt
dt

2
d2
d2
2
d
d2 w

2
hxi
=
x

2
hxi
hxi
(4)
dt2
dt2
dt
dt2


We need to compute the time derivatives of hxi and x2 . The relevant
equation is


d
1
F
hF i =
hF H HF i +
dt
i
h
t
dw
dt

for any operator F . For a free particle, the Hamiltonian is H = p2 /2m, and
the all-important commutation relation is px = xp i
h. We can use this to
calculate the time derivatives:
d
hxi
dt

=
=
=
=

1
h[x, H]i
i
h

1
2
xp p2 x
2im
h

1
2
xp p(xp i
h)
2im
h

1
2
xp pxp + i
hp
2im
h
2


1
2
xp (xp i
h)p + i
hp
2im
h
1
=
h2i
hpi
2im
h
hpi
(5)
=
m
d2
1 d
hxi =
hpi = 0
(6)
dt2
m dt

d
2
1
2
x
=
[x , H]
dt
i
h

1
2 2
x p p 2 x2
=
2im
h

1
2 2
=
x p p(xp i
h)x
2im
h

1
2 2
=
x p pxpx + i
hpx
2im
h

1
2 2
=
x p (xp i
h)2 + i
h(xp i
h)
2im
h

1
2 2
=
x p xpxp + 2i
hxp + h
2 + h
2 + i
hxp
2im
h

1
2 2
=
x p x(xp i
h)p + 3i
hxp + 2
h2
2im
h

1
2
=
2
h + 4i
hxp
2im
h
i
h
2
= + hxpi
(7)
m m
2 d
d2
2
x
=
hxpi
2
dt
m dt
2
=
h[xp, H]i
i
hm

1
3
=
xp p2 xp
i
hm2

1
3
=
xp p(xp i
h)p
i
hm2

1
3
=
xp pxp2 + i
hp2
i
hm2

1
3
=
xp (xp i
h)p2 + i
hp2
i
hm2

1

=
2i
hp2
i
hm2
2
2
p
(8)
=
m2

d3
2
2
2
x
=
[p , H] = 0
(9)
3
2
dt
i
hm


Now that weve computed all time derivatives of hxi and x2 , its time to
=

plug them into (3) and (4) to compute the time derivatives of w.
dw
dt

=
=
=
=
=

d2 w
dt2

=
=

d
2
d
x 2 hxi hxi
dt
dt
i
h
2
2
+
hxpi
hxi hpi
m
 m
 m
i
h
2
2
+ xp
hxi hpi
m
2
m


2 px xp
2
+ xp
hxi hpi
m
2
m


2
2 px + xp

hxi hpi
m
2
m
2

d2
2
d2
d
hxi
hxi
x

2
hxi
dt2
dt
dt2
2
2
2
2
p 2 hpi2 = 2 (p)2
m2
m
m

(10)

(11)

Finally, we plug these into the original equation (2) to find




2 1
(p)2 2
w(t) = w(0) +
hpx + xpi hxi hpi t +
t.
m 2
m2
The other portion of this problem, the constancy of (p)2 , is trivial, since
(p)2 contains expectation values of p and p2 , which both commute with H.

Problem 3.3
Consider a linear harmonic oscillator with Hamiltonian
H =T +V =

p2
1
+ m 2 x2 .
2m 2

(a) Derive the equation of motion for the expectation value hxi t ,
and show that it oscillates, similarly to the classical oscillator,
as
hpi0
sin t.
hxit = hxi0 cos t +
m
(b) Derive a second-order differential equation of motion for the
expectation value hT V it by repeated application of (3.44)
and use of the virial theorem. Integrate this equation
and,


remembering conservation of energy, calculate x2 t .
(c) Show that


2
(x)2t x2 t hxit

(p)2
= (x)20 cos2 t + 2 20 sin2 t
m


sin 2t
1
+
hxp + pxi0 hxi0 hpi0
2
m

Verify that this reduces to the result of Problem 2 in the limit


0.
(d) Work out the corresponding formula for the variance (p)2t .

(a) Again I like to use slightly different notation: e(t) = hxi t . Then
d
e(t) =
dt
=
=
=
=
=
d2
e(t) =
dt2

1
hxH Hxi
i
h

1
2
xp p2 x
2i
hm

1
2
xp p(xp i
h)
2i
hm

1
2
xp (xp i
h)p + i
hp
2i
hm
1
h2i
hpi
2i
hm
hpi
.
m
d hpi
dt m
5

=
=
=
=
=
=

1
hpH Hpi
i
hm
2

px2 x2 p
2i
h

2

(xp i
h)x x2 p
2i
h

2

x(xp i
h) i
hx x2 p
2i
h
2
h2i
hxi
2i
h
2 hxi .

So we have

d2
e(t) = 2 e(t)
dt2
with general solution e(t) = A cos t + B sin t. The coefficients are determined
by the boundary conditions:
e(0) = hxi0
hpi0
e0 (0) =
m

A = hxi0
hpi0
B=
.
m

(b) Lets define v(t) = hT V it . Then


1
h(T V )H H(T V )i
i
h
1
=
h(T V )(T + V ) (T + V )(T V )i
i
h
2
=
hT V V T i
i
h
2

p 2 x2 x 2 p 2 .
=
2i
h
We already worked out this commutator in Problem 2:

2 2


p x x2 p2 = 4i
hxp + 2
h2
d
v(t)
dt

so

d
v(t) = 2 2 hxpi + i
h 2 .
dt
= 2 2 hxpi + 2 hxp pxi
= 2 hxp + pxi

(12)

Next,
d2
2 2
v(t)
=

hxpH Hxpi
dt2
i
h 

m 2


2 2 1
3
=
xp p2 xp +
xpx2 x3 p
i
h 2m
2
6

(13)

The bracketed expressions are

3

xp p2 xp =

=
=

xpx x p

=
=
=

xp3 p(xp i
h)p


xp3 (xp i
h)p2 + i
hp2


2i
hp2


x(xp i
h)x x3 p

2

x (xp i
h) i
hx2 x3 p


2i
hx2

and plugging these back into (13) gives

d2
v(t) = = 4 2
dt2

#
"

p2
m 2
2

x
m
2

= 4 2 v(t)
with solution

v(t) = A cos 2t + B sin 2t.

(14)

Evaluating at t = 0 gives
A = hT i0 hV i0 .
Also, we can use (12) evaluated at t = 0 to determine B:
2 hxp + pxi0 + i
h 2 = 2B
so

hxp + pxi0
.
B=
2

2
The next task is to compute x t :

x2

=
=

2
hV it
m 2
1
hH (T V )it
m 2
1
[hHit v(t)] .
m 2

Since H does not depend explicitly on time, hHi is constant in time. For v(t)
we can use (14):



2
hxp + pxi0
1
sin
2t
x t =
hT
i
+
hV
i

[hT
i

hV
i
]
cos
2t
+
0
0
0
0
m 2
2


hxp + pxi0
1
2
2
=
2
hT
i
sin
t
+
2
hV
i
cos
t
+
sin
2t
0
0
m 2
2


p 0
1
sin 2t
sin2 t + x2 0 cos2 t + hxp + pxi0
.
(15)
=
m2 2
2
m
7

(c) Earlier we found that


hxit
hxi2t

hpi0
sin t
m
2
hpi
= hxi20 cos2 t + 2 02 sin2 t + hxi0 hpi0 sin 2t.
m
= hxi0 cos t +

Subtracting from (15) gives


h

i


x2 0 hxi20 cos2 t
(x)2t = x2 hxi2 =
i
1 h

+ 2 2 p2 0 hpi20
m


1
hxp + pxi0 hxi0 hpi0 sin 2t
+
2


1
sin 2t
(p)2
= (x)20 cos2 t + 2 20 sin2 t +
hxp + pxi0 hxi0 hpi0
.
m
2
m
As 0, cos2 t 1, (sin2 t/ 2 ) 1, and (sin 2t/) 2, as needed
to ensure matchup with the result of Problem 2.

Problem 3.4
Prove that the probability density and the probability current density at position r0 can be expressed in terms of the operators r and
p as expectation values of the operators
(r0 ) (r r0 )

j(r0 )

1
[p(r r0 ) + (r r0 )p] .
2m

Derive expressions for these densities in the momentum representation.


The first one is trivial:
Z
h(r r0 )i = (r)(r r0 )(r)dr = (r0 )(r0 ) = (r0 ).

For the second one,

1
i
h
hp(r r0 ) + (r r0 )pi =
2m
2m

[ (r r0 ) + (r r0 )] dr

The gradient operator in the first term operates on everything to its right:
Z
i
h
=
[ (r r0 ) + 2(r r0 ) ] dr.
2m
8

Here we can use the identity

f (x) 0 (x a)dx = f 0 (a) :

i
h
|( ) + 2 |r=r0
2m
i
h
=
| |r=r0
2m
= j(r0 ).
=

Problem 3.5
For a system described by the wave function (r0 ), the Wigner
distribution function is defined as
 


1
r00
r00
0
0
0
0
0 00

W (r , p ) =
r +
dr00 .
exp(ip x /
h) r
(2
h)3
2
2
(a) Show that W (r0 , p0 ) is a real-valued function, defined over the
six-dimensional phase space (r0 , p0 ).
(b) Prove that

W (r0 , p0 )dp0 = |(r0 )|2

and that the expectation value of a function of the operator


r in a normalized state is
Z Z
hf (r)i =
f (r0 )W (r0 , p0 )dr0 dp0 .
(c) Show that the Wigner distribution function is normalized as
Z
W (r0 , p0 )dr0 dp0 = 1.
(d) Show that the probability density (r0 ) at position r0 is obtained from the Wigner distribution function with
(r0 ) f (r) = (r r0 ).
(a)

Solutions to Problems in Merzbacher,


Quantum Mechanics, Third Edition
Homer Reid
June 24, 2000

Chapter 5

Problem 5.1
Calculate the matrix elements of p2x with respect to the energy eigenfunctions of the
harmonic oscillator and write down the first few rows and columns of the matrix.
Can the same result be obtained directly by matrix algebra from a knowledge of
the matrix elements of px ?
For the harmonic oscillator, we have
H=

1 2 1
p + m 2 x2
2m x 2

so
p2x = 2mH m2 2 x2
and

1
< n |p2x |k >= 2m
h(n + )nk m2 2 < n |x2 |k > .
2
The nth eigenfunction is
(x) =

1
n
2 n!

1/2 

m 2
m 1/4
exp(
x )Hn (
h

2
h

(1)

m
x).
h

The matrix element of x2 is then


r
r
1/2 

Z
m 1/2 2
m 2
m
m
1
2
x exp(
< n |x |k >=
x )Hn (
x)Hk (
x) dx.
n+k
2
n!k!
h

Homer Reids Solutions to Merzbacher Problems: Chapter 5

The obvious substitution is u = (m/


h)x, with which we obtain
< n |x2 |k >=

1
2n+k n!k!

1/2 

Z

u2 eu Hn (u)Hk (u) du.

(2)

The integral is what Merzbacher calls Inkp with p = 2. The useful formula is
X

n,k,p

Inkp

2
sn tk (2)p
= e +2(s+t)+2st .
n! k! p!





1
1
1
1 + 2 + 4 +
1 + 2(s + t) + (2)2 (s + t)2 +
1 + (2st) + (2st)2 +
2
2
2
(3)
There are two ways to get a 2 term out of this. One way is to take the 2 term
from the first series and the 1 from the second series, together with any term
from the last series. The second way is to take the 1 from the first series and
the 2 term from the second series, along with any term from the last series.
Writing down only terms obtainable in this way, we have




1
= + 2 1 + 2(s + t)2 1 + (2st) + (2st)2 + +
2

X 1

(2st)j +
= + 2 1 + 2s2 + 2t2 + 4st
j!
j=0

 j
j+1
X
2
2 j j 2
2j+1 j j+2 2j+2 j+1 j+1
j+2 j
= +
s t +
s t +
s t
+
s t
+
j!
j!
j!
j!
j=0
=

Comparing termwise with (3), we can read off

, n=k2
(n + 2)!2

n!2n1 (1 + 2n) , n = k
Ink2 =

0
, otherwise.

Plugging this into (2), we have

< n+2 |x2 |n > =


< n |x2 |n > =

1
1/2
[(n + 2)(n + 1)]
2


h

1
(2n + 1)
.
2
m

Finally, from (1),


1
1/2
h)
< n+2 |p2 |n > = [(n + 2)(n + 1)] (m
2
1
< n |x2 |n > =
(2n + 1)(m
h).
2

Homer Reids Solutions to Merzbacher Problems: Chapter 5

I find it kind of confusing that the matrix element for p2 comes out negative
in the first case. It would be absurd for the expectation value (i.e., diagonal
matrix element) of the square of an observable operator to come out negative.
In this case it is less absurd since theres no classical interpretation of the offdiagonal matrix elements of an operator, but its still weird.
However, in another sense it seems inescapable that p2 should have a negative
off-diagonal matrix element here, because the off-diagonal matrix elements of H
must vanish in the energy eigenfunction basis, but x2 has a nonvanishing matrix
element, and H is just a sum of x2 and p2 terms, so p2 must have a negative
matrix element to cancel out the positive matrix element of x2 .

Problem 5.2
Calculate the expectation values of the potential and kinetic energies in any stationary state of the harmonic oscillator. Compare with the results of the virial theorem.

The potential energy operator is U = m 2 x2 /2. We found the expectation


values of x2 in the last problem, so

h
1

1
hU i = m 2 x2 =
(n + )
2
2
2
which is just half the energy expectation value. The kinetic energy expectation
value must of course make up the difference, so we have hT i = hU i.
On the other hand, the virial theorem is supposed to be saying


d
2 hT i = x V (x) .
dx
In this case,

d
V (x) = m 2 x,
dx

so the virial theorem says that


hT i =

1
m 2 x2
2

= hU i

in accord with what we concluded earlier.

Problem 5.3
Calculate the expectation value of x4 for the nth energy eigenstate of the harmonic
oscillator.


n x4 n

r
Z
1  m 1/2 4
m
m 2 2
x exp(
x )Hn (
x) dx
n
n! 2
h

Homer Reids Solutions to Merzbacher Problems: Chapter 5

n! 2n

2 Z

u4 eu Hn2 (u) du

(4)

For the integral we want to use (3) again, but this time well need to write
out the expansion a little further than before.
X

2
sn tk (2)p
= e +2(s+t)+2st .
(5)
Inkp
n! k! p!
n,k,p

=
=
=
=
=


X

1 4
1
1
(2st)j
2
2
4
4
1+ + +
1 + + (2) (s + t) + + (2) (s + t) +
2
2
4!
j!
j=0



 X (2st)j

1
1 + 2 + 4 + 1 + + 42 st + + 44 (st)2 +
2
j!
j=0

 j1

X
2
2j1
2j2
(4 )(st)j
+
+
+4
+4
j!
(j 1)!
(j 2)!
j=0



j
X
1
4
j2
( )(st)
+
+ 2j + j(j 1)
j! 2
j=0



X
2j 1
(4 )(st)j
+ j + j2
+
j! 2
j=0

In the first line, I only wrote out terms that can be combined to give a factor of
4 . In the second line, I further limited it to terms that also contain the same
number of powers of s as t. Equating powers in (5),
1
3
Inn4 = 2n n! ( + n + n2 ),
2
2
so (4) is

3
n x4 n =
2

2

1
( + n + n2 ).
2

Problem 5.4

For the energy eigenstates with n=0, 1, and 2, compute the probability that the
coordinate of a linear harmonic oscillator in its ground state has a value greater
than the amplitude of the classical oscillator of the same energy.
p
The classical amplitude is A = (2E)/(m 2 ). The probability of finding
the particle with coordinate greater than this is
Z A
Z
P (|x| > A) =
2n (x) dx +
2n (x) dx

Homer Reids Solutions to Merzbacher Problems: Chapter 5

= 2

=
=
=

2n (x) dx

r
Z
2  m 1/2
m 2 2
m
exp(
x )Hn (
x) dx
n
n! 2
h

A
Z
2
2

eu Hn2 (u) du
n! 2n 2E/h
Z
2
1

eu Hn2 (u) du

n1
n! 2

2n+1

In going from the first line to the second we invoked the fact that n has either
even or odd parity, so 2n has even parity. In going from the second to last line
to the last line, we noted that the energy of the nth eigenstate is h
(n + 1/2).
In particular,

Pn=0 (|x| > A) =


=
=
=

1
(
Z

eu du

e
0

p2 /2

dp

e
0

p2 /2

dp


r

2 erf( 2)
2
2

1
erf( 2) 0.31
2

Problem 5.5
Show that if an ensemble of linear harmonic oscillators is in thermal equilibrium,
governed by the Boltzmann distribution, the probability per unit length of finding
a particle with displacement x is a Gaussian distribution. Plot the width of the
distribution as a function of temperature. Check the results in the classical and
low-temperature limits. [Hint: Equation (5.43) may be used.]
Suppose we denote the number of oscillators in the nth energy state by Nn .
If the ensemble is in thermal equilibrium, the ratio of the number of oscillators
in the n0 th state to the number of oscillators in the nth state is
0
Nn 0
= e(n n)h/kT .
Nn

In particular, for any n,


Nn = N0 enh/kT .

Homer Reids Solutions to Merzbacher Problems: Chapter 5

The probability of finding a particle between x and dx is


P (x)dx

n=0

= C0

Cn |n (x)|2 dx

 m 1/2
h

exp(

m 2 X enh/kT 2
x )
Hn (
h

n! 2n
n=0

m
x)dx
h

This can be summed using the Mehler formula with t = exp(


h/kT ) :





 m 1/2
m 2
m 2
1
2t
P (x) = C0
x )
x
exp
exp(
h

1+t
h

1 t2



 
 m 1/2 
1
1 t m 2

exp
x
= C0
h

1+t
h

1 t2
This is a Gaussian distribution with variance


h

1+t
2 =
2m 1 t


1 + eh/kT
h

=
2m 1 eh/kT


h

coth
=
2m
2kT

Solutions to Problems in Merzbacher,


Quantum Mechanics, Third Edition
Homer Reid
June 24, 2000

Chapter 6

Problem 6.1
Obtain the transmission coefficient for a rectangular potential barrier of width 2a
if the energy exceeds the height V0 of the barrier. Plot the transmission coefficient
as a function E/V0 (up to E/V0 = 3), choosing (2ma2 V0 )1/2 = (3/2)~.
In the text, Merzbacher treats this problem for the case where the particles
energy is less than the potential barrier. He obtains the result


i
M11 = cosh 2a + sinh 2a e2ika
(1)
2
where
=
and

2m(V0 E)
~2

k
.
(2)
k

We can re-use the result (1) for the case where the energy is greater than
the potential barrier. To do this we note that becomes imaginary in this case,
and we write
r
2m(E V0 )
= i = i
~2
so that (2) becomes


i
k
k

=
i

=i
+
k
i
k

=

Homer Reids Solutions to Merzbacher Problems: Chapter 6

Plugging into (1) and noting that cosh ix = cos x, sinh ix = i sin x we have


i
sin 2a e2ika .
M11 = cos 2a
2
and
1/2
2
2
|M11 | = cos 2a +
sin a
4

 2

1/2

= 1+
1 sin2 a
4

1/2

 4
1
+ k4
2
sin
a

= 1+
4 2 k 2
2

1/2

 2
2 2
( k )
2
sin a
= 1+
4 2 k 2

1/2


V02
2
= 1+
sin a
4E(E V0 )



1/2
1
= 1+
sin2 a
4( 1)

1/2
4( 1) + sin2 a
=
4( 1)


where = E/V0 .
We have
1/2
2m
a
(E

V
)
0
h2

1/2
2mV0 a2
=
( 1)1/2
h2
3
( 1)1/2
=
2

a =

so the transmission coefficient is


"
#
1
4( 1)


T =
=
1/2
|M11 |2
4( 1) + sin2 3
2 ( 1)
This is plotted in Figure 1.

Homer Reids Solutions to Merzbacher Problems: Chapter 6

1
0.9
0.8

Transmission Coefficient

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0

0.5

1.5

2
E/V0

2.5

3.5

Figure 1: Transmission coefficient versus E/V0 for Problem 6.1.

Homer Reids Solutions to Merzbacher Problems: Chapter 6

Problem 6.2
Consider a potential V = 0 for x > a, V = V0 for a x 0, and V = +
for x < 0. Show that for x > a the positive energy solutions of the Schr
odinger
equation have the form
ei(kx+2) eikx

Calculate the scattering coefficient |1 e2i |2 and show that it exhibits maxima
(resonances) at certain discrete energies if the potential is sufficiently deep and
broad.
We have

with

0,

(x) = Aeik1 x + Beik1 x ,

ik2 x
Ce
+ Deik2 x ,

x 0

0 x
a

a
x

r
2m(E + V0 )
2mE
k2 =
.
k1 =
~2
~2
Applying the requirement that be continuous at x = 0, we see we must take
A = B, so (x) = sin k1 x for 0 x a. The other standard requirement,
that the derivative of also be continuous, does not hold at x = 0 because the
potential is infinite there. Hence is undetermined as yet. Eventually, we could
apply the normalization condition on to find if we wanted to.
Next applying continutity of and its derivative at x = a, we obtain
sin k1 a = Ceik2 a + Deik

a
2

k1 cos k1 a = ik2 [Ceik2 a Deik a ]


Combining these yields


1 ik2 a k1
e
cos k1 a + i sin k1 a
2i
k2


1
k1
D = e+ik2 a
cos k1 a i sin k1 a
2i
ik2
C=

(3)
(4)

Its now convenient to write


k1
cos k1 a + i sin k1 a = ei
k2
where
2 =

k1
k2

2

cos2 k1 a + sin2 k1 a

(5)

(6)

Homer Reids Solutions to Merzbacher Problems: Chapter 6

and
= tan

k2
tan k1 a
k1

so that contains magnitude information, while represents phase information.


Then we can rewrite (3) and (4) as
ik2 a i
e
e
2i

D = e+ik2 a ei
2i
C=

Then the expression for the wavefunction to the left of x = a becomes


(x) = Ceik2 x + Deik2 x
(x > a)
i
h ik2 (xa) i
ik2 (xa) i
e
e e
e
=
2i
i
i h ik2 (xa) 2i
e
=
e
e eik2 (xa) .
2i

Using (5) and (6), the scattering coefficient is



2
 2


2
k1
k1
2


cos
k
a
sin
k
a

sin
k
a
cos
k
a
+
2i
1
1
1
1
k2
k2


2i 2
|1 e | = 1

 2


2
k1


cos2 k1 a + sin k1 a
k2

2
h
i


2 sin k1 a sin k1 a i kk1 cos k1 a
2


=  2



k1
cos2 k1 a + sin2 k1 a

k2
We have

=


4 sin2 k1 a

k1
k2

2

2

(7)

cos2 k1 a + sin2 k1 a



E + V0
1
= 1+
E

r
2ma2 (E + V0 )
k1 a =
= ( + 1)1/2
~2

k1
k2

where = E/V0 and = (2ma2 V0 /~2 )1/2 in Merzbachers notation. Then the
scattering coefficient (7) is
scattering coefficient =

(1 +

1 2
)

4 sin2 [( + 1)1/2 ]
cos2 [( + 1)1/2 ] + sin2 [( + 1)1/2 ]

In Figure 6.2 I have plotted this for = 25.

Homer Reids Solutions to Merzbacher Problems: Chapter 6

3.5

Scattering coefficient

2.5

1.5

0.5

0
0

0.5

1.5
E/V0

2.5

Figure 2: Scattering coefficient versus E/V0 for Problem 6.2.

Homer Reids Solutions to Merzbacher Problems: Chapter 6

Problem 6.3
A particle of mass m moves in the one-dimensional double well potential
V (x) = g(x a) g(x + a).
If g > 0, obtain transcendental equations for the bound-state energy eigenvalues of
the system. Compute and plot the energy levels in units of ~2 /ma2 as a function of
the dimensionless parameter mag/~2 . Explain the features of the plot. In the limit
of large separation, 2a, between the wells, obtain a simple formula for the splitting
E between the ground state (even parity) energy level, E+ , and the excited (odd
parity) energy level, E .
In this problem, we can divide the x axis into three regions. In each region,
the wavefunction is just the solution to the free-particle Schr
odinger equation,
but
with
energy
E
<
0
since
were
looking
for
bound
states.
Putting k =
p
2mE/~2 , we have

x
a
Aekx + Bekx ,
Cekx + Dekx ,
a x a
(x) =

Eekx + F ekx ,
a x.

Now, first of all, the wavefunction cant blow up at infinity, so B = E = 0.


Also, since the potential in this problem has mirror-reversal symmetry, the
wavefunction will have definite parity. Considering first the even parity solution,

x
a
Aekx ,
B cosh(kx), a x a
(x) =
(8)

Aekx ,
a x.
Matching the value of the wavefunction at x = a gives
Aeka = B cosh(ka).

(9)

Since the potential becomes infinite at x = a, the normal derivative-continuity


condition doesnt hold there. Instead, we can write down the Schr
odinger equation,
d2
2m
2m
(x) = 2 V (x)(x) 2 E(x),
2
dx
~
~
then integrate from a  to a +  and take the limit as  0. This gives
a+
d
2mg
= 2 (a).
(10)
dx a
~
Applying this condition to the wavefunction (8) yields
kB sinh(ka) kAeka =

2mg
B cosh(ka).
~2

Homer Reids Solutions to Merzbacher Problems: Chapter 6

Substituting from (9),


kB sinh(ka) kB cosh(ka) =

2mg
B cosh(ka)
~2

or
tanh(ka) =

2mg
1=
1
~2 k
ka

with = 2mag/~2 . This equation determines the energy eigenvalue of the evenparity state, which will be the ground state. On the other hand, the odd parity
state looks like

x
a
Aekx ,
B sinh(kx), a x a
(11)
(x) =

Aekx ,
a x.
Matching values at x = a gives

Aeka = B sinh(ka)
and applying condition (10) gives
2mg
B sinh(ka)
~2
2mg
kB cosh(ka) + kB sinh(ka) = 2 B sinh(ka)
~

1
coth(ka) =
ka
kB cosh(ka) kAeka =

so this is the condition that determines the energy of the odd parity state.
In Figure (3) I have plotted tanh(ka), coth(ka), and /(ka) 1 for the case
= 3. As expected, the coth curve crosses the /(ka) 1 curve at a lower value
of ka than the tanh curve; that means that the energy eigenvalue for the odd
parity state is smaller in magnitude (less negative) than the even parity state.

Problem 6.4
Problem 3 provides a primitive model for a one-electron linear diatomic molecule
with interatomic distance 2a = |X|, if the potential energy of the molecule is
taken as E (|X|), supplemented by a repulsive interaction g/|X| between the wells
(atoms). Show that, for a sufficiently small value of , the system (molecule)
is stable if the particle (electron) is in the even parity state. Sketch the total
potential energy of the system as a function of |X|.

Homer Reids Solutions to Merzbacher Problems: Chapter 6

PSfrag replacements

2
tanh(ka)
coth(ka)
/(ka) 1

1.5

0.5

0
1

1.2

1.4

1.6

1.8

ka

Figure 3: Graphical determination of energy levels for Problem 6.3 with = 3.

10

Homer Reids Solutions to Merzbacher Problems: Chapter 6

Problem 6.5
If the potential in Problem 3 has g < 0 (double barrier), calculate the transmission
coefficient and show that it exhibits resonances. (Note the analogy between the
system and the Fabry-Perot etalon in optics.)
Now were assuming that the energy E is positive, so
ikx
ikx
x a

Ae + Be
ikx
ikx
(x) = Ce + De
axa

ikx
ikx
Ee + F e
ax

with k =

p
2mE/~2 . Matching values at x = a, we have
Aeika + Beika = C ika + Dika

(12)

(13)

Also, as before, we have the derivative condition


x=a+
d
2mg
= 2 (a)
dx x=a
~

where g is now negative. Applying this to the wavefunction in (12), we have


ik[Ceika Deika Aeika + Beika ] =

2mg
[Aeika + Beika ].
~2

Combining (13) and (14) yields




2ka

e B
A+
C = 1+
ika
ika


2ka

D=
B
e
A+ 1
ika
ika

(14)

(15)
(16)

with = mag/~2 as before.


Now, applying the matching conditions to the wavefunction at x = +a will
give two equations exactly like (13) and (14), but with the substitutions A C,
B D, C E, D F , and a a. Making these substitutions in (15) and
(16) we obtain



2ka
e D
(17)
E = 1
C
ika
ika


2ka

F =+
D
(18)
e C + 1+
ika
ika

PSfrag replacements

11

Homer Reids Solutions to Merzbacher Problems: Chapter 6

1
0.9
0.8
0.7

T (k)

0.6
0.5
0.4
0.3
0.2
0.1
0
0

10

15

20

25

ka

Figure 4: Transmission coefficient in Problem 6.4 with = 15.

Combining equations (15) through (18), we have


#
"


2



2

4ika
(e
1) A + +
E = 1+
1
sin 2ka B
ika
ka
ika
#
"





2
2

4ika
(e
1) B
F =
1+
sin 2ka A + 1 +
ka
ika
ika
This is the M matrix, and the transmission coefficient is given by T = 1/|M11 |2 ,
or
1

T =
 2  2

+ 1 (1 cos(4ka))
1 + 2 ka
ka

In Figure 4 I have plotted this for = 15.

30

Homer Reids Solutions to Merzbacher Problems: Chapter 6

12

Problem 6.6
A particle moves in one dimension with energy E in the field of a potential defined
as the sum of a Heaviside step function and a delta function:
V (x) = V0 (x) + g(x)

(with V0 and g > 0)

The particle is assumed to have energy E > V0 .


(a) Work out the matrix M , which relates the amplitudes of the incident and
reflected plane waves on the left of the origin (x < 0) to the amplitudes on
the right (x > 0).
(b) Derive the elements of the matrix S, which relates incoming and outgoing
amplitudes.
(c) Show that the S matrix is unitary and that the elements of the S matrix satisfy
the properties expected from the applicable symmetry considerations.
(d) Calculate the transmission coefficient for particles incident from the right and
for particles incident from the left, which have the same energy (buf different
velocities).
We have
(x) =
with

Aeik1 x + Beik1 x ,
Ce

ik2 x

2m
E
k1 =
~2
Matching values at x = 0 gives

+ De

k2 =

ik2 x

x0

x0

2m
(E V0 ).
~2

C +D =A+B

(19)

Also, the delta function at the origin gives rise to a discontinuity in the derivative
of the wavefunction as before:
0+
d
2mg
= 2 (0)
dx 0
~

so

ik2 (C D) ik1 (A B) =

or
C D =

2mg
(A + B)
~2

k1
2mg
(A B) +
(A + B).
k2
ik2 ~2

(20)

Homer Reids Solutions to Merzbacher Problems: Chapter 6

13

Adding and subtracting (19) and (20), we can read off






1
2mg
2mg
1
k1
k1
A+
B
+
+
C=
1+
1
2
k2
ik2 ~2
2
k2
ik2 ~2




2mg
2mg
1
1
k1
k1

D=
A 1+
B.
1
2
k2
ik2 ~2
2
k2
ik2 ~2
We could also write this as

 
C
M11
=

D
M12

M12

M11



A
B

Or instead of the M matrix we could use the S matrix, which is defined by




 

A
S11 S12
B
=
S21 S22
D
C
Since we already know the M coefficients, we can calculate the elements of the
S matrix from the formula
S11
S21

= M12

11
1
= M
11

S12 = M1
11
M12
S22 = + M

11

However, this is tedious and long and boring and I dont want to do it.

Solutions to Problems in Merzbacher,


Quantum Mechanics, Third Edition
Homer Reid
April 5, 2001

Chapter 7
Before starting on these problems I found it useful to review how the WKB
approximation works in the first place. The Schr
odinger equation is

~2 d 2
(x) + V (x)(x) = E(x)
2m dx2

or

d2
(x) + k 2 (x)(x) = 0,
dx2
We postulate for the functional form

k(x)

2m
[E V (x)].
~2

(x) = AeiS(x)/~
in which case the Schr
odinger equation becomes
i~S 00 (x) = [S 0 (x)]2 ~2 k 2 (x).

(1)

This equation cant be solved directly, but we obtain guidance from the observation that, for a constant potential, S(x) = kx, so that S 00 vanishes. For a
nonconstant but slowly varying potential we might imagine S 00 (x) will be small,
and we may take S 00 = 0 as the seed of a series of successive approximations
to the exact solution. To be specific, we will construct a series of functions
S0 (x), S1 (x), , where S0 is the solution of (1) with 0 on the left hand side;
S1 is a solution with S000 on the left hand side; and so on. In other words, at
the nth step in the approximation sequence (by which point we have computed
Sn (x)), we compute Sn00 (x) and use that as the source term on the LHS of (1)
to calculate Sn+1 (x). Then we compute the second derivative of Sn+1 (x) and
use this as the source term for calculating Sn+2 , and so on ad infinitum. In

Homer Reids Solutions to Merzbacher Problems: Chapter 7

symbols,
0 = [S00 (x)]2 ~2 k 2 (x)

i~S000
i~S100

[S10 (x)]2
[S20 (x)]2

(2)

2 2

(3)

2 2

(4)

~ k (x)

~ k (x)

Equation (2) is clearly solved by taking


S00 (x)

= ~k(x)

S0 (x) = S00 ~

k(x0 )dx0

(5)

for any constant S00 . Then S000 (x) = ~k 0 (x), so (3) is


p
S10 (x) = ~ k 2 (x) ik 0 (x).

With the two signs here, we appear to have four possible choices for S10 . But
lets think a little about the signs in this equation. The sign under the
radical comes from the two choices of sign in (5). But if we chose, say, the plus
sign in that equation, so that S00 > 0, we would also expect that S10 > 0. Indeed,
if we choose the plus sign in (5) but the minus sign in (3), then S00 and S10 have
opposite sign, so S10 differs from S00 by an amount at least as large as S00 , in
which case our approximation sequence S0 , S1 , has little hope of converging.
So we choose either both plus signs or both minus signs in (3), whence our two
choices are
p
p
S10 = +~ k 2 (x) + ik 0 (x) or S10 = ~ k 2 (x) ik 0 (x).
(6)

If V (x) is constant, k(x) is constant, and, as we observed before, the sequence


of approximations terminates at 0th order with S0 being an exact solution. By
extension, if V (x) is not constant but changes little over one particle wavelength,
we have k 0 (x)/k 2 (x)  1, so we may expand the radicals in (6):




ik 0 (x)
ik 0 (x)
S10 ~k(x) 1 + 2
or S10 ~k(x) 1 2
2k (x)
2k (x)
or
i~k 0 (x)
S10 ~k(x) +
.
(7)
2k(x)
Integrating,
Z
i~ x k 0 (u) 0
dx
2 a k(u)
Za x
i~ k(x)
= S1 (a) ~
k(u)du +
ln
2
k(a)
a

S1 (x) = S1 (a) ~

k(u)du +

where a is some point chosen such that the approximation (7) is valid in the
full range a < x0 < x. We could go on to compute S2 , S3 , etc., but in practice
it seems the approximation is always terminated at S1 .

Homer Reids Solutions to Merzbacher Problems: Chapter 7

The wavefunction at this order of approximation is


1/2



Rx
(x) = exp(iS1 (x)/~) = eiS1 (a)/~ ei a k(u)du eln k(x)/k(a)
s
k(a) i R x k(u)du
= (a)
e a
k(x)
= (a)G (x; a)
where
(8)
G (x; a)

k(a) i R x k(u)du
e a
.
k(x)

(9)

We have written it this way to illustrate that the function G(x, a) is kind of like
a Greens function or propagator for the wavefunction, in the sense that, if you
know what is at some point a, you can just multiply it by G (x; a) to find
out what is at x. But this doesnt seem quite right: Schr
odingers equation
is a second-order differential equation, but (8) seems to be saying that we need
only one initial conditionthe value of at x = ato find the value of at
other points. To clarify this subtle point, lets investigate the equations leading
up to (8). If the approximation (7) makes sense, then there are two solutions
of Schr
odingers equation at x = a, one whose phase increases with increasing
x (positive derivative), and one whose phase decreases. Equation (8) seems to
be saying that we can use either G+ or G to get to (x) from (a); but the
requirement the d/dx be continuous at x = a means that only one or the
other will do. Indeed, in using (8) to continue from a to x we must choose
the appropriate propagatoreither G+ or G , according to the derivative of
at x = a; otherwise the overall wave function will have a discontinuity in its
first derivative at x = a. So to use (8) to obtain values for at a point x, we
need to know both and 0 at a nearby point x = a, as should be the case for
a second-order differential equation.
If we want to de-emphasize this nature of the solution with the propagator
we may write
s
1 i R x k(u)du
(x) = C
e a
(10)
k(x)
p
where C = (a) k(a). In regions where V (x) > E, k(x) is imaginary, so its
useful to define
r
2m
(x) = ik(x) =
[V (x) E]
(11)
~2
and
s
1 R x (u)du
e a
.
(12)
(x) = C
(x)

Homer Reids Solutions to Merzbacher Problems: Chapter 7

If we have a region of space in which the WKB approximation is valid,


knowing the value of (and its derivative) at one point within the region is
equivalent to knowing it everywhere, because we can use the propagator (9)
to get from that one point to every other point within the region. The WKB
method, however, gives us no way of determining the value of at that one
starting point. Furthermore, even if we know at one point within a region of
validity, we cant use (8) to determine in other, nonadjacent regions, because
we cant carry the propagator across regions of invalidity. So basically what we
need is a way of finding one starting value for (x) in every region of validity
of the WKB approximation.
How do we find such points? Well, one sure-fire way to get starting points
in regions of validity is to identify regions of invalidity, of which there will be
at least one adjacent to each region of validity, and then get values of at the
boundaries of the regions of invaliditywhich will also count as values in the
regions of validity. So we need to identify the regions of invalidity of the WKB
approximation and do a more accurate solution of the Schr
odinger equation
there.
The WKB approximation breaks down when k 0 /k 2  1 ceases to hold, which
is true when k 0 but k 0 6= 0, which happens near a classical turning point of
the motioni.e., a point x0 at which V (x0 ) = E. But near such a point we may
expand V (x) E in a Taylor series around the point x0 ; if we keep only the
first (linear in x) term in the series, we arrive at a Schr
odinger equation which
we can solve exactly in the vicinity of x0 . To do this, suppose the point x0 is a
classical turning point of the motion, so that V (x0 ) = E. In the neighborhood
of x0 we may expand V (x):
V (x) = E + (x x0 )V 0 (x0 ) +

(13)

Then the Schr


odinger equation becomes
d2
2m
(x) 2 V 0 (x0 )(x x0 )(x) = 0.
2
dx
~
The useful substitution here is

1/3
2m 0
u(x) = (x x0 )

V (x0 )
~2

so

x(u) =

u
+ x0 .

If we define
(u) = (x(u))
then
d
d dx
1
=
= 0 (x(u))
du
dx du

1 00
d2
= 2 (x(u))
du2

(14)

Homer Reids Solutions to Merzbacher Problems: Chapter 7

so (14) becomes
2

d2
(u) 3 (x x0 )(u) = 0
du2

or

d2
(u) u(u) = 0.
du2
The solution to this differential equation is
(u) = 1 Ai(u) + 2 Bi(u)

(15)

so the solution to the Schr


odinger equation (14) is




(x) = 1 Ai (x x0 ) + 2 Bi (x x0 ) .

For (x x0 )  1 we have the asymptotic expression




3/2
2
1/4 1 23 |(xx0 )|3/2
(x) 1/2 [(x x0 )]
e
+ 2 e+ 3 |(xx0 )|
2

(16)

and for (x x0 )  1 we have


(x) 1/2 |(x x0 )|

1/4

2


3/2
|(x x0 )|
3
4
2
i
3/2
2 sin
.
|(x x0 )|
3
4

1 cos

(17)

To simplify these, we need to consider two possible kinds of turning point.


Case 1: V 0 (x0 ) > 0.
In this case the potential is increasing through the turning point at x0 , which
means that V (x) < E for x < x0 , and V (x) > E for x > x0 . Hence the region
to the left of the turning point is the classically accessible region, while the right
of the turning point is classically forbidden. Since V 0 (x0 ) > 0, > 0, so for
x < x0 (??) holds. For points close to the turning point on the left side,
r

1/2
2m 0
2m
[E

V
(x)]

V
k(x) =
(x0 x)1/2 = 3/2 (x0 x)1/2
~2
~2
so
|(x x0 )|

1/4

and
Z

x0

k(u)du =
x

3/2

k(x)

x0
x

(x0 x)1/2 du =

2
= |(x x0 )|3/2 .
3

(18)

2 3/2
(x0 x)3/2
3
(19)

On the other hand, for points close to the turning point on the left side we
have x > x0 , so (x x0 ) > 0. In this region,

Homer Reids Solutions to Merzbacher Problems: Chapter 7

(x) =

1/2


2m 
2m 0
= 3/2 (x x0 )1/2
V (x0 )(x x0 )
V (x) E
~2
~2

so, for x near x0 ,


Z x
Z x
2
(u)du = 3/2
(u x0 )1/2 du = 3/2 (x x0 )3/2
3
x0
x0
and also
|(x x0 )|

1/4

.
(x)

(20)

(21)

(22)

Using (18) and (19) in (17), and (21) and (22) in (16), the solutions to the
Schr
odinger equation on either side of a classical turning point x0 at which
V 0 (x0 ) > 0 are
s
 Z x0
1 h

k(u)du
21 cos
(x) =
k(x)
4
x
 Z x0
i
k(u)du
2 sin
,
x < x0
(23)
4
x
s
i
R
1 h Rxx (u)du
x (u)du
(x) =
1 e 0
+ 2 e x0
,
x < x0
(24)
(x)
(we redefined the constants slightly in going to this equation).
Case 2: V 0 (x0 ) < 0.
In this case the potential is decreasing through the turning point, so the
classically accessible region is to the right of the turning point, and the forbidden
region to the left. Since V 0 (x0 ) < 0, < 0. That means that the regions of
applicability of (16) and (17) are on opposite sides of the turning points as they
were in the previous case. The solutions to the Schr
odinger equation on either
side of the turning point are
s
i
R
1 h R x0 (u)du
x (u)du
1 e x
+ 2 e x0
,
x < x0
(25)
(x) =
(x)
s
Z x

1 h
k(u)du
(x) =
21 cos
k(x)
4
x0
Z x
i
2 sin
,
x > x0
(26)
k(u)du
4
x0
(27)

So, to apply the WKB approximation to a given potential V (x), the first
step is to identify the classical turning points of the motion, and to divide space

Homer Reids Solutions to Merzbacher Problems: Chapter 7

up into regions bounded by turning points, within which regions the WKB
approximation (7) is valid. Then, for each turning point, we write down (23)
and (24) (or (25) and (26)) at nearby points on either side of the turning point,
and then use (10) to evolve the wavefunction from those points to other points
within the separate regions.
We should probably quantify the meaning of nearby in that last sentence.
Suppose x0 is a classical turning point of the motion, and we are looking for
points x0  at which to make the handoff from approximations (16) and
(17) to the WKB approximation These points must satisfy several conditions.
First, the approximate Schr
odinger equation (14) is only valid as long as we can
neglect the quadratic and higher-order terms in the expansion (13), so we must
have

0
V (x0 )
.
(28)
 |V 00 (x0 )|  |V 0 (x0 )|   00
V (x0 )
But at the same time,  must be sufficiently greater than 1 to justify the
approximation (16) (or sufficiently less than -1 to justify (17)); the condition
here is


1/3
1/3
2m 0
2m 0








1



.
(29)
V
(x
)
V
(x
)
0
0
~2
~2
Finally, the points x must be sufficiently far away from the turning points that
the approximation (7) is valid for the derivative of the phase of the wavefunction;
the condition for this to be the case was

0

 2 

0
k (x)
~
V
(x

)
1



1
(30)

 1.
k 2 (x)
2 2m [E V (x )]3/2

If there are no points x0  satisfying all three conditions, the WKB approximation cannot be used.

To apply all of this to the problem of bound states in a potential well,


consider a potential like that shown in Figure 1, with two classical turning points
at x = a and x = b. Although there are no discontinuities in the potential
here, the problem may be analyzed in a manner similar to that used in the
consideration of one-dimensional piecewise constant potentials, as in Chapter
6: we divide space into a number of distinct regions, obtain solutions of the
Schr
odinger equation in each region, and then match values and derivatives at
the region boundaries.
To divide space into distinct regions in this case, we begin by identifying
narrow regions around the turning points a and b in which the linear approximation (13) is valid. In the narrow region around x = a, we may use (25)
and (26); around x = b we may use (23) and (24). Let the narrow such region
around a be a 1 < x < a + 1, and that around b be b 2 < x < b + 2. Then
space divides naturally into five regions: (a) x < a 

In this region we are far enough to the left of the turning point that the WKB
approximation is valid, and the wavefunction takes the form (10). However, we

Homer Reids Solutions to Merzbacher Problems: Chapter 7

PSfrag replacements
V (x)
E

Figure 1: A potential V (x) with two classical turning points for an energy E.

Homer Reids Solutions to Merzbacher Problems: Chapter 7

must throw out the term that grows exponentially as x , so we are left
with
s
1 R (a) (u)du
e x
,
x < a .
(31)
(x) = A
(x)
(b) a  < x < a + 
In this region we are close enough to the turning point that (13) is valid, so
(25) and (26) may be used.
s

Ra
1  R a (u)du
(x) =
1 e x
+ 2 e+ x (u)du ,
x < (a ).
(32)
(x)
From (31) and (32) we see that continuity of both the value and first derivative of
(x) at x = a  requires taking 1 = A, 2 = 0. With this choice of constants,
we achive continuity not only of the value and first derivative of but also of
all higher derivatives, as must be the case since there is no discontinuity in the
potential.
But now that we know the value of at x = a , we also know it at
x = a + , because of course the solution of the Schr
odinger equation in the
narrow strip around a (to which (32) is an asymptotic approximation for x < a)
is valid throughout the strip; the same solution thats valid at x = a  is valid
at x = a + . With 1 = A and 2 = 0, (26) becomes

(x) = 2A
=A


1
cos
k(x)

x
a

k(u)du


4

i
Rx
1 h +i(R x k(u)du/4)
a
+ ei( a k(u)du/4) ,
e
k(x)

x = (a + )
(33)

(c) a +  < x < b 


In this region the WKB approximation (7) is valid, so we may use (8) to find
the wavefunction at any point within the region. Using the expression (33) for
the wavefunction at x = a + , integrating from a +  to x in the propagator (9),
and using G+ and G , respectively, to propagate the first and second terms in

Homer Reids Solutions to Merzbacher Problems: Chapter 7

10

(33), we obtain for the wavefunction at a point x in this region


s
 R
R


Rx
Rx
1
i a(a+) k(u)du+ (a+)
/4
+i a(a+) k(u)du+ (a+)
k(u)du/4
+e
(x) = A
e
k(x)
s
i
Rx
1 h +i(R x k(u)du/4)
a
=A
e
+ ei( a k(u)du/4)
k(x)
s

Z x
1

= 2A
cos
k(u)du
k(x)
4
a
s
!
Z b
Z b
1

,
a+<x <b
cos
= 2A
k(u)du
k(u)du
k(x)
4
a
x
(34)

Okay, I have now carried this analysis far enough to see for myself exactly
where the Bohr-Sommerfeld quantization condition


Z b
1
k(u)du = n +
,
n = 1, 2,
(35)
2
a
comes from, which was my original goal, so I am now going to stop this exercise
and proceed directly to the problems.

Problem 7.1
Apply the WKB method to a particle that falls with acceleration g in a uniform
gravitational field directed along the z axis and that is reflected from a perfectly
elastic plane surface at z = 0. Compare with the rigorous solutions of this problem.
Well start with the exact solution to the problem. The requirement of
perfect elastic reflection at z = 0 may be imposed by taking V (x) to jump
suddenly to infinity at z = 0, i.e.
(
mgz, z > 0
V (x) =
,
z 0.
For z > 0, the Schr
odinger equation is
d2
(x) +
dx2
d2
= 2 (x) +
dx
d2
= 2 (x)
dx

0=

2m
[E mgz] (x)
~2 

2m2 g E

z
(x)
~2
mg
2m2 g
[z z0 ] (x)
~2

(36)

11

Homer Reids Solutions to Merzbacher Problems: Chapter 7

where z0 = E/mg. With the substitution


u = (z z0 )

2m2 g
~2

1/3

and taking (u) = (x(u)), we find that (36) is just the Airy equation for (u),
d2
(u) u(u) = 0
du2
with solutions
(u) = 1 Ai(u) + 2 Bi(u).
Since we require a solution that remains finite as z , we must take 2 = 0.
The solution to (36) is then

(x) = 1 Ai (z z0 ) ,
(z > 0).
(37)

For z < 0, I wasnt quite sure how to account for the infinite potential
jump at z = 0, so instead I supposed the potential for z < 0 to be a constant,
V (z) = V0 , where eventually Ill take V0 . Then the Schr
odinger equation
for z < 0 is
d2
2m
(z) 2 [V0 E] (z) = 0
2
dz
~
with solution
r
2m
kz
[V0 E].
(38)
(z) = Ae
,
k=
~2
Matching values and derivatives of (37) and (38) at z = 0, we have
1 Ai(z0 ) = A
1 Ai0 (z0 ) = kA
Dividing, we obtain
1 Ai(z0 )
1
=
Ai0 (z0 )
k
Now taking V0 , we also have k , so the RHS of this goes to zero;
thus the condition is that z0 be a zero of the Airy function, which means the
energy eigenvalues En are given by


2m2 g
~2

1/3

En
= xnm En =
mg

where xn is the nth root of the equation


Ai(xn ) = 0.

mg 2 ~2
2

1/3

xn

(39)

Homer Reids Solutions to Merzbacher Problems: Chapter 7

12

So thats the exact solution. In the WKB approximation, the spectrum of


energy eigenvalues is determined by the condition (35). In this case the classical
turning points are at z = 0 and z = z0 , so we have


Z z0
1
n+
=
k(z) dz
2
0
r
Z
2m z0
=
[E mgz]1/2 dz
~2 0
r
Z
2m2 g z0
[z0 z]1/2 du
=
~2
0
r

z0

2m2 g 2
3/2
(z0 z)
=

2
~
3
0
r

3/2
2
2 2m g E
=
3
~2
mg
so the nth eigenvalue is given by

En =

mg 2 ~2
2

1/3  
 2/3
1
3
n+

.
2
2

Solutions to Problems in Merzbacher,


Quantum Mechanics, Third Edition
Homer Reid
May 13, 2001

Chapter 8

Homer Reids Solutions to Merzbacher Problems: Chapter 8

Problem 8.1
Apply the variational method to estimate the ground state energy of a particle
confined in a one-dimensional box for which V = 0 for a < x < a, and (a) = 0.
(a) First, use an unnormalized trapezoidal trial function which vanishes at a and
is symmetric with respect to the center of the well:
(
(a |x|),
bxa
t (x) =
(a b),
|x| b.
(b) A more sophisticated trial function is parabolic, again vanishing at the end
points and even in x.
(c) Use a quartic trial function of the form
t (x) = (a2 x2 )(x2 + ),
where the ratio of the adjustable parameters and is determined variationally.
(d) Compare the results of the different variational calculations with the exact
ground state energy,
R a and, using normalized wave functions, evaluate the mean
square deviation a |(x) t (x)|2 dx for the various cases.

(e) Show that the variational procedure produces, in addition to the approximation
to the ground state, an optimal quartic trial function with nodes between the
endpoints. Interpret the corresponding stationary energy value.

First lets observe that the exact expressions for the ground state wavefunction
and energy are
1
n (x) = cos(kn x),
a

kn =

n
,
2a

En = n 2

~2
~2 2

1.23
.
8ma2
ma2

(a) We need first to normalize the trial wavefunction. Taking

t (x) =

(a |x|),
(a b),

b |x| a
x| b.

Homer Reids Solutions to Merzbacher Problems: Chapter 8

we have
Z

2t (x)dx = 2

= 2

2t (x)dx
0
(
Z

(a b)

dx +
0

a
b

(a x) dx


1
3
= 2 b(a b) + (a b)
3


1 3
3
2
2
2
2
2
= 2 b(a + b 2ab) + (a b ) a b + b a
3

2
= 2 a3 + 2b3 3ab2
3
2

so t is normalized by taking

2 =

3
2

1
a3 + 2b3 3ab2

(1)

Now we can compute the energy expectation value of t :


Z a
~2
d2
< t |H|t > =
t (x) 2 t (x) dx
2m a
dx
Integrating by parts,
=

~2
2m

Z
a

t (x)0t (x)
a

02
t (x) dx
a

(the first integral vanishes since t vanishes at the endpoints)


=+

~2
m

a
0

02
t (x) dx

Z
~2 2 a

dx
m
b
~2 2
(b a).
=
m
=

Using (1), this is


< H >=

3~2
2m

(b a)
a3 + 2b3 3ab2

To find the optimal value of b, we zero the derivative of this with respect to b:
0=

(a3

1
6b2 (b a)
6ab(b a)
3
+ 3
3
2
3
2
2
+ 2b 3ab ) (a + 2b 3ab )
(a + 2b3 3ab2 )2

= 4b3 + 9b2 a 6a2 b + a3

Homer Reids Solutions to Merzbacher Problems: Chapter 8

(b) For a parabolic trial function we take


t (x) = (a2 x2 ).
The normalization integral is
Z a
Z a
(a2 x2 )2 dx
2t (x) dx = 2 2
0
a
Z a
2
(a4 + x4 2a2 x2 )dx
= 2
0

1
2
2
= 2 a5 + a5 a5
5
3
16 2 5
a
=
15
so t (x) is normalized by taking
2 =

15
.
16a5

The expectation value of the energy is


Z a
d2
~2
t (x) 2 t (x) dx
<H >=
2m a
dx
Z a
2
~
= 2 2
(a2 x2 )dx
m
0
4 ~2 2 3
a
=
3m
~2
5 ~2

1.25
.
=
4 ma2
ma2
So this is in good agreement with the exact ground state energy.
(c) In this case we have
t (x) = (a2 x2 )(x2 + )

= [x4 + (a2 )x2 + a2 ]

The kinetic energy is

~2
~2 d 2
t (x) = [6x2 (a2 )].
2
2m dx
m

The expectation value of the energy is

Homer Reids Solutions to Merzbacher Problems: Chapter 8

Problem 8.2
Using scaled variables, as in Section 5.1, consider the anharmonic oscillator Hamiltonian,
1
1
H = p2 + 2 + 4
2
2
where is a real-valued parameter.
(a) Estimate the ground state energy by a variational calculation, using as a trial
function the ground state wave function for the harmonic oscillator
H0 () =

1 2 1 2 2
p +
2 2

where is an adjustable variational parameter. Derive an equation that


relates and .
(b) Compute the variational estimate of the ground state energy of H for various
positive values of the strength .
(c) Note that the method yields answers for a discrete energy eigenstate even if
is slightly negative. Draw the potential energy curve to judge if this result
makes physical sense. Explain.

(a) To find the ground state eigenfunction of the Hamiltonian Merzbacher proposes, its convenient to rewrite it:
1 2 1 2 2
p +
2 2
1
1 2
+ 2 2
=
2
2
2

H0 () =

Upon substituting u = 1/2 we obtain


1 2
1
=
+ u2
2 u2 2


and now this is just the ordinary harmonic oscillator Hamiltonian, scaled by a
constant factor , with ground-state eigenfunction
() = Ceu

/2

= Ce

/2

Adding the normalization constant,


() =

 1/4

/2

Homer Reids Solutions to Merzbacher Problems: Chapter 8

Now we want to treat as a parameter and vary it until the energy expectation
value of () is minimized. The energy expectation value is
h| H |i = h| T |i + h| V |i
where T = p2 /2 and V = 2 /2 + 4 . Lets compute the two expectation values
separately. First of all, to compute the expectation value of T , we need to know
the result of operating on () with p2 :
p2 ()


 1/4  
2 /2
=
e



i
 1/4 h
2 /2
e
=

 1/4 

2
+ 2 2 e /2
=

Then for the expectation value of T we have


Z
1
()p2 ()d
h| T |i =
2
r Z

2
1
=
+ 2 2 e d
2
r  r
r 
1

=
+ 2
2
2
3

= .
4

(2)

On the other hand, for the expectation value of V we have


(3)
r

 Z

Z
1 2 2
4 2
exptwoV =
e
d +
e
d
2

r  r
r 
3
1

=
+
3
4
4
5
1
3
=
+
.
4 4 2

(4)

Adding (2) and (4),


exptwoH =



1
3
1
+ + 2 .
2

To minimize this with respect to we equate its derivative to 0:


0=1

1
6
3
2

(5)

Homer Reids Solutions to Merzbacher Problems: Chapter 8

or
3 6 = 0.

(6)

We could then solve this equation for in terms of to obtain the energyminimizing value of for a given perturbing potential strength . But writing
down the full solution would be tedious. Instead lets see what happens when
is small.
Evidently, when = 0 the Hamiltonian in this problem degenerates to the
normal harmonic oscillator Hamiltonian, for which the energy is minimized by
the (unscaled) ground state harmonic oscillator wavefunction, i.e. () with
= 1. We can thus imagine that, for small , the energy-minimizing value of
will be close to 1, and we may write () 1 +  for some small . Inserting
this in (6),
(1 + 3 + 32 + 3 ) (1 + ) = 6
Keeping only terms of zeroth or first order in the small quantity  (which is
equivalent to keeping terms of lowest order in the perturbing potential strength
) we obtain from this
 3,
so for  0 the minimizing value of is
1 + 3.
Inserting this estimate into (5) and again keeping only terms of lowest order in
we find
(7)
exptwoH =


1
(1 + ) + (1 + )1 + 3(1 + )2
4

1
(1 + 3) + (1 + 3)1 + 3(1 + 3)2
4
1
[(1 + 3) + (1 3) + 3(1 6)]
4
1 3
+ .
2 4

(8)

Since the 1/2 term is the normal (unperturbed) energy of the state, the energy
shift caused by the perturbing potential is
E =

3
.
4

(9)

Homer Reids Solutions to Merzbacher Problems: Chapter 8

Problem 8.3
In first-order perturbation theory, calculate the change in the energy levels of a
linear harmonic oscillator that is perturbed by a potential gx4 . For small values of
the coefficient, compare the result with the variational calculation in Problem 2.
The energy shift to first order is
E = exptwon (x)|gx4 |n (x) = g hn | x4 |n i .
I worked out this expectation value in Problem 5.3:
3g
E = gexptwon x n =
2
4

~
m

2 


1
+ n + n2
2

In particular, the energy shift of the ground state is


3g
E0 =
4

~
m

2

which agrees with () (the difference in the factor (~/m)2 just represents the
fact that in Problem 8.2 we used scaled variables, whereas in this problem we
inserted the units explicitly).

Problem 8.4
2

Using a Gaussian trial function, ex , with an adjustable parameter, make a variational estimate of the ground state energy for a particle in a Gaussian potential
well, represented by the Hamiltonian
H=

2
p2
V0 ex
2m

(V0 > 0, > 0).

For notational simplicity, I like to put /2 = . Then


(x) = Cex

/2

and the normalization constant is determined by


1=C

x2

dx

 1/4

C=
.

Homer Reids Solutions to Merzbacher Problems: Chapter 8

The kinetic energy operator operating on this state yields





p2
~2
T =
(x) =
(x)
2m
2m x x
 1/4 2
i
2
~ h

xex /2
=

2m x
 1/4 2

2
~ 

=
+ 2 x2 ex /2

2m
and its expectation value is
 1/2 2  r
r 
2
~

hT i =

2m

2
3
~2
.
4m
The expectation value of the potential energy is
Z
2
hV i = V0
2 (x)ex dx

r Z
2
2

ex ex dx
= V0

r r

= V0
( + )
s

.
= V0
( + )

(10)

(11)

Combining (10) and (11),


~2
exptwo|H| = h| T |i + h| V |i =
V0
2m

.
( + )

(12)

To minimize with respect to we equate the first derivative of this to zero:


"
#

V0
1

~2
p

p
0=
2m
2
( + )
( + )3

1/2
~2
2
V0
=

2m
2 ( + )3

2
mV0
3
= ( + )
~2
2

mV0
= 4 + 3 3 + 3 2 2 + 3
~2
2

mV0
= x4 + 3x3 + 3x2 + x
~2

Homer Reids Solutions to Merzbacher Problems: Chapter 8

10

where I put x = /. In theory we could write down an explicit expression for


the roots of this quartic in terms of mV0 /~2 , and then insert said expression
into (12) to obtain the lowest energy attainable with this form of trial wave
function. In practice, however, this would be a mess, and I cant see any way
to proceed other than numerically. Am I missing some kind of trick here?

Problem 8.5
Show that as inadequate a variational trial function as

( 
C 1 |x|
|x| a
a
(x) =
0
|x| > a
yields, for the optimum value of a, an upper limit to the ground state energy of the
linear harmonic oscillator, which lies within less than 10 percent of the exact value.

The first task is to evaluate the normalization constant C.


Z a
2
1=C
(x)2 dx
a
Z a
x 2
2
1
= 2C
dx
a
0

Z a
x x2
2
12 + 2
= 2C
a
a
i
h0
a
= 2C 2 a a +
3
so

C=

3
.
2a

The harmonic oscillator hamiltonian is


E =T +V =

p2
m 2 x2
+
.
2m
2
(13)

~
2m

~2
=
2m

exptwoT =

(x)
a

(x) dx
x2

Integrating by parts,
a
2 )
Z a


(x) dx
(x) (x)
x
a x
a

(14)

Homer Reids Solutions to Merzbacher Problems: Chapter 8

11

The first term vanishes...


~2
=
2m

3~2
=
2ma2

3
2a

Z

a
a

1
dx
a2
(15)

(16)
exptwoV =

m
2

x2 (x)2 dx

a
a

x2 (x)2 dx

 Z a
2x3
x4
3
2
2
x
+ 2 dx
= m
2a
a
a
0
a
  3
4
5
x
x
3 x

+ 2
= m 2

2a
3
2a 5a 0
= m

m 2 a2
20

3~2
m 2 a2
+
.
2
2ma
20
To minimize with respect to a we set the a derivative of this to zero:
exptwo|H| = h| T |i + h| V |i =

0=

(17)

(18)

m 2 a
3~2
+
ma3
10

or
30~2
m2 2

~
.
a2 = 30
m

a4 =

Inserting into (18),


3
exptwoH = ~ 0.547 ~.
30
Of course the actual ground state energy is 0.5 ~, so the fractional error is
0.047/0.5 < 10%.

Homer Reids Solutions to Merzbacher Problems: Chapter 8

12

Problem 8.6
A particle of mass m moves in a potential V (r). The n th discrete energy eigenfunction of this system, n (r), corresponds to the energy eigenvalue En . Apply the
variational principle by using as a trial function,
t (r) = n (r),
where is a variational (scaling) parameter, and derive the virial theorem for stationary states.

You might also like