0% found this document useful (0 votes)
473 views23 pages

Chapter 9 Townsend Solutions

The document contains solutions to problems about quantum mechanics from the textbook "A Modern Approach to Quantum Mechanics" by Townsend. It discusses: 1) Applying translational and rotational operators to show that the commutator of momentum operators is equal to zero. 2) Deriving the position and momentum representation of quantum states in 3 dimensions. 3) Showing that the Hamiltonian operator commutes with the time evolution operator, demonstrating conservation of energy.

Uploaded by

sepi khandan
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
473 views23 pages

Chapter 9 Townsend Solutions

The document contains solutions to problems about quantum mechanics from the textbook "A Modern Approach to Quantum Mechanics" by Townsend. It discusses: 1) Applying translational and rotational operators to show that the commutator of momentum operators is equal to zero. 2) Deriving the position and momentum representation of quantum states in 3 dimensions. 3) Showing that the Hamiltonian operator commutes with the time evolution operator, demonstrating conservation of energy.

Uploaded by

sepi khandan
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 23

A Modern Approach to Quantum

Mechanics by Townsend - Solutions


Solutions by: GT SPS

Contributors: Charles Cardot, Cody McCord, Robin Glefke,


Kiernan Reising, Kevin Jin, Alexander George-Kennedy, William
Benston

Last Updated: January 3, 2021

Contents

9 Translational and Rotational Symmetry in the Two-Body Prob-


lem 2
9.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
9.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
9.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
9.4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
9.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
9.6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
9.7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
9.8 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
9.9 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
9.10 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
9.11 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
9.12 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
9.13 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
9.14 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
9.15 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
9.16 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
9.17 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
9.18 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
9.19 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
9.20 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
9.21 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
9.22 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
9.23 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

1
9 Translational and Rotational Symmetry in the
Two-Body Problem

9.1
First we expand both Tˆx Tˆy and Tˆy Tˆx , keeping only up to second order terms.

!
p̂2 a2 p̂2y a2y
 
ip̂x ax ip̂y ay
T̂x T̂y = 1− − x 2x 1− −
~ 2~ ~ 2~2
i ax ay 1
(p̂y ay + p̂x ax ) − 2 p̂x p̂y − 2 p̂2y a2y + p̂2x a2x + ...

=1−
~ ~ 2~

!
p̂2y a2y p̂2 a2

ip̂y ay ip̂x ax
T̂y T̂x = 1− − 1− − x 2x
~ 2~2 ~ 2~
i ay ax 1
(p̂x ax + p̂y ay ) − 2 p̂y p̂x − 2 p̂2x a2x + p̂2y a2y + ...

=1−
~ ~ 2~
Here we see that, when we subtract the two products, everything cancels out
except one term, which must equal zero.

ay ax ay ax
[T̂x , T̂y ] = T̂x T̂y − T̂y T̂x = − (p̂x p̂y − p̂y p̂x ) = − 2 [p̂x , p̂y ] = 0
~2 ~
This implies that [p̂x , p̂y ] = 0.

9.2
Z ∞
1 −ipx
hp|Ψ|p|Ψi = √ e ~ hx|Ψ|x|Ψi (6.57a)
−∞ 2~π
Z ∞
1 ipx
hx|Ψ|x|Ψi = √ e ~ hp|Ψ|p|Ψi (6.57b)
−∞ 2~π
In 3-dimensions,

|~ri = |x, y, zi , |~
pi = |px , py , pz i ,
   
1 −ipx x 1 −ipy y 1 −ipz z 1 −i~
p·~
r
h~r|~ pi = √
p|~r|~ e ~ √ e ~ √ e ~ = 3 e
~
2~π 2~π 2~π (2~π) 2
(9.25)

2
Therefore,
Z ∞ Z ∞
1 −i~
p·~
r
h~
p|Ψ|~
p|Ψi = h~
p|~r|~
p|~ri h~r|Ψ|~r|Ψi d~r = 3 e ~ h~r|Ψ|~r|Ψi d~r
−∞ −∞ (2~π) 2

Z ∞ Z ∞
1 p·~
i~ r
h~r|Ψ|~r|Ψi = h~r|~
p|~r|~
pi h~
p|Ψ|~
p|Ψi d~r = 3 e ~ h~
p|Ψ|~
p|Ψi d~r
−∞ −∞ (2~π) 2

9.3
Conservation of energy arises from invariance under time translations. Consider
−iĤt/~

h (t) = ei , the generator of finite time translations. We wish to show
Ĥ, Û (t) = 0.

" #
h
−iĤt/~
i X (−iĤt/~)k
Ĥ, e = Ĥ,
k!
k=0
h i

X Ĥ, (−iĤt/~)k
= =0
k!
k=0

It is clear thathĤ commutes


i with any exponential of itself times some scalars,
and so indeed Ĥ, Û (t) = 0.

9.4
To make life a little easier, let’s define two constants, ci = m1m
+m2 , where i = 1, 2
i

and c1 + c2 = 1. We now expand out P̂j , p̂j , X̂i , and x̂i . Note that R̂i → X̂i
and r̂i → x̂i .

P̂j = p̂1j + p̂2j


p̂j = c2 p̂1j − c1 p̂2j
X̂i = c1 x̂1i + c2 x̂2i
x̂i = x̂1i − x̂2i

3
[x̂i , P̂j ] = [x̂1i − x̂2i , p̂1j + p̂2j ]
= [x̂1i , p̂1j ] + [x̂1i , p̂2j ] + [−x̂2i , p̂1j ] + [−x̂2i , p̂2j ]
= [x̂1i , p̂1j ] − [x̂2i , p̂2j ]
= i~δij − i~δij
=0

[x̂i , p̂j ] = [x̂1i − x̂2i , c2 p̂1j − c1 p̂2j ]


= [x̂1i , c2 p̂1j ] + [x̂1i , −c1 p̂2j ] + [−x̂2i , c2 p̂1j ] + [−x̂2i , −c1 p̂2j ]
= [x̂1i , c2 p̂1j ] + [x̂2i , c1 p̂2j ]
= c2 i~δij + c1 i~δij
= i~δij

[X̂i , P̂j ] = [c1 x̂1i + c2 x̂2i , p̂1j + p̂2j ]


= [c1 x̂1i , p̂1j ] + [c1 x̂1i , p̂2j ] + [c2 x̂2i , p̂1j ] + [c2 x̂2i , p̂2j ]
= [c1 x̂1i , p̂1j ] + [c2 x̂2i , p̂2j ]
= c1 i~δij + c2 i~δij
= i~δij

[X̂i , p̂j ] = [c1 x̂1i + c2 x̂2i , c2 p̂1j − c1 p̂2j ]


= [c1 x̂1i , c2 p̂1j ] + [c1 x̂1i , −c1 p̂2j ] + [c2 x̂2i , c2 p̂1j ] + [c2 x̂2i , −c1 p̂2j ]
= [c1 x̂1i , c2 p̂1j ] + [c2 x̂2i , −c1 p̂2j ]
= c1 c2 i~δij − c1 c2 i~δij
=0

9.5

P̂ 2 = p̂21 + p̂1 p̂2 + p̂2 p̂1 + p̂22

2
p̂22 p̂2
  
µ µ p̂1 p̂2 p̂2 p̂1
p̂2 = p̂1 − p̂2 = µ2 2 − − + 12
m1 m2 m2 m1 m2 m1 m2 m1

4
P̂ 2 p̂2 1
p̂21 + p̂1 p̂2 + p̂2 p̂1 + p̂22

+ =
2M 2m 2M 
µ p̂22 p̂21

p̂1 p̂2 p̂2 p̂1
+ − − +
2 m22 m1 m2 m1 m2 m2
   1
1 µ 1 µ
= + p̂21 + + p̂22
2M 2m21 2M 2m22
   
1 µ 1 µ
+ − p̂1 p̂2 + − p̂1 p̂2
2M m1 m2 2M 2m1 m2

A little algebra reveals that the first two terms in parenthesis reduce to
1/2m1 and 1/2m2 , respectively, and the second two terms reduce to 0. Hence,
we have

P̂ 2 p̂2 p̂2 p̂2


+ = 1 + 2
2M 2m 2m1 2m2

9.6

9.7

X
L̂2 = L̂ · L̂ = L̂i L̂i
i
  !
X X
=  εijk x̂j p̂k  εilm x̂l p̂m
ijk ilm
!
X X
= εijk εilm x̂j p̂k x̂l p̂m
jklm i
X
= (δjl δkm − δjm δkl ) x̂j p̂k x̂l p̂m
jklm
X
= δjl δkm x̂j p̂k x̂l p̂m − δjm δkl x̂j p̂k x̂l p̂m
jklm

5
Before we continue we should notice two things. In the first expression, we
can rewrite x̂l p̂m as p̂m x̂l + i~δlm from the commutation relation. Also, we can
rearrange the second expression by pulling the second Kronecker delta into the
middle, sum over k,l (which equates to r̂ · p̂), and do the same thing with j,m.

δjm δkl x̂j p̂k x̂l p̂m = δjm x̂j (δkl p̂k x̂l ) p̂m
= δjm x̂j (r̂ · p̂) p̂m
= r̂ · p̂ (δjm x̂j p̂m )
2
= (r̂ · p̂)

X 2
= δjl δkm x̂j p̂k (p̂m x̂l + i~δlm ) − (r̂ · p̂)
jklm

X : δjk
 2
= δjl δkm x̂j p̂k p̂m x̂l i~δlm + i~
δil
δkm lm x̂j p̂k − (r̂ · p̂)
δ
jklm
X 2
= (δjl x̂j x̂l ) (δkm p̂k p̂m ) + i~(δjk x̂j p̂k ) − (r̂ · p̂)
jklm

From here it’s easy to see how the sums reduce all expressions.

2
L̂2 = r̂2 p̂2 − (r̂ · p̂) + i~ (r̂ · p̂)

9.8

First we note that L̂i = x̂j p̂k − x̂k p̂j and L̂j = x̂k p̂i − x̂i p̂k . Thus we have

[L̂i , L̂j ] = [x̂j p̂k − x̂k p̂j , x̂k p̂i − x̂i p̂k ]
= [x̂j p̂k , x̂k p̂i ] − [x̂j p̂k , x̂i p̂k ] − [x̂k p̂j , x̂k p̂i ] + [x̂k p̂j , x̂i p̂k ]
=A−B−C +D

9.9

6
9.10
a) What is the wavelength of a photon emitted in a vibrational
transition?

The energy of each photon is equal to the spacing of each energy, or the
transition energy:

hc
E = ~ω =
λ
Therefore the photon wavelength is given by:

eV
hc 1240 nm
λ= = = 3351 nm
E .37 eV

b) What is the effective spring constant k for this molecule?

Recall from classical mechanics that the angular frequency of this


vibration can be written:

r s
k k
w= =
m µ

Thus, combining this with Equation 9.152 (a) we can find the spring
constant to be:
 2
2 2πc
k=w µ= µ
λ
2
2π 3 × 108 ms

kg N
= 1.62 × 10−27 kg = 513 2 = 513
3351 × 10−9 m s m

c) What resolution is required for a spectrometer to resolve the


presence of H 35 Cl and H37 Cl molecules in the vibrational spec-
trum?

Following the derivation (Townsend 9.101 - 9.103) the generalized energy


spacing is:

! 12  21 
d2 V me e4
   
1 me
E=~ ∼
µ drr r=r0 MN (4π0 ~)2

7
 4

where m~e2e is the approximate energy scale of an electron. For each
molecule, the transition energy is:
1 
me e4
 
me 2
E H35 Cl ∼
MN (4π0 ~)2
  12
−31
(9.11 × 10 kg)
= −4 kg

(7.07×10 mol )
atoms
(6.02×1023 mol )
!
(9.11 × 10−31 kg)(1.602 × 10−19 C)4
× s2 C2 kgm2 2
(4π 8.85 × 10−12 kgm3 1.05 × 10−34 s )
kg m2
= 1.22 × 10−19
s2
= .764 eV

  12
−31
(9.11 × 10 kg)
E H37 Cl '  kg

(7.27×10−4 mol )
(6.02×1023 atoms
mol )
!
(9.11 × 10−31 kg)(1.602 × 10−19 C)4
× s2 C2 kgm2 2
(4π 8.85 × 10−12 kgm3 1.05 × 10−34 s )
kg m2
= 1.21 × 10−19
s2
= .754 eV

And now if we follow the same procedure as in (a), we can find the re-
spective wavelengths:
eV
hc 1240 nm
λ H35 Cl = = = 1621 nm
E .764 eV
eV
hc 1240 nm
λ H37 Cl = = = 1644 nm
E .754 eV
And the difference between these wavelengths is the minimum size required
to resolve transitions from both molecular isotopes.

∆λ = 23 nm

The resolution is then


λ (3351 nm)
= = 145
∆λ (23 nm)

8
9.11
a) Show that the population of rotational energy levels first in-
creases and then decreases with increasing l.

The degeneracy for each level increases as the energy increases. The
energy of each level depends on the quantum number l as:

l(l + 1)~2
El =
2I
where I is the moment of inertia of the molecule defined in terms of the
reduced mass (µ) and position of minimum potential (r0 ) as I = µr02 . We
can expand ?? as a power series about l such that

l2 5l3
 
nl 4
∼ (2l + 1) 1 − l − − − O(l )
n0 2 6
3l2 l3
=1+l− − − O(l4 )
2 6
Therefore at first order the ratio of states increases linearly with l, but
quickly higher order terms are much larger and negative, so the ratio will
rapidly decrease as l increases (see Figure 1).

Figure 1: Sketch of the ratio of molecules with energy El to energy E0 as a


function of l.

b) Which energy level will be occupied by the largest number of


molecules for HCl at room temperature? Compare your result
with the intensities of the absorption spectrum in Fig. 9.9. What
do you deduce about the temperature of the gas?

9
The maxima of Equation ?? will give the energy level with the most
number of molecules:

( 2 !)
~2 (1 + 2l)2 E0 − l(l+1)~
 
d nl 2I
= 2− exp =0
dl n0 2Ikb T kb T

Which has real solutions


1 Ikb T
l=− ± =
q 2 ~
m2 kg
1 (2.64 × 10−47 kg m2 )(1.38 × 10−23 s2 K )(293 K)
− ± m2 kg
2 1.054 × 10−34 s
' 3 and − 4

We only care about the positive solution since the quantum number l
must be greater than or equal to zero. As seen in Figure 2, the spectra
is centered on the zeroth transition with transition levels of the same
quantum number reflections of one another about this center. Therefore
it is clear that Figure 9.9 in the text has maximum transmission for the
l = 3 states. The HCl must then have a temperature of 293 K, which
which we have calculated to produce maxma at l = 3 transitions.

Figure 2: Total HCl spectra in the mid-infrared with both vibrational and
rotational components

9.12
The first four spherical harmonics (l = 0, 1) can be written in Cartesian
coordinates as

10
r
1
Y0,0 (θ, φ) =

r r
3 ±iφ 3 (x ± iy)
Y1,±1 (θ, φ) = ∓ e sin θ = ∓
8π 8π r
r r
3 3 z
Y1,0 (θ, φ) = cos θ =
4π 4π r
with linear combinations
r r
(Y1,−1 − Y1,1 ) 3 x i(Y1,1 + Y1,−1 ) 3 y
√ = √ =
2 4π r 2 4π r
If we sum together the Y1,0 harmonic and linear combinations in the correct
ratios we can produce the wave function
r r r r !
4π 3 z 3 y 3 z
Ψ(r) = (x + y + z)f (r) = + + rf (r)
3 4π r 4π r 4π r
r  
4π (Y1,−1 − Y1,1 ) i(Y1,1 + Y1,−1 )
= √ + √ + Y1,0 (θ, φ) rf (r)
3 2 2

r
4π  
= Y1,−1 (1 + i) + Y1,1 (i − 1) + 2Y1,0 (θ, φ) rf (r)
6
Now to normalize this wavefunction we must force it to follow the
normalization condition
Z ∞
∞|Ψ(r)|2 dr = 1

For which we simply choose any f (r) which satisfies that condition
L2 : Because Ψ(r) is made up of only Y1,m spherical harmonics, a measurement
of L2 could only yield l(l + 1)~2 = 2~2 with a 100% probability.
L̂z : Ψ(r) is made up of Y1,1 , Y1,0 , Y1,−1 spherical harmonics, which means that
the possible measurements of L̂z are m~ = ~, 0, and −~, each with a 1/3
probability
   r
1
Ψ(r)|L̂z |Ψ(r) = Ψ(r) L̂z Y1,−1 (1 + i)
6
r r 
1 1
+ Y1,1 (i − 1) + Y1,0 (θ, φ)
6 3
     
1 1 1
= ~− ~+ 0~
3 3 3

11
9.13
We can begin by finding the appropriate expectation values in terms of the
angular momentum raising and lowering operators

∆Lx = hl, m| (L̂x − hLx |Lx i) |l, mi


 
1 1
= hl, m| (L̂+ + L̂− ) − hl, m| (L̂+ + L̂− ) |l, mi |l, mi
2 2
∆Ly = hl, m| (L̂y − hLy |Ly i) |l, mi
 
1 1
= hl, m| (L̂+ − L̂− ) − hl, m| (L̂+ − L̂− ) |l, mi |l, mi
2i 2i
where,
p
L̂+ = ~ l(l + 1) − m(m + 1) |l, m + 1i
p
L̂− = ~ l(l + 1) − m(m − 1) |l, m − 1i
The state |l, mi must be a superposition for ∆Lz and ∆Ly to be non-zero and
to satisfy the uncertainty relation.
Or we can recall from Chapter 3 Section 5 that [Â, B̂] = iĈ implies
∆A∆B ≥ 12 | hC|Ci |. starting with the Schwartz inequality

hLx |Lx |Lx |Lx i hLy |Ly |Ly |Ly i ≥ | hLx |Ly |Lx |Ly i |2
where,
D E
hLx |Lx |Lx |Lx i = (∆Lx )2 hLy |Ly |Ly |Ly i = (∆Ly )2 hLx |Ly |Lx |Ly i = l, m|Ô|l, m l, m|Ô|l, m

F̂ iĜ
And because any operator can be written as Ô = 2 + 2 , assuming F̂ and Ĝ
have real expectation values, we find

1D E iD E2
| hLx |Ly |Lx |Ly i |2 = l, m|F̂ |l, m l, m|F̂ |l, m + l, m|L̂z |l, m l, m|L̂z |l, m
2 2
D E2 D E2
l, m|F̂ |l, m l, m|F̂ |l, m l, m|L̂z |l, m l, m|L̂z |l, m |hLz |Lz i|
2
= + ≥
4 4 2
Combining these definitions with the Schwartz inequality

~
(∆Lx )2 (∆Ly )2 ≥ | hLz |Lz i |2
4
~
∆Lx ∆Ly ≥ | hLz |Lz i |
4

12
9.14

9.15
Using the general solution to the spherical harmonics

s
(−1)l (2l + 1)(l + m)! imφ 1 dl−m
Yl,m (θ, φ) = l e sin2l θ
2 l! 4π(l − m)! sinm θ d(cos θ)(l−m)

Then applying the parity operator, which is achieved by inverting the


wavefunction coordinates, we find

s
(−1)l (2l + 1)(l + m)! im(φ+π) 1 dl−m
P̂ Yl,m (θ, φ) = l e m sin2l (π − θ)
2 l! 4π(l − m)! sin (π − θ) d(cos(π − θ))(l−m)
s
(−1)l (2l + 1)(l + m)! 1 dl−m
= l (−1)m eimφ m sin2l θ
2 l! 4π(l − m)! sin θ d(− cos θ)(l−m)
s
(−1)2l (2l + 1)(l + m)! imφ 1 dl−m
= l
e m sin2l θ = (−1)l Yl,m (θ, φ)
2 l! 4π(l − m)! sin θ d(cos θ)(l−m)

where,

eiπ = −1
cos(π − θ) = − cos θ
sin(π − θ) = sin θ

9.16
a) We already know that
r
3 iφ
Y1,1 (θ, φ) = − e sin(θ) = hθ, φ|1, 1|θ, φ|1, 1i , (9.152 (a))

 
~ ±iφ ∂ ∂
L̂± → e ±i − cot(θ) . (9.142)
i ∂θ ∂φ

13
Applying the lowering operator as instructed,
r  
D E ~ 3 −iφ ∂ ∂
θ, φ|L̂− |1, 1 θ, φ|L̂− |1, 1 = − e −i − cot(θ) eiφ sin(θ)
i 8π ∂θ ∂φ
r
~ 3 −iφ
=− e (−2i cos(θ))eiφ
i 8π
r
3
=~ cos(θ).

According to Eq. 9.147,



L̂− |1, 1i = 2~ |1, 0i .

Together with the above calculation, this implies that


r
3
2~Y1,0 (θ, φ) = ~ cos(θ)

r
3
⇒ Y1,0 (θ, φ) = cos(θ)

b)
In position space, the Laplacian L̂2 takes the form

∂2
   
1 ∂ ∂ 1
L̂2 → −~2 sin(θ) + (9.129)
sin(θ) ∂θ ∂θ sin2 (θ) ∂φ2

so applying this to the state |1, 1i we get


r  
ˆ 2 3 1 ∂ 1
sin(θ) cos(θ)eiφ + 2 iφ

2
hθ, φ| L |1, 1i = ~ (i e sin(θ))
8π sin(θ) ∂θ sin2 (θ)
r
− sin2 (θ) + cos2 (θ) iφ eiφ
  
2 3
=~ e −
8π sin(θ) sin(θ)
r 2 2
3 iφ − sin (θ) + cos2 (θ) − sin (θ) − cos2 (θ)
 
= ~2 e
8π sin(θ)
r
3 iφ
= −2~2 e sin(θ)

= 2~2 hθ, φ|1, 1|θ, φ|1, 1i

9.17

14
To begin this problem we will perform a substitution of u = cos θ in the
eigenvalue equation, where du = − sin θdθ. We can first expand the derivatives

δ δ δu δ p
= =− 1 − u2
δθ δu δθ δu
Which we can then substitute in the equation, with the second order term
ignored, in order to produce Legendre’s equation

  p 
−1 p δ p δ
− √ 1 − u2 − 1 − u2 1 − u2 Θλ,m (θ) = λΘλ,m (θ)
1 − u2 δu δu
  
δ δ
− (1 − u2 ) Θλ,m (θ) = λΘλ,m (θ)
δu δu
dΘλ,0 dΘλ,0
(1 − u2 ) − 2u + λΘλ,0 = 0
du2 du
Now to do the power series. The first and second derivatives of the possible
power series solution are

∞ ∞
dΘλ,0 X d2 Θλ,0 X
= (k + 1)ak+1 uk 2
= (k + 1)(k + 2)ak+2 uk
du du
k=0 k=0

Which we can now plug into the Legendre equation


X ∞
X ∞
X
(1 − u2 ) (k + 1)(k + 2)ak+2 uk − 2u (k + 1)ak+1 uk + λ a k uk = 0
k=0 k=0 k=0

X ∞
X
(k + 1)(k + 2)ak+2 uk − (k + 1)(k + 2)ak+2 uk+2
k=0 k=0

X ∞
X
−2 (k + 1)ak+1 uk+1 + λ ak uk = 0
k=0 k=0


X ∞
X
2a2 u0 + 6a3 u1 + (k + 1)(k + 2)ak+2 uk − (k − 1)(k)ak uk
k=2 k=2

X ∞
X
−2a1 u1 − 2 (k)ak uk + λa0 u0 + λa1 u1 + λ ak uk = 0
k=2 k=2


X
u0 (2a2 + λa0 ) + u1 (6a3 − 2a1 + λa1 ) + [(k + 1)(k + 2)ak+2
k=2
−(k − 1)kak − 2kak + λak ]uk = 0

15
We now have a representation of the coefficients of each u term, which we
know must equal zero. Thus we can set up a system of equations

2a2 + λa0 = 0
6a3 − 2a1 + λa1 = 0

X
[(k + 1)(k + 2)ak+2 − (k + 1)kak + λak ] = 0
k=2

We then know that your choice of a0 and a1 is arbitrary, but we can create a
chart of the rest of the solutions

k equation
2 12a4 − 6a2 + λa2 = 0
3 20a5 − 12a3 + λa3 = 0
4 30a6 − 20a4 + λa4 = 0
.̇ .̇

The series of l(l + 1) goes 1, 2, 6, 12, 20. And unless λ is one of these series, it
is not possible for all the the coefficients to be zero, and therefore the series
will diverge.

The solutions to this eigenvalue problem are in fact the Legendre polynomials

P0 = 1
P1 = cos θ
1
P2 = (cos2 θ − 1)
2
which have exactly the same form as the spherical harmonics, times their
respective normalizations which are a function of the degeneracy of their
energy level

r
1
Y0,0 (θ, φ) =

r
3
Y1,0 (θ, φ) = cos θ

r
5
Y2,0 (θ, φ) = (3 cos2 θ − 1)
16π

16
9.18
D E p
θ, φ|L̂− |l, l θ, φ|L̂− |l, l = ~ l(l + 1) − l(l − 1)Yl,l−1

= ~ 2lYl,l−1
r
~ (−1)l (2l + 1)! −iφ
 
∂ ∂
= e −i − cot(θ) (eilφ sinl (θ))
i 2l l! 4π ∂θ ∂φ
r
~ (−1)l (2l + 1)! −iφ
 
l−1 ilφ cos(θ) ilφ l
= e −il sin (θ) cos(θ)e − (ile ) sin (θ)
i 2l l! 4π sin(θ)
r
l
~ (−1) (2l + 1)! iφ(l−i)
−2il sinl−1 (θ) cos(θ)

= l
e
i 2 l! 4π
r
(−1)l (2l + 1)! iφ(l−i)
sinl−1 (θ) cos(θ)

= −2l~ l e
2 l! 4π
r
√ (−1)l (2l + 1)! iφ(l−i)
sinl−1 (θ) cos(θ)

⇒ Yl,l−1 = − 2l l e
2 l! 4π
We now want to compare this result to the general expression;
s
(−1)l (2l + 1)(l + m)! imφ dl−m
 
1 2l
Yl,m = l e sin (θ)
2 l! 4π(l − m)! sinm (θ) d(cos(θ))l−m
s !
d 2l
(−1)l (2l + 1)! iφ(l−1) 1 dθ sin (θ)
⇒ Yl,l−1 = l e
2 l! (2l)4π sinl−1 (θ) d
dθ cos(θ)
s !
(−1)l (2l + 1)! iφ(l−1) 1 2l sin2l−1 (θ) cos(θ)
= l e
2 l! (2l)4π sinl−1 (θ) − sin(θ)
r
√ (−1)l (2l + 1)! iφ(l−1)
= − 2l l e (sinl−1 (θ) cos(θ))
2 l! 4π
The two results agree, as expected.

9.19

9.20
a) What is hθ, φ|Ψ(t)i?

Recall that the time evolution of a wavefunction is given by:

iĤt/~
|Ψ(t)i = e |Ψ(0)i

17
We must then write this in terms of the spherical harmonics so we know
how to act the Hamiltonian upon the state. First, note the Euler
identity defines that

eiθ − e−iθ
sin θ =
2i
We can use this to express the wavefunction in terms of the spherical
harmonics

r r r
3 3 eiφ 3 e−iφ
hθ, φ|Ψ(0)i = sin θ sin φ = sin θ − sin θ
4π 4π 2i 4π 2i
r r
3 −iφ i sin θ 3 iφ i sin θ i(Y1,1 + Y1,−1 )
= e √ − e √ = √
8π 2 8π 2 2

This is now in a form we can act on with the Hamiltonian to produce the
state’s eigenenergies, and thus the time-evolved state


iL̂2 t/2I~ i(Y1,1 + Y1,−1 )
il(l+1)~t/2I
|Ψ(t)i = e |Ψ(0)i = e √
2
i i~t/I i i~t/I
=√ e |Y1,1 i + √ e |Y1,−1 i
2 2
The initial state is an energy eigenstate, so the state just picks up an
overall phase as time progress. For t = 0, we recover the state vector at
t = 0.
b) What values of Lz will be obtained if a measurement is carried
out and with what probability will these values occur?

Recall that L̂z acting on some state is given by

L̂z |Ψl,m i = m~ |Ψl,m i

Therefore for our system, a measurement will produce values of ±~ with


sign corresponding to the m value. The probabilities are given by

P = |cΨ |2

where cΨ is the coefficient of each state. Therefore, both measurements


have a 1/2 probability of occurring.

c) What is hLx i for this state?

18
Using bra-ket notation I will express the operator in terms of the raising
and lowering operators.
p
L̂+ |l, mi = (L̂x + iL̂y ) |l, mi = ~ l(l + 1) − m(m + 1) |l, m + 1i
p
L̂− |l, mi = (L̂x − iL̂y ) |l, mi = ~ l(l + 1) − m(m − 1) |l, m − 1i
1
L̂x |l, mi = (L̂− + L̂+ ) |l, mi
2
Therefore the expectation value is
 
1
hΨ|Lx |Ψi = Ψ (L̂− + L̂+ ) Ψ
2

~ p p 
= Ψ l(l + 1) − m(m − 1) + l(l + 1) − m(m + 1)
2
√ √
 
i i~t/I i i~t/I
√ e |Y1,1 i + √ e |Y1,−1 i = 2~ − 2~ = 0
2 2
d) If a measurement of Lx is carried out, what results(s) will be
obtained? With what probability?
We can use the expressions for states |1, mix

1 2 1
|1, 1ix = |Y1,1 i + |Y1,0 i + |Y1,−1 i
2 √ 2 √2
2 2
|1, 0ix = |Y1,1 i − |Y1,−1 i
2√ 2
1 2 1
|1, −1ix = |Y1,1 i − |Y1,0 i + |Y1,−1 i
2 2 2
to rewrite the state in terms making the measurement results more easily
calculable.

i i i i
√ |Y1,1 i + √ |Y1,−1 i = √ |1, 1ix + √ |1, −1ix
2 2 2 2
Recall that L̂x acting on some state is given by

L̂x |Ψl,m i = m~ |Ψl,m i


and the probabilities are given by

P = |cΨ |2
where cΨ is the coefficient of each state. Therefore such a measurement
would produce values ±~ with probability 1/2 for both.

19
9.21
We need to first rewrite hθ, φ|ψ(0)|θ, φ|ψ(0)i in terms of the orthogonal states
Yl,m .
r
3
hθ, φ|ψ(0)|θ, φ|ψ(0)i = sin (θ) sin (φ)

r
3 eiφ − eiφ
= sin (θ)
4π 2i
√ "r r #
2 3 iφ 3 −iφ
= e sin (θ) − e sin (θ)
2i 8π 8π

2
=i (Y1,1 + Y1,−1 )
√2
2
=i (|1, 1i + |1, −1i)
2
We need to figure out how the Hamiltonian behaves on both basis states.
Remember that L̂2 |l, mi = l(l + 1)~2 |l, mi and L̂z |l, mi = m~ |l, mi.
 
~
Ĥ |1, 1i = + ω0 |1, 1i
I
 
~
Ĥ |1, −1i = − ω0 |1, −1i
I

hθ, φ|ψ(t)|θ, φ|ψ(t)i = hθ, φ| Ĥ |ψ(0)i


√ 
2 −i()t 
=i e |1, 1i + e−i()t |1, −1i
√2
2 −i~t/I −iω0 t
|1, 1i + eiω0 t |1, −1i

=i e e *
2
√ " r r #
2 −i~t/I 3 i(φ−ω0 t) 3 −i(φ−ω0 t)
=i e − e sin (θ) + e sin (θ)
2 8π 8π
r
− e−i(φ−ω0 t)
 i(φ−ω0 t) 
3 −i~t/I e
= e sin (θ)
4π 2i

This last term just reduces to a cosine function, so we are left with
r
3 −i~t/I
hθ, φ|ψ(t)|θ, φ|ψ(t)i = e sin (θ) cos (φ − ω0 t)

To find the expectation value of the x-component of angular momentum, L̂x ,


we will refer to the equation marked with * from above. Remember also that
L̂x can be written in terms of the raising and lowering operators and takes the

20
form L̂x = L̂+ + L̂− . We√have that L̂+ |1, 1i = L̂− |1, −1i = 0 and
L̂− |1, 1i = L̂+ |1, −1i = 2~ |1, 0i.

  2 −i~t/I √ √ 
L̂+ + L̂− |ψ(t)i = i e 2~e−iω0 t |1, 0i + 2~eiω0 t |1, 0i
2
= i~e−i~t/I e−iω0 t + eiω0 t |1, 0i


The expectation value, hL̂x i, is the inner product of hψ(t)| with the last
expression. However, we see that this simply evaluates to 0 because
h1, 1|1, 0|1, 1|1, 0i = h1, −1|1, 0|1, −1|1, 0i = 0.

hL̂x i = 0

9.22
We first write the Hamiltonian in terms of L̂x , L̂y , and L̂z . However, remember
that L̂2 = L̂2x + L̂2y + L̂2z → L̂2x + L̂2y = L̂2 − L̂2z , meaning we can transform the
Hamiltonian into a form in which the eigenstates are known.

L̂2x L̂2y L̂2 L̂2 L̂2x + L̂2y L̂2 L̂2 − L̂2z


Ĥ = + + z = z + = z +
2I3 2I1 2I1 2I3 2I1 2I3 2I1

I1 − I3 1
Ĥ = α1 L̂2z + α2 L̂2 , where α1 = and α2 =
2I1 I3 2I1

In order to show that Ĥ and L̂z commute, we use a commutator identity and
the commutation relations of L̂2 and L̂z (i.e. L̂z commutes with itself and L̂2 ).

[Ĥ, L̂z ] = [α1 L̂2z + α2 L̂2 , L̂z ]


= α1 [L̂2z , L̂z ] + α2 [L̂2 , L̂z ]
=0

Because the Hamiltonian is composed of L̂2 and L̂2z , it shares eigenstates with
these two operators. Thus, we only need to know how L̂2 and L̂2z act on the
state vector.

Ĥ |l, mi = α1 L̂2z |l, mi + α2 L̂2 |l, mi


= α1 m2 ~2 |l, mi + α2 l(l + 1) |l, mi
= (α1 m2 ~2 + α2 l(l + 1)) |l, mi

The eigenstates of the Hamiltonian are |l, mi (or Yl,m , the Legendre polynomi-
als/spherical harmonics) and it’s eigenvalues are
l(l + 1) ~2
 
I1 − I3
El,m = + .
I1 I3 I1 2

21
To find the time evolved solution, we simply need to determine how the Hamil-
tonian acts on the individual basis states.
2 ~2
 
I1 − I3
Ĥ |0, 0i = 0 and Ĥ |1, 1i = + |1, 1i
I1 I3 I1 2
Finally, we have
α~
e− 2 t ~2
 
1 I1 − I3 2
|ψ(t)i = √ |0, 0i + √ |1, 1i , where α= +
2 2 I1 I3 I1 2

9.23
The exact position along the z-axis is arbitrary and the potential is only a
function of ρ, the distance from the central axis. Thus, we know that the
Hamiltonian must commute with the z-component of linear momentum, p̂z .
Additionally, since the potential is only dependent on the magnitude of
distance from the central axis, |ρ|, the Hamiltonian must be invariant under
rotations about this axis. Thus, the Hamiltonian should commute with the
z-component of the angular momentum, L̂z .
Because the Hamiltonian, z-component of angular momentum, and
z-component of linear momentum all commute, they must share simultaneous
eigenstates, |ψi = |E, m, pz i.

hφ|E, m, pz |φ|E, m, pz i = Φ(φ)


hρ|E, m, pz |ρ|E, m, pz i = R(ρ)
hz|E, m, pz |z|E, m, pz i = Z(z)

hρ, φ, z|E, m, pz |ρ, φ, z|E, m, pz i = R(ρ)Φ(φ)Z(z)
Consider Z(z) first (see Townsend chapter 6 section 6):

1 2
/2a2
Z(z) = p√ e−z
πa
The z-component of the wavefunction is simply a Gaussian wave-packet along
it’s central axis. Next, consider the angular part of the wavefunction, Φ(φ).

We can use the fact that L̂z |ψi = m~ |ψi and L̂z |ψi = ~i ∂φ |ψi.

~ ∂
hφ| L̂z |ψi → hφ| |ψi = hφ| m~ |ψi
i ∂φ
~ ∂
hφ|ψ|φ|ψi = m~ hφ|ψ|φ|ψi
i ∂φ

22
Solving this simple differential equation yields the angular equation,

Φ(φ) = eimφ

It will be useful to go ahead of find the second derivative of Φ(φ) for the next
part of the question.

∂2Φ
= −m2 Φ(φ)
∂φ2
Now we want to find the radial equation. For this we will consider
hρ, φ, z| Ĥ |ψi = E hρ, φ, z|ψ|ρ, φ, z|ψi. We can rewrite rewrite this equation in
the form that is presented on page 330, equation 9.130, which makes use of the
Laplacian.

hρ, φ, z| Ĥ |ψi = E hρ, φ, z|ψ|ρ, φ, z|ψi


p̂2
hρ, φ, z| + V (ρ̂) |ψi = ER(ρ)Φ(φ)Z(z)

 2 
~
− ∇2 + V (ρ̂) R(ρ)Φ(φ)Z(z) = ER(ρ)Φ(φ)Z(z)

We can now simply plug in for the cylindrical Laplacian.

 2
1 ∂2 ∂2
   
~ 1 ∂ ∂
− ρ + 2 2 + 2 + V (ρ̂) R(ρ)Φ(φ)Z(z) = ER(ρ)Φ(φ)Z(z)
2µ ρ ∂ρ ∂ρ ρ ∂φ ∂z

At this point we can separate out the z-component of the equation and
distribute the R(ρ)Φ(φ).

~2 1 ∂2
   
1 ∂ ∂
− Φ(φ) ρ R(ρ) + R(ρ) 2 2 Φ(φ) +R(ρ)Φ(φ)V (ρ̂) = ER(ρ)Φ(φ)
2µ ρ ∂ρ ∂ρ ρ ∂φ

Recall from above that ∂ 2 Φ/∂φ2 = −m2 Φ(φ), meaning we can substitute this
into our equation and factor/cancel out Φ(φ). This leaves us with the final
form of the radial equation,

 2
m2
 
~ 1 ∂ ∂
− ρ + 2 + V (ρ̂) R(ρ) = ER(ρ)
2µ ρ ∂ρ ∂ρ ρ

23

You might also like