You are on page 1of 22

Physics 215 Solution Set 2 Winter 2018

1. At time t = 0, the wave function of a free particle moving in a one-dimension is given


by, Z +∞
ψ(x, 0) = N e−|k|/k0 eikx dk , (1)
−∞
where N and k0 are real positive constants.

Before attacking this problem, we shall first evaluate the integral that appears in eq. (1)
and determine the value of the constant N. Since k0 > 0 the integral is convergent, and it
follows that
(Z Z ∞ )
0     
1 1
ψ(x, 0) = N dk exp k + ix +N dk exp −k − ix
−∞ k0 0 k0
" −1  −1 #
1 1 2Nk0
=N + ix + − ix = .
k0 k0 1 + k02 x2

The normalization constant N is determined by demanding that


Z ∞
|ψ(x, 0)|2 dx = 1 . (2)
−∞

Hence,

dx
Z
4|N|2 k02 = 1. (3)
−∞ (1 + k02 x2 )2
The integral in eq. (3) is easily computed if you remember that
Z ∞ ∞
dx −1

I ≡ k0 2 2
= tan (k0 x) = π . (4)
−∞ (1 + k0 x ) −∞

If we multiply eq. (4) by k0 and define q ≡ 1/k02 , then eq. (4) becomes,
Z ∞
dx π
k0 I = =√ .
−∞ q + x q
2

Finally, taking the derivative of the above result with respect to q yields
Z ∞ Z ∞
d dx 4 dx π
(k0 I) = − = −k 0 2
= − = − 21 πk03 ,
dq −∞ (q + x )
2 2
−∞ (1 + k0 x )
2 2 2q 3/2

after making use of q 1/2 = 1/k0 . We therefore end up with,


Z ∞
dx π
2 2 2
= .
−∞ (1 + k0 x ) 2k0

1
Inserting this last result into eq. (3) yields,

2π|N|2 k0 = 1 . (5)

Since N is a real positive constant by assumption (one is always free to rephase the wave
function such that N > 0), it follows that
r
2k0 1
ψ(x, 0) = . (6)
π 1 + k02 x2
REMARKS:
1. A good check of the results above is to make use of dimensional analysis. If x has
dimensions of length, then k0 has dimensions of inverse length. We can indicate this by
writing [x] = L and [k] = L−1 . Then, eq. (3) implies that [N] = L1/2 and [ψ(x, 0)] = L−1/2 .
In addition, the parameter q = 1/k02 has dimension [q] = L2 . Using these results, one can
verify that all the equations above are dimensionally correct. Such a check helps immensely
in avoiding algebraic mistakes.
2. Actually, there is a much quicker way to derive Req. (5). Using the results of problem
+∞
3(b) of Problem Set 1, it follows that if f (x) = √12π −∞ a(k)eikx dk , then
Z +∞ Z +∞
2
|f (x)| dx = |a(k)|2 dk .
−∞ −∞

Applying this result to the evaluation of eq. (2), we can identify a(k) = N 2π e−|k|/k0 and
f (x) = ψ(x, 0). Hence, Z ∞
2π|N| 2
e−2|k|/k0 dk = 1 .
−∞

The above integral is elementary, and we immediately recover eq. (5).

(a) What is the probability that a measurement of the momentum performed at time
t = 0 will yield a result between −p1 and p1 ? How does the probability change if the
measurement is performed instead at time t? Explain your result.

Applying the postulates of quantum mechanics, the probability that a measurement of the
momentum performed at time t = 0 will yield a result between −p1 and p1 , denoted by
P (p1 , 0) below, is given by
Z p1 Z p1
2
P (p1 , 0) = |hp|ψi| dp = |φ(p, 0)|2 dp , (7)
−p1 −p1

where the momentum-space wave function at t = 0 is the Fourier transform of ψ(x, 0),
Z ∞
1
φ(p, 0) = √ e−ipx/~ ψ(x, 0) dx ,
2π~ −∞

2
Inverting the Fourier transform yields,

1
Z
ψ(x, 0) = √ eipx/~ φ(p, 0) dx .
2π~ −∞

If we write p = ~k and compare with eq. (1), we can identify


r

φ(p, 0) = Ne−|p|/p0 ,
~

where we have defined p0 ≡ ~k0 . Hence, using N = (2πp0 /~)1/2 [cf. eq. (5)], it follows that
1
φ(p, 0) = √ e−|p|/p0 . (8)
p0

Finally, inserting this result into eq. (7) yields

1 p1 −2|p|/p0 2 p1 −2p/p0
Z Z
P (p1 , 0) = e dp = e dp = 1 − e−2p1 /p0 .
p0 −p1 p0 0

Next, we evaluate P (p1, t). The wave function evolves according to,

|ψ(t)i = U(t) |ψ(0)i ,



where the time evolution operator is given by U(t) = exp −iHt/~ . For a free particle
moving in one dimension, the Hamiltonian is H = P 2 /(2m). Hence,

iP 2 t
 
U(t) = exp − . (9)
2m~

We proceed to evaluate the momentum space wave function,


Z ∞ Z ∞
′ ′ ′
φ(p, t) = hp|ψ(t)i = hp| U(t) |p i hp | ψ(0)i dp = hp| U(t) |p′ i φ(p′ , 0) dp′ , (10)
−∞ −∞

after inserting a complete set of momentum eigenstates. Using eq. (9) and P |pi = p |pi,
where hp|p′i = δ(p − p′ ),

iP 2 t i~p2
   
′ ′
hp| U(t) |p i = hp| exp − |p i = exp − δ(p − p′ ) .
2m~ 2m~

Plugging this result into eq. (10), the integration over p′ is now trivial due to the delta
function. Hence,
i~p2
 
φ(p, t) = exp − φ(p, 0) . (11)
2m~
It therefore follows that |φ(p, t)|2 = |φ(p, 0)|2. That is, |φ(p, t)|2 is independent of time.
This result immediately implies that the probability P (p, t) is also independent of time.

3
This result is expected since for a free particle, the energy eigenstates are also momentum
eigenstates. Since the Hamiltonian, H = P 2 /(2m) is time-independent, it follows that the
energy (and consequently the momentum) eigenstates are stationary states. In particular,
the probability of finding a particle with a given range of momenta must also be time-
independent.

(b) What is the form of the (position space) wave packet at time t = 0? Calculate the
product ∆X∆P at time t = 0. Describe qualitatively the subsequent evolution of the wave
packet.

The form of the wave packet was explicitly obtained in eq. (6). Using this result as well
as the corresponding result for the momentum space wave function obtained in eq. (8), we
compute the following expectation values.
Z ∞
2k0 ∞ x dx
Z
2
hXi = x|ψ(x, 0)| dx = = 0,
−∞ π −∞ (1 + k02 x2 )2
Z ∞
1 ∞ −2|p|/p0
Z
2
hP i = p|φ(p, 0)| dp = pe dp = 0 ,
−∞ p0 −∞

2k0 ∞ x2 dx 1
Z Z

2 2 2
X = x |ψ(x, 0)| dx = 2 2 2
= 2,
−∞ π −∞ (1 + k0 x ) k0
Z ∞
1 ∞ 2 −2|p|/p0
Z

2 2 2
P = p |φ(p, 0)| dp = pe dp = 12 p20 ,
−∞ p0 −∞

(12)

where p0 ≡ ~k0 . It then follows that


1/2 1
X − hXi2

2
∆X = = , (13)
k0
1/2 ~k0
P − hP i2

2
∆P = = √ . (14)
2

Hence, ∆X∆P = 21 2 ~, which is consistent with the Heisenberg uncertainty principle,
∆X∆P ≥ 12 ~ .
Finally, as time evolves, the wave packet in coordinate space spreads out. In principle,
one can demonstrate this behavior by computing,
Z ∞ Z ∞
p2 t
  
1 ipx/~ 1 i
ψ(x, t) = √ φ(p, t)e dp = √ exp px − e−|p|/p0 dp .
2π~ −∞ 2π~p0 −∞ ~ 2m
Unfortunately, the above integral cannot be evaluated in terms of elementary functions
alone [one would need to employ the error function erf(x)]. However, if one simply wants

4
to demonstrate the qualitative behavior of the spreading wave packet, one can make use of
the uncertainty relation obtained in eq. (2.2.30) on p. 85 of Sakurai and Napolitano,
~t
∆X(0)∆X(t) ≥ .
2m
Using ∆X(0) = 1/k0 obtained in eq. (13), it follows that ∆X(t) ≥ ~k0 t/(2m), which
indicates that the width of the wave packet increases (at least as fast as) linearly in the
time t.

2. (a) Consider a quantum mechanical ensemble characterized by a density matrix ρ.


Suppose that the system is governed by a Hamiltonian H (which may be time dependent).
Show that the time evolution of ρ (in the Schrödinger picture) is given by:
∂ρ i
= − [H, ρ] .
∂t ~

The density operator for a quantum ensemble made up of N representatives is defined by


X
ρ= pi |ψi i hψi | ,
i

where the state |ψi i appears ni times in the ensemble and pi ≡ ni /N. That is, pi is the
probability that an arbitrarily chosen element of the ensemble is in the state |ψi i.
The time evolution of the state |ψi i (in the Schrödinger picture) is given by the Schrödinger
equation,
d
i~ |ψi i = H |ψi i . (15)
dt
Under the assumption that the Hamiltonian is a self-adjoint operator, the adjoint of eq. (15)
is given by,
d
−i~ hψi | = hψi | H . (16)
dt
Noting that the probabilities pi do not depend on the time t, it follows that
   
∂ρ X ∂ ∂
= pi |ψi i hψi | + |ψi i hψi |
∂t i
∂t ∂t

1 X   i 
= pi H |ψi i hψi | − |ψi i hψi | H = − H , ρ .
i~ i ~

It follows that
∂ρ i
+ [H, ρ] = 0 ,
∂t ~
which is called the von Neumann equation.

5
(b) Let U(t, t0 ) be the time evolution operator. Find a general expression for ρ(t) in
terms of ρ(t0 ) and U.

Starting from
X X
ρ(t) = pi |ψi (t)i hψi (t)| , ρ(t0 ) = pi |ψi (t0 )i hψi (t0 )| ,
i i

with |ψi (t)i = U(t, t0 ) |ψ(t0 )i, it immediately follows that

ρ(t) = U(t, t0 )ρ(t0 )U † (t, t0 ) . (17)

REMARKS:

Note that U(t, t0 ) satisfies the differential equation


i~ U(t, t0 ) = H(t)U(t, t0 ) , (18)
∂t
which is equivalent to the Schrödinger equation given by eq. (15). Taking the adjoint of
this equation, under the assumption that the Hamiltonian is a self-adjoint operator,
∂ †
−i~ U (t, t0 ) = U † (t, t0 )H(t) . (19)
∂t
Taking the time derivative of eq. (17) and employing eqs. (18) and (19), one again derives
the von Neumann equation obtained in part (a).

(c) Prove that Tr ρ2 is time-independent. Hence, show that a pure state cannot evolve
into a mixed state.

Using eq. (17),

Tr ρ2 (t) = Tr U(t, t0 )ρ(t0 )U † (t, t0 ) U(t, t0 )ρ(t0 )U † (t, t0 )


  

= Tr U(t, t0 )ρ2 (t0 )U † (t, t0 ) = Tr ρ2 (t0 ) .




In the penultimate step above, we cyclically permuted the arguments of the trace (which
does not change its value) and put U † U = I after noting that U is unitary. It follows that
Tr ρ2 is time-independent.
We now recall that a pure state satisfies Tr ρ2 = 1, whereas a mixed state satisfies
Tr ρ2 < 1. We can therefore conclude that pure states always evolve into pure states,
whereas mixed states always evolve into mixed states.

6
3. In a one-dimensional problem, consider a particle of potential energy V (x) = −f x,
where f > 0.
(a) Write Ehrenfest’s theorem for the mean values of the position hXi and the momen-
tum hP i of the particle. Integrate these equations and compare with the classical motion.

Ehrenfest’s theorem states that


   
d ∂H d ∂H
hXi = , hP i = − . (20)
dt ∂P dt ∂X

For this problem, the Hamiltonian is given by

P2
H= − fX ,
2m
where f > 0. Using ∂H/∂P = P/m and ∂H/∂X = −f I (where I is the identity operator),
eq. (20) yields,
d 1 d
hXi = hP i , hP i = f . (21)
dt m dt
In obtaining the second result above, we used the fact that the expectation value of the
identity operator I with respect to normalized state is equal to 1. Integrating eq. (21) then
yields
t f t2
hP i = hP i0 + f t , hXi = hXi0 + hP i0 + , (22)
m 2m
where hP i0 ≡ hP it=0 and hXi0 ≡ hXit=0 . As expected, eq. (22) is identical in form to the
corresponding classical mechanics results.

(b) Show that the root-mean-square deviation ∆P does not vary with time.

The root-mean square deviation satisfies the following relation,

(∆P )2 = P 2 − hP i2 .


(23)

To evaluate hP 2i, we use the fact that d hHi /dt = 0, which is a consequence of the fact
that H has no explicit time dependence.1 It then follows that
1 d
2 d
P + f hXi = 0 .
2m dt dt
1
Recall that the generalized Ehrenfest theorem is given by
 
d i
∂Ω
hΩi = − [Ω , H] + ,
dt ~ ∂t

for any operator Ω. In the case of Ω = H, we have d hHi /dt = h∂H/∂ti (since H commutes with itself).
For a time-independent Hamiltonian, we have ∂H/∂t = 0. It then follows that d hHi /dt = 0.

7
Using eq. (21) for dhXi/dt, we immediately obtain,
d
2
P = −2f hP i . (24)
dt
Next, we make use of eq. (21) for dhP i/dt to conclude that
d d
hP i2 = 2 hP i P = 2f hP i . (25)
dt dt
Hence, the time derivative of eq. (23) yields
d
(∆P )2 = −2f hP i + 2f hP i = 0 ,
dt
after using eqs. (24) and (25). Hence, ∆P does not vary in time.

(c) Write the Schrödinger equation in the p-representation and deduce a relation between
∂ ∂
| hp|ψ(t)i |2 and | hp|ψ(t)i |2
∂t ∂p
Solve the equation thus obtained and give a physical interpretation.

The Schrödinger equation [cf. eq. (15)] for this problem is,
 2 
d P
i~ |ψ(t)i = − f X |ψ(t)i . (26)
dt 2m
In the p-representation,
 2 
∂ P
i~ hp|ψ(t)i = hp| − f X |ψ(t)i . (27)
∂t 2m

Using the adjoint of P |pi = p |pi recalling that P is self-adjoint and hence its eigenvalues
are real), and recalling that2

hp| X |ψ(t)i = i~ hp|ψ(t)i , (28)
∂p
it then follows from eq. (27) that

p2
 
∂φ(p, t) ∂
i~ = − i~f φ(p, t) , (29)
∂t 2m ∂p

where φ(p, t) ≡ hp|ψ(t)i is the wave function in the p-representation.


2
Due to eq. (28), one says that the operator X is represented by the differential operator i~ ∂/∂p in the
p-representation.

8
Using eq. (29),
∂ ∂φ(p, t) ∂φ∗ (p, t)
|φ(p, t)|2 = φ∗ (p, t) + φ(p, t)
∂t ∂t ∂t
 2   2 
∗ ip ∂ ip ∂
= −φ (p, t) +f φ(p, t) + φ(p, t) −f φ∗ (p, t)
2m~ ∂p 2m~ ∂p
∂φ∗ (p, t)
 
∗ ∂φ(p, t) ∂
= −f φ (p, t) + φ(p, t) = −f |φ(p, t)|2 .
∂p ∂p ∂p

Denoting P(p, t) ≡ |φ(p, t)|2 , the above equation reads,


∂P(p, t) ∂P(p, t)
= −f . (30)
∂t ∂p
To solve eq. (30), it is convenient to change variables by introducing

r ≡ p − ft , s ≡ p + ft .

Then
∂P ∂P ∂P ∂P ∂P ∂P
= + , = −f +f .
∂p ∂r ∂s ∂t ∂r ∂s
Inserting these results into eq. (30) yields
∂P
= 0.
∂s
That is P(r, s) is a function of r alone. This means that

P(p, t) = f (p − f t) ,

where f is an arbitrary function. In particular, it follows that

P(p, t) = P(p − f t, 0) .

The interpretation of this result is that the probability distribution in p moves according
to the classical equation of motion, p = p0 + f t.

(d) Write the Schrödinger equation in the x-representation. What are the energy eigen-
functions?

The Schrödinger equation in the x-representation is obtained from eq. (26) by multiplying
from the left by hx|. The wave function in the x-representation is ψ(x, t) = hx|ψ(t)i. After
writing ψ(x, t) = ψ(x)e−iEt/~ , we obtain the time-independent Schrödinger equation in the
x-representation,
~2 d 2
 
− − f x ψ(x) = Eψ(x) ,
2m dx2

9
which we can rewrite as
d2 ψ
 
2mf 2mE
+ x+ 2 ψ = 0. (31)
dx2 ~ 2 ~
It is convenient to introduce a new variable,
 −2/3  
2mf 2mf 2mE
y≡ x + . (32)
~2 ~2 ~2

By changing variables from x to y, eq. (31) simplifies to

d2 ψ
+ yψ = 0 . (33)
dy 2

Consulting, e.g., N.N. Lebedev, Special Functions and Their Applications (Dover Publica-
tions, Inc., New York, 1972), pp. 136–139, we see that the solutions to eq. (33) are the Airy
functions,
ψ(y) = c1 Ai(−y) + c2 Bi(−y) , (34)
where c1 and c2 are arbitrary constants are determine by the boundary conditions relevant
for the problem.

(e) Is the energy spectrum continuous or discrete? What is the behavior of the energy
eigenfunctions as |x| → ∞ ? If the potential were replaced by V (x) = f |x|, how would
your answer change?

To see whether the energy spectrum is discrete or continuous, let us examine the asymptotic
behavior of the Airy functions. On p. 138 of Lebedev (op. cit.), the following results are
given:
1 −1/4

exp − 23 x3/2 ,

 2 √π x for x → ∞,


Ai(x) ∼
1
 √ (−x)−1/4 cos 23 (−x)3/2 − 14 π ,

for x → −∞,


π

1 −1/4

exp 23 x3/2 ,

 √π x for x → ∞,


Bi(x) ∼
1
− √ (−x)−1/4 sin 23 (−x)3/2 − 41 π ,

for x → −∞.


π

10
Noting that f > 0, the potential V (x) = −f X is exhibited below.

V (x)

Since V (x) > E as x → −∞, we demand that limx→−∞ ψ(x) = 0. Thus, we must choose
c2 = 0 in the solution given in eq. (34). We conclude that ψ(y) = cAi(−y), where y
is given in terms of x by eq. (32) and the constant c is determined by an appropriate
normalization condition. Note that as x → ∞, we expect oscillatory behavior of the
wave function (in analogy to the wave function ψ(x) ∼ eikx of a free particle). This is
indeed the case as is evident from the asymptotic form for Ai(−x) as x → ∞ given above.
Note that no condition on E needs to be imposed to satisfy the boundary conditions (i.e.,
an exponentially vanishing wave function as x → −∞ and an oscillatory wave function as
x → ∞). Thus, the energy spectrum is continuous (and there are no bound state solutions).
In contrast, consider the potential V (x) = f |x|, with f > 0. The previous figure is now
modified as follows.

V (x)

In this case, V (x) > E as x → ±∞. The spectrum must consist entirely of bound states,
and the energy spectrum is discrete. To solve for the bound state energies, we must solve

d2 ψ
 
2mf 2mE
+ − 2 |x| + 2 ψ = 0. (35)
dx2 ~ ~

11
separately for x > 0 and x < 0. Following our previous analysis, we obtain,
(
c Ai(−y) , for x < 0,
ψ(x) = ′ ′
(36)
c Ai(y ) , for x > 0,

where y is given in terms of x by eq. (32) and


 −2/3  
′ 2mf 2mf 2mE
y ≡ x− 2 .
~2 ~ 2 ~

Note that the wave function given by eq. (36) exponentially approaches zero as x → ±∞, as
required by the bound state solutions to this problem. To determine the relation between
the constants c and c′ , we require that both ψ(x) and dψ/dx are continuous functions at
x = 0. These two conditions can only be satisfied for a discrete set of energy eigenvalues E
with a fixed value of c′ /c. There remains one overall undetermined coefficient (which
depends on the bound state energy E), which is determined by requiring that the norm of
the bound state wave functions are unity.

4. Consider a particle in three dimensions whose Hamiltonian is given by:


~2
P
H= ~ .
+ V (X) (37)
2m
~ ·P
By calculating the commutator, [X ~ , H], derive the quantum Virial Theorem,
* +
d D ~ ~E P~2 D E
X ·P = − X ~ · ∇V
~ . (38)
dt m

To identify eq. (38) with the quantum mechanical analog of the Virial Theorem, it is
essential that the left-hand side of eq. (38) vanish. Under what condition does this happen?

We shall make use of the generalized Ehrenfest theorem,


 
d i
∂Ω
hΩi = − [Ω , H] + , (39)
dt ~ ∂t
~ ·P
Using Ω = X ~ , it follows that

d D ~ ·P
~
E i D ~ ~ E
X =− X ·P , H . (40)
dt ~
For H given by eq. (37),
~2 ~ 2 
~ ·P
~,H = X ~ , P + V (X)
~ ·P ~ = X ~, P + X
~ ·P ~ ·P
~ , V (X)
~ .
     
X (41)
2m 2m
12
We evaluate separately the two commutators on the right hand side of eq. (41) First,

~ 2
P 1  1  1
 
~ ~
     
X ·P , = Xi P i , P j P j = Xi , P j P j P i = Pj Xi , Pj Pi + Xi , Pj ]Pj Pi ,
2m 2m 2m 2m
 
where
 there is
 an implicit
 sum over the pairs of repeated indices. Using X i , P j = i~δij
and Xi , Xj = Pi , Pj = 0, it follows that

~ 2  i~ 2
P
~ ~ ~ .

X ·P , = P (42)
2m m
Second, we evaluate
~ ·P
~ , V (X)
~ = Xi Pi , V (X)
~ .
   
X (43)
In light of problem 5(b) on Problem Set 1,
  ∂F
P , F (X) = −i~ .
∂X
The extension to three dimensions is straightforward,
  ∂F
Pi , F (X) = −i~ .
∂Xi
Hence, using eq. (43) we obtain

~ ·P
~ , V (X) ~ = −i~Xi ∂V = −i~X
~ = Xi Pi , V (X) ~ · ∇V
~ .
   
X (44)
∂Xi
Combining eqs. (42) and (44), we end up with
" #
~2
P ~2
P
~ ·P
~,H = X~ ·P
~, ~ = i~ ~ · ∇V
~
   
X + V (X) −X .
2m m

Thus, eq. (40) yields * +


d D ~ ~E ~2
P D E
X ·P = ~ · ∇V
− X ~ . (45)
dt m
To identify eq. (45) with the quantum mechanical analog of the Virial Theorem, it is
essential that the left-hand side of eq. (45) vanish. This will happen if the expectation
values are taken with respect to a stationary state. In this case,
D E
X~ ·P
~ ≡ hψ(t)| X ~ ·P
~ |ψ(t)i = hψ(0)| eiHt/~ X
~ ·P
~ e−iHt/~ |ψ(0)i
t
D E
~ ·P
= hψ(0)| eiEt/~ X ~ e−iEt/~ |ψ(0)i = hψ(0)| X
~ ·P
~ |ψ(0)i = X~ ·P
~ .
0

13
D E
That is, ~ ·P
X ~ is time-independent when taken with respect to a stationary state. In
this case,
d D ~ ~E
X ·P = 0 , (46)
dt
and we arrive at the quantum mechanical virial theorem,
* +
~2
P D E
= X ~ · ∇V
~ . (47)
m

REMARKS:

A technical remark is in order. The classical mechanical virial theorem is derived


under the assumption that all the particle orbits are bounded. In the quantum mechanical
context, this condition translates into the requirement that the stationary states used in
computing the expectation values above are bound states. That is, the states used to
evaluate expectation values must be discrete energy levels rather than continuous energy
levels. To see where this condition arises in the above analysis, we note that if Ω is time-
independent, then eq. (39) yields,

d i

hΩi = − [Ω , H] .
dt ~
With respect to an energy eigenstate,


[Ω , H] = hE| ΩH − HΩ |Ei = E hE| Ω |Ei − E hE| Ω |Ei = 0 , (48)

after using H |Ei = E |Ei and the corresponding adjoint relation. But this last computa-
tion is correct as long as hE| Ω |Ei is finite. Note that for discrete energy levels, we have
hE|E ′ i = δEE ′ . Thus, we expect that hE| Ω |Ei is also finite, in which case eq. (48) is valid.
In contrast, for energy eigenvalues in the continuum corresponding to scattering states,
we have hE|E ′ i = δ(E − E ′ ). That is, for continuum energy levels,
hE|Ei = ∞. Likewise
hE| Ω |Ei is also infinite, so we cannot conclude as in eq. (48) that [Ω , H] = 0. Hence,
eq. (46) is not expected to be valid for expectation values with respect to stationary scat-
tering states (in the continuum). That is, the quantum mechanical virial theorem given by
eq. (47) applies only for the vacuum expectation values with respect to stationary bound
states.

5. In this problem, you are asked to derive the Feynman-Hellmann Theorem.


(a) If the Hamiltonian H(λ) depends on a real parameter λ, i.e., H(λ) |ψi = E(λ) |ψi,
then show that:  
∂E ∂H
= ψ ψ .
∂λ ∂λ

14
Consider E(λ) ≡ hψ| H(λ) |ψi. For a normalized state, hψ|ψi = 1. Then,
     
dE d dψ dψ ∂H
= hψ| H(λ) |ψi = H(λ) ψ + ψ H(λ) + ψ ψ
dλ dλ dλ dλ ∂λ
     
dψ dψ ∂H
= E(λ) ψ + ψ + ψ
ψ . (49)
dλ dλ ∂λ
Note that if one differentiates the normalization condition, hψ|ψi = 1 with respect to λ, we
immediately obtain,    
d dψ dψ
0= hψ|ψi = ψ + ψ .
dλ dλ dλ
Using this last result in eq. (49), we end up with
 
∂E ∂H
= ψ ψ , (50)
∂λ ∂λ
which is the Feynman-Hellman theorem.

(b) Consider the one-dimensional problem with the Hamiltonian:


~2 d 2
H=− + V (x) ,
2m dx2
where V is independent of the parameter m. Suppose one finds that this Hamiltonian
possesses a particular energy eigenstate with energy eigenvalue E. Describe the behavior
of E as m decreases.

If we rewrite the Hamiltonian in terms of the momentum operator P and the position
operator X, then
P2
H= + V (X) .
2m
Since V (x) is independent of m, it follows that
∂H P2
=− 2.
∂m 2m
Hence, using eq. (50),
 
dE ψ = − 1 hψ| P 2 |ψi < 0 ,
∂H
= ψ
dλ ∂m 2m2
since the operator P 2 is non-negative.3 More explicitly,
  Z ∞ 2
∂H ~2 dψ
ψ ψ =− < 0.
∂m 2m −∞ dx
We conclude that ∂E/∂m < 0. That is, as m decreases, the bound state energies increase.
3
If we define |φi ≡ P |ψi where P is self-adjoint (P = P † ), then hψ| P 2 |ψi = hψ| P † P |ψi = hφ|φi > 0
since the norm of any non-zero state vector is positive.

15
6. Consider the operator:
mωX + iP
√ , a= (51)
2m~ω
where X and P are the position and momentum operators, respectively.
(a) Let |zi be an eigenvector of a with eigenvalue z. This state is called a coherent state.
Compute hx|zi. Using this result, show that hz|z ′ i =
6 0. Why does the lack of orthogonality
of these states not violate any of our quantum mechanics postulates?

The eigenvalue equation for the operator a is


 
mωX + iP
a |zi = √ |zi = z |zi .
2m~ω
Consider the eigenvalue equation with respect to the x-basis. Multiplying on the left by
hx| then yields,    
mωX + iP
x √ z = z |zi .
2m~ω
To evaluate this equation, we make use of
d
hx| X |zi = x hx|zi , hx| P |zi = −i~ hx|zi .
dx
It then follows that  mω 1/2  
~ d
x+ hx|zi = z hx|zi .
2~ mω dx
This is a differential equation fro hx|zi. Denoting ψz (x) ≡ hx|zi, we proceed to solve
"  1/2 #
d mωx 2mω
+ − z ψz (x) = 0 .
dx ~ ~
The solution to this equation is
"  1/2 #
mω 2 2mω
hx|zi ≡ ψz (x) = Cz exp − x + zx , (52)
2~ ~
where Cz is a normalization constant (which will depend on the eigenvalue z).
We can check to see whether states of different values of z are orthogonal by computing
hz|z ′ i. After inserting a compete set of position eigenstates,,
Z ∞ "  1/2 # "  1/2 #
′ ∗ mω 2 2mω mω 2 2mω
hz|z i = Cz Cz ′ exp − x + zx exp − x + z′x
−∞ 2~ ~ 2~ ~
" 1/2 #
∞ 
mω 2 2mω
Z
= Cz∗ Cz ′ exp − x + (z ′ + z ∗ )x
−∞ ~ ~

 1/2
π~
Cz∗ Cz ′ exp 21 (z ′ + z ∗ )2 .
 
= (53)

16
For z 6= z ′ , it is clear that hz|z ′ i =
6 0. Hence, the eigenstates of the operator a are not
mutually orthogonal. This does not violate any of the quantum mechanics postulates since
a is not a self-adjoint (nor is it an hermitian) operator. Thus, its eigenvalues need not be
real and its eigenstates need not be mutually orthogonal.

(b) Consider the operator a in the context of the one-dimensional harmonic oscillator.
Compute hn|zi, where |ni is the nth energy eigenstate. (Assume that |zi is normalized to
unity.) Given a coherent state |zi, find the most probable value of n (and corresponding
energy E).

Repeated application of a† |ni = n + 1 |n + 1i yields
(a† )n
|ni = √ |0i . (54)
n!
Taking the adjoint of eq. (54), it then follows that
 n 
a zn
hn|zi = 0 √ z = √ h0|zi . (55)
n! n!
Since {|ni} are complete set of states, we can use the completeness relation to obtain
∞ ∞
X X zn
|zi = |ni hn|zi = h0|zi √ |ni . (56)
n=0 n=0 n!
Using eq. (56), we can compute the norm of |zi,

X |z|2n 2
hz|zi = |h0|zi|2 = |h0|zi|2 e|z| .
n=0
n!

We shall normalize |zi to unity, i.e. we set hz|zi = 1. It then follows that
2 /2
h0|zi = e−|z| , (57)
after (conventionally) setting an overall phase factor to 1. Inserting this last result back
into eq. (55) then yields,
zn 2
hn|zi = √ e−|z| /2 .
n!
Given the state |zi, the probability of measuring a value n is given by
|z|2 −|z|2
|hn|zi|2 = e .
n!
The most probable value of n depends on z. For |z| ≤ 1, we see that n = 0 is the most
probable value. As |z| becomes larger, so does the most probable value of n. For |z| ≫ 1
we can make use of Stirling’s approximation,

n! ≃ 2πn nn e−n .

17
Thus, for |z| ≫ 1, we approximately have,
n
|z|2

1 2
2
|hn|zi| ≃ √ en−|z| .
2πn n
We would like to maximize the function,
  2 n 
1 |z| n−|z|2
f (n) = ln √ e = n − |z|2 + n ln |z|2 − (n + 21 ) ln n .
n n
Taking a derivative and setting f ′ (n) = 0 yields ln(|z|2 /n) = 1/(2n). For |z| ≫ 1, it is clear
that n ≃ |z|2 is an approximate solution to f ′ (n) = 0. Moreover, f ′′ (n) ≃ −1/n < 0 at the
extremum, which indicates that n ≃ |z|2 is a maximum.
We conclude that the maximum value of |hn|zi|2 occurs for n ≃ |z|2 , That is, for
large |z|, the most probable energy is given by E ≃ ~ω|z|2 .

2 /2 †
(c) Prove that the normalized coherent state can be written as, |zi = e−|z| eza |0i.

Combining eqs. (56) and (57), it follows that


∞ ∞
−|z|2 /2
X zn −|z|2 /2
X zn † n
|zi = e √ |ni = e (a ) |0i , (58)
n=0 n! n=0
n!

after making use of eq. (54) to obtain the final result above. The sum over n can now be
performed. The end result is
2 †
|zi = e−|z| /2 eza |0i .

(d) Prove that the coherent state is a state of minimum uncertainty, i.e., ∆X∆P = ~/2.

We first compute the expectation values of X, X 2 P and P 2 with respect to the coherent
state |zi. From eq. (51) and its adjoint, we can solve for the operators X and P ,
 1/2  1/2
~  †
 ~  † 
X= a+a , P =i a −a , (59)
2m0 ω 2m0 ω
where we have used the fact that X and P are self-adjoint operators. It then follows that
~ ~  2
X2 = (a + a† )2 = a + (a† )2 + 2aa† + 1 ,

(60)
2m0 ω 2m0 ω
~ ~  2
P2 = − (a† − a)2 = − a + (a† )2 − 2aa† − 1 .

(61)
2m0 ω 2m0 ω
Employing the above results and making use of,

a |zi = z |zi , hz| a† = hz| z ∗ , (62)

18
it follows that
 1/2
~ ~ 
z + z∗ , (z + z ∗ )2 + 1 ,

2
X ≡ hz| X 2 |zi =
 
hXi ≡ hz| X |zi =
2mω 2mω
 1/2
h~ω
z∗ − z , P ≡ hz| P 2 |zi = − 12 m~ω (z ∗ − z)2 − 1 .

2  
hP i ≡ hz| P |zi = i
2
Using the above results, we obtain
2 ~ 2
(∆X)2 = X 2 − hXi = (∆P )2 = P 2 − hP i = 21 m~ω .



,
2mω
Hence, (∆X)2 (∆P )2 = 14 ~2 . Taking the positive square root yields
∆X∆P = 21 ~ ,
corresponding to a state of minimum uncertainty.

(e) Consider coherent states in which the parameter z is a real positive number much
larger than 1. Evaluate the expectation value of the quantum Hamiltonian with respect
to a coherent state with |z| ≫ 1. In what way is this state a good approximation to the
classical limit of the harmonic oscillator?

Recall that the quantum Hamiltonian of the harmonic oscillator can be written in terms of
the raising and lowering operators as
H = ~ω a† a + 21 ,


where the identity operator multiplying the factor of 12 has been suppressed in the notation
above. Taking the expectation value of H with respect to the coherent state |zi and making
use of eq. (62),
hHi ≡ hz| H |zi = ~ω hz| a† a |zi + 12 = ~ω |z|2 + 21 .
  
(63)
In the case of |z| ≫ 1, we have
hHi ≃ ~ω|z|2 , for |z| ≫ 1. (64)

REMARK: One can also compute hHi starting from


P2
H= + 1 mω 2 X 2 ,
2m 2
Using the results of part (d), it follows that
1
2 1 2 2
2
hHi ≡ hz| H |zi = P + 2m ω X
2m
= − 14 ~ω ((z ∗ − z)2 − 1 + 14 ~ω (z + z ∗ )2 + 1
   

= ~ω |z|2 + 12 ,
 

thereby reproducing the result of eq. (63).

19
In order to appreciate why |z| ≫ 1 corresponds to the classical limit, we need to
examine the time dependence of the coherent states. Suppose that we examine the state |zi
[cf. eq. (58)] at time t = 0.

2 /2
X zn
|zi ≡ |z, t = 0i = e−|z| √ |ni .
n=0 n!

Then at time t,

−|z|2 /2
X zn
|z, ti = e √ e−iEn t/~ |ni ,
n=0 n!
where En = 21 ~ω(n + 12 ) are the eigenvalues of the harmonic oscillator Hamiltonian. That
is,

2
X (ze−iωt )n
|z, ti = e−|z| /2 e−iωt/2 |ni = e−iωt/2 ze−iωt .

√ (65)
n=0 n!
The wave function corresponding to the coherent state in the x-representation is

ψz (x, t) ≡ hx|z, ti = e−iωt/2 x|ze−iωt .




(66)

The wave function hx|zi was previously obtained in eq. (52). We can now evaluate Cz by
requiring that hz|zi = 1. Using eq. (53), it follows that
 mω 1/4
exp − 14 (z + z ∗ )2 ,
 
Cz =
π~
where we have (conventionally) set the arbitrary phase of Cz to unity. Inserting this back
into eq. (52) yields
"  1/2 #
 mω 1/4 mω 2 2mω
hx|zi = exp − x + zx − 14 (z + z ∗ )2 . (67)
π~ 2~ ~

Replacing z with ze−iωt in eq. (67) and using the resulting expression in eq. (66) yields
"  1/2 #
 mω 1/4 mω 2mω
ψz (x, t) = e−iωt/2 exp − x2 + xze−iωt − 14 (ze−iωt + z ∗ eiωt )2 .
π~ 2~ ~

We now evaluate the squared magnitude of ψ(z, t) which yields the probability density
of the coherent state wave packet. Remarkably, it takes on a very simple form,
 mω 1/2 h mω i
2 2
|ψz (x, t)| = exp − (x − xc ) , (68)
π~ ~
where  1/2
~
ze−iωt + z ∗ eiωt .

xc ≡ (69)
2mω

20
We see that the center of the wave packet, x = xc , oscillates with the frequency of the
harmonic oscillator. But the shape of the wave packet is time-independent! In particular,
this wave packet does not spread. Indeed, we can easily repeat the calculation of part (d)
by taking expectation values with respect to |z, ti. The end result is the same, namely
∆X∆P = 21 ~, independently of the time t. Thus, the coherent state wave packet is a
minimum uncertainty wave packet for all times t.
Using the results of part (d), we can easily work out the expectation values of X and P
with respect to |z, ti,
 1/2
2~
hXi ≡ hz, t| X |z, ti = Re(ze−iωt ) , (70)

1/2
hP i ≡ hz, t| P |z, ti = 2m~ω Im(ze−iωt ) . (71)
It is convenient to write z = |z|eiδ . The above equations then take the form,
hXi = A cos(ωt − δ) , hP i = −mωA sin(ωt − δ) , (72)
where 1/2

2~
A≡ |z| . (73)

As expected from Ehrenfest’s theorem, the mean position and momentum of the coherent
state wave packet follows precisely the position and momentum of the classical harmonic
oscillator as a function of time.4 Moreover, for |z| ≫ 1, we have hHi = ~ω|z|2 [cf. eq. (64)],
which can be rewritten as
hHi ≃ 21 mω 2 A2 , for |z| ≫ 1,
which is the expected behavior of the classical harmonic oscillator (where quantum fluctu-
ations are negligible).5
Hence, the coherent state wave packet follows the classical motion of the harmonic
oscillator without changing its form. Moreover as suggested above, the relative size of the
quantum fluctuations vanish in the limit of |z| → ∞. For example, using eq. (65) and the
results of part (d), we easily compute
~
X 2 = A2 cos2 (ωt − δ) +

2
P = m2 ω 2A2 sin2 (ωt − δ) + 21 ~mω .


,
2mω
It then follows that6
2 ~ 2
(∆X)2 = X 2 − hXi = (∆P )2 = P 2 − hP i = 21 ~mω .



,
2mω
4
This should be compared with the expectation values of X and P with respect to the energy eigen-
states |ni. Indeed, in contrast to eqs. (70) and (71), hn| X |ni = hn| P |ni = 0 since X and P are parity-odd
operators and the |ni are states of definite parity. Moreover, these expectation values are time-independent
since the |ni are stationary states.
For the classical variables, H = P 2 /(2m)+ 21 mω 2 X 2 = 12 mω 2 A2 cos2 (ωt−δ)+sin2 (ωt−δ) = 21 mω 2 A2 .
5
 
6
As asserted below eq. (69), one immediately obtains ∆X∆P = 12 ~, independently of the time t.

21
Comparing ∆X to the amplitude A of harmonic motion [given by eq. (73)],

∆X 1
= ≪ 1,
A 2|z|

in the limit of |z| ≫ 1. Likewise,

∆P 1
= ≪ 1.
mωA 2|z|

Hence, the quantum fluctuations are suppressed in the limit of |z| ≫ 1.


A similar conclusion can be obtained by computing the root-mean square of the Hamil-
tonian.
2
H = hz| H 2 |zi = hz| a† a + 21 |zi = ~2 ω 2 hz| a† a† aa + 2a† a + 41 |zi

2 

2
= ~2 ω 2 |z|4 + 2|z|2 + 41 = hHi + ~2 ω 2 |z|2 ,
  

after making use of [a , a† ] = 1 and employing eq. (63). Hence, it follows that
2
∆H 2 = H 2 − hHi = ~2 ω 2|z|2 ,



and
∆H 1
≃ ≪ 1,
hHi |z|
in the limit of |z| ≫ 1.
In conclusion, a coherent state with |z| ≫ 1 is a very good approximation to the classical
limit of the harmonic oscillator.

22

You might also like