Quantum Mechanics II Lecture Notes
Quantum Mechanics II Lecture Notes
Contents
tes
1 Quantum Dynamics 1
1.1 Time Dependence of Operators (Heisenberg Picture) . . . . . . . . . . . 1
1.1.1 Time Evolution Operator . . . . . . . . . . . . . . . . . . . . . . 1
1.1.2 Heisenberg Equation of Motion . . . . . . . . . . . . . . . . . . . 2
1.2 Evolution of x̂ and p̂ in the Harmonic Oscillator . . . . . . . . . . . . . . 3
1.2.1 Evolution of â and ↠. . . . . . . . . . . . . . . . . . . . . . . . . 3
No
1.2.2 Evolution of x̂ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2.3 Evolution of p̂ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Coherent States of the Harmonic Oscillator . . . . . . . . . . . . . . . . . 5
1.3.1 Expansion of Coherent States in terms of Number States . . . . . 5
1.3.2 Non Orthogonality . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.3.3 Expectation values of x̂ and p̂, and Uncertainty . . . . . . . . . . 8
1.3.4 Time Evolution of Coherent States . . . . . . . . . . . . . . . . . 10
1.4 Chapter Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
ure
2 Approximation Methods 12
2.1 The Variational Method . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.2 Perturbation Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2.1 Time Independent Perturbation Theory . . . . . . . . . . . . . . . 15
2.2.2 Time Dependent Perturbation Theory . . . . . . . . . . . . . . . 23
2.3 Scattering, Born Approximation and its Applications . . . . . . . . . . . 29
2.3.1 Scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.3.2 Born Approximation . . . . . . . . . . . . . . . . . . . . . . . . . 32
ct
3 Angular Momentum 46
3.1 Orbital Angular Momentum . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.1.1 Commutation Relations of Angular Momentum Components . . . 47
Le
tes
ˆ
3.3.1 Lowering and Raising Operators of J⃗ . . . . . . . . . . . . . . . . 63
3.4 Addition of Angular Momentum . . . . . . . . . . . . . . . . . . . . . . . 65
3.5 Chapter Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
No
Eigenfunctions) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
4.2 Quantum Numbers l, n and ml . . . . . . . . . . . . . . . . . . . . . . . 93
4.2.1 Principal Quantum Numbers (n) . . . . . . . . . . . . . . . . . . 93
4.2.2 Angular Momentum/Azimuthal/Orbital Quantum Number (l) . . 94
4.2.3 Magnetic Quantum Number ml . . . . . . . . . . . . . . . . . . . 95
4.3 The Spin Quantum Number (s) and the Pauli Exclusion Principle . . . . 95
4.3.1 The Spin Quantum Number (s) . . . . . . . . . . . . . . . . . . . 95
4.3.2 The Pauli Exclusion Principle . . . . . . . . . . . . . . . . . . . . 95
4.4 Spin-Orbit Coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
ure
4.5 The Zeeman Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
4.5.1 Anomalous Zeeman Effect . . . . . . . . . . . . . . . . . . . . . . 101
4.5.2 Strong Zeeman Effect . . . . . . . . . . . . . . . . . . . . . . . . . 102
4.6 Chapter Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
ii
References 127
tes
No
ct ure
Le
iii
1 Quantum Dynamics
tes
1.1.1 Time Evolution Operator
If the initial state of a system is given by |ψ(t0 )⟩ and after a time t, the state changes to
|ψ(t)⟩, we can examine how these states evolve in time.
The two states can be related by a linear operator Û (t, t0 ) such that;
No
|ψ(t)⟩ = Û (t, t0 ) |ψ(t0 )⟩ (1.1)
Û (t, t0 ) = Iˆ (1.2)
∂
iℏ |ψ(t)⟩ = Ĥ |ψ(t)⟩ (1.3)
∂t
∂ Û (t, t0 ) i
= − Ĥ Û (t, t0 ) (1.5)
∂t ℏ
If Ĥ doesn’t depend on time, integrating equation (1.5) gives;
Ĥ
Û (t, t0 ) = e−i(t−t0 ) ℏ (1.6)
Le
Ĥ
|ψ(t)⟩ = e−i(t−t0 ) ℏ |ψ(t0 )⟩ (1.7)
1
For t = t0 , Û (t0 , t0 ) = 1 hence the constant of integration must be equal to 1. Therefore,
the state vector at the finite time t is related to the state vector at t = 0 by the equation;
Ĥ
|ψ(t)⟩ = eit ℏ |ψ(0)⟩ (1.8)
tes
Note that Û (t, t0 ) is a unitary operator since Û (t, t0 )Û † (t, t0 ) = Iˆ
If Ĥ depends on time, the integration of equation (1.5) becomes less trivial.
No
Ĥt Ĥt
Â(t) = ei ℏ Ae−i ℏ (1.9)
iĤ iÂ
= Â − Ĥ (1.12)
ℏ ℏ
dÂ(t) ih i
= Ĥ, Â(t) (1.13)
dt ℏ
This is the Heisenberg equation of motion. It leads to equations that are formally quite
simillar to the corresponding classical equations.
ct
Heisenberg discovered that it is possible to describe nature with classical equations but
we must realise that [x̂, p̂] = iℏ and it is this little correction to the classical way of doing
things that leads to wonderful phenomena of Quantum Mechanics.
Le
2
1.2 Evolution of x̂ and p̂ in the Harmonic Oscillator
1.2.1 Evolution of â and â†
tes
† 1
Ĥ = ℏω â â + (1.14)
2
No
Equation (1.15) is the Hamiltonian equation for the Harmonic oscillator in the Heisenberg
picture. Further more, the commutation relations of â and ↠with one another and with
Ĥ retain the same form;
â(t), ↠(t) = 1
(1.16)
h i
Ĥ, â(t) = −ℏωâ(t) (1.17)
h i
Ĥ, ↠(t) = ℏω↠(t) (1.18)
d↠(t)
= iω↠(t) (1.20)
dt
Solving equation (1.20) gives;
↠(t) = eiωt ↠(0) (1.21)
ct
3
1.2.2 Evolution of x̂
iĤt iĤt
x̂e−
tes
x̂(t) = e ℏ ℏ (1.25)
Using,
h i 1 h h ii 1 h h h iii
e B̂e− = B̂ + Â, B̂ + Â, Â, B̂ + Â, Â, Â, B̂ + ... (1.26)
2! 3!
2 h h
it h i 1 it ii
x̂(t) = x̂ + Ĥ, x̂ + Ĥ, Ĥ, x̂ + ... (1.27)
ℏ 2! ℏ
No
t (ωt)2 (ωt)3 1 (ωt)4 (ωt)5 1
= x̂ + p̂ − x̂ − p̂ + x̂ + p̂ + ... (1.28)
m 2! 3! mω 4! 5! mω
(ωt)2 (ωt)4 (ωt)3 (ωt)5
1
= x̂ 1 + + + ... + p̂ (ωt) − + + ... (1.29)
2! 4! mω 3! 5!
1
x̂(t) = x̂ cos ωt + p̂ sin ωt (1.30)
mω
dx̂(t)
For dt
,
dx̂(t) ih i
= Ĥ, x̂(t) (1.31)
dt ℏ
ure
i itĤ h i −itĤ
= e ℏ Ĥ, x̂ e ℏ (1.32)
ℏ
dx̂(t) 1
= p̂(t) (1.33)
dt m
1.2.3 Evolution of p̂
2 h h
it h i 1 it ii
p̂(t) = p̂ + Ĥ, p̂ + Ĥ, Ĥ, p̂ + ... (1.34)
ℏ 2! ℏ
−iℏmω 2
dp̂(t) iĤt iĤt
= e ℏ x̂e− ℏ (1.39)
dt iℏ
tes
dp̂(t)
= −mω 2 x̂(t) (1.40)
dt
No
â |α⟩ = α |α⟩ (1.41)
Coherent states play an important role in quantum optics, especially in laser physics. By
giving a good overview of coherent states of laser beams, the states describing a laser
beam can be briefly characterized as having;
2. A precisely defined phase in contrast to a state with fixed particle number when
ure
the phase is completely random.
A coherent state |α⟩ can be expressed in terms of the number states |n⟩ by;
∞
X
|α⟩ = cn |n⟩ (1.42)
n=0
ct
5
We obtain;
∞
X √
αcn − cn+1 n + 1 |n⟩ = 0 (1.45)
n=0
tes
√
αcn − cn+1 n + 1 = 0 (1.46)
√
αcn = cn+1 n + 1 (1.47)
No
α3
α
c3 = c = c
√ √
3 2 6 0
Generally,
αn
cn = √ c0 (1.48)
n!
Equation (1.42) becomes;
∞
X αn
|α⟩ = c0 √ |n⟩ (1.49)
n=0 n!
Applying the normalizing condition ⟨α|α⟩ gives;
ure
X (α∗ )m αn
|c0 |2 √ ⟨m|n⟩ = 1 (1.50)
n,m m!n!
X |α|2n
|c0 |2 =1 (1.51)
n
n!
2
|c0 |2 e|α| = 1 (1.52)
2
|c0 |2 = e−|α| (1.53)
ct
This gives,
1 2
c0 = e− 2 |α| ei(arbitraryphase) (1.54)
1 2
c0 = e− 2 |α| (1.55)
6
Therefore, equation (1.49) becomes;
∞
− 12 |α|2
X αn
|α⟩ = e √ |n⟩ (1.56)
n n!
tes
Where α ranges freely on the complex plane.
Equation (1.57) gives the probability for obtaining a particular energy level n when the
system is in a quantum coherent state.
< n >n
Pα (n) = | ⟨n|α⟩ |2 = e−<n> (1.57)
n!
No
Where < n >= ⟨α| ↠â |α⟩ = |α|2 . Also, ∆n2 = |α|2 . This shows that the coherent states
have a poissonian distribution.
Coherent states |α⟩ do not form a proper basis, since they are eigen vectors of a non-
Hermitian operator. In particular, they are not orthogonal.
Let us compute the inner product of two coherent states |α⟩ and |β⟩
ure
∞
|α|2 +|β|2
− 2
X (α∗ )n β m
⟨α|β⟩ = e √ ⟨n|m⟩ (1.58)
n,m m!n!
∞
|α|2 +|β|2
− 2
X (α∗ β)n
=e (1.59)
n
n!
|α|2 +|β|2 −2α∗ β
− 2
⟨α|β⟩ = e (1.60)
ct
2
| ⟨α|β⟩ |2 = e−|α−β| (1.62)
Le
7
Because of this close relation, any state can be written in terms of coherent state super-
position.
tes
For the coherent state |α⟩, the expectation value of x̂ is defined by;
Recall that, r
ℏ
â + â†
x̂ = (1.65)
2mω
No
Therefore,
λ
< x̂ >= √ ⟨α| â + ↠|α⟩
(1.66)
2
q
ℏ
Where λ = mω
λ
= √ (α ⟨α|α⟩ + α∗ ⟨α|α⟩) (1.68)
2
λ
< x̂ >= √ (α + α∗ ) (1.69)
2
√
< x̂ >= 2λ Re{α} (1.70)
ct
Recall that,
Le
r
mℏω †
p̂ = i â − â (1.72)
2
Therefore,
iℏ
< p̂ >= √ ⟨α| ↠− â |α⟩
(1.73)
λ 2
8
iℏ
= √ ⟨α| ↠|α⟩ − ⟨α| â |α⟩
(1.74)
λ 2
iℏ
= √ (α∗ ⟨α|α⟩ − α ⟨α|α⟩) (1.75)
λ 2
tes
iℏ
= √ (α∗ − α) (1.76)
λ 2
iℏ
= − √ (α − α∗ ) (1.77)
λ 2
√
i 2ℏ
< p̂ >= − (α − α∗ ) (1.78)
2λ
√
No
i 2ℏ
< p̂ >= Im{α} (1.79)
λ
Equations (1.69) and (1.78) gives;
1 1 λ
α= √ < x̂ > +i < p̂ > (1.80)
2 λ ℏ
λ2 2
x̂ 2
= ⟨α| â + ↠|α⟩ (1.81)
2
ure
λ2
⟨α| â2 + 2↠â + 1 + â†2 |α⟩
= (1.82)
2
λ2
⟨α| â2 |α⟩ + ⟨α| 2↠â + 1 |α⟩ + ⟨α| â†2 |α⟩
= (1.83)
2
λ2 2
α + 2α∗ α + 1 + α∗2
= (1.84)
2
λ2
x̂2 = (α + α∗ )2 + 1
(1.85)
2
ct
q
From ∆x = ⟨x̂2 ⟩ − ⟨x̂⟩2 ,
s 2
λ2 λ
(α + α∗ )2 + 1 +
∆x = √ (α + α∗ ) (1.86)
2 2
Le
λ
∆x = √ (1.87)
2
Exactly the same variance as the ground state |n = 0⟩.
9
For ⟨p̂2 ⟩,
−ℏ2 † 2
p̂2 =
⟨α| â − â |α⟩ (1.88)
2λ2
−ℏ2
â2 − 2↠â − 1 + â†2 |α⟩
= (1.89)
tes
2λ2
−ℏ2 2
α − 2α∗ α − 1 + α∗2
= 2
(1.90)
2λ
−ℏ2
p̂2 = (α − α∗ )2 − 1
2
(1.91)
2λ
q
But ∆p = ⟨p̂2 ⟩ − ⟨p̂⟩2 ,
No
√
v !2
u
u −ℏ2 − 2iℏ
∆p = t 2 (α − α∗ )2 − 1 −
(α − α∗ ) (1.92)
2λ 2λ
ℏ
∆p = √ (1.93)
λ 2
The uncertainty relation ∆x∆p is given by;
λ ℏ
∆x∆p = √ . √ (1.94)
2 λ 2
ure
ℏ
∆x∆p = (1.95)
2
We can see that all coherent states (no matter what complex values α takes on) are
minimum uncertainty states.
∞
−|α|2 X an
|ψ(0)⟩ = e 2 √ 0 |n⟩ (1.96)
n=0 n!
∞
−|α|2 an 1
√ 0 e−iω(n+ 2 )t |n⟩
X
|ψ(t)⟩ = e 2 (1.97)
n!
Le
n=0
∞
−iωt −|α|2 X an
=e 2 e 2 √ 0 e−iωnt |n⟩ (1.98)
n=0 n!
∞ n
−iωt −|α|2 X (a0 e−iωt )
|ψ(t)⟩ = e 2 e 2 √ |n⟩ (1.99)
n=0 n!
10
Let α(t) = α0 e−iωt
ψ(t) = |α(t)⟩ (1.100)
By this, it remains in a coherent state. But, the value of parameter α changes in time.
tes
1.4 Chapter Exercises
1. Show that coherent states are minimum uncertainty states
2. Evaluate |⟨β|α⟩|2 for coherent states |α⟩ and β. Show that the states |α⟩ and β
become completely orthogonal in the limit |α − β| >> 1.
No
3. A coherent state of a one-dimensional harmonic oscillator is defined to be the eigen-
state of the (non-Hermitian) annihilation operator â;
â |α⟩ = α |α⟩
11
2 Approximation Methods
Most problems in quantum mechanics can not be solved exactly. Exact solutions of the
Schrodinger equation exists only for a few idealized systems hence a need for approxima-
tes
tion methods.
In this chapter, we are going to discuss the variational method, the perturbation theory,
the Born approximation and partial waves analysis approximation methods of solving
quantum systems.
No
2.1 The Variational Method
The variational method provides an estimation of the eigenvalues of the ground state of
a system for which one has only a qualitative idea about the form of the wave function.
This method is based on a simple fact that the average energy of the Hamiltonian Ĥ in
a state represented by an arbitrary wave function has to be greater than or equal to the
ground state energy of the system.
X
ψ= an ϕn (2.3)
n
X
|ψ⟩ = an |ϕn ⟩ (2.4)
n
Le
If |ψ⟩ is not normalized, the expectation value of Ĥ in the state is given as;
D E ⟨ψ| Ĥ |ψ⟩
Ĥ = (2.5)
⟨ψ|ψ⟩
12
D E P |a |2 E
n n n
Ĥ = P 2
(2.6)
n |an |
tes
|an |2
D E P
Ĥ = E0 Pn 2
= E0 (2.7)
n |an |
1. Choose judiciously a trial wave function that contain a whole bunch of free param-
eters αi ;
ψ = ψ (α1 , α2 , ....) (2.8)
No
2. Vary the parameter αi until the expectation value of Ĥ is minimized;
∂ < Ĥ >
(2.9)
∂αi
The minimum value determined in this way provides an upper bound to E0 , which will
be close to the actual value if the trial function has a form closely resembling that of the
actual ground state function ψ0 .
ure
Example 2.10
2 ℏ2 d2
Consider a trial function ψ = Ae−αx for a harmonic oscillator Hamiltonian Ĥ = − 2m dx2
+
1 2
2
kx where α is the adjustable parameter. Use the variational method to calculate the
ground state energy.
−∞ −∞
r
2 π
A =1 (2.11)
2α
14
2α
Le
A= (2.12)
π
D E
We now compute Ĥ
D E
Ĥ = ⟨ψ| Ĥ |ψ⟩ (2.13)
13
21 Z ∞
ℏ2 d2
2α −αx2 1 2 −αx2
= e − + kx e dx (2.14)
π −∞ 2m dx2 2
D E ℏ2 α mω 2
Ĥ = + (2.15)
2m 8α
tes
We now minimize Ĥ by differentiating equation (2.15) with respect to the parameter α
and equating it to zero.
d < Ĥ > ℏ2 mω 2
= − (2.16)
dα 2m 8α2
mω
α= (2.17)
2ℏ
Substituting equation (2.17) into equation (2.15) gives;
No
ℏ2 mω mω 2
D E 2ℏ
Ĥ = + (2.18)
min 2m 2ℏ 8 mω
1 1
= ℏω + ℏω (2.19)
4 4
D E 1
Ĥ = ℏω (2.20)
min 2
To tell the appropriateness of the variational method above, we recall that the energy of
the harmonic oscillator is quantized and it is given by;
ure
1
En = ℏω n + , n = 0, 1, 2, 3, ... (2.21)
2
It should be noted that while the variational method exactly determines the ground state
energy above, it is not always the case. It remains an approximation which is exact foe
ct
It is also necessary to note that the choice of the form of the trial wave function is very
important and decisive in the appropriateness of the approximation of the variational
method.
Le
14
2.2 Perturbation Theory
Perturbation theory is based on the assumption that the problem we wish to solve is in
some sense only slightly different from a problem that can be solved exactly.
tes
In the case where the deviation between the two problems is small, perturbation is suit-
able fro calculating the contribution associated with this deviation; this contribution is
then added as a correction to the energy and the wave function of the exactly solvable
Hamiltonian.
Therefore, perturbation theory builds on the known exact solutions to obtain approximate
No
solutions.
Where Ĥ1 is very small compared to Ĥ0 . Ĥ0 is the Hamiltonian of the unperturbed
system and λĤ1 is a small perturbation. λ is a real parameter (0 ≤ λ ≤ 1) used to
ure
distinguish between the different orders of perturbation.
Where |ψn ⟩ are the eigen states of the complete Hamiltonian Ĥ and En are the corre-
sponding eigenvalues.
ct
From equation (2.23), it follows that we are to consider two separate cases depending on
whether the exact solutions of Ĥ0 are non degenerate or degenerate.
(0)
For non degenerate
E energy eigenvalues, for every En , there corresponds only one eigen
(0)
state ψn
Ĥ0 ψn(0) = En(0) ψn(0) (2.24)
15
E
(0) (0)
Where the exact eigenvalues En and eigen functions ψn are known.
The basic idea of the perturbation is assuming that the perturbed eigenvalues and eigen
states can both be expanded in power series in the parameter λ.
tes
En = En(0) + λEn(1) + λ2 En(2) + ..., (2.25)
No
Substituting equation (2.25) and (2.26) in equation (2.23) gives;
ψn(0) + λ ψn(1) + λ2 ψn(2) + ... =
Ĥ0 + λĤ1
(2.27)
En(0) + λEn(1) + λ2 En(2) + ... ψn(0) + λ ψn(1) + λ2 ψn(2) + ...
Zero order
ure
λ0 ; Ĥ0 ψn(0) = En(0) ψn(0) (2.28)
First order
Second order
λ2 ; Ĥ0 ψn(2) + Ĥ1 ψn(1) = En(0) ψn(2) + En(1) ψn(1) + En(2) ψn(0)
ct
(2.30)
Third order
λ3 ; Ĥ0 ψn(3) + Ĥ1 ψn(2) = En(0) ψn(3) + En(1) ψn(2) + En(2) ψn(1) + En(3) ψn(0) (2.31)
Le
Ĥ0 ψn(j) + Ĥ1 ψn(j−1) = En(0) ψn(j) + En(1) ψn(j−1) + .... + En(j) ψn(0) (2.32)
E
(1) (2) (1)
Let us now proceed to determine the eigenvalues En , En and ψn from equations
16
(2.29) and (2.30).
tes
To obtain the first order
D correction to the energy eigenvalues En , we simply both sides of
(0)
equation (2.29) by ψn .
ψn(0) Ĥ0 ψn(1) + ψn(0) Ĥ1 ψn(0) = ψn(0) En(0) ψn(1) + ψn(0) En(1) ψn(0) (2.33)
ψn(0) Ĥ0 ψn(1) + ψn(0) Ĥ1 ψn(0) = En(0) ψn(0) ψn(1) + En(1) ψn(0) ψn(0) (2.34)
No
ψn(0) Ĥ1 ψn(0) = En(1) (2.35)
D E D E D E
(0) (0) (0) (1) (0) (0) (1)
Since ψn ψn , ψn Ĥ0 ψn and En ψn ψn are both equal to zero. Equation
(2.35) states the first order correction to the energy En ; This implies that the first order
correction is just the expectation value of the perturbing interaction potential in the state
under correction.
!
X
ψn(1) = (0)
ψm (0)
ψm ψn(1) (2.37)
m
ct
X
ψn(1) = (0)
ψm ψn(1) (0)
ψm (2.38)
m̸=n
D E
(0) (1)
The term m = n does not contribute since ψm ψn = 0.
Le
D E
(0) (1)
The coefficient ψm ψn can be inferred from equation (2.29) by multiplying both sides
D
(0)
by ψm ;
D E
(0) (0)
ψm Ĥ1 ψn
(0)
ψm ψn(1) = (0) (0)
(2.39)
En − Em
17
Substituting equation (2.39) into (2.38) gives;
D E
(0) (0)
X ψm Ĥ1 ψn
ψn(1) = (0) (0)
(0)
ψm (2.40)
En − Em
tes
m̸=n
Therefore, the eigen functions |ψn ⟩ of Ĥ to first order in the λĤ1 can be obtained by
substituting equation (2.40) into equation (2.26)
D (0) (0)
E
X ψm Ĥ1 ψn
|ψn ⟩ = ψn(0) +
(0) (0)
(0)
ψm (2.41)
m̸=n En − Em
No
Second order correction
(2)
To obtain the second
D order correction En to the energies En , we multiply both sides of
(0)
equation (2.30) by ψn .
ψn(0) Ĥ0 ψn(2) + ψn(0) Ĥ1 ψn(1) = ψn(0) En(0) ψn(2) + ψn(0) En(1) ψn(1) + ψn(0) En(2) ψn(0)
(2.42)
(0) (1) (2) (0) (0)
ψn Ĥ1 ψn = En ψn ψn (2.43)
ure
En(2) = ψn(0) Ĥ1 ψn(1) (2.44)
The eigen energy to second order in λĤ1 is obtained by substitution equation (2.45) into
equation (2.25)
ct
D E 2
(0) (0)
X ψm Ĥ1 ψn
En = En(0) + λ ψn(0) Ĥ1 ψn(0) + λ2 (0) (0)
+ ... (2.46)
m̸=n En − Em
Le
E
(2)
We could go on to calculate the second order correction to the state vector ψn , the
third order correction to the energy and so on, but in practice, equation (2.45) is ordi-
narily as far as it is useful to peruse this method.
18
Note; For perturbation theory to work, the corrections it produces must be small.
D E 2
(0) (0)
ψm λĤ1 ψn
(0) (0)
≪ 1, (m ̸= n) (2.47)
tes
En − Em
(0) (0)
If the unperturbed energy levels En and Em were equal(degenerate), then the condition
(2.47) would break down.
Example 2.11 E q
(0) 2 nπ
The unperturbed wave functions for the finite square well are ψn = a
sin a
x .
No
Suppose we perturb the system by simply raising the floor of the well by a constant
amount V0 .
ure
Figure 2.1: Perturbed Finite Square Potential Well
En(1) = V0 (2.51)
(2.52)
They are simply lifted by the amount V0 . The only surprising thing is that in this case,
the first order yields the exact answer.
19
Example 2.12
Calculate the first order correction to the ground state energy of an anharmonic oscil-
lator of mass m and angular frequency ω subjected to a potential V (x) = 12 mω 2 x2 +
bx4 , where b is a parameter independent of x. The ground state wave function is
tes
−mωx2
1
E
(0)
ψ0 = mω
πℏ
4
e 2ℏ .
Sln;
(1)
E0 = ψn(0) Ĥ1 ψn(0) (2.53)
(1)
E0 = ψn(0) bx4 ψn(0) (2.54)
No
mω 21 Z ∞
−mωx2
4
= b xe ℏ
(2.55)
πℏ −∞
Using;
Z ∞ √
4 3 π 12
x exp(−ax )dx = 5 ,
0 8 a2
mω 21 3√π 1
(1)
E0 = b .2. 5 (2.56)
πℏ 8 a2
(1) 3bℏ2
ure
E0 = (2.57)
4m2 ω 2
Example 2.13
A particle of mass m0 and charge q oscillates a long the x-axis in a one-dimensional har-
monic potential with an angular frequency ω. If an electric filed ϵ is applied along the
x-axis, evaluate the first and second order corrections to the energy of the nth state.
The potential energy due to the field is V = −qϵx̂ and hence the perturbation is Ĥ1 = qϵx̂.
(1)
The first order correction En ,
ct
21
ℏ
En(1) = qϵ ⟨n| â + ↠|n⟩ = 0 (2.60)
2m0 ω
En(1) = 0 (2.61)
20
(2)
The second order correction En ,
2
X ⟨m| Ĥ1 |n⟩
En(2) = (0) (0)
(2.62)
tes
m̸=n En − Em
21
But ⟨m| Ĥ1 |n⟩ = qϵ ⟨m| â + ↠|n⟩. Here, m can take all integral values except
ℏ
2m0 ω
n. Therefore, the non vanishing terms of m are (n + 1) and (n − 1).
(0) √
Since En = n + 21 ℏω and using the relations ⟨n + 1| â + ↠|n⟩ = n + 1 and
√ (0) (0) (0) (0)
⟨n − 1| â + ↠|n⟩ = n, also En − En−1 = ℏω, En − En+1 = −ℏω.
No
2 2
!
q 2 ϵ2 ℏ ⟨n + 1| â + ↠|n⟩ ⟨n − 1| â + ↠|n⟩
En(2) = (0) (0)
+ (0) (0)
(2.63)
2m0 ω En − En+1 En − En−1
√ 2 √ !
q 2 ϵ2 ℏ n+1 ( n)
= + (2.64)
2m0 ω −ℏω ℏω
q 2 ϵ2 ℏ n + 1
n
= + (2.65)
2m0 ω −ℏω ℏω
q 2 ϵ2 ℏ
= (−n − 1 + n) (2.66)
ure
2m0 ω 2 ℏ
q 2 ϵ2
En(2) = − (2.67)
2m0 ω 2
Hence the energy is given to the second order by;
q 2 ϵ2
1
En = n + ℏω − (2.69)
2 2m0 ω 2
ct
Example 2.14
A particle of mass m moves in a one-dimensional potential well defined by;
0
f or − 2a < x < −a, a < x < 2a
Le
Treating V0 for −a < x < a as perturbation on the flat bottom box V (x) = 0 for
−2a < x < 2a and V (x) = ∞ otherwise, calculate the energy of the ground state cor-
rected up to the first order.
21
Sln;
The unperturbed energy and wave function of the ground state are;
tes
(0) π 2 ℏ2 (0) 1 πx
E1 = , ψ = √ cos
32ma2 1 2a 4a
(1)
The first order correction En ,
No
D E
(1) (0) (0)
E1 = ψ1 V0 ψ1 (2.71)
Z a 2
(1) 1 πx
E1 = V0 √ cos2 dx (2.72)
−a 2a 4a
V0 a 1
Z πx
= 1 + cos dx (2.73)
2a −a 2 2a
V0 a V0 2a πx
= x|−a + . sin |a (2.74)
4a 4a π 2a −a
(1) V0 V0 1 1
E1 = + = V0 + (2.75)
ure
2 π 2 π
Therefore, the corrected ground state energy is;
π 2 ℏ2
1 1
E1 = + V0 + (2.76)
32ma2 2 π
ct
Le
22
2.2.2 Time Dependent Perturbation Theory
We now suppose a time dependent external perturbation Ĥ1 (t) to the unperturbed Hamil-
tonian Ĥ0 which yields a total time dependent Hamiltonian Ĥ(t) of a system.
tes
Ĥ(t) = Ĥ0 + λĤ1 (t) (2.77)
Let the eigen states of Ĥ0 take the form |ψn ⟩ such that;
If |ψn (t)⟩ denote the quantum state of the perturbed system at time t, the corresponding
No
Schrodinger equation is,
∂ |ψn (t)⟩
iℏ = Ĥ0 + λĤ1 (t) |ψn (t)⟩ (2.79)
∂t
How does Ĥ1 affect the system?. When the system interacts with Ĥ1 , it either absorbs
or emits energy. This process inevitably causes the system to undergo transitions from
one unperturbed state to another.
The main aim of the time dependent perturbation theory consists of answering this ques-
ure
tion; If the system is initially in an unperturbed state |ψn ⟩ of Ĥ0 , what is the probability
that the system will be found at a later time in another unperturbed state |ψm ⟩.
We now need to look for solutions to equation (2.79) in order to prepare ground for an-
swering this question.
The method to solve equation (2.79) is to expand |ψn (t)⟩ in terms of the expansion
coefficients cn (t).
ct
X iEn t
|ψn (t)⟩ = cn (t)e− ℏ |ψn ⟩ (2.80)
n
X
dcn (t) − iEn t
cn (t)En + iℏ e ℏ |ψn ⟩ =
Le
n
dt
(2.81)
X iEn t
cn (t)e− ℏ En + λĤ1 |ψn ⟩
n
23
Equating the right hand sides of equation (2.81) gives;
tes
Multiplying equation (2.82) by ⟨ψm | from the left gives;
X iEn t X iEn t
iℏ ċn (t) ⟨ψm |ψn ⟩ e− ℏ =λ cn (t) ⟨ψm | Ĥ1 (t) |ψn ⟩ e− ℏ (2.83)
n n
X iEm t X iEn t
No
iℏ ċm (t)e− ℏ =λ cn (t)Ĥmn (t)e− ℏ (2.84)
m n
X X −i(En −Em )t
iℏ ċm (t) = λ cn (t)Ĥmn (t)e ℏ (2.85)
m n
X
iℏċm (t) = λ cn (t)Ĥmn (t)eiωmn t (2.86)
n
(Em −En )
Where Ĥmn (t) = ⟨ψm | Ĥ1 (t) |ψn ⟩ and ωmn = ℏ
. The quantity ωmn is called the
Born angular frequency.
ure
Equation (2.86) is a set of coupled first order differential equations similar to the Schrodinger
equation.
Assuming that the perturbation λĤ1 is weak, we can expand the coefficients cn in powers
of the parameter λ as;
Substituting equation (2.87) into equation (2.86) and equating corresponding powers of
λ gives;
ċ(0)
m = 0 (2.88)
1 X
ċ(1)
m = Ĥmn (t)eiωmn t c(0)
n (2.89)
iℏ n
Le
1 X
ċ(2)
m = Ĥmn (t)eiωmn t c(1)
n (2.90)
iℏ n
Generally,
1 X (s)
ċ(s+1)
m = Ĥmn (t)eiωmn t ck (2.91)
iℏ n
24
These equations can now in principle be integrated successively to any given order in the
(0)
perturbation. Equation (2.88) confirms that cn are time independent. It also tells us
that when there is no time dependence on the Hamiltonian, the Scrodinger equation has
a solution given by,
tes
X
ψ= ψn e−iωn t (2.92)
n
Transition Probability
Suppose that the system is initially in a particular unperturbed state |ψa ⟩ of energy Ea ,
Thus we write,
c(0)
n = δna (2.93)
No
Substituting equation (2.93) into equation (2.89) gives
1 X
ċ(1)
m = Ĥma (t)eiωma t (2.94)
iℏ a
(Em −Ea )
Where ωma = ℏ
Integrating equation (2.94) gives the solution to the first order equation as
t
ure
Z
1 ′ ′ ′
c(1)
m = Ĥma (t )eiωma t dt (2.95)
iℏ 0
Z t
i ′ ′ ′
c(1)
m =− Ĥma (t )eiωma t dt (2.96)
ℏ 0
Hence, to the first order, the probability of finding the system in some final energy eigen
state labeled f at time t, given that it is definitely in a different initial state i at a time
t = 0 is;
(1) 2
Pf i = c1f (t) (2.97)
ct
2
i t
Z
′ ′ ′
(1)
Pf i = − Ĥf i (t )eiωf i t dt , (f ̸= i) (2.98)
ℏ 0
(Ef −Ei )
1 ′
Where ωf i = ℏ
= ℏ
⟨ψf | Ĥ0 |ψ f ⟩ − ⟨ψ i | Ĥ 0 |ψ i ⟩ and Ĥf i (t ) = ⟨ψf | Ĥ1 |ψi ⟩
Le
For most problems in Atomic and Nuclear Physics, the first order (2.98) is usually suffi-
cient.
25
Example 2.15
tes
2
− t2
Ĥ1 = −x̂e t0
Calculate the probability for transition from the ground state, given that
∞
Z r 2
(−αt2 +iωt) π −ω
e dt = −i e 4α
0 α
Sln
No
The probability that a transition to a state f = 1 has occurred is;
(1) 2
P10 = c11 (t) (2.99)
But
Z t
(1) i ′ ′ ′
c1 =− Ĥ10 (t )eiω10 t dt (2.100)
ℏ 0
But
ure
′
Ĥ10 (t ) = ⟨f | Ĥ1 |i⟩ = ⟨1| Ĥ1 |0⟩ (2.101)
2
′ − t2
Ĥ10 (t ) = − ⟨1| x̂ |0⟩ e t0
(2.102)
r 2
′ ℏ − tt2
Ĥ10 (t ) = − e 0 (2.103)
2mω
Substituting equation (2.103)into equation (2.100)
ct
r ′2
Z t
(1) i ℏ −t2 ′ ′
c1 = e t0
eiω10 t dt (2.104)
ℏ 2mω 0
r Z t t′ 2 ′
(1) i ℏ − 2 +iωt ′
c1 = e t0 dt , ω10 = ω (2.105)
ℏ 2mω 0
Le
r
i
q
ℏ −ω 2 2
= . − i πt20 e 4 t0 (2.106)
ℏ 2mω
r
πt20 −ω2 t20
= e 4 (2.107)
2mℏω
(1) 2
P10 = c11 (t) (2.108)
26
(1) πt20 −ω2 t20
P10 = e 2 (2.109)
2mℏω
Example 2.16
tes
The time varying Hamiltonian Ĥ1 (t) induces transitions between states |j⟩ and |k⟩. Using
the time dependent perturbation theory, show that the probability for a transition from
state |j⟩ to state |k⟩ is the same as the probability of transition from state |k⟩ to state |j⟩.
Sln
The probability of transition from state |j⟩ to state |k⟩ at a time t is,
No
2
(1) (1)
Pkj = ck (t) (2.110)
Z t 2
i ′ ′ ′
= − Ĥkj (t )eiωkj t dt (2.111)
ℏ 0
Z t 2
(1) i ′ ′
Pkj = − ⟨k| Ĥ1 (t) |j⟩ eiωkj t dt (2.112)
ℏ 0
2
(1) (1)
Pjk = cj (t) (2.113)
ure
Z t 2
i ′ ′ ′
= − Ĥjk (t )eiωjk t dt (2.114)
ℏ 0
Z t 2
(1) i ′ ′
Pkj = − ⟨j| Ĥ1 (t) |k⟩ eiωjk t dt (2.115)
ℏ 0
Since Ĥ1 is Hermitian, ⟨k| Ĥ1 (t) |j⟩ = ⟨j| Ĥ1 (t) |k⟩. Also, it follows that ℏωkj = Ek −Ej =
−ℏωjk .
ct
As the integrand of the second integral is the complex conjugate of that of the first one.
We have,
2 2
(1) (1)
ck (t) = cj (t) (2.116)
Le
(1) (1)
Pkj = Pjk (2.117)
27
Example 2.17
A quantum system is initially in the ground state |0⟩. At t = 0, a perturbation of the
form Ĥ1 (t) = Ĥ0 e−αt , where α is a constant, is applied. Show that the probability that
the system is in a state |1⟩ after a long time is;
tes
2
⟨1| Ĥ0 |0⟩ E1 − E0
P10 = 2
, W here ω10 =
ℏ2 (α2 + ω10 ) ℏ
Rt ′
(1) ′ ′
From cf = − ℏi 0
Ĥf i (t )eiωf i t dt ,
′
Ĥf i (t ) = ⟨ψf | Ĥ1 (t) |ψi ⟩ (2.118)
No
Ĥ10 (t) = ⟨1| Ĥ0 e−αt |0⟩ = ⟨1| Ĥ0 |0⟩ e−αt (2.119)
i t
Z
′ ′ ′
(1)
c1 = − ⟨1| Ĥ0 |0⟩ e−αt eiω10 t dt (2.120)
ℏ 0
As t −→ ∞, equation (2.120) becomes;
Z ∞
i ′ ′
=− ⟨1| Ĥ0 |0⟩ e−(α−iω10 )t dt (2.121)
ℏ 0
i −1 ′
ure
= − ⟨1| Ĥ0 |0⟩ e−(α−iω10 )t |∞
0 (2.122)
ℏ (α − iω10 )
i ⟨1| Ĥ0 |0⟩
=− (2.123)
ℏ (α − iω10 )
2
(1) (1)
P10 = c1 (t) (2.124)
2
⟨1| Ĥ0 |0⟩
(1)
P10 = 2
(2.125)
ℏ2 (α2 + ω10 )
ct
Le
28
2.3 Scattering, Born Approximation and its Applications
2.3.1 Scattering
tes
In a typical scattering experiment, a beam of particles often from an accelerator is pro-
jected at a fixed target composed of other particles; Some particles are reflected and
others transmitted.
In three dimensions, there is more than reflection and transmitted. The experimentalist
counts the number of particles scattered through angles θ and φ that enter a detector
that subtends a certain solid angle as indicated in Figure 2.2.
No
ure
Figure 2.2: Scattering between an incident beam of particles and a fixed target: the scattered
particles are detected within a solid angle dΩ along the direction (θ, φ)
The number of particles scattered into an element of solid angle dΩ is proportional to the
differential cross-section.
ct
dσ
The differential cross-section dΩ is defined as the number of particles scattered into an
element of solid angle dΩ in the direction (θ, φ) per unit time and incident flux.
dσ(θ, φ) 1 dN (θ, φ)
= (2.126)
dΩ jinc dΩ
Le
29
But dΩ = sin θdθdφ Z π Z 2π
dσ(θ, φ)
σ= sin θdθ dφ (2.128)
0 0 dΩ
Note; The dimension of the total cross-section is area. This area may have a simple sig-
tes
nificance in some cases. For example, in a nuclear scattering experiment with neutrons
as the particle projectiles and a nucleus as the target, the total cross section is one the
order of the size of the nucleus since the nuclear force is shorted ranged and neutrons
that strike the nucleus are likely to interact with it.
On the other hand, if neutrinos are the projectiles and the target is a nucleus, the cross
section is many orders of magnitude smaller. This is not because the nucleus has sud-
No
denly shrunk in size but because neutrinos interact so weakly with the nucleus that most
of them pass right through the nucleus without scattering at all.
Scattering Amplitude
⃗
The wave describing the incident particle is ϕinc (⃗r) = Aeik0 .⃗r and scattered wave is
i⃗
k.⃗
r
ϕsc (⃗r) = Af (θ, φ) e r where f (θ, φ) is called the scattering amplitude, ⃗k is the wave
number associated with the scattered particle and θ is the angle between k⃗0 and ⃗k.
ure
After scattering has taken place, the total wave is given by a superposition of the incident
plane wave and the scattered wave.
⃗
ik⃗0 .⃗
r eik.⃗r
= Ae + Af (θ, φ) (2.130)
r
⃗
!
ik⃗0 .⃗
r eik.⃗r
ct
Differential Cross-section
Le
Incident flux
iℏ
jinc = (ϕinc ∇ϕ∗inc − ϕ∗inc ∇ϕinc ) (2.132)
2µ
30
⃗
From ϕinc (⃗r) = Aeik0 .⃗r ,
|A|2 ℏk0
ȷinc = (2.133)
µ
where µ is the effective mass.
tes
Scattered flux
iℏ
jsc = (ϕsc ∇ϕ∗sc − ϕ∗sc ∇ϕsc ) (2.134)
2µ
i⃗
k.⃗
r
From ϕsc = Af (θ, φ) e r ,
|A|2 ℏk
jsc = |f (θ, φ)|2 (2.135)
No
µr2
The number of particles dN (θ, φ) scattered into an element of solid angle dΩ in the di-
rection (θ, φ) and passing through a surface element dA = r2 dΩ per unit time is given as;
dσ(θ, φ) k
|f (θ, φ)|2
ct
= (2.141)
dΩ k0
Since the normalization factor A doesn’t contribute to the differential cross-section, we
will take it equal to 1. For elastic scattering, k0 = k hence,
dσ
= |f (θ, φ)|2 (2.142)
Le
dΩ
dσ
The problem of determining the differential cross-section dΩ
reduces to that of obtaining
the scattering amplitude f (θ, φ).
31
2.3.2 Born Approximation
A particularly good approach for calculating the scattering amplitude when the energy
of the incident beam is large in comparison with the potential energy is the Born approx-
tes
imation.
The scattering amplitude f (θ, φ) can be determined from the solutions of the Schrodinger
equation;
ℏ2 2
− ∇ ψ(⃗r) + V (r)ψ(⃗r) = Eψ(⃗r) (2.143)
2m
Equation (2.143) can be written as;
No
∇2 + k 2 ψ(⃗r) = Q
(2.144)
√
2mE 2m
where k = ℏ
and Q = ℏ2
V ψ(⃗r).
Suppose we could find a function G(⃗r) that solves the Helmholtz equation with a delta
function ”source”.
∇2 + k 2 G(⃗r) = δ 3 (⃗r)
(2.145)
ure
We could express ψ(⃗r) as an integral.
Z
ψ(⃗r) = G (⃗r − ⃗r0 ) Q (⃗r0 ) d3⃗r0 (2.146)
Z
= δ 2 (⃗r − ⃗r0 ) Q (⃗r0 ) d3⃗r0 (2.148)
∇2 + k 2 ψ(⃗r) = Q (⃗r)
(2.149)
What we are trying to do is expressing the Schrodinger equation in integral form rather
Le
G (⃗r) is called the Green’s function for the Helmholtz equation. Derivations (see reference
32
textbooks) show that;
eik|⃗r−r⃗0 |
G (⃗r) = − (2.150)
4π |⃗r − r⃗0 |
We can add it to any function that satisfies the Homogeneous Helmholtz equation.
tes
∇2 + k 2 G0 (⃗r) = 0
(2.151)
eik|⃗r−r⃗0 |
Z
m
ψ(⃗r) = − V (⃗r0 ) ψ (⃗r0 ) d3⃗r0 (2.152)
2πℏ2 4π |⃗r − r⃗0 |
No
Now we arrive at the integral form of the Schrodinger equation;
eik|⃗r−r⃗0 |
Z
m
ψ(⃗r) = ψ0 (⃗r) − V (⃗r0 ) ψ (⃗r0 ) d3⃗r0 (2.153)
2πℏ2 4π |⃗r − r⃗0 |
Where ψ0 (⃗r) satisfies the free particle Schrodinger equation (∇2 + k 2 ) ψ0 (⃗r) = 0.
Assuming that |⃗r| >> |⃗r| for all points that contribute to the integral in equation (2.153),
2 2 2⃗r.⃗
r0
|⃗r − r⃗0 | = r + r02 − 2⃗rr⃗0 ≈ r 2
1− 2 (2.154)
r
ct
And hence;
|⃗r − r⃗0 |2 ≈ r − r̂.⃗r0 (2.155)
⃗
Let ⃗k = kr̂, then eik|⃗r−⃗r0 | ≈ eikr .e−ik.r⃗0 and therefore,
≈ e (2.156)
|⃗r − r⃗0 | r
eikr −i⃗k.⃗r0
Z
m
ψ(⃗r) = ψ0 (⃗r) − .e V (⃗r0 ) ψ (⃗r0 ) d3⃗r0 (2.157)
2πℏ2 r
33
In the case of scattering, ψ0 (⃗r) = Aeikz , the incident plane wave. Therefore, our wave
function in (2.157) can be written as;
eikr
Z
ikz m −i⃗k.⃗
r0 3
ψ(⃗r) = Ae +A − e V (⃗r0 ) ψ (⃗r0 ) d ⃗r0 (2.158)
tes
r 2πAℏ2
Therefore,
Z
m ⃗
f (θ, φ) = − e−ik.⃗r0 V (⃗r0 ) ψ (⃗r0 ) d3⃗r0 (2.159)
2πAℏ2
is the scattering amplitude.
No
Let us assume that the potential doesn’t significantly alter the wave function;
⃗′
ψ(⃗r0 ) ≈ ψ0 (⃗r0 ) = Aeikz0 = Aeik .⃗r0 (2.160)
′
For ⃗k = kẑ Then;
Z
m −i⃗k.⃗r0
′
i⃗k .⃗
r0 3
f (θ, φ) = − e V (⃗r 0 ) Ae d ⃗r0 (2.161)
2πAℏ2
Z
m −i⃗k.⃗
r0
′
i⃗k .⃗
r0 3
=− 2
e V (⃗r 0 ) e d ⃗r0 (2.162)
2πℏ
ure
Z
m ′
i(⃗k −⃗k).⃗ r0
f (θ, φ) = − e V (⃗r0 ) d3⃗r0 (2.163)
2πℏ2
′ ′
Let us clarify what are the vectors ⃗k and ⃗k. ⃗k points in the incident direction while ⃗k
in the scattering direction.
ct
Le
34
In particular, for low energy (long wavelength) scattering, the exponential factor in equa-
tion (2.163) is essentially constant over the scattering region and the Born approximation
simplifies to;
tes
Z
m
f (θ, φ) = − V (⃗r) d3⃗r (low energy) (2.164)
2πℏ2
′
⃗ ⃗
If the potential V (⃗r) is spherically symmetric, V (⃗r0 ) = V (r0 ) and k − k .⃗r0 = qr0 cos(θ0 ).
Z
m
f (θ, φ) ≈ − eiqr0 cos(θ0 ) V (r0 ) r02 sin(θ0 )dr0 dθ0 dφ0 (2.165)
2πℏ2
Z
2m
f (θ) ≈ − 2 r0 V (r0 ) sin(qr0 )dr0 (2.166)
No
ℏq
Dropping the subscript on r, we get;
Z
2m
f (θ) = − 2 rV (r) sin(qr)dr (Symmetric potential) (2.167)
ℏq
′
Note; from q = ⃗k − ⃗k ,
′ 2 ′ ′
q 2 = ⃗k − ⃗k = k 2 + k 2 − 2k k cos(θ) (2.168)
ure
′
For quantum scattering, k = k
2 2 θ 2 2
q = 2k (1 − cos(θ)) = 4k sin (2.169)
2
θ
q = 2ksin (2.170)
2
This is referred to as the momentum transfer, which is the amount of momentum one
particle gives to another.
ct
Example 2.18
Use low-energy scattering in Born approximation to calculate the differential and total
cross-section for the low energy soft-sphere scattering.
Le
(
V0 ;r ≤ a
V (⃗r) =
0 ;r > a
Sln;
35
m
R
From f (θ, φ) = − 2πℏ2 V (⃗r) d3⃗r,
Z a
m
f (θ, φ) = − 4πV0 r2 dr (2.171)
2πℏ2 0
tes
m a3
=− (4πV0 ) (2.172)
2πℏ2 3
2ma3
f (θ, φ) = − V0 (2.173)
3ℏ2
dσ
= |f (θ, φ)|2 (2.174)
dΩ
dσ 4m2 a6 2
No
= V (2.175)
dΩ 9ℏ4 0
dσ
R
Total cross-section σ = dΩ
dΩ. But dΩ = sin(θ)dθdφ
Z π Z 2π
dσ
σ= sin θ dφ (2.176)
0 0 dΩ
π 2π
4m2 a6 2
Z Z
= sin θ V dφ (2.177)
0 0 9ℏ4 0
2 6
4m a 2
= 4π V (2.178)
9ℏ4 0
ure
16πm2 a6 V02
σ= (2.179)
9ℏ4
Example 2.19
The scattering amplitude for the first born approximation is given by;
Z ∞
2µ ′
′ ′ ′
f (θ) = − 2 r V r sin qr dr
ℏq 0
ct
e−γr
Le
V (r) = β
r
a) Show that the differential scattering cross-section for the scattering vector k is given
by;
36
2
dσ 2µβ
=
dΩ ℏ (k 2 + γ 2 )
2
b) Show that the differential scattering cross-section for alpha particles of energy E
tes
and incident on a nucleus of atomic number Z is given by;
!2
dσ Ze2
=
8πϵ0 E sin2 θ
dΩ 2
Sln; (a)
dσ
= |f (θ)|2 (2.180)
No
dΩ
Z ∞ 2
4µ2 ′
′ ′ ′
= r V r sin qr dr (2.181)
ℏ4 q 2 0
−γr
For V (r) = β e r
∞ ∞
e−γr
Z ′ ′ ′ Z
′
r V r sin qr dr = rβ sin(qr)dr (2.182)
0 0 r
Z ∞
= βe−γr sin(qr)dr (2.183)
0
ure
Let u = sin(qr) and dv
dr
= e−γr , du
dr
= q cos(qr) and v = − γ1 e−γr
Therefore;
Z ∞ Z ∞
βq 1 −γr q −γr
Le
−γr ∞
βe sin(qr)dr = − e cos(qr)|0 − e sin(qr)dr (2.186)
0 γ γ 0 γ
q ∞ −γr
Z
βq 1
= − − e sin(qr)dr (2.187)
γ γ γ 0
37
∞
q2
Z
−γr βq
βe sin(qr)dr 1 + 2 = 2 (2.188)
0 γ γ
Z ∞
βq
βe−γr sin(qr)dr = (2.189)
0 γ2 + q2
tes
From equation (2.181)
2
dσ 4µ2 βq
= 4 2 2 (2.190)
dΩ ℏ q γ + q2
4µ2 β 2 q 2
= (2.191)
ℏ4 q 2 (γ 2 + q 2 )2
2
dσ 2µβ
= (2.192)
No
dΩ ℏ2 (γ 2 + q 2 )
θ
. For θ = π3 , q = k
But q = 2k sin 2
2
dσ 2µβ
= (2.193)
dΩ ℏ2 (γ 2 + k 2 )
(b)
′ Z1 Z2 e2 2Ze2
From potential energy V (r ) = 4πϵ0 r
= 4πϵ0 r
ure
ℏ2 k2 1 2µ
Since the energy of the incident particle is proportional to k i.e. E = 2µ
, E = ℏ2 k2
From dσ
dΩ
= |f (θ)|2 ,
∞ 2
4µ2
Z
dσ ′
′ ′ ′
= 4 2 r V r sin qr dr (2.194)
dΩ ℏq 0
But;
∞ ∞
2Ze2
Z ′ ′ ′ Z
′
ct
2Ze2
= cos(qr)|∞
0 (2.197)
4πϵ0 q
Le
2Ze2
= (2.198)
4πϵ0 q
2
dσ 4µ2 2Ze2 16µ2 Z 2 e4
= 4 2 = (2.199)
dΩ ℏ q 4πϵ0 q 16ℏ4 q 4 π 2 ϵ20
38
2
µZe2
= (2.200)
ℏ2 q 2 πϵ0
θ
and q 2 = 4k 2 sin2 ( 2θ )
But q = 2k sin 2
tes
!2
dσ µZe2
= (2.201)
4k 2 sin2 ( 2θ ) πϵ0
dΩ ℏ2
!2
µZe2
= (2.202)
4ℏ2 k 2 sin2 ( 2θ )πϵ0
1 µ
But 2E
= ℏ2 k2
No
!2
dσ Ze2
= (2.203)
dΩ 8πϵ0 E sin2 ( 2θ )
Example 2.20
The scattering amplitude for the first Born approximation is given by;
Z ∞
2m ′
′ ′ ′
f (θ) = − 2 r V r sin qr dr
ℏq 0
β −r
e a V (r) =
r
where β and a are constants. Show that the differential cross section is given by;
!
dσ 4m2 β 2 1
=
ct
2
dΩ ℏ4 1
+ q2
a2
b) Show that in the limit of large incident energies where q 2 >> a−2 , the Yakuwa
scattering approximates the Rutherford scattering given by the cross section;
Le
Ze2
dσ 1
=
dΩ 2E sin4 ( 2θ )
Where θ is the scattering angle and E is the kinetic energy of the incident particle.
39
Sln;
(a)
dσ
= |f (θ)|2 (2.204)
tes
dΩ
Z ∞ 2
4m2 ′
′ ′ ′
= r V r sin qr dr (2.205)
ℏ4 q 2 0
r
But V (r) = βr e− a Z ∞
r
= βe− a sin(qr)dr (2.206)
0
r r
Let u = sin(qr) and dv
dr
= e− a , du
dr
= q cos(qr) and v = −ae− a
No
Solving the integral (equation (2.206)) gives;
Z ∞ Z ∞
− ar − ar r
βe sin(qr)dr = −βae sin(qr)|∞
0 + −βae− a .q cos(qr)dr (2.207)
0 0
Z ∞
r
= βaqe− a cos(qr)dr (2.208)
0
r r
Also, let Let u = cos(qr) and dv
dr
= e− a , du
dr
= −q sin(qr) and v = −ae− a
Therefore;
ure
Z ∞ Z ∞
− ar − ar ∞ − ar
βe sin(qr)dr = βaq −ae cos(qr)|0 − aqe sin(qr)dr (2.209)
0 0
Z ∞
− ar
= βaq a − aq e sin(qr)dr (2.210)
0
Z ∞
r
βe− a sin(qr)dr 1 + a2 q 2 = a2 qβ
(2.211)
0
∞
a2 qβ
Z
ct
r
βe− a sin(qr)dr = (2.212)
0 (1 + a2 q 2 )
Therefore;
∞
a2 qβ
Z ′ ′ ′
′
r V r sin qr dr = (2.213)
(1 + a2 q 2 )
Le
And;
2
dσ 4m2 a2 qβ
= 4 2 (2.214)
dΩ ℏ q (1 + a2 q 2 )
4m2 a4 q 2 β 2
= . (2.215)
ℏ4 q 2 (1 + a2 q 2 )2
40
4m2 a4 β 2 4m2 a4 β 2
= = (2.216)
ℏ4 (1 + a2 q 2 )2
2
ℏ4 a12 + q 2 a4
dσ 4m2 β 2
= 2 (2.217)
dΩ ℏ4 a12 + q 2
tes
(b)
Z1 Z2 e2 2Ze2 β V0
For a coulomb potential, V (r) = r
. For Z1 = Z2 , V (r) = r
= r
= r
. This
implies that β = V0 = 2Ze2
θ
and q 2 = 4k 2 sin2 ( 2θ )
Also, q = 2k sin 2
From;
No
dσ 4m2 β 2
= 2
dΩ ℏ4 a12 + q 2
dσ dσ
= lim (2.218)
dΩ Rutherf ord a−→∞ dΩ Y akuwa
4m2 V02
= (2.219)
ℏ4 (q 2 )2
4m2 V02
= 2 (2.220)
ℏ4 4k 2 sin2 ( 2θ )
ure
16m2 Z 2 e4
= (2.221)
16ℏ4 k 4 sin4 ( 2θ )
m2 Z 2 e4
= (2.222)
ℏ4 k 4 sin4 ( 2θ )
ℏ2 k2
But E = 2m
m 2 Z 2 e4
= . 4 θ (2.223)
ℏ2 k 2 sin ( 2 )
ct
2 2
dσ Ze 1
= . 4 θ (2.224)
dΩ Rutherf ord 2E sin ( 2 )
Individual Assignment;
Read and make notes about Partial Wave Analysis
Le
41
2.4 Chapter Exercises
1. The Schrodinger equation of a particle confined to the positive x-axis is;
ℏ d2 ψ
tes
− + mgxψ = Eψ
2m dx2
with ψ(0) = 0, ψ(x) −→ 0 as x −→ ∞ and E is the energy eigen value. Use the
trial function xe−ax and obtain the best value of a
2
2. A particle of mass m moves in the attractive central potential V (r) = − g3 , where
r2
4 12 kr
k
e− 2 as the trial function,
No
g is a constant. Using the normalized function 8π
estimate the upper bound to the energy of the lowest state. Assume;
g2
ℏ 1 d 2 d
Ĥ = − r − 3
2m r2 dr dr r2
Also;
∞ √ ∞
Z r Z
−ax 1 x n −ax n!
xe dx = ; x e dx = if n is positive and a > 0
0 2a a 0 an+1
3. Evaluate the ground state energy of a harmonic oscillator of mass m and angular
ure
momentum ω using a trial function;
(
πx
cos 2a
, −a ≤ x ≤ a
ϕ(x) =
0 |x| > a
5. Evaluate the first order correction to the energy of the n = 1 state of an oscillator
of mass m and angular frequency ω subjected to a potential
1 1
V (x) = mω 2 x2 + bx, bx << mω 2 x2
2 2
Le
6. Use the variational method for solving the Schrodinger equation for the truncated
harmonic oscillator potential;
1
V (x) = kx2 , f orx > 0
2
42
= ∞, f orx < 0
Use the trial function ψ = xe−bx where b is the variational parameter, to calculate
an approximate value for the ground state energy.
tes
7. A simple harmonic oscillator of mass m and angular frequency ω is perturbed by
an additional potential 12 bx2 . Obtain the first and second order corrections to the
ground state energy.
8. Calculate the first order energy shift of the first three states of a particle in a one
dimensional box problem (walls at x = 0 and x = a) due to the potential;
V1 x
(a)
No
a
x 2
(b)V1 a
, where V1 is a constant
(c) What is the condition under which the perturbation condition is valid?
(0)
9. Calculate the first order correction to E3 for a particle in a one dimensional box
with walls at x = 0 and x = L due to the following perturbations.
(i) Ĥ1 = 10−3 E1 Lx
x 2
(ii) Ĥ1 = 10−3 E1
L
10. A particle which is initially (t = 0) in the ground state of an infinite one dimensional
potential box with walls at x = 0 and x = a is subjected for 0 ≤ t ≤ α to a
t
perturbation Ĥ1 (t) = x̂2 e− τ . Calculate to the first order, the probability of finding
the particle in its first excited state for t ≥ 0.
r !
n2 π 2 ℏ2 2 nπx
En = , ψn (x) = sin
2ma2 a a
ct
11. A particle initially( i.e. t −→ −∞ )in its ground state in an infinite potential well
whose walls are located at x = 0 and x = a is subjected at a time t = 0 to a time
2
dependent perturbation V̂ (t) = ϵx̂2 e−t where ϵ is a real number. Calculate to the
probability that the particle will be found in its first excited state after a sufficiently
Le
43
12. A particle of mass m in the ground state of a one-dimensional harmonic oscillator is
t
placed in a perturbation V̂ (t) = −V0 x̂2 e− τ . Calculate to the first order probability
theory the probability of finding the particle in its first excited state after a long
time.
tes
13. Consider a one dimensional harmonic oscillator with angular frequency ω and elec-
tric charge e at a time t = 0 in the ground state. An electric field is applied for a
time t = τ so that the perturbation is;
No
= 0, else where
14. Find the differential cross section in the first born approximation for neutron-
r
neutron scattering in the case where the potential is approximately V (r) = V0 e− a .
ure
r2
15. Consider the scattering from the potential V (r) = V0 e− a2 . Find
(a) the differential cross section in the first born approximation.
(b) the total cross section.
16. Calculate the born approximation differential cross section for the following poten-
tial (
V0 , r≤a
V (r) =
ct
0 r>a
17. Use the born approximation to calculate the differential cross section for scattering
by the central potential V (r) = rα2 where α is a constant
Z ∞
sin ax π
Le
Given dx = ; f or a>0
0 x 2
2V0 eαr
18. Using the born approximation approach for the potential V (r) = r
44
(a) Write down the expression for the scattering amplitude and show that;
4mV0
f (θ) =
ℏ2 (α2 + q 2 )
tes
Z ∞
αr q θ
Hint; e sin(qr)dr = and q = 2k sin
0 (α2 + q 2 ) 2
32πm2 V02
σtotal =
α2 ℏ4 (α2 + 4k 2 )
No
19. With the aid of the born approximation, find the scattering amplitude f 9θ for the
truncated coulomb potential,
1 1
V (r) = c − r < a, and V (r) = 0 f or r ≥ a
a r
Find the expression for the scattering amplitude at very small angles of θ and hence
deduce the approximate value of sin 2θ at which the differential cross section has
2m ∞
Z
ure
T ake; f (θ) = − 2 rV (r) sin(qr)dr
ℏq 0
ct
Le
45
3 Angular Momentum
tes
⃗ = ⃗r × p⃗ ; where ⃗r
Classically, the orbital angular momentum of a particle is given by L
and p⃗ are the position vector and momentum of the particle, respectively.
⃗r = xî + y ĵ + z k̂ (3.1)
No
p⃗ = px î + py ĵ + pz k̂ (3.2)
⃗ is ;
the angular momentum L
î ĵ k̂
⃗ = x y z
L (3.3)
px p y pz
⃗ = î y z − ĵ x z + k̂ x y
L (3.4)
py pz px pz px p y
ure
= î (ypz − zpy ) − ĵ (xpz − zpx ) + k̂ (xpy − ypx ) (3.5)
But;
⃗ = Lx î + Ly ĵ + Lz k̂
L (3.7)
Lx = (ypz − zpy )
ct
The corresponding quantum operators are obtained by the standard prescription; p̂x =
∂ ∂ ∂
−iℏ ∂x , p̂y = −iℏ ∂y and p̂z = −iℏ ∂z . Therefore;
Le
∂ ∂
L̂x = (ŷ p̂z − ẑ p̂y ) = −iℏ y ∂z − z ∂y
∂ ∂
L̂y = (ẑ p̂x − x̂p̂z ) = −iℏz ∂x − x ∂z
(3.9)
∂ ∂
L̂z = (x̂p̂y − ŷ p̂x ) = −iℏ x ∂y − y ∂x
46
⃗ˆ are each Hermitian
These individual components of the angular momentum operator L
⃗ˆ its self.
and so, correspondingly is the operator L
tes
Here, we are to compute the commutation relations involving these component operators
using the fundamental commutators;
No
[Â ± B̂, Ĉ] = [Â, B̂] ± [B̂, Ĉ] (3.11)
= ŷẑ[p̂z , p̂x ] + ŷ[p̂z , ẑ]p̂x + ẑ[ŷ, p̂x ]p̂z + [ŷ, ẑ]p̂x p̂z
= ẑ 2 [p̂y , p̂x ] + ẑ[p̂y , ẑ]p̂x + ẑ[ẑ, p̂x ]p̂y + [ẑ, ẑ]p̂x p̂y
47
= ẑ x̂[p̂y , p̂z ] + ẑ[p̂y , x̂]p̂z + x̂[ẑ, p̂z ]p̂y + [ẑ, x̂]p̂z p̂y
Substituting equations (3.15), (3.16), and (3.17) into equation (3.14) gives;
tes
[L̂x , L̂y ] = −iℏŷ p̂x + 0 + 0 + iℏx̂p̂y (3.19)
No
[L̂y , L̂z ] = iℏL̂x (3.22)
The commutation relations in equations (3.21, 3.22, and 3.23) can be combined together
into a single vector equation.
⃗ˆ × L
L ⃗ˆ = iℏL
⃗ˆ (3.24)
Now, let us consider the operator corresponding to the square of the orbital angular
momentum;
L̂2 = L̂2x + L̂2y + L̂2z (3.25)
ure
⃗
and compute the commutators of L̂2 with the components of L.
= L̂y [L̂y , L̂x ] + [L̂y , L̂x ]L̂y + L̂z [L̂z , L̂x ] + [L̂z , L̂x ]L̂z
48
[L̂2 , L̂z ] = 0 (3.29)
tes
⃗ˆ we can find
Since L̂2 commutes with all the components of the angular momentum L,
⃗ˆ
simultaneous eigenstates of L̂2 and any one component of L.
Let us now introduce the ladder operator technique similar to the one we applied to solve
the harmonic oscillator problem. Let the operators be defined as;
No
Lowering Operator ; L̂− = L̂x − iL̂y (3.31)
†
L̂†+ = L̂x + iL̂y = L̂− (3.34)
ure
From the above equations we realize that these two operators are not Hermitian but
instead are mutually ad-joint. Manipulating equations (3.31) and (3.32) results into;
L̂+ + L̂−
L̂x = (3.35)
2
L̂+ − L̂−
L̂y = (3.36)
2i
ct
By using the standard commutation relations, we can go a head and compute the com-
mutator of L̂z and L̂±
[L̂z , L̂+ ] = [L̂z , L̂x + iL̂y ] (3.37)
= iℏL̂y + i(−i)ℏL̂x
= ℏ L̂x + iL̂y
49
Also,
[L̂z , L̂− ] = [L̂z , L̂x − iL̂y ] (3.39)
tes
= iℏL̂y − i(−i)ℏL̂x
= iℏL̂y − ℏL̂x = −ℏ L̂x − iL̂y
Hence the Lowering and Raising operators don’t commute with L̂z . However, they com-
mute with L̂2 .
No
[L̂2 , L̂± ] = [L̂2x + L̂2y + L̂2z , L̂± ] (3.41)
From equation (3.30), since L̂2 commutes with L̂z , these two operators have simultaneous
eigen functions.
Suppose that the simultaneous eigen states are given by |l, m⟩, we can write the eigenvalue
equations as;
L̂2 |l, m⟩ = l(l + 1)ℏ2 |l, m⟩ (3.43)
ct
Where l(l + 1)ℏ2 and mℏ are eigenvalues of L̂2 and L̂z respectively.
L̂± L̂2 |l, m⟩ = L̂± l(l + 1)ℏ2 |l, m⟩
(3.45)
50
From equation (3.42), L̂± L̂2 = L̂2 L̂± , therefore, equation (3.46) becomes;
tes
L̂2 L̂± |l, m⟩ = l(l + 1)ℏ2 L̂± |l, m⟩ (3.48)
From equation (3.48) we realize that the action of the operator L̂± on the simultaneous
eigen state |l, m⟩ of L̂2 is to form a new eigen state L̂± |l, m⟩. Thus L̂± |l, m⟩ are also
eigen states of L̂2 with eigenvalues l(l + 1)ℏ2 .
No
L̂± L̂z |l, m⟩ = L̂± (mℏ |l, m⟩) (3.49)
We see that L̂+ |l, m⟩ is an eigen state of L̂z with corresponding eigenvalues (m + 1)ℏ.
51
From the commutation relation [L̂− , L̂z ] = ℏL̂− ,
tes
Substituting equation (3.59) into equation (3.52) gives;
L̂z L̂− + ℏL̂− |l, m⟩ = mℏL̂− |l, m⟩ (3.60)
No
L̂z L̂− |l, m⟩ = (m − 1) ℏL̂− |l, m⟩ (3.63)
L̂z L̂− |l, m⟩ = (m − 1) ℏ L̂− |l, m⟩ (3.64)
We as well see that L̂− |l, m⟩ is an eigen state of L̂z with corresponding eigenvalues
(m − 1)ℏ.
2
L̂+ |l, m⟩ = ⟨l, m| L̂†+ L̂+ |l, m⟩ = ⟨l, m| L̂− L̂+ |l, m⟩ (3.66)
But
L̂− L̂+ = L̂2 − L̂2z − ℏL̂z (3.67)
2
L̂+ |l, m⟩ = ⟨l, m| L̂2 − L̂2z − ℏL̂z |l, m⟩ (3.68)
2 ∗
L̂− |l, m⟩ = L̂− |l, m⟩ L̂− |l, m⟩ (3.71)
52
2
L̂− |l, m⟩ = ⟨l, m| L̂†− L̂− |l, m⟩ = ⟨l, m| L̂+ L̂− |l, m⟩ (3.72)
But
L̂+ L̂− = L̂2 − L̂2z + ℏL̂z (3.73)
tes
Substituting equation (3.73) into equation (3.72) gives;
2
L̂− |l, m⟩ = ⟨l, m| L̂2 − L̂2z + ℏL̂z |l, m⟩ (3.74)
No
From equations (3.58) and (3.64) we conclude that L̂+ |l, m⟩ is proportional to |l, m + 1⟩
and L̂+ |l, m⟩ is proportional to |l, m − 1⟩. Thus;
L̂+ |l, m⟩ ≈ |l, m + 1⟩ and L̂− |l, m⟩ ≈ |l, m − 1⟩
Recall that the norm of |l, m − 1⟩ and L̂− |l, m⟩ is L hence from equations (3.70) and
(3.76)
p
L̂+ |l, m⟩ = ℏ (l(l + 1) − m(m + 1)) |l, m + 1⟩ (3.77)
p
L̂− |l, m⟩ = ℏ (l(l + 1) − m(m − 1)) |l, m − 1⟩ (3.78)
ure
Equations (3.77) and (3.78) give the normalised eigen functions/states resulting from the
action of the raising and the lowering operators respectively.Wherem = −l, −l +1, ....., l −
1, l for l = 0, 1, 2, ...
ct
Le
53
3.2 Spin
In addition to orbital angular momentum, some particles such as electrons possess an
intrinsic (spin) angular momentum which, unlike the orbital angular momentum has
tes
nothing to do with the spatial degrees of freedom i.e. spin angular momentum doesn’t
relate to a particle’s coordinates or momenta, nor are the eigen states of spin dependent
on the boundary conditions imposed in coordinate space.
The spin angular momentum (Ŝ) operator and its cartesian components are Hermitian.
The algebraic theory of spin is a carbon copy of that of the orbital angular momentum
discussed in subsection 3.1 hence satisfying the following commutation relations.
No
[Ŝx , Ŝy ] = iℏŜz (3.79)
Also,
Ŝ 2 = Ŝx2 + Ŝy2 + Ŝz2 (3.82)
As earlier obtained in subsection 3.1, the z-component of the spin angular momentum
ure
commutes with the operator Ŝ 2 i.e.;
Therefore, Ŝ 2 and Ŝz have simultaneous eigen functions |s, ms ⟩ with the following eigen-
vale problems;
Ŝ 2 |s, ms ⟩ = s(s + 1)ℏ2 |s, ms ⟩ (3.84)
Le
In nature, every fundamental particle has a specific spin. Some particles have integer
spin s = 0, 1, 2, ... (πmesons have s = 0, photons have s = 1 and so on) and others have
half-odd-integer spins s = 12 , 23 , 52 , ... (electrons, protons, and neutons have spin 21 , the
deltas have spin 32 and so on).
54
Particles with half-odd-integer spins are called fermions (quarks, electrons, protons, neu-
trons, etc) and those with integer spins are called bosons (pions, photons, gravitons, etc).
tes
Similarly we have;
p
Ŝ± |s, ms ⟩ = ℏ s(s + 1) − ms (ms ± 1) |s, ms ± 1⟩ (3.86)
No
2 2
D E D E
Where  denotes  = ⟨s, ms |  |s, ms ⟩.
s
X
|s, ms ⟩ ⟨s, ms | = Iˆ (3.89)
ms =−s
ure
Where Iˆ is a unit matrix.
1
3.2.1 Spin 2
and Matrix Elements of Spin Angular Momentum Components
For a Particle with spin 12 , the quantum number ms takes on only two values; ms = ± 12 .
The particle can be found in either of the following two states;
1 1 1 1
|s, ms ⟩ = , and |s, ms ⟩ = , − (3.90)
2 2 2 2
ct
in the Ŝ 2 and Ŝz basis. There are two eigen states, meaning that the Hilbert space is two
dimensional. The eigen states of Ŝ 2 and Ŝz are given by;
2 1 1 3 1 1
Ŝ ,± = ℏ2 , ± (3.91)
2 2 4 2 2
Le
1 1 ℏ 1 1
Ŝz , ± =± ,± (3.92)
2 2 2 2 2
Let us now study the matrix representation of spin s = 12 .
55
The two dimensional Hilbert space implies that the matrix that act on this space is a
2 × 2 form. Therefore, the spin operators for an electron are a 2 × 2 matrix.
′ ′
Evoking a bra s , ms onto equation (3.85) gives;
tes
D ′ ′
D ′ ′ E
s , ms Ŝz |s, ms ⟩ = ms ℏ s , ms s, ms (3.93)
No
!
ℏ 1 0
= (3.96)
2 0 −1
′ ′
Multiplying equation (3.84) by s , ms gives;
ure
D ′ ′
D ′ ′ E
s , ms Ŝ 2 |s, ms ⟩ = s(s + 1)ℏ2 s , ms s, ms (3.98)
3ℏ2 1 0
Ŝ 2 = (3.101)
4 0 1
To obtain the matrix elements of the operator Ŝ+ , we use the expression in equation
′ ′
(3.86). Evoking a bra s , ms on equation (3.86) gives;
Le
D ′ ′ 1
D ′ ′ E
s , ms Ŝ+ |s, ms ⟩ = ℏ (s(s + 1) − ms (ms + 1)) 2 s , ms s, ms + 1 (3.102)
56
s can only take one value (s = 12 ) while ms = ± 21 , thus;
21 ! 12 !
1 3 1 3 0 0 1 3 1 1 0 1
Ŝ+ = ℏ . − +ℏ . + (3.103)
2 2 2 2 0 0 2 2 2 2 0 0
tes
!
0 1
Ŝ+ = ℏ (3.104)
0 0
1
Where the first and second terms in equation (3.103) corresponds to ms = 2
and ms = − 21
respectively.
Also, to obtain the matrix element of the operator Ŝ− we use the expression in equation
No
′ ′
(3.86) by evoking a bra s , ms on both sides of it which gives;
D ′ ′ 1
D ′ ′
E
s , ms Ŝ− |s, ms ⟩ = ℏ (s(s + 1) − ms (ms − 1)) 2 s , ms s, ms − 1 (3.105)
1
= ℏ (s(s + 1) − ms (ms − 1)) 2 δs′ s δm′s ms −1 (3.106)
21 ! 12 !
1 3 1 1 0 0 1 3 1 3 0 0
Ŝ− = ℏ . − − +ℏ . + − (3.107)
ure
2 2 2 2 1 0 2 2 2 2 0 0
!
0 0
Ŝ− = ℏ (3.108)
1 0
1
Since Ŝx = 2
Ŝ+ + Ŝ− ,
" ! !#
1 0 1 0 0
Ŝx = ℏ +ℏ (3.109)
2 0 0 1 0
ct
!
ℏ 0 1
Ŝx = (3.110)
2 1 0
1
Also, Ŝy = 2i
Ŝ+ − Ŝ−
Le
" ! !#
1 0 1 0 0
Ŝy = ℏ −ℏ (3.111)
2i 0 0 1 0
57
!
ℏ 0 1
= (3.112)
2i −1 0
!
ℏ 0 −i
Ŝy = (3.113)
tes
2 i 0
⃗ˆ2 and Ŝz are expressed in terms of two element column matrices
The joint eigen vectors of S
known as spinors. !
1 1 1
χ+ = , = (3.114)
2 2 0
!
1 1 0
χ− = , − = (3.115)
No
2 2 1
Let us now verify that these eigen vectors form a complete basis;
1 ! !
2
X 1 1 0 1
, ms , ms = 0 1 + 1 0 (3.116)
2 2 1 0
ms =− 21
!
1 0
= (3.117)
0 1
ure
Also, orthonormal; !
1
1 1 1 1
, ,= 1 0 =1 (3.118)
2 2 2 2 0
!
1 1 1 1 0
,− ,− = 0 1 =1 (3.119)
2 2 2 2 1
1 1 1 1 1 1 1 1
, ,− = ,− , =0 (3.120)
2 2 2 2 2 2 2 2
ct
Le
58
3.2.2 Pauli Matrices
Since the matrix elements Ŝz , Ŝx and Ŝy in equations (3.97), (3.110) and (3.113) respec-
tively carry a factor ℏ2 , we observe that they can be generally written as;
tes
ℏ
Ŝj = σ̂j , j = x, y, z (3.121)
2
No
!
0 −i
σ̂y = (3.123)
i 0
!
1 0
σ̂z = (3.124)
0 −1
These matrices satisfy the following properties;
ˆ
σ̂j2 = I, (j = x, y, z) (3.125)
ure
σ̂j σ̂k + σ̂k σ̂j = 0, (j ̸= k) (3.126)
Where Iˆ is a 2 × 2 matrix.
Also,
T rσ̂x = T rσ̂y = T rσ̂z = 0 (3.127)
and the determinants of these matrices are all equal to −1 i.e. detσ̂x = detσ̂y = detσ̂z =
−1
ct
2 2 2
σ̂x = Ŝx , σ̂y = Ŝy , σ̂z = Ŝz , (3.128)
ℏ ℏ ℏ
59
Also,
σ̂x2 + σ̂y2 + σ̂z2 = 3Iˆ (3.132)
tes
Let us consider the action of the operators Ŝx and Ŝy on the eigen states of Ŝz ; ± 21
1 ℏ 1
Ŝx + = σ̂x + (3.133)
2 2 2
! !
ℏ 0 1 1
= (3.134)
2 1 0 0
No
!
ℏ 0 ℏ 1
= = − (3.135)
2 1 2 2
1 ℏ 1
Ŝx + = − (3.136)
2 2 2
Also,
1 ℏ 1
Ŝx − = σ̂x − (3.137)
2 2 2
! !
ℏ 0 1 0
= (3.138)
ure
2 1 0 1
!
ℏ 1 ℏ 1
= = + (3.139)
2 0 2 2
1 ℏ 1
Ŝx − = + (3.140)
2 2 2
For Ŝy + 12 ,
1 ℏ 1
Ŝy + = σ̂y + (3.141)
ct
2 2 2
! !
ℏ 0 −i 1
= (3.142)
2 i 0 0
!
ℏ 0
Le
= (3.143)
2 i
1 iℏ 1
Ŝy = − (3.144)
2 2 2
60
For Ŝy − 12 ,
1 ℏ 1
Ŝy − = σ̂y − (3.145)
2 2 2
! !
ℏ 0 −i 0
tes
= (3.146)
2 i 0 1
!
ℏ −i
= (3.147)
2 0
1 −iℏ 1
Ŝy − = + (3.148)
2 2 2
The results in equations (3.136) and (3.140) show the action of the x-component of the
No
spin angular momentum Ŝx on the eigen states of Ŝz is to make a transition between the
two states; This also applies to Ŝy as seen in equations (3.144) and (3.148)
3.2.4 Action of the Lowering and Raising Operators of Ŝ on the Eigen States
of Ŝz
The lowering and raising operators of the spin angular momentum operator Ŝ are defined
as;
Ŝ+ = Ŝx + iŜy (3.149)
ure
Ŝ− = Ŝx − iŜy (3.150)
Now, let us consider the action of Ŝ+ on the eigen state + 21 and − 12 .
1
1
Ŝ+ + = Ŝx + iŜy + (3.151)
2 2
1 1
= Ŝx + + iŜy + (3.152)
2 2
ct
ℏ 1 iℏ 1
= − + i. − (3.153)
2 2 2 2
ℏ 1 ℏ 1
= − − − (3.154)
2 2 2 2
Le
1
Ŝ+ + =0 (3.155)
2
Similarly,
1 1
Ŝ+ − = Ŝx + iŜy − (3.156)
2 2
61
1 1
= Ŝx − + iŜy − (3.157)
2 2
ℏ 1 −iℏ 1
= +i (3.158)
2 2 2 2
tes
ℏ 1 ℏ 1
= + (3.159)
2 2 2 2
1 1
Ŝ+ − =ℏ (3.160)
2 2
Following the same procedure, it can be shown that;
1 1
No
Ŝ− + =ℏ − (3.161)
2 2
1
Ŝ− − =0 (3.162)
2
We observe that the highest state that the system can ever find its self in is the state
+ 12 and thus, further action of the raising operator on it gives zero (0).
We therefore conclude that the only fruitful action of the lowering operator is to act
on + 12 to produce − 21 and the only meaningful action of the raising operator is to
ure
transform − 21 to + 21 . Hence, there are only two possible choices to take and are con-
ventionally been considered as ”spin-up” (f or + 21 ) and ”spin-down” (f or − 12 ).
Jˆ = L̂ + Ŝ (3.163)
ct
For an isolated system of n-particles, the total angular momentum operator is the sum
of the angular momentum operators of the individual particle.
n
X
Jˆ = Jˆi
Le
i=1
It is defined by its three components Jˆx ,Jˆy , and Jˆz , which satisfy the following commu-
tation relations;
[Jˆx , Jˆy ] = iℏJˆz (3.164)
62
[Jˆy , Jˆz ] = iℏJˆx (3.165)
or equivalently by;
tes
Jˆ × Jˆ = iℏJ⃗ (3.167)
Since Jˆx ,Jˆy , and Jˆz don’t commute, they can not be simultaneously diagonalized i.e. they
don’t posses common eigen states.
ˆ
J⃗2 = Jˆx2 + Jˆy2 + Jˆz2
No
(3.168)
This is a scalar operator, hence it commutes with Jˆx ,Jˆy , and Jˆz .
ˆ
[J⃗2 , Jˆi ] = 0 (i = x, y, z) (3.169)
All the operators; Jˆx ,Jˆy , and Jˆz are Hermitian and hence their eigen values are real.
ˆ
From equation (3.169), J⃗2 and Jˆz commute; Hence, they have simultaneous eigen func-
tions/vectors |j, m⟩ with the following eigen value problems;
ure
Jˆ2 |j, m⟩ = j(j + 1)ℏ2 |j, m⟩ (3.170)
ˆ
3.3.1 Lowering and Raising Operators of J⃗
ct
63
Some properties of these operators in relation to Jˆz and Jˆ2 are;
tes
[Jˆz , Jˆ− ] = −ℏJˆ− (3.176)
No
[Jˆz , Jˆ+ ] = [Jˆz , Jˆx ] + i[Jˆz , Jˆy ]
= iℏJˆx + i(−iℏJˆy )
= ℏ(Jˆx + iJˆy )
We can as well prove the other relations in (3.176), (3.177) and (3.178).
ure
Let us see the effect of Jˆ+ and Jˆ− on the eigen states of Jˆz and Jˆ2 ; i.e. |j, m⟩.
From equations (3.177) and (3.178) Jˆ± Jˆ2 = Jˆ2 Jˆ± ; therefore, equation (3.182) becomes;
ct
64
Jˆ+ Jˆz |l, m⟩ = mℏJˆ+ |j, m⟩ (3.186)
From equation (3.175), Jˆ+ Jˆz = Jˆz Jˆ+ − ℏJˆ+ . Therefore, equation (3.186) becomes;
Jˆz Jˆ+ − ℏJˆ+ |l, m⟩ = mℏJˆ+ |j, m⟩
tes
(3.187)
Equation (3.190) asserts that Jˆ+ |j, m⟩ is also an eigen function of Jˆz with the corre-
No
sponding eigenvalues (m + 1)ℏ. Hence Jˆ+ is a raising operator. It raises the eigenvalues
of Jˆz by one unit of ℏ.
Acting Jˆ− on both sides of equation (3.171) and making use of equation (3.176) yields;
Jˆz Jˆ− |l, m⟩ = (m − 1)ℏ Jˆ− |j, m⟩ (3.191)
Showing that Jˆ− |l, m⟩ is an eigen state of Jˆz with eigenvalues (m − 1)ℏ. Hence Jˆ− is a
lowering operator.
ure
3.4 Addition of Angular Momentum
ˆ ˆ
Consider two angular momenta J⃗1 and J⃗2 which belong to two different subspaces 1 and
ˆ ˆ
2. For J⃗1 and J⃗2 either referring to two distinct particles or two different properties of the
same particle.
ˆ ˆ ˆ
The total angular momentum J⃗2 can be expressed in terms of J⃗1 and J⃗2 as;
ct
ˆ ˆ ˆ
J⃗ = J⃗1 + J⃗2 (3.192)
ˆ ˆ
We assume that J⃗1 and J⃗2 are independent angular momenta, i.e. each satisfies the usual
angular momentum commutation relations.
Le
Where n = 1, 2 and i = x, y, z.
65
Also, any component of Jˆ1 commutes with any component of Jˆ2 ;
tes
so that the two angular momenta are compatible.
Now, denoting the joint eigstates of Jˆ12 and Jˆ1z by |j1 m1 ⟩, we have;
No
Also, denoting the joint eigstates of Jˆ22 and Jˆ2z by |j2 m2 ⟩, we have;
The four operators Jˆ12 , Jˆ22 ,Jˆ1z and Jˆ2z form a complete set of commuting operators; they
can thus jointly be diagonalised by the same states.
Denoting their joint eigenstates by |j1 j2 ; m1 m2 ⟩, we can write them as direct products of
ure
|j1 m1 ⟩ and |j2 m2 ⟩
|j1 j2 ; m1 m2 ⟩ = |j1 m1 ⟩ |j2 m2 ⟩ (3.199)
Because the coordinates of Jˆ1 and Jˆ2 are independent, we can rewrite equations (3.195)
to (3.198) as;
Jˆ12 |j1 j2 ; m1 m2 ⟩ = j1 (j1 + 1)ℏ2 |j1 j2 ; m1 m2 ⟩ (3.200)
! !
X X X
|j1 j2 ; m1 m2 ⟩ ⟨j1 j2 ; m1 m2 | = |j1 m1 ⟩ ⟨j1 m1 | |j2 m2 ⟩ ⟨j2 m2 | (3.204)
m1 m2 m1 m2
66
and since {|j1 m1 ⟩} and {|j2 m2 ⟩} are complete, also the basis {|j1 j2 ; m1 m2 ⟩} is complete.
j1 j2
X X
|j1 j2 ; m1 m2 ⟩ ⟨j1 j2 ; m1 m2 | = 1 (3.205)
tes
m1 =−j1 m2 =−j2
The basis {|j1 j2 ; m1 m2 ⟩} clearly spans the total space which is made of subspaces 1 and
No
2. From equation (3.199), we see that the dimensions N of this space is equal to the
product of dimensions of two subspaces spanned by {|j1 m1 ⟩} and {|j2 m2 ⟩};
Let us introduce the step operators Jˆ1± = Jˆ1x + iJˆ1y and Jˆ2± = Jˆ2x + iJˆ2y whose action
on |j1 j2 ; m1 m2 ⟩ gives;
1
Jˆ1± |j1 j2 ; m1 m2 ⟩ = ℏ ((j1 ∓ m1 )(j1 ± m1 + 1)) 2 |j1 j2 ; m1 ± 1, m2 ⟩ (3.209)
ure
1
Jˆ2± |j1 j2 ; m1 m2 ⟩ = ℏ ((j2 ∓ m2 )(j2 ± m2 + 1)) 2 |j1 j2 ; m1 , m2 ± 1⟩ (3.210)
ˆ ˆ
The problem of adding two angular momenta J⃗1 and J⃗2 consists of finding the eigenvalues
and eigenvectors of Jˆ2 and Jˆz in terms of the eigenvalues and eigenvectors of Jˆ12 , Jˆ22 ,Jˆ1z
and Jˆ2z .
67
From (3.215) |j1 j2 ; m1 m2 ⟩ is also an eigenstate of Jˆz with corresponding eigenvalues
(m1 + m2 ) ℏ.
tes
From (3.192),
ˆ ˆ ˆ 2
J⃗2 = J⃗1 + J⃗2 (3.216)
ˆ ˆ ˆ ˆ
= J⃗12 + J⃗22 + 2J⃗1 .J⃗2 (3.217)
ˆ ˆ
Since all the components of J⃗1 commutes with all those of J⃗2 and since;
No
[Jˆ12 , Jˆ1 ] = [Jˆ22 , Jˆ2 ] = 0 (3.218)
Also,
ˆ ˆ
2J⃗1 .J⃗2 = Jˆ1+ Jˆ2− + Jˆ1− Jˆ2+ + 2Jˆ1z Jˆ2z (3.219)
ˆ
Applying the operator J⃗2 to the states |j1 j2 ; m1 m2 ⟩
ˆ ˆ ˆ ˆ ˆ
ure
J⃗2 |j1 j2 ; m1 m2 ⟩ = J⃗12 + J⃗22 + 2J⃗1 .J⃗2 |j1 j2 ; m1 m2 ⟩ (3.220)
ˆ ˆ ˆ
J⃗2 |j1 j2 ; m1 m2 ⟩ = J⃗12 + J⃗22 + Jˆ1+ Jˆ2− + Jˆ1− Jˆ2+ + 2Jˆ1z Jˆ2z |j1 j2 ; m1 m2 ⟩ (3.221)
Using expressions in (3.200)...(3.203), (3.209) and (3.210), equation (3.221) can be sim-
plified.
⃗ˆ = 0
2. Prove that [L̂2 , L]
68
4. Prove the following commutation relations
(i) [L̂+ , L̂− ] = 2ℏL̂z
(ii) [L̂z , L̂+ ] = ℏL̂+
tes
(iii) [L̂z , L̂− ] = −ℏL̂−
5. Show that;
1
(a) L̂2 = 2
L̂+ L̂− + L̂− L̂+
6. Applying the operators L̂+ = L̂x + iL̂y and L̂− = L̂x − iL̂y on the eigenstates of L̂2
No
and L̂z (|l, m⟩) and interpret the physical meaning of the results.
(a) Calculate the norm of L̂+ |l, m⟩ and L̂− |l, m⟩.
(b) Calculate the eigenvalues of L̂2 and L̂z for the states L̂+ |l, m⟩ and L̂− |l, m⟩.
8. Evaluate;
(a) [L̂2 , p⃗ˆ]
(b) [L̂, p⃗ˆ2 ]
ure
⃗ˆ × L]
(c) [L̂, L ⃗ˆ
10. Calculate [L̂2x , L̂y ],[L̂2z , L̂y ] and [L̂2 , L̂y ], then show that ⟨l, m| L̂2x |l, m⟩ = ⟨l, m| L̂2y |l, m⟩.
h i
2 2 2 2
11. Show that L̂ , [L̂, x̂] = 2ℏ x̂L̂ + L̂ x̂ .
D E D E
12. Find L̂x and L̂y in the state |ψ⟩ = √13 (|1, 1⟩ + |1, 0⟩ + |1, −1⟩) where the first
Le
69
14. Prove the following relations;
(a) Ŝ+ + 21 = 0
(b) Ŝ+ − 12 = ℏ
2
+ 12
tes
(c) Ŝ− + 21 = ℏ − 21
(d) Ŝ− − 12 = 0
ℏ2
(e) Ŝx2 = Ŝy2 = Ŝz2 = 4
No
(i) [σ̂x , σ̂x ] = 2iσ̂z
(ii) [σ̂y , σ̂z ] = 2iσ̂x
(iii) [σ̂z , σ̂x ] = 2iσ̂y
(iii) Ŝy + 21 = iℏ
2
− 21
(iv) Ŝy − 12 = − iℏ2 + 21
19. Prove that the operators Ŝ+2 and Ŝ−2 are zero for the spin - 12 space.
Le
70
(b) Find the expectation values of Ŝx ,Ŝy and Ŝz
(c) Find the uncertainties ∆Ŝx , ∆Ŝy and ∆Ŝz .
tes
!
1 i
χ= √
5 2
No
h D E i
Hint; Ŝi = χ† Ŝi χ, i = x, y, z
ˆ ˆ
25. Let the operators J⃗1 and J⃗2 be the respective angular momenta of the individual
ˆ ˆ ˆ
components of a two system. The total system has angular momentum J⃗ = J⃗1 + J⃗2 .
ct
Show that;
ˆ ⃗ˆ
(a) J1 .J2 = 2 J1+ J2− + J1− J2+ + Jˆ1z Jˆ2z
⃗ 1 ˆ ˆ ˆ ˆ
ˆ2 ⃗ˆ2 ˆ ˆ
ˆ2 ⃗ ˆ ˆ ˆ
(b) J = J1 + J2 + J1+ J2− + J1− J2+ + 2J1z J2z ˆ
Le
71
4 The Hydrogen Atom
The hydrogen atom involves two particles (the electron and the proton). The potential
energy of the electron in the electrostatic field of the proton is;
tes
e2 1 e2
V (r) = −k =− (4.1)
r 4πϵ0 r
were e is the electron charge and r is the distance of separation between the electron and
the proton.
Let ψ be the wave function that describes all the states of the electron around the nucleus
No
such that the three dimension Schrodinger equation for the system is;
ℏ2 2
− ∇ ψ + V (x, y, z)ψ = Eψ (4.2)
2µ
were ∇2 is the Laplacian operator, E is the total energy and µ is the reduced mass.
me M
µ= (4.3)
me + M
were me is the mass of the electron and M is the mass of the nucleus.
ure
Since the nucleus mass M is mach larger than the electron mass me , the reduced mass µ
is very close to the electron mass me .
We can now take advantage of symmetry and replace the Cartesian coordinates with the
spherical coordinates r, θ and ϕ, since the potential V (r) is a central potential i.e. it only
depends on distance not direction.
ct
72
tes
No
Figure 4.1: Transformation of Cartesian Coordinates to Spherical
From triangle AOP in Figure 4.1, the side OA is simply the z-coordinate and can be
ure
obtained as;
OA
= cos θ, z = r cos θ (4.4)
OP
Similarly, from triangle AOP ,
AP
= sin θ, AP = OP cos θ, OQ = r sin θ (4.5)
OP
From triangle BOQ , the side OB is the x-coordinate and it can be obtained as;
OB
ct
Since the side BQ is equal to OC, BQ also represent the y-coordinate and it can be
obtained as;
BQ
= sin ϕ, BQ = OQ sin ϕ (4.8)
OQ
This implies that;
y = r sin θ sin ϕ (4.9)
73
Using equations (4.4), (4.7) and (4.9), equation (4.2) can be expressed in spherical polar
coordinates as;
ℏ2 1 ∂ ∂2
2 ∂ 1 ∂ ∂ 1
− r + 2 sin θ + 2 2 ψ + V (r)ψ = Eψ (4.10)
tes
2µ r2 ∂r ∂r r sin θ ∂θ ∂θ r sin θ ∂ϕ2
were ψ = ψ(r, θ, ϕ)
No
Therefore, ψ(r, θ, ϕ) is an eigenfunction.
Equation (4.10) is separable and the plan is to look for a variable separable solution such
that;
ψ(r, θ, ϕ) = R(r)f (θ)g(ϕ) (4.12)
∂ 2 ∂ 1 ∂ ∂
r R(r)f (θ)g(ϕ) + sin θ R(r)f (θ)g(ϕ)+
∂r ∂r sin θ ∂θ ∂θ
(4.14)
1 ∂2 2µr2 e2
R(r)f (θ)g(ϕ) + 2 E+k R(r)f (θ)g(ϕ) = 0
sin2 θ ∂ϕ2 ℏ r
∂ 2 ∂ 1 ∂ ∂
r Rf g + sin θ Rf g+
∂r ∂r sin θ ∂θ ∂θ
(4.15)
1 ∂2 2µr2 e2
Rf g + 2 E+k Rf g = 0
sin2 θ ∂ϕ2 ℏ r
74
1 ∂ 2 g 2µr2 e2
∂ 2 ∂ 1 ∂ ∂f
fg r + Rg sin θ + 2 + 2 E+k Rf g = 0 (4.16)
∂r ∂r sin θ ∂θ ∂θ sin θ ∂ϕ2 ℏ r
Dividing equation (4.16) by Rf g gives
tes
1 ∂ 2 g 2µr2 e2
1 ∂ 2 ∂R 1 ∂ ∂f
r + sin θ + + 2 E+k =0 (4.17)
R ∂r ∂r f sin θ ∂θ ∂θ g sin2 θ ∂ϕ2 ℏ r
1 ∂ 2 g 2µr2 e2
1 ∂ 2 ∂R 1 ∂ ∂f
r + sin θ + + 2 E+k =0 (4.18)
R ∂r ∂r f sin θ ∂θ ∂θ g sin2 θ ∂ϕ2 ℏ r
2µr2 e2 1 ∂ 2g
1 ∂ 2 ∂R 1 ∂ ∂f
r + 2 E+k =− sin θ − (4.19)
R ∂r ∂r ℏ r f sin θ ∂θ ∂θ g sin2 θ ∂ϕ2
No
Equation (4.19) holds true if we equate both sides of it to a constant α
2µr2 e2
1 ∂ 2 ∂R
r + 2 E+k =α (4.20)
R ∂r ∂r ℏ r
1 ∂ 2g
1 ∂ ∂f
sin θ + = −α (4.21)
f sin θ ∂θ ∂θ g sin2 θ ∂ϕ2
Equation (4.20) contains only r variable and therefore, it is called a radial equation.
However, equation (4.21) still contains two variables and therefore further separation is
needed.
ure
Multiplying through equation (4.21) by sin2 θ gives;
1 ∂ 2g
sin θ ∂ ∂f
sin θ + = −α sin2 θ (4.22)
f ∂θ ∂θ g ∂ϕ2
1 ∂ 2g
sin θ ∂ ∂f
sin θ + α sin2 θ = − (4.23)
f ∂θ ∂θ g ∂ϕ2
Equation (4.23) holds true if we equate both sides of it to a constant m2l such that;
ct
sin θ ∂ ∂f
sin θ + α sin2 θ = m2l (4.24)
f ∂θ ∂θ
1 ∂ 2g
= −m2l (4.25)
g ∂ϕ2
Le
Let us call equation (4.24) are theta equation and equation (4.25) a phi equation.
75
4.1.1 Solutions of R(r), f (θ) and g(ϕ) Equations (Radial and Angular Eigen-
functions)
tes
From equation (4.25),
∂ 2g
+ gm2l = 0 (4.26)
∂ϕ2
The general solution of such equation is
No
were A represents the normalization constant, with ml = 0, ±1, ±2, .... The quantity ml
is called the magnetic quantum number.
Replacing ϕ with ϕ + 2π, the position of the point under consideration should remain the
same i.e.
g(ϕ + 2π) = g(ϕ) (4.28)
ure
Therefore,
Aeiml (ϕ+2π) = Aeiml ϕ (4.29)
ei2πml = 1 (4.30)
ct
From Euler’s expansion, eix = cos x + i sin x. Using this, equation (4.30) takes the form
76
(ii) The Normalization constant of g(ϕ)
Z 2π
g ∗ gdϕ = 1 (4.32)
0
tes
Z 2π
A 2
e−iml ϕ eiml ϕ dϕ = 1
0
Z 2π
2
A dϕ = 1
0
A2 (2π) = 1
r
1
No
A= (4.33)
2π
Therefore, equation (4.27) becomes;
r
1 iml ϕ
g(ϕ) = e (4.34)
2π
m2l
1 ∂ ∂f
sin θ + αf − f =0 (4.36)
sin θ ∂θ ∂θ sin2 θ
ct
m2l
1 ∂ ∂f
sin θ + α− f =0 (4.37)
sin θ ∂θ ∂θ sin2 θ
∂y
Let y = cos θ, such that sin2 θ = (1 − y 2 ) and ∂θ
= − sin θ. Also,
Le
∂f ∂f ∂y
= (4.38)
∂θ ∂y ∂θ
∂f ∂f
= − sin θ (4.39)
∂θ ∂y
77
Multiplying both sides of equation (4.39) by sin θ gives;
∂f ∂f
sin θ = − sin2 θ (4.40)
∂θ ∂y
tes
∂f ∂f
sin θ = − 1 − y2 (4.41)
∂θ ∂y
Substituting equation (4.39) and (4.41) into equation (4.37) gives;
m2l
1 ∂f 2 ∂f
− sin θ − 1−y + α− f =0 (4.42)
sin θ ∂y ∂y 1 − y2
m2l
∂f ∂f
No
2
1−y + α− f =0 (4.43)
∂y ∂y 1 − y2
Equation (4.43) is a Legendre’s polynomial and it has physical significance only in the
range of y = +1 to y = −1.
Therefore, considering the one more form of f function so that this condition is satisfied
gives;
ml
f (θ) = 1 − y 2 2 Y (y) (4.44)
∂f ml ml dY
= −ml y(1 − y 2 ) 2 −1 Y + 1 − y 2 2 (4.45)
∂y dy
∂
Multiplying equation (4.45) by (1 − y 2 ) and ∂y
gives;
∂ ∂f2 ∂ m
2 2l 2 2 +1 dY
ml
1−y = −ml y(1 − y ) Y + 1 − y (4.46)
ct
∂y ∂y ∂y dy
h
2
ml m2l −1 i h m2l i ′
= −ml y(1 − y ) 2 + m2l y 2 1−y 2
Y − 2y(ml + 1) 1 − y 2
Y +
h m i
l +1 ′′ (4.47)
1 − y2 2 Y
Le
h ml +1 i ′′
m2l
ml (4.48)
1 − y2 2 Y + α− 1 − y 2 2
Y =0
1 − y2
78
ml
Dividing through equation (4.48) by (1 − y 2 ) 2
gives;
′′ ′
1 − y 2 Y + 2(ml + 1)yY + [α − ml (ml + 1)] Y = 0 (4.49)
tes
′′ ′
1 − y 2 Y + 2βyY + λY = 0 (4.50)
Y = a0 + a1 y + a2 y 2 + a3 y 3 + ... (4.51)
No
Such that;
′
Y = a1 + 2a2 y + 3a3 y 2 + 4a4 y 3 + ... (4.52)
′′
Y = 2a2 + 6a3 y + 12a4 y 2 + 20a5 y 3 + ... (4.53)
Substituting equations (4.51), (4.52) and (4.53) into equation (4.50) gives;
The above equation is satisfied only if each term on the left hand side is individually
equal to zero ie., the coefficients vanish.
The general expression for the coefficients follow the condition below;
For n = 0, 1, 2, 3, ......
2nβ + n(n − 1) − λ
an+2 = an (4.57)
(n + 1)(n + 2)
Substituting the values of β and λ in equation (4.57) gives
Le
an+2 (n + ml )(n + ml + 1) − α
= (4.58)
an (n + 1)(n + 2)
In order to obtain a valid wave function, the power series must contain a finite number of
79
terms number of terms which is only possible if the numerator in equation (4.58) becomes
zero.
(n + ml )(n + ml + 1) − α = 0 (4.59)
tes
α = (n + ml )(n + ml + 1) (4.60)
Since n and ml are both whole numbers, their sum must also be a whole number. Let
(n + ml ) = l such that;
α = l(l + 1) (4.61)
No
Putting the value of α into equation (4.43) we get;
m2l
∂f ∂f 2
1−y + l(l + 1) − f =0 (4.62)
∂y ∂y 1 − y2
f = Bpm ml
l (y) = Bpl (cos θ)
l
(4.63)
were B is the normalization constant and pm l (y) is the associated Legendre function
l
ure
defined as;
ml dml p (y)
l
pm
l
l
(y) = (1 − y) 2
m
(4.64)
dy l
were pl (y) is the Legendre polynomial given by;
1 dl (y 2 − 1)l
pl (y) = (4.65)
2l l! dy l
Z +1
2 (l + ml )!
pml ml
k (y)pl (y) = if k = l (4.67)
−1 2l + 1 (l − ml )!
(ii) Normalization of f .
80
Z +1
∗
fm l l fm ll
dθ = 1 (4.68)
−1
Z +1
2
pm ml
tes
B k (y)pl (y) = 1
l
(4.69)
−1
2 (l + ml )!
B2 =1 (4.70)
2l + 1 (l − ml )!
s
2l + 1 (l − ml )!
B= (4.71)
2 (l + ml )!
Substituting equation (4.71) into equation (4.63) gives;
No
s
2l + 1 (l − ml )! ml
fl,ml (θ) = Bpm
l (y) =
l
p (cos θ) (4.72)
2 (l + ml )! l
e2
1 ∂ 2 ∂R 2µ l(l + 1)
r + 2 E+k R= R (4.74)
r2 ∂r ∂r ℏ r r2
1 2 ∂ 2R e2
∂R 2µ l(l + 1)
r + 2r + 2 E+k − R=0 (4.75)
ct
r2 ∂r2 ∂r ℏ r r2
∂ 2 R 2 ∂R 2µE 2µke2 l(l + 1)
+ + + 2 − R=0 (4.76)
∂r2 r ∂r ℏ2 ℏr r2
Assume that the electron moving around the nucleus is some how bound. i.e.
Le
2µE µke2
− = α2 and =λ (4.77)
ℏ2 ℏ2 α
∂ 2 R 2 ∂R
2 2αλ l(l + 1)
+ + −α + − R=0 (4.78)
∂r2 r ∂r r r2
81
Let us now define a new variable ρ = 2αr such that
∂ρ
= 2α (4.79)
∂r
tes
It follows that;
∂R ∂R ∂ρ ∂R
= . = 2α (4.80)
∂r ∂ρ ∂r ∂ρ
Also,
∂ 2R
∂ ∂R ∂ ∂R ∂ ∂ρ ∂R ∂ρ ∂ ∂R
2
= = 2α = 2α = . 2α (4.81)
∂r ∂r ∂r ∂r ∂ρ ∂ρ ∂r ∂ρ ∂r ∂ρ ∂ρ
No
∂ 2R ∂ 2R
∂ ∂R
= 2α 2α = 4α2 2 (4.82)
∂r2 ∂ρ ∂ρ ∂ρ
Substituting equations (4.80), (4.82) into equation (4.78) gives;
2
R 2
2∂ ∂R 2 2αλ l(l + 1)
4α + (2ρ) + −α + − R=0 (4.83)
∂ρ2 r ∂ρ r r2
∂ 2R
1 ∂R 1 λ l(l + 1)
+ + − + − R=0 (4.84)
ure
∂ρ2 αr ∂ρ 4 2αr 4α2 r2
∂ 2 R 2 ∂R
1 λ l(l + 1)
+ + − + − R=0 (4.85)
∂ρ2 ρ ∂ρ 4 ρ ρ2
∂ 2R 1
− R=0 (4.86)
ct
∂ρ2 4
ρ
R(ρ) = e 2 (4.87)
Le
ρ
R(ρ) = e− 2 (4.88)
ρ
The function R(ρ) = e 2 ←→ ∞ as ρ ←→ ∞. Hence, it is not acceptable. Therefore,
ρ
R(ρ) = e− 2 (4.89)
82
The acceptable solution is only valid for large values of ρ. For this reason we have to
think of a pre exponential part to be part of this solution so that it is valid for all values
of ρ.
tes
By incorporating some ρ -dependent unknown function F (ρ), equation (4.89) becomes;
ρ
R(ρ) = F (ρ)e− 2 (4.90)
Differentiating equations (??) w.r.t ρ at first and second order and then substituting the
2
values of R(r), ∂R
∂ρ
and ∂∂ρR in equation (4.85)
No
∂ 2F
2 ∂F 1 λ l(l + 1)
+ −1 + − + − F =0 (4.91)
∂ρ2 ρ ∂ρ ρ ρ ρ2
Hence, the problem has been reduced to determining the solution of F which can be
assumed a;
G(ρ) = a0 + a1 ρ + a2 ρ2 + a3 ρ3 + ....
ure
f or a0 ̸= 0 (4.93)
j=∞
X
G(ρ) = aj ρ j (4.94)
j=0
′ ′
F (ρ) = sρs−1 G + ρs G (4.95)
ct
′′ ′ ′′
F (ρ) = s(s − 1)ρs−2 G + 2sρs−1 G + ρs G (4.96)
Substituting equations (4.92), (4.95) and (4.96) into equation (4.91) gives;
Le
s−2 s−1 ′ s 2′′ ′
s(s − 1)ρ G + 2sρ G +ρ G + − 1 sρs−1 G + ρs G +
ρ
(4.97)
1 λ l(l + 1) s
− + − ρ G(ρ) = 0
ρ ρ ρ2
83
Multiplying through equation (4.97) by 4ρ2 we get;
2 s−2 2 s−1 ′ 2 s ′′ 2 2 ′
4ρ s(s − 1)ρ G + 8sρ ρ G + 4ρ ρ G + 4ρ − 1 sρs−1 G + 4ρ2 ρs G +
ρ
(4.98)
tes
2 1 λ l(l + 1) s
4ρ − + − ρ G(ρ) = 0
ρ ρ ρ2
No
Dividing through equation (4.99) by ρs gives;
′
[4s(s − 1) + 8s − 4sρ − 4ρ + 4λρ − 4l(l + 1)] G + 8sρ + 8ρ − 4ρ2 G +
′′
(4.100)
4ρ2 G = 0
Since a0 ̸= 0,
ure
[4s(s − 1) + 8s − 4l(l + 1)] = 0 (4.102)
′
[4l(l − 1) + 8l − 4lρ − 4ρ + 4λρ − 4l(l + 1)] G + 8lρ + 8ρ − 4ρ2 G +
′′
(4.107)
4ρ2 G = 0
′ ′′
[−4lρ − 4ρ + 4λρ] G + 8lρ + 8ρ − 4ρ2 G + 4ρ2 G = 0
(4.108)
84
Dividing through equation (4.108) by 4ρ gives;
′ ′′
[−l − 1 + λ] G + [2l + 2 − ρ] G + ρG = 0 (4.109)
tes
Differentiating equation (4.94) at first and second order gives;
∞
′
X
G (ρ) = aj .j.ρj−1 (4.110)
j=0
∞
′′
X
G (ρ) = aj .j.(j − 1)ρj−2 (4.111)
j=0
No
Substituting equations (4.94), (4.110) and (4.111) into equation (4.109) gives;
∞
X ∞
X ∞
X
j j−1
[−l − 1 + λ] aj ρ + [2l + 2 − ρ] aj .j.ρ +ρ aj .j.(j − 1)ρj−2 = 0 (4.112)
j=0 j=0 j=0
The above equation holds true only if the coefficients of individual powers of ρ become
zero. Simplifying equation (4.112) for two summation terms aj and aj+1 we get;
85
Since the series G(ρ) consists of finite number of terms, the function F (ρ) becomes finite
at very large values of j and consequently, the function R(ρ) will also become infinite if
the number of the terms is not limited to finite values.
tes
To break off the series to a finite number of terms, we equate the numerator in equation
(4.119) to zero.
l+1−λ+j =0 (4.120)
λ=l+1+j (4.121)
No
λ=l+1+j =n (4.122)
Moreover, as n > l + 1, the largest value that l can take is n − 1. Hence, the values of l
has a domain ranging from 0 to n − 1.
′ ′′
[q − p] G + [p + 1 − ρ] G + ρG = 0 (4.124)
The solutions of the above equation is the associated laguerre polynomial multiplied by
a factor i.e.
ct
The constant C can be set as normalized constant and the associated laguerre polynomial
is;
n−l−1
2l+1
X (−1)j+1 [(n + 1)!]2 ρj
Le
ρ
R(ρ) = ρs G(ρ)e− 2 (4.127)
86
Substituting equation (4.125) into equation (4.127) gives;
ρ
R(ρ) = Ce− 2 ρl Lpq (ρ) (4.128)
tes
Since s = l
n−l−1
− ρ2 l
X (−1)j+1 [(n + 1)!]2 ρj
Rn,l (ρ) = Ce ρ (4.129)
j=0
(n − l − 1 − j)!(2l + 1 + j)!j!
No
Let us put the squared integral over whole configuration space sd unity
Z ∞
2
Rn,l (r).r2 dr = 1 (4.130)
0
The r2 factor is introduced to convert the length dr into a volume around the center of
the nucleus.
But ρ = 2αr
1
−2µE 2
ure
ρ=2 r (4.131)
ℏ2
1
−8µE 2
= r (4.132)
ℏ2
were E is the kinetic energy of the electron.
21
2µe2
ρ= r (4.133)
n(4πϵ0 )ℏ2
ct
2
ρ= r (4.134)
na0
4πϵ0 ℏ2
where a0 = me2
is the Bohr radius.
2
dρ = dr (4.135)
na0
This implies that;
na0
dr = dρ (4.136)
2
87
Substituting equation (4.128) and (4.136) into equation (4.130) gives;
Z ∞ 2 h na0 i2 na0
2
eρ ρ2l Lpq (ρ)
C dρ = 1 (4.137)
0 2 2
tes
h na i3 Z ∞ 2
2 0
eρ ρ2l Lpq (ρ) dρ = 1
C (4.138)
2 0
" #
h na i3 2n [(n + l)!]3
0
C2 =1 (4.139)
2 (n − l − 1)!
s 3
2 (n − l − 1)!
C= (4.140)
na0 2n [(n + l)!]3
No
The complete normalized radial eigenfunction for the hydrogen atom is;
s 3
2 (n − l − 1)! − ρ l p
Rn,l (ρ) = e 2 ρ Lq (ρ) (4.141)
na0 2n [(n + l)!]3
The first three normalized radial eigen functions of the hydrogen atom are;
32
1 − ar
R10 (r) = 2 e 0 (4.142)
a0
ure
3
1 2 r − r
R20 (r) = 2− e 2a0 (4.143)
2a0 a0
3
1 1 2 r − r
R20 (r) = √ e 2a0 (4.144)
3 2a0 a0
Figure 4.2 shows the graphs of these three radial functions. Note that only for the
s-states (l = 0) are the radial functions different from zero at r = 0. This is due to the
presence of the factor rl in the expression (4.141).
ct
Note. The ground state of hydrogen is designated as the 1s state where ′ 1′ indicates the
energy level (n = 1) and ′ s′ indicates the orbital angular momentum state (l = 0)
Le
88
tes
length is a0
From λ = µke2
ℏ2 α
,
No
Figure 4.2: Radial eigen functions R10 (r), R20 (r) and R21 (r) for hydrogen atom. The unit of
ure
µ2 k 2 e4
λ2 = (4.145)
ℏ4 α2
−2µE
Substituting the value of α2 = ℏ2
in equation (4.145) gives;
µ2 k 2 e4 −ℏ2 µk 2 e4
2
λ = =− (4.146)
ℏ4 2µE 2Eℏ2
But λ = n. Therefore,
µk 2 e4
n2 = −
ct
(4.147)
2En ℏ2
2
µe4
1
En = 2 2 (4.148)
n ℏ 4πϵ0
µe4
En = (n = 1, 2, 3, ...) (4.149)
Le
32π 2 ϵ20 ℏ2 n2
This is the famous Bohr formula. Bohr could not quite anticipate the correct assignment
of angular momentum to the energy levels.
89
The complete normalized eigen functions for the hydrogen atom are;
tes
For,
n = 1, 2, 3, 4, ......
l = 0, 1, 2, ....(n − 1) (4.151)
ml = −l, −l + 1, ...., 0, ....l − 1, l
The complete eigen functions for the lowest few states of hydrogen are;
No
32
1 1 − ar
ψ100 =√ e 0 (4.152)
π a0
32
1 1 r − 2ar
ψ200 =√ 1− e 0 (4.153)
π 2a0 2a0
5
11 2 − 2ar
ψ210 =√ re 0 cos θ (4.154)
2a0
π
52
1 1 − r
ψ21±1 = √ re 2a0 sin θe±iϕ (4.155)
ure
8 π a0
Example 4.10
Calculate the most probable distance of the electron in the ground state of the hydrogen
atom. What is the radial probability density of that distance?
Sln.
0
a0
32 2
2 1 − ar
P10 (r) = r 2 e 0 (4.158)
a0
4 2 − a2r
= r e 0 (4.159)
a30
90
dP10 (r)
P10 (r) will be maximum when dr
=0
2 − 2r − 2r
− r2 e a0 + 2re a0 = 0 (4.160)
a0
tes
r = a0 (4.161)
This is the most probable distance. The maximum value of the radial probability density
is;
4 −2
.a
(P10 )max = 3 (a0 )2 .e a0 0 (4.162)
a0
4 −2
(P10 )max = e (4.163)
a0
No
Example 4.11
Calculate 1r for an electron in the ground state of an hydrogen atom and use the results
to calculate.
1 1
= |R10 (r)|2 r2 . dr (4.166)
r 0 r
Z ∞
4 −2r
= 3 re a0 dr (4.167)
a0 0
1 1
= (4.168)
Le
r a0
(a)
−e2 −e2
1
⟨V (r)⟩ = =− (4.169)
4πϵ0 r 4πϵ0 r
91
−e2 1
= . (4.170)
4πϵ0 a0
−e2 me2
= . (4.171)
4πϵ0 4πϵ0 ℏ2
tes
2 2
m e
⟨V (r)⟩ = − 2 (4.172)
ℏ 4πϵ0
⟨V (r)⟩ = −mc2 α2 (4.173)
(b)
⟨T ⟩ = E − ⟨V (r)⟩ (4.174)
2 2 " 2 2 #
No
m e m e
=− 2 − − 2 (4.175)
2ℏ 4πϵ0 ℏ 4πϵ0
2
e2
m
= 2 (4.176)
2ℏ 4πϵ0
mc2 α2
⟨T ⟩ = (4.177)
2
Note that from the above, ⟨T ⟩ = − 12 ⟨V (r)⟩.
Example 2.12
ure
The ground state wave function for the hydrogen atom is
1 − r
ψ(r) = p 3 e a0
πa0
a0
(b) between 2
and 2a0 from the nucleus.
Sln.
Le
(a)
Z a0
P1 = |ψ(r)|2 dτ, dτ = 4πr2 dr (4.178)
0
92
!2
Z a0
1 − ar
2
P1 = p e 4πr 0dr (4.179)
0 πa30
4 a0 2 − a2r
Z
= 3 r e 0 dr (4.180)
tes
a0 0
But;
a0 r2 a0 r2 a30
Z
2r
2 − a0
r e dr = − − − (4.181)
2 2 4
Taking the limit from 0 to a0 ;
a0
5a30 −2 a30
Z
2r
2 − a0
r e dr = − e + (4.182)
0 4 4
No
a30
1 − 5e−2
= (4.183)
4
a3 4
P1 = 0 3 1 − 5e−2 = 0.323
(4.184)
4 a0
(b) Z 2a0
4 − a2r
P2 = 3 r2 e 0 dr (4.185)
a0 a0
2
5
= e−1 − 13e−4 (4.186)
ure
2
P2 = 0.682 (4.187)
The principal quantum number (n) results from the solution of the radial wave function
R(r) in equation (4.73). It specifies the energy of the electron in the atom and the average
distance (r) between the electron and the nucleus.
93
13.6eV
En = − (4.188)
n2
2
n
r= 0.529Å (4.189)
tes
Z
It also represent the maximum number of electrons in an orbit as n2 . The principal takes
on integral values.
n = 1, 2, 3, ....... n>0 (4.190)
The orbital quantum number l is associated with the R(r) and f (θ) parts of the wave
No
function f (r, θ, ϕ). It determines the number of sub shells or sub levels to which the elec-
tron belongs. It also specifies the shape of an orbital with a particular principal quantum
number.
The quantum number l is related to the magnitude of the orbital angular momentum L
by
p
L = l(l + 1)ℏ (4.191)
This quantum result disagrees with the elementary Bohr theory of the hydrogen atom
ure
p
where L = nℏ. This is obvious in the state l = 0, L = 0(1)ℏ. Based on these results,
we will have to discard Bohr’s semiclassical ’planetary’ model of electrons orbiting the
nucleus.
l: 0 1 2 3 4 5 ...(n − 1)
(4.192)
Letter : s p d f g h ...
ct
Atomic states are normally reflected to their n and l numbers. Thus, a state/sub shell with
n = 2 and l = 1 is called a 2p state/sub shell. Examples of other various atomic states/sub
shells are 1s(n = 1, l = 0), 2s(n = 2, l = 0), 4d(n = 4, l = 2) and 6g(n = 6, l = 4).
Le
A state/sub shell such as 2d is not possible because this refers to n = 2 and l = 2; our
boundary condition requires that n > l.
The value of l also has a slight effect on the energy of the sub shell. i.e. the energy of
the sub shell increases with l (s < p < d < f ).
94
4.2.3 Magnetic Quantum Number ml
The angle ϕ is a measure of the rotation about the z-axis. The solution to g(ϕ) in
equation (4.34) specifies that ml is an integer and relate to the z-component of the
tes
angular momentum L ⃗ by;
Lz = ml ℏ (4.193)
The value of ml varies from −l to l through zero and it specifies the orientation in space
of an orbital of a given energy and shape.
This number divides the sub shells into individual orbits which hold the electrons; there
are 2l + 1 orbitals in each sub shell. Thus, the s sub shell has only one orbital, the p sub
No
shell has three orbitals and so on.
4.3 The Spin Quantum Number (s) and the Pauli Exclusion
Principle
4.3.1 The Spin Quantum Number (s)
The spin quantum number s describes the intrinsic angular momentum of the electron
within the orbital and it gives the projection of the spin angular momentum along the
ure
specific axis.
The value of s is ± 12 which signifies the spin or rotation or direction of an electron on its
axis during motion.
The Pauli exclusion principle states that no two electrons in an atom can have the same
set of quantum numbers (n, l, ml , ms ).
ct
This means that no more than two electrons can occupy the same orbital and that two
electrons in the same orbital must have opposite spins.
Le
Because an electron spins, it creates a magnetic field which can be oriented in one or
two directions. For two electrons in the same orbital, the spins must be opposite to each
other; the spins are said to be paired. These substances are not attracted to magnets
and are said to diamagnetic. Atoms with more electrons that spin in one direction than
another contains unpaired electrons. These substances are weakly attracted to magnets
95
and are said to be paramagnetic.
The exclusion principle plays an important role in the structure of atoms. It has a direct
effect on the spatial distribution of fermions (all particles of half-integer spin).
tes
Read about electron configuration and the periodic table.
No
ure
⃗ due to the
The single electron of the hydrogen atom can feel an internal magnetic field B
ct
proton. A careful examination of this effect shows that the spins of the electron and the
orbital angular momentum interact, an effect called the spin-orbital coupling .
The magnetic field B⃗ exerts a torque on the spinning electron, tending to align its mag-
netic moment (⃗µ) along the direction of the field.
Le
The electron in its rest frame sees the proton moving at a speed u in a circular path.
96
From classical electrodynamics, the magnetic field experienced by the electron is;
B ⃗ = − 1 p⃗ × E
⃗ = − 1 ⃗u × E ⃗ = 1 E ⃗ × p⃗ (4.195)
c me c me c
tes
⃗ is the electric field generated by the pro-
where p⃗ = m⃗u is the linear momentum and E
ton’s coulomb potential field.
For an hydrogen atom where an electron moves in a central potential of a nucleus V (r) =
−eϕ(r), the electric field is;
⃗ r) = −∇ϕ(r) = 1 ∇V (r) = 1 . ⃗r dV
E(⃗ (4.196)
No
e e r dr
e 2
But V (r) = − 4πϵ0r
e2
⃗ r) 1 . ⃗r
E(⃗ =
e
⃗r (4.197)
e r 4πϵ0 r2 4πϵ0 r3
Substituting equation (4.197) into (4.195) gives;
⃗ = 1
B
e
⃗r × p⃗ =
e ⃗
L (4.198)
me c 4πϵ0 r3 4πϵ0 me cr3
⃗ BY;
ure
The magnetic moment µ
⃗ of the electron is related to its spin S
e ⃗
µ
⃗= S (4.199)
me c
e2
1 ⃗L ⃗
Ĥ = S. (4.201)
4πϵ0 me c2 r3
2
This expression is obtained by considering the rest frame of the electron; actually, the
electron accelerates as it orbits around the nucleus. To correct this, let us make an
appropriate kinematic correction known as the Thomas precession by throwing in a factor
Le
e2
1 ⃗L ⃗
ĤSO = S. (4.202)
8πϵ0 me c2 r3
2
97
For quantum mechanics, we deal with operators, hence equation (4.202) becomes;
e2
1 ⃗ˆ L
⃗ˆ
ĤSO = S. (4.203)
tes
8πϵ0 m e c2 r 3
2
We can use the perturbation theory to calculate the contribution of spin -orbit interaction
in the hydrogen atom. The Hamiltonian of the system is;
p⃗ˆ2 e2 e2
Ĥ = −k + ⃗ˆ L
S. ⃗ˆ (4.204)
2me r 2 2
8πme c r 3
No
Ĥ = Ĥ0 + ĤSO (4.205)
where Ĥ0 is the unperturbed Hamiltonian and ĤSO is the perturbed Hamiltonian.
⃗ and
In the presence of spin-orbit coupling, the Hamiltonian no-longer commutes with L
⃗ so, the spin and orbital are not separately conserved.
S,
However, Ĥ0 commutes with L2 , S 2 and the total angular momentum J⃗ given by;
J⃗ = L
⃗ +S
⃗ (4.206)
ure
Hence, these quantities are conserved.
Considering the perturbation, the eigenstates of Lz and Sz are not good states to use but
the eigenstates of L2 , S 2 , J 2 and Jz are.
Now,
⃗ˆ L
⃗ˆ
2 ⃗ ⃗ ⃗ ⃗
J = L + S . L + S = L̂2 + Ŝ 2 + 2S. (4.207)
ct
Also,
⃗ˆ L
⃗ˆ = L. ⃗ˆ = 1 Jˆ2 − L̂2 − Ŝ 2
⃗ˆ S
S. (4.208)
2
⃗ˆ S,
Let |nljm⟩ be the eigenstates of L. ⃗ˆ their corresponding eigenvalues are;
Le
2
⃗ˆ S
⟨nljm| L. ⃗ˆ |nljm⟩ = ℏ [j(j + 1) − l(l + 1) − s(s + 1)] (4.209)
2
98
For s = 12 , equation (4.209) becomes;
ℏ2
ˆ ⃗ˆ
⃗ 3
⟨nljm| L.S |nljm⟩ = j(j + 1) − l(l + 1) − (4.210)
2 4
tes
The eigen values of Ĥ (4.205) are given to the first order correction by;
13.6eV (0)
Enlj = − 2
+ ESO (4.212)
n
(0)
where − 13.6eV
n2
are the energy levels of hydrogen and ESO is the energy due to the spin-
No
orbit interaction.
e2 ℏ2
(0) 3 1
ESO = ⟨nljm| ĤSO |nljm⟩ = j(j + 1) − l(l + 1) − ⟨nl| |nl⟩ (4.213)
16πϵ0 m2e c2 4 r3
But,
1 2
⟨nl| |nl⟩ = 3 (4.214)
r 3 n l(l + 1)(2l + 1)a30
Substituting equation (4.214) into equation (4.213) gives;
e2 ℏ2
(0) 3 2
ESO = j(j + 1) − l(l + 1) − (4.215)
ure
2
16πϵ0 me c 2 4 n l(l + 1)(2l + 1)a30
3
e2 ℏ2 j(j + 1) − l(l + 1) − 34
(0)
ESO = (4.216)
8πϵ0 m2e c2 n3 l(l + 1)(2l + 1)a30
Using the fine structure constant.
e2
α= (4.217)
4πϵ0 ℏc
Equation (4.216) becomes;
ct
j(j + 1) − l(l + 1) − 34
(0) αℏ
ESO = (4.218)
2m2e c2 n3 l(l + 1)(2l + 1)a30
(0)
Expanding En in terms of α gives;
Le
e4 me
En(0) =− (4.219)
32π 2 ϵ20 n2 ℏ2
e2 e2
1 me
=− (4.220)
2 4πϵ0 ℏ 4πϵ0 ℏ n2
99
1 me
= − (αc) (αc) 2 (4.221)
2 n
α2 c2 me
En(0) = − (4.222)
2n2
tes
4πϵ0 ℏ2
From a0 = m e e2
,
1 4πϵ0 ℏ ℏ
a0 = ℏ= (4.223)
me e2 cαme
This implies that;
ℏ3
a30 = 3 3 3 (4.224)
c α me
Substituting equation (4.224) into equation (4.218) gives;
No
3
c3 α3 m3e j(j + 1) − l(l + 1) −
(0) αℏ 4
ESO = (4.225)
2m2e c2 ℏ3 n3 l(l + 1)(2l + 1)
2 α4 c4 m2e
En(0) = (4.227)
4n4
ure
2
(0)
4 En n4
α4 = (4.228)
c4 m2e
Substituting equation (4.228) into (4.226) gives;
2
(0)
2n En
j(j + 1) − l(l + 1) − 43
(0)
ESO = (4.229)
c3 ℏ2 me l(l + 1)(2l + 1)
ct
Le
100
4.5 The Zeeman Effect
The zeeman effect is the splitting of atomic energy levels and consequently the associated
spectral lines when atoms are placed in an external magnetic field. This effect provides
tes
a strong experimental confirmation of the quantization of angular momentum and also
tells why ml is called the magnetic quantum number.
Depending on the strength of the external magnetic field, the zeeman effect manifests its
self differently in terms of the number of spectral lines into which the original spectral
line splits. For weak magnetic fields(B ≈ 0.8T or less), it is divided into two classes;
the Normal Zeeman effect and Anomalous Zeeman effect.
No
In the normal zeeman effect, the splitting of spectral lines is determined by the orbital
angular momentum alone.i.e. zero spin, while in the Anomalous zeeman effect, it cannot
be explained in terms of orbital angular momentum alone., i.e. the spin angular momen-
tum is included.
The generic nature of the anomalous zeeman effect let to the discovery of the spin angular
momentum.
ure
4.5.1 Anomalous Zeeman Effect
The interaction of the external magnetic field B ⃗ ext with the electrons orbital and spin
magnetic dipole moments; µ⃗ L and µ
⃗ S give rise to two energy terms;
⃗ ext
VL = −⃗µL .B (4.230)
⃗ ext
VS = −⃗µS .B (4.231)
ct
⃗ ext − µ
ĤZ = −⃗µL .B ⃗ ext
⃗ S .B (4.232)
The magnetic dipole moment associated with the electron’s spin is µ ⃗ while
⃗ S = − mee c S
⃗ Therefore, equation
Le
101
⃗ ext to be along the z-direction such that B
Considering B ⃗ ext = Bext ẑ, equation (4.233)
becomes;
eBext
ĤZ = L̂z + 2Ŝz (4.234)
2me c
tes
The Hamiltonian of the system when placed in the external magnetic field becomes;
eBext
Ĥ = Ĥ0 + Ĥf s + L̂z + 2Ŝz (4.235)
2me c
The nature of the zeeman splitting depends critically on the strength of the external mag-
No
netic field in comparison with the internal field (4.198) that give rise to the spin-orbit
coupling.
⃗ ext << B
If B ⃗ , the fine structure dominates and Ĥz can be treated as a small perturbation.
⃗ ext >> B,
If B ⃗ the zeeman effect dominates and the fine structure becomes the perturba-
tion.
We take the ’unperturbed’ Hamiltonian to be (Ĥ + Ĥz ) and the perturbation to be Ĥf s .
Since Ĥ commutes with Ĥ0 (because Ĥ0 commutes with L̂z and Ŝz ), they can be diago-
nalized by a common set of states |nlml ms ⟩ such that;
eBext
Ĥ = Ĥ0 + L̂z + 2Ŝz (4.237)
2me c
102
Substituting equation (4.237) into (4.236) gives;
eBext
Ĥ0 + L̂z + 2Ŝz |nlml ms ⟩ = Enlml ms |nlml ms ⟩ (4.238)
2me c
tes
eBext
Ĥ0 |nlml ms ⟩ + L̂z + 2Ŝz |nlml ms ⟩ = Enlml ms |nlml ms ⟩ (4.239)
2me c
eBext ℏ
En(0) |nlml ms ⟩ + (ml + 2ms ) |nlml ms ⟩ = Enlml ms |nlml ms ⟩ (4.240)
2me c
(0) eBext ℏ
En + (ml + 2ms ) |nlml ms ⟩ = Enlml ms |nlml ms ⟩ (4.241)
2me c
This implies that;
No
eBext ℏ
Enlml ms = En(0) + (ml + 2ms ) (4.242)
2me c
13.6eV eBext ℏ
Enlml ms = − 2
+ (ml + 2ms ) (4.243)
n 2me c
13.6eV
Enlml ms = − + µB Bext (ml + 2ms ) (4.244)
n2
eℏ
where µB = 2m ec
is a constant known as Bohr Magneton. From (4.244) we see that the
(0)
energy levels En are shifted by the amount ∆E = µB Bext (ml + 2ms ). This is known as
the Paschen-Back shift,
ure
The states |nlml ms ⟩ are degenerate since the energy doesn’t depend on l.
For the strong field zeeman effect, the external magnetic field overpowers the the spin-
orbit interaction effect and decouples L ⃗ and S⃗ so that they press about Bext nearly
independently; thus, the projection of L ⃗ behave as if S
⃗ = 0 and the effect reduces to
three lines, each of which is a closely spaced doublet.
ct
Le
103
4.6 Chapter Exercises
1. From x = r sin θ cos ϕ, y = r sin θ sin ϕ and z = rcosθ. Find the expression for r, θ
and ϕ.
tes
2. Find the spherical coordinates of the point (x, y, z) = (−1, 1, 2) in the Cartesian
coordinate system.
3. Calculate the cartesian coordinates of point whose spherical coordinates are (r, θ, ϕ) =
√
( 6, 450 , 1350 )
4. After separating the schrodinger equation using ψ = R(r)f (θ)g(ϕ), the equation
No
for ϕ is
1 d2 g
= k2
g dϕ2
where k is a constant. Solve for g(ϕ) in this equation and apply the appropriate
boundary conditions. Show that k must be 0 or positive or negative (k = ml , the
magnetic quantum number).
7. Using the fact that the radial momentum operator is given by p̂r = −iℏ 1r ∂r ∂
r.
Calculate the commutator [r̂, p̂r ] between the position operator r̂ and the radial
ct
momentum operator.
8. Find the expected value of r and r2 in the ground state of the hydrogen atom.
9. Consider a hydrogen atom which is in its ground state; the ground stste wave
function is given by;
Le
1 − r
ψ(r, θ, ϕ) = p 3 e a0
πa0
where a0 is the Bohr radius
(a) Find the most probable distance between the electron and the proton when
the hydrogen atom is in its ground state.
104
(b) Find the average distance between the electron and the proton.
10. An electron in the coulomb field of a proton is in a state described by the wave
function;
tes
3
α 2 − α2 r 2
ψ(r) = √ e 2
π
What is the probability that it will be found in the ground state of the hydrogen
atom.
11. Show that the magnetic moment of an orbital electron is given by;
e ⃗
⃗ =−
µ L
No
2me
12. Calculate to first order perturbation theory the contributions due to the spin-orbit
interaction for an hygrogen-like ion having Z protons.
13. What is meant by Zeeman effect?. Discuss in details how it is expalined using the
quantum theory.
14. Distinguish between the normal and anomalous Zeeman effect. Discuss the theory
ure
of the anomalous Zeeman effect.
ct
Le
105
5 Introduction to Two Level Systems
tes
nite Dimensions
From the Schrodinger equation;
∂ |ψ⟩
iℏ = Ĥ |ψ⟩ (5.1)
∂t
Replacing the partial time derivatives by the total time derivatives in equation (5.1) gives;
No
d |ψ⟩
iℏ = Ĥ |ψ⟩ (5.2)
dt
This is true, since we are only interested in the matrix representation of this equation
where Ĥ is a matrix, not a differential operator.
Let |j⟩ be our base states such that multiplying equation (5.2) from the left with ⟨j|
gives;
d X
iℏ ⟨j|ψ⟩ = ⟨j| Ĥ |ψ⟩ = ⟨j| Ĥ |i⟩ ⟨i|ψ⟩ (5.3)
dt i
ure
Let cj = ⟨i|ψ⟩ such that equation (5.3) becomes;
d X
iℏ cj = Hji cj (5.4)
dt i
where Hji = ⟨j| Ĥ |i⟩ Equation (5.4) is a matrix equation with c a column matrix and
H a square matrix; Therefore, this is the finite matrix represenation of the Schrodinger
equation.
ct
d
iℏ c1 = H11 c1 + H12 c2 (5.5)
dt
d
iℏ c2 = H21 c1 + H22 c2 (5.6)
dt
106
which can also be written as;
! ! !
d c1 H11 H12 c1
iℏ = (5.7)
dt c2 H21 H22 c2
tes
The two state schrodinger equation help us to understand some important basic physics
regarding the ammonia molecule.
Suppose that the off diagonal matrix elements H12 = H21 = 0, equation (5.7) gets
decoupled and it becomes;
dc1
iℏ = H11 c1 (5.8)
dt
No
dc2
iℏ = H21 c2 (5.9)
dt
Integrating equations (5.8) and (5.9) gives;
iE1 t
c1 (t) = Ae− ℏ (5.10)
iE2 t
c1 (t) = Be− ℏ (5.11)
X
= |i⟩ ci
i
X
= ci |i⟩
i
The amplitudes in equation (5.14) corresponds to stationary states i.e, the probabilities
107
are independent of time.
In order to obtain the eigen states of the system, we have to diagonalise the matrix
elements of H.
tes
5.3 The Ammonia Molecule
In the ammonia molecule, N H3 , the hydrogen atoms are symmetrically located in a plane,
and the lone nitrogen atom can lie either above or below the symmetry plane of the H ′ s
as shown in Figure 5.1
No
ure
|1⟩=|N above the plane⟩ and |2⟩=|N below the plane⟩ and construct the matrix rep-
Le
resentation of the Hamiltonian using these two states depicted in Figure 5.1 as basis
states.
108
5.3.1 Assumptions
1. We shall assume that the states |1⟩ and |2⟩ are the base states needed to calculate
the ground states (lowest energy states of the N H3 molecule)
tes
2. There are more complex possibilities such as rotation, vibrations and electronic
motions but they all require considerably more energy and we shall ignore them for
simplicity.
By symmetry, the energy of the system in the two configurations must be the same. If
E0 is this energy, then;
H11 = H22 = E0 (5.15)
No
As such, the matrix elements H12 and H21 represent the interaction that mixes the two
base states. Physically, although the nitrogen is strongly repelled from the symmetry
plane of the hydrogen atoms due to the steep potential barrier, there is still a small prob-
ability that nitrogen tunnels through this symmetry plane.
For convenience, let us choose that interaction matrix element to be a negative number
say −A with A being a real number since H must be Hermitian. Therefore,
Thus, the energy matrix is parameterized in terms of two real quantities E0 and A.
From equation (5.7), ! ! !
d c1 E0 −A c1
iℏ = (5.18)
dt c2 −A E0 c2
Le
Therefore; !
E0 −A
H= (5.19)
−A E0
109
5.3.2 Diagonalizing the Energy Matrix H
To diagonalize the energy matrix (5.19), we need to first solve the secular equation;
tes
|H − λI| (5.20)
No
E0 − λ −A
=0 (5.22)
−A E0 − λ
λ = E0 ± A (5.25)
This gives the eigenvalues of H. Labeling the values with subscripts I and II gives;
ure
EI = E0 + A (5.26)
EII = E0 − A (5.27)
The Hamiltonian in equation (5.19) can be written in terms of the identity matrix I and
the pauli matrix σx as
H = E0 I − Aσx (5.28)
! !
ct
1 0 0 1
where I = and σx = The two terms in equation (5.28) commute.
0 1 1 0
Also, the first term is already a diagonalized matrix with constant entries E0 . We are
therefore required to diagonalize the second term and to do that, we should note that
this term has eigenvalues ±1 (eigenvalues of σx are ±1).
Le
Recall that; once diagonalized, the matrix has eigenvalues on its major diagonal. This
implies that equation (5.28) becomes;
H = E0 I − A(±)I (5.29)
110
H = (E0 ∓ A)I (5.30)
tes
H11 = E0 − A (5.31)
H22 = E0 + A (5.32)
The diagonalization procedure described above suggests that both H and the Pauli ma-
trix σx have the same diagonalizing matrix.
No
! !
+ √12 √1
2
⟨I|1⟩ ⟨I|2⟩
U= = (5.33)
√1 − 12
√ ⟨II|1⟩ ⟨II|2⟩
2
We can verify by direct multiplication that U given by equation (5.33) and H given by
equation (5.9), it is indeed the case that;
!
E0 − A 0
ure
U HU −1 = U HU † = (5.34)
0 E0 + A
Let us now write the new kets; |I⟩ and |II⟩ in terms of the old ones; |1⟩ and |2⟩. With
the use of the matrix elements of U . E.g;
X
|I⟩ = |i⟩ ⟨i|I⟩ (5.35)
i
X
|II⟩ = |i⟩ ⟨i|II⟩ (5.37)
i
111
1
|II⟩ = √ (|1⟩ − |2⟩) (5.38)
2
We can easily verify that the states |I⟩ and |II⟩ are an orthonormal set; they are our new
base states.
tes
The existence of tunneling between the states |1⟩ and |2⟩ splits the energy states of the
molecule into two states with different energies, one with energy E0 − A and the other
with energy E0 + A (see Figure 5.2).
No
Figure 5.2: The two energy levels of the ammonia molecule
Let us now proceed to find out how the new base states |I⟩ and |II⟩ relate to the sta-
tionary states of the problem.
ure
The amplitudes of a general state ψ which are also the probability of the system to be
found in the new base states are;
cI = ⟨I|ψ⟩
(5.39)
cII = ⟨II|ψ⟩
iℏ = (5.40)
dt cII 0 E0 + A cII
−i(E0 +A)t
cII = De ℏ (5.42)
Where C and D are constants of integration which are essentially normalization constants
of the eventual wave function.
112
It follows that the general solution |ψ⟩ can be expressed as a linear combination of these
two functions.
−i(E0 −A)t −i(E0 +A)t
|ψ⟩ = Ce ℏ |I⟩ + De ℏ |II⟩ (5.43)
tes
|ψ⟩ = C |ψI ⟩ + D |ψII ⟩ (5.44)
Where |ψI ⟩ and |ψII ⟩ are the stationary states of the system. We therefore conclude that
both bases yield the same form of solutions.
We conclude that because there is some chance that the nitrogen atom can flip from one
position to another, the energy of the molecule is not just E0 as we would expect but
there are two energy levels (E0 − A) and (E0 + A). Every one of the possible states of
No
the molecule splits into two levels, whatever energy it has. The splitting of the energy
levels of the ammonia molecule is strictly a quantum effect.
For the maser to operate, there should be away of getting ammonia molecules predom-
inantly populating the excited state. The conditions should also allow for their de-
excitation and eventually donating radiation to a cavity so that microwave energy can be
ct
built in it.
The first step involves the interaction of an ammonia molecule with a static electric field
and the second, its interaction with a time dependent electric field.
Le
113
5.4.1 Ammonia Molecule in an External Static Electric Field
In the ammonia molecule, the valence electrons have a slight preference for the nitrogen
(N) atoms over the hydrogen atoms (H). As a result, the valence electrons stay on average
tes
a little closer to the nitrogen atom. This causes an imbalance in the charge distribution
in the ammonia molecule leading to the creation of an electric dipole moment.
The created electric dipole moment is directed from the nitrogen atom towards the hy-
drogen atom plane.
When there is an external electric field, there is an additional contribution to the Hamil-
No
tonian resulting from the interaction of the dipole with the electric field.
It is important to note that when the nitrogen flips over to the other side, i.e. from state
|1⟩ to state |2⟩, the electric dipole moment also flips over (Figure 5.3)
ct ure
Figure 5.3: The two base states of the ammonia molecule give rise to opposite interaction with
the static filed. These states have an electric dipole moment µ
⃗
Consequently, the dipole interaction adds with a different sign in the two states;
Le
114
Since there is only a diagonal contribution to the energy matrix from the dipole interaction
term, we have the Hamiltonian in the |1⟩ − |2⟩ basis as;
! !
⟨1| Ĥ |1⟩ ⟨1| Ĥ |2⟩ E0 + µε −A
tes
Ĥ = = (5.47)
⟨2| Ĥ |2⟩ ⟨2| Ĥ |2⟩ −A E0 − µε
!
E0 + µε −A
Ĥ = (5.48)
−A E0 − µε
We assume that the external electric field is sufficiently weak that is doesn’t affect the
amplitude for the nitrogen atom to tunnel through the barrier.
No
The two states |1⟩ and |2⟩ are not stationary; there are non zero off-diagonal matrix
elements of Ĥ in this basis. The eigenvalues of (5.48) are determined by solving the
secular equation;
ˆ =0
|Ĥ − λI| (5.49)
! !
E0 + µε −A 1 0
−λ =0 (5.50)
−A E0 − µε 0 1
E0 + µε − λ −A
=0 (5.51)
−A E0 − µε − λ
ure
(E0 + µϵ − λ)(E0 − µε − λ) − A2 = 0 (5.52)
λ2 − 2E0 λ + E02 − µ2 ε2 − A2 = 0
(5.53)
Equation (5.53) can be solved using the Buldoza method since it is in the form ax2 +
bx + c = 0. From which; √
−b ± b2 − 4ac
x= (5.54)
2
Using equation (5.54),
ct
p
2E0 ± 4E02 − 4 (E02 − µ2 ε2 − A2 )
λ= (5.55)
2
p
2E0 ± 2 µ2 ε2 + A2
= (5.56)
Le
2
1
λ1 = E0 + µ2 ε2 + A2 2 (5.57)
1
λ2 = E0 − µ2 ε2 + A2 2 (5.58)
Let us call the stationary state |α⟩ and |β⟩; their energy eigenvalues Eα and Eβ are given
115
from the solutions above;
21
Eα = E0 + µ2 ε2 + A2
12 (5.59)
Eβ = E0 − µ2 ε2 + A2
tes
These two energies are plotted as a function of the electric strength ε in Figure ( 5.4)
No
ure
Figure 5.4: Energy Levels of Ammonia Molecule in an Electric Field
The curves in Figure 5.4 are hyperbolic; for large fields, the splitting becomes propor-
tional to ε as the curves approach their asymptotes; the splitting doesn’t depend on A
ct
For small fields, the effect of tunneling interaction, A dominates and their energies vary
with the electric field via a term ε2
Le
116
5.4.2 Effect of the Electric Field
Suppose a beam of ammonia molecules is passed through slits into a region of a strong
transverse electric filed as shown in Figure 5.5
tes
No
Figure 5.5: Separation of the two tates of an Ammonia beam by an Electric Field with a gradient
in a direction perpendicular to the beam
The electrodes to produce the field are shaped so that the electric field varies rapidly
across the beam. Then, the square of the electric field ε2 will have a large gradient per-
pendicular to the beam.
ure
Now, a molecule in state |α⟩ has an energy which increases with ε2 and therefore this part
of the beam will be deflected towards the region of lower ε2 in order to minimize its energy.
A molecule in a state |β⟩ will on the other hand be deflected towards the region of larger
ε2 since its energy decreases as ε2 increases. Therefore, molecules in the different states
will be deflected in opposite directions and can be separated.
The energy µϵ is always smaller than A. Therefore, the square root term in equation
ct
µ 2 ε2
Eα = E0 + A +
2A (5.61)
µ 2 ε2
Eβ = E0 − A +
2A
117
These energies vary approximately linearly with ε2 . The force on the molecule is then;
µ2 ⃗
F⃗ = ∇ε (5.62)
2A
tes
Many molecules have energy in the electric filed which is proportional to ε2 . The coef-
ficient is the polarizability of the molecule. Ammonia has unusually high polarizability
because of the small value of A in the denominator. Thus, Ammonia molecules are
unusually sensitive to an electric field.
No
In the Ammonia maser, the beam with molecules in the state |α⟩ is sent through a reso-
nant cavity with a time-varying electric field. ε = ε0 cos(ωt). See figure 5.6
ure
Figure 5.6: In the ammonia maser, molecules in a higher state |α⟩ get de-excited to the lower
state |β⟩, building up the electromagnetic energy inside a resonant cavity
By interacting with the time-varying electric field, the molecules surrender their excita-
ct
tion energy (Eα − Eβ ) to the cavity as they de-excite to the state |β⟩.
Energy builds up in the electromagnetic filed of the maser cavity due to resonant absorp-
tion or emission of electromagnetic energy. Molecular energy is converted into the energy
Le
Note; Before the beam enters the cavity, we have to use filters which separate the beam
so that only the upper states enters.
118
5.5 Magnetic Resonance
Consider the motion of a particle with just a spin degree of freedom in a magnetic field.
For example;
tes
• Protons placed in a magnetic field as in the phenomenon of Nuclear Magnetic
Resonance (NMR).
In both examples, the particles interact with the external magnetic field by virtue of their
intrinsic magnetic moment arising from spin;
No
⃗ = qg S
M ⃗ (5.63)
2mc
where q is the charge and g is the gyromagnetic ratio of the particle. For a proton, q = e
while for an electron, q = −e.
In the absence of the magnetic field, the spin up and spin down states of the proton are
degenerate. The field removes this degeneracy.
If B0 is the magnitude of the field along the z-axis, equation (5.64) becomes;
ct
ge geℏ
Ĥz = − Ŝz B0 = − B0 σ̂z = −µp B0 σ̂z (5.65)
2mp c 4mp c
geℏ
where mp is the proton mass, µp = 4mp c
and σ̂z is a Pauli matrix.
From equation (5.65), we can obtain the eigenvalues of Ĥz for the corresponding eigen-
Le
states;
Ĥz |+z⟩ = −µp B0 σ̂z |+z⟩ = −µp B0 |+z⟩ (5.66)
The eigenvalues are −µp B0 and +µp B0 respectively. The states |+z⟩ and |−z⟩ are sta-
119
tionary states of the Hamiltonian.
From equations (5.66) and (5.67), there is a Zeeman splitting of the energy level.
tes
Figure 5.7: Energy diagram with zeeman splitting; The lower energy level; −µp B0 (state |+z⟩).
No
The upper energy level; +µp B0 (state |−z⟩).
For Nuclear Magnetic Resonance, the field in the z-direction is static, but an additional
time varying field is applied along the x-direction (see Figure 5.8 )
Bx = B1 cos ωt (5.71)
ct
solenoid can be oriented in such a fashion that the RF magnetic field is along the x-axis.
The Hamiltonian of interaction of the protons with the radio-frequency magnetic field is
given by;
Ĥx = −µp Bx σ̂x = −µp B1 cos ωtσ̂x (5.72)
120
tes
Figure 5.8: Experimental arrangement of NMR. There is a time-dependent magnetic field in
No
addition to a constant field in the z-direction and perpendicular to it.
Ĥ = Ĥz + Ĥx
(5.73)
= −µp B0 σ̂z − µp Bx σ̂x
ℏ ω0 B1
= − ω0 σ̂z + cos(ωt)σ̂x (5.78)
2 B0
ℏ
Ĥ = − [ω0 σ̂z + ω1 cos(ωt)σ̂x ] (5.79)
2
ω0 B1 geB0 geB1
where ω1 = B0
= 2m pc
. B 1
B0
= 2m pc
.
121
The Hamiltonian is time dependent, this implies that the application of an oscillating
magnetic field transverse to the z-axis, can induce transitions between the states |+z⟩
and |−z⟩. This can be done properly by adjusting the frequency of the transverse mag-
tes
netic field (B1 ).
To determine how these states evolve in time, we return to the Schrodinger equation
(5.2). We are going to work in the z-basis and express |ψ(t)⟩ in this basis by;
!
a(t)
|ψ(t)⟩ = a(t) |+z⟩ + b(t) |−z⟩ = (5.80)
b(t)
No
!
1
With the initial state; |ψ(0)⟩ = Using the Schrodinger equation;
0
d |ψ(t)⟩
iℏ = Ĥ |ψ(t)⟩ (5.81)
dt
da(t) db(t)
Where ȧ(t) = dt
and ḃ(t) = dt
.
This coupled set of first order differential equations can not be solved exactly.
In practice, however, the transverse field B1 is significantly weaker than the field B0 in
the z-direction and therefore, the frequency ω1 is considerably smaller than ω0 . Using
ct
(5.83)
Le
iω0
t
b(t) = b(0)e 2
122
Where we expect that we have included the major part of the time dependence in the
exponentials.
tes
! !
ċ(t) ω1 c(t)eiω0 t
i = cos ωt
˙
d(t) 2 d(t)e−iω0 t
! ! (5.85)
ei(ω0 +ω)t + ei(ω0 −ω)t d(t)
ċ(t) ω1
i =
˙ ei(ω−ω0 )t + e−i(ω0 +ω)t c(t)
d(t) 4
Unless ω is chosen to be very near to ω0 , both exponentials in the second line of (5.85) are
No
rapidly oscillating functions that when multiplied by a more slowly oscillating function
such as c(t) and d(t), whose time scale is set by ω1 will cause the right-hand side of (5.85)
to average to zero.
However, if ω is near ω0 , the terms oscillating at (ω0 + ω) can be neglected with respect to
those oscillating at (ω0 − ω), and these latter terms are now oscillating sufficiently slowly
that c(t) and d(t) vary with time.
For our case, we shall consider and solve the time dependence when ω = ω0 ; the resonant
ure
condition.
Setting ω = ω0 and neglecting the exponentials oscillating at 2ω0 , equation (5.85) becomes
! !
ċ(t) ω1 d(t)
i = (5.86)
˙
d(t) 4 c(t)
ct
iω1
ċ(t) = − d(t)
4 (5.87)
˙ = − iω1 c(t)
d(t)
4
Le
iω1 ˙
c̈(t) = −d(t)
4 (5.88)
¨ = − iω1 ċ(t)
d(t)
4
123
Using equation (5.87), equation (5.88) becomes;
ω 2
1
c̈(t) = −c(t)
4 (5.89)
¨ = − ω1 d(t)
2
tes
d(t)
4
No
4i ω1 ω
1 ω1 ω
1
d(t) = − A1 sin t + A2 cos t (5.91)
ω1 4 4 4 4
ω ω
1 1
d(t) = i −A1 sin t + A2 cos t (5.92)
4 4
Applying the boundary conditions; c(0) = 1 and d(0) = 1 gives A1 = 1 and A2 = 0.
Therefore, we have
ω
1
c(t) = cos t
4
ω (5.93)
1
d(t) = −i sin t
ure
4
Using the above equations, the solutions to a(t) and b(t) become;
! ω0 !
a(t) cos ω41 t ei 2 t
= ω0 (5.94)
b(t) −i sin ω41 t ei 2 t
Thus, the probability of finding the particle in the state |−z⟩ at a time t is given by;
(5.95)
ω1
P− (t) = sin2 ( t) (5.96)
4
Thus, the probability of finding the particle in the state |+z⟩ at a time t is given by;
Le
ω1
P+ (t) = cos2 (t) (5.98)
4
Note that these probabilities add up to one since these two states form a complete set
and the probability is conserved in time.
124
If a particle is initially in a state |−z⟩ and makes a transition to a state |−z⟩, the energy
of the spin system is reduced by E+ − E− = ℏω0 . Assuming ω0 > 0, this energy is added
to the electromagnetic energy of the oscillating field that is stimulating the transition.
tes
For 0 < t < ω2π1 , the probability of making a transition to the lower energy state grows
until b∗ (t)b(t) = 1 and a∗ (t)a(t) = 0. Then the particle is in the state |−z⟩.
For ω2π1 < t < ω4π1 , the probability of being in the lower energy state decreases and that
of being in the higher energy state grows as the system absorbs energy back from the
electromagnetic field. This cycle of emission and absorption continues indefinitely (see
No
Figure 5.9)
ure
Figure 5.9: The probabilities |⟨+z|ψ(t)⟩|2 (solid line) and |⟨−z|ψ(t)⟩|2 (dashed line) for a spin-
half particle that is in a state |+z⟩ at t = 0 when the time dependent magnetic field in the
x-direction is tuned to be resonant frequency.
ct
Le
125
5.6 Chapter Exercises
1. Write the two-state Hamiltonian matrix in a certain basis |1⟩, |2⟩ in a general form
as;
tes
!
H11 H12
H21 H22
Impose hermiticity of H
(a) Find the eigenstates and the unitary transformation that diagonalizes the
Hamiltonian
(b) Express the eigenstates in terms of the old base states
No
2. Find the eigenvalues and eigen vectors of aσ̂z + bσ̂x + cIˆ where a, b, and c are
constants. σ̂z , σ̂x are Pauli matrices and Iˆ is a 2 × 2 matrix. Hence find the
eigenstates |α⟩ and |β⟩ of the ammonia molecule in a state electric field ⃗ε.
126
References
Goswami, R. (1997). Quantum Mechanics, (2nd Edition), Waveland Press, Inc. Long-
Grove, Illinios, USA.
tes
Griffiths, D., J., and Schroeter, D., F. (2018). Introduction to Quantum Mechanics,
(3rd Edition), Cambridge University Press, London, UK.
No
Zettilli, N. (2009). Quantum Mechanics; Concepts and Applications, (2nd Edition), John
Willey and Sons, West Sussex, UK.
ct ure
Le
127