Lecture Notes in Quantum Mechanics: June 2017
Lecture Notes in Quantum Mechanics: June 2017
Lecture Notes in Quantum Mechanics: June 2017
net/publication/318899706
CITATIONS READS
0 36,582
1 author:
Salwa Alsaleh
King Saud University
24 PUBLICATIONS 85 CITATIONS
SEE PROFILE
Some of the authors of this publication are also working on these related projects:
All content following this page was uploaded by Salwa Alsaleh on 04 August 2017.
Salwa Alsaleh
2 Mathematical preliminaries 5
2.1 Abstract vector spaces . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.2 Functions as vectors . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.3 Dual spaces and inner product . . . . . . . . . . . . . . . . . . . . . 7
2.3.1 The Bra-Ket notation . . . . . . . . . . . . . . . . . . . . . 7
2.3.2 Normed spaces . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.4 Hilbert space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.4.1 Basis of a Hilbert space . . . . . . . . . . . . . . . . . . . . 8
2.5 The space L2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.6 Outer product . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.7 Linear operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.7.1 The algebra of operators . . . . . . . . . . . . . . . . . . . . 11
2.8 Spectral theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.9 Projection operators . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.9.1 The identity operator . . . . . . . . . . . . . . . . . . . . . . 13
2.9.2 Projection operators . . . . . . . . . . . . . . . . . . . . . . 13
2.10 Unitary operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.11 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.11.1 Rotation in the Euclidean 2-D space . . . . . . . . . . . . . 14
2.11.2 The differential operator . . . . . . . . . . . . . . . . . . . . 14
2.12 Commutators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.13 Function of operator . . . . . . . . . . . . . . . . . . . . . . . . . . 16
iii
iv CONTENTS
4 The Wavefunction 25
4.1 Position representation . . . . . . . . . . . . . . . . . . . . . . . . 25
4.2 Separation of variables in Schrödinger’s equation . . . . . . . . . . . 26
4.3 The free-particle solution . . . . . . . . . . . . . . . . . . . . . . . . 27
4.4 Probability flux and density . . . . . . . . . . . . . . . . . . . . . . 28
4.5 The Born conditions . . . . . . . . . . . . . . . . . . . . . . . . . . 30
4.6 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
7 Stationary States 45
7.1 Particle in a box . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
7.1.1 Solution to Schrödinger’s equation . . . . . . . . . . . . . . . 46
7.1.2 Momentum eigenfucntions . . . . . . . . . . . . . . . . . . . 48
7.2 Simple Harmonic Oscillator . . . . . . . . . . . . . . . . . . . . . . 49
7.2.1 Quantization of the SHO Hamiltonian . . . . . . . . . . . . 49
CONTENTS v
8 Angular Momentum 55
8.1 The Classical angular momentum . . . . . . . . . . . . . . . . . . . 55
8.2 Quantisation of the angular momentum . . . . . . . . . . . . . . . 56
8.3 The spherical harmonics . . . . . . . . . . . . . . . . . . . . . . . . 57
8.4 Properties of the spherical harmonics . . . . . . . . . . . . . . . . . 57
8.5 Angular momentum eigenstates . . . . . . . . . . . . . . . . . . . . 59
8.6 The spectrum of angular momentum observable . . . . . . . . . . . 60
8.7 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
10 Spin 69
10.1 Transformation of the wavefunction . . . . . . . . . . . . . . . . . . 69
10.2 External and internal degrees of freedom for a system . . . . . . . 70
10.3 Mathematical description for internal degrees of freedom . . . . . . 71
10.4 Discovery of electron’s magnetic dipole moment . . . . . . . . . . . 72
10.5 Electron’s spin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
10.6 Infel-van der Warden Symbols . . . . . . . . . . . . . . . . . . . . . 76
10.7 Matrix representation of spin states . . . . . . . . . . . . . . . . . . 77
10.8 Geometric representation . . . . . . . . . . . . . . . . . . . . . . . . 79
10.9 Spin in constant magnetic field . . . . . . . . . . . . . . . . . . . . 80
10.9.1 Stationary states . . . . . . . . . . . . . . . . . . . . . . . . 81
10.10Electron paramagnetic resonance EPR . . . . . . . . . . . . . . . . 81
10.11Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
1.1.1 Examples
1. The Hamiltonian for a free particle is given by:
p2
H= (1.3)
2m
2. In a central potential, the Hamiltonian takes the form :
p2 Q1 Q2
H= +g (1.4)
2m r
Where g is a constant, called coupling constant, and Q1 Q2 are charges/
masses. Depending on the type of interaction.
1
recall that H = T + V
2
Note that p~ here is not always m~v !
3
recall that L = T − V
1
2 CHAPTER 1. REVIEW OF CLASSICAL MECHANICS
H=µ ~
~ ·B (1.5)
4
4. The Hamiltonian for a free rotating mass is:
L2
H= (1.6)
2I
Here, L is the angular momentum of the system.
P2 mω 2 q 2
H= + . (1.7)
2m 2
∂H
= −ṗi (1.8)
∂q i
∂H
= q̇ i (1.9)
∂pi
Some dynamical systems however, does not obey these equations. These systems
are known as Hamiltonian constraint systems.
1.4 Problems
1. Show that the Hamiltonian for a particle of mass m, orbiting a mass M , and
they interact gravitationally is gien by:
L2 Mm
H= +G
2I r
6
One needs as many of them as the number of degrees of freedom f for the system
4 CHAPTER 1. REVIEW OF CLASSICAL MECHANICS
Figure 1.1: (a) The phase space for free SHO. (B) The phase space of a damped
SHO
Then derive the equations of motion for this system, comment on your re-
sults.
2. Derive and solve the equation of motion for a 2-D SHO, with m = 1 and
ω = 1.
3. Show that :
∂pj
{pj , H} =
∂t
4. Draw the shape of the phase space for a particle free-falling from altitude y0
Mathematical preliminaries
is also an element of that vector space, where α and β are complex or real numbers.
The expression (2.1) is called the superposition of the vectors |ψi and |φi.
The dimension of V could either be finite , countably infinite or uncountably infinite
( see next section). For finite dimensional- or countably infinite- vector spaces. It
is possible to represent a vector |ψi as a column matrix :
f1
f2
|f i ⇔ f (2.2)
3
..
.
5
6 CHAPTER 2. MATHEMATICAL PRELIMINARIES
Figure 2.1: Spike digram of a vector in 3 dimensional vector space, and 30 dimen-
sional one.
we use the spike diagrams discussed above.This time using a continuous parameter
x taking real-number values instead of the discrete index i. The spike digram
for such vector would look like: In fact, this spike digram looks familiar to the
single vector in an infinite dimensional space. It should be noted that to make the
transition to infinite dimensions mathematically meaningful, you need to impose
some smoothness constraints on the function. Typically, it is required that the
function is continuous, or at least integrable in some sense. These details are not
important for our purpose, thus we shall not discuss them further.
2.3. DUAL SPACES AND INNER PRODUCT 7
We can easily show that the set of such maps form a vector space themselves, sometimes they
which we call the dual vector space and denoted by V ∗ . Moreover, the operation are called linear
between a vector and a (dual) vector is called inner product . functionals
Note that the norm is always a real number . Vectors with a unit norm is called
normal vectors .
Instead of using the Bra-Ket notation for functions, we shall only denote them by
φ(x).
Note that we need some-sort of structure in the Hilbert space to insure that (2.7)
converges if the sum is infinite. We call the Hilbert space in with the series of
this type the `2 space, or the space of square-summable sequences. Moreover, the
integrals used in this course - and in quantum mechanics in general- are known
as the Lebesgue integrals , they are different from Riemann integrals defined in
calculus courses.
Here, we have illustrated the most relevant properties of Hilbert spaces that con-
cern us in the study of introductory quantum mechanics. A lot of mathematical
details and rigour has been spared in this lecture. It is urged from the reader to
conduct a further reading in the theory of Hilbert space; please consult: Functional
Analysis by M. Reed and B. Simon.
For the norm to exist, the function φ(x) needs to be square integrable on the
interval [a, b]. It is not hard to show that the set of square-integrable functions
on the same interval form a Hilbert space. This Hilbert space is known as the
L2 (R; dµ) space. It reads; the space of square-integrable function on the inter-
val/Region R 5 , with respect to the measure dµ. By measure we mean the volume
element that we integrate over. In 1-D case dµ = dx. Sometimes, one wishes to
define a weight w(x) for the space, but this is out of the scope of this course.
The space L2 can have basis of orthogonal, and normalised functions (u1 (x), u2 (x), . . . )
depending on the interval of interest. For example the classical orthogonal poly-
nomials including:
• Hermite polynomials:
2 dn −x2
Hn (x) = Kn−1 ex e . (2.11)
dxn
5
Functions of L2 could be of several variables, real or complex-valued
10 CHAPTER 2. MATHEMATICAL PRELIMINARIES
They could form basis for the space L2 (−∞, +∞; dx).
• Legendre polynomials:
dn
Lνn = Kn−1 x− νex xν+1 ex .
n
(2.12)
dx
2
They could form basis for the space L (0, +∞; dx).
• Legendre polynomials ( of the first kind):
dn
Pn = Kn−1 1 − x2
n
(2.13)
dx
They could form basis for the space L2 (0, 1; dx).
The second example ofL2 spaces, are the ones used in Fourier analysis. Where
cos(nkx) and sin(nkx) form an orthonormal basis for a given L2 space with an
interval L. Recall that we can expand any function in a Fourier series :
∞
X nπx nπx
f (x) = An cos( ) + Bn sin( ) (2.14)
n=0
L L
Where :
Z L
nπx
An = f (x) cos(
)dx (2.15)
0 L
Z L
nπx
Bn = f (x) sin( )dx (2.16)
0 L
Or expanding the function in a continuous basis ( Fourier integral).
Note Advanced readers might not find the discussion in this lecture formal nei-
ther accurate enough, as discussing mathematical rigour of Hilbert spaces is very
distant from the course aims.
Moreover:
Â(α|ψi + β|φi) = α(Â|ψi) + β(Â|φi) (2.18)
A result from above : X
Â|ψi = hi|ψi Â|ii (2.19)
i
2. Eigenvalue:
A scalar λ is called an eigenvalue if it satisfied the equation:
Called the eigenvalue equation, and the vector |ψi is called an eigenket/
eigenvector. This equation is equivalent to :
det  − λIˆ = 0 (2.21)
For Iˆ or just I being the identity operator. An important result from this
property is the spectral decomposition, that we shall discuss later in this
lecture.
The hermitian conjugate acts on the dual space of H ( acts on the Bras).
The following properties for the hermitian conjugation are listed below ( for
reminding)
†
• † =  – involutiveness
• Antilinearity:
†
α + β B̂ = α∗ † + β ∗ B̂ †
• (ÂB̂)† = B̂ † † .
† = Â
In other words, it acts both on the Kets and on the Bras. An important
theorem for self-adjoint operators is stated below:
All the eigenvalues for a self-adjoint operator are real
Other properties of the matrices can be revised from a linear algebra book.
Note that the outer product |ui ihui | is gives the identity matrix / operator I. 7
Hence, we may diagonalise the operator  if we have found all of its eigenvalues.
7
If two eigenkets have different eigenvalues they ought to be orthogonal
2.9. PROJECTION OPERATORS 13
where, dµ(λ) is the integration measure that depends on the nature of spectrum
for the measurement outcomes in the theory. We are not going to go further in
the details, as they are beyond the scope of our course.
must be the identity operator, which sends each vector to itself. This can be
inserted in any expression without affecting its value, for example
X X X
hv|wi = hv| |ei ihei |wi = hv| |ei ihei | |ej ihej |wi = hv|ei ihei |ej ihej |wi
i∈N i∈N j∈N
(2.26)
2.11 Examples
2.11.1 Rotation in the Euclidean 2-D space
A vector in the 2-D plane is represented by :
x
|ri = (2.30)
y
We can define a rotation operator R̂(ϑ), that acts on the vector |ri by rotating it
with an angle ϑ. This operator has a matrix representation:
cos ϑ − sin ϑ
R̂ = (2.32)
sin ϑ cos ϑ
Figure 2.4: The rotation in 2-D space, carried put by the rotation operator.
The most famous operators which act on function space are the differential oper-
ators; denoted by L̂. There are a variety of differential operators. They play an
important rôle in the theory of differential equations. In fact, most of the prob-
lems in quantum mechanics are related to the analysis of the differential operators
related to dynamical observables; as we shall see.
Take the function f (x) = eλx . It is the eigenfunction of the differential opera-
d
tor , with an eigenvalue λ. Hence, we conclude that eλx , is a solution to the
dx
differential equation:
d
(f (x) = λf (x) (2.33)
dx
d2
The operator 2
has two eigenfunctions e+λx and e−λx they resemble solutions
dx
for the differential equation :
d2
(f (x) = λf (x) (2.34)
dx2
And by the superposition principle, a general solution would be:
2.12 Commutators
Just like ordinary matrix multiplication, the product between operators is generally
non-commutative. In fact, this particular property of operator multiplication is
behind the unfamiliar phenomena observed in quantum mechanics, thus generally
we have :
ÂB̂ 6= B̂ Â (2.36)
We define the commutator between two operators as:
Provided that
[[B̂, Â], Â] = 0
Another important formula to learn is the Hadamard Lemma:
1
e B̂e− = B̂ + [Â, B̂] + [Â, [Â, B̂]] + . . . (2.40)
2!
The eigenkets form a mutual eigenbasis for the states Â|ψi and B̂|ψi. Hence one
can simultaneously diagonalise both operators.
2.15 Problems
1. Find the eigenvalues of the operator :
1 i 0
 = −i 2 −i
0 i 1
Can we diagonalise it ?.
2.15. PROBLEMS 17
2. Is the operator d/dx acting on the L2 Hilbert space hermitian ? How about
−id/dx ?
4. Is the function cos kx an eigenfunction for the operator d/dx ? What about
the operator d2 /dx2 ?
eiĤt Âe−iĤt
7. Recall that we defined the identity operator as the outer product of the basis
: X
I= |iihi|
i
Using this definition, show that this operator sends the ket |ψi to itself (
does not change the ket), then show this for the Bra vector hψ|.
8. Discuss why if [Â, B̂]|ψi 6= 0 one cannot find a mutual eigenbasis to expand
|ψi with ?
18 CHAPTER 2. MATHEMATICAL PRELIMINARIES
Chapter 3
3.1 Introduction
In modern physics course and in modern physics lab, we have seen the clear motiva-
tion for using quantum mechanics in place of classical physics for the fundamental
description of nature. Classical physics is merely an approximation for the real
picture of the quantum world .
One of most important lessons learnt from quantum mechanics is the fact that mea-
surement affects the system, no matter how hard ones tries to avoid such effect, it
shall remain present ( there are exceptions known as weak measurements). That
implies that the order of measuring dynamical quantities of the system matters,
hence non-commutativity is the heart and soul of quantum mechanics. Max
Born tried to make an interpretation for these observable facts. His interpreta-
tion states that the quantum system is non-deterministic and each outcome of the
measurement has its own probability. Moreover, the system before measurement
takes a superposition of all of its possible states. The experimental motivation
for these statements are assumed to be known by the reader.
19
20 CHAPTER 3. POSTULATES OF QUANTUM MECHANICS
M −→ H
The state of the system becomes a ‘Ket’|ψi in H instead of a vector in the phase
space as we have seen in lecture (1). This abstract vector is unlike the state vector
for the classical system has no direct physical meaning . As one can multiply this
vector by any complex/real number and get the same state for physical system
. 1 . Hence if |ψi describes the system , then a|ψ for a a complex/real number
describes the same state for the physics system, This however has rare exceptions
( known as Berry phase). Therefore, one can normalise the state vector. such that
it satisfies:
hψ|ψi = 1 (3.1)
X
⇔ P (i) = 1 (3.2)
i
Surely the sum of all the probabilities of the possible measurable quantities for the
system is ought to equal one. This postulates implies what Max Born suggested for
the statistical nature of quantum mechanics, or more generally the superposition
principle ( the quantum system takes all of its possible configurations when not
Recall measured ). .
Schrödinger’s
cat!
3.2.2 The second postulate
Any dynamical observable for a classical system ω(p, q) is defined to be a func-
tion on the phase space. Upon quantisation, these observables will be resembled
by a linear self-adjoint (hermitian) operators acting on the Hilbert space
Ω̂. Measurement is expressed mathematically by acting the operator correspond-
ing to the physical observable on the state vector. The possible values (outcomes)
for a measurement is the set of eigenvalues for that operator. This a direct physical
result from the spectral theorem.
1
In fact, the physical state is described by a ray in H, not a one vector.
3.2. THE POSTULATES OF QUANTUM MECHANICS 21
We have learnt that if two operators do not commute, then one cannot have a
mutual set of complete eigenbasis to simultaneously diagonalise them . This phys-
ically means that one cannot measure both observables with absolute accuracy.
Leading to the uncertainty principal.
We have used the word ’measurement’ a lot in this lecture, it may be confusing for
the reader what is really meant by it within the context of quantum mechanics.
One may picture a ’physicist’ in a lab who intend to measure a quantum system
when thinking of the word measurement. However, measurement may not involve
any experiment of actual detectors, rather simply it is any interaction between the
quantum system and a classical object, which we call the apparatus.
In pure quantum mechanical view, there is no meaning for a path of a particle, as
which is -in fact- schrödinger’s equation. We can write the Hamiltonian in matrix
form :
E1 0 0 ... 0 ...
0 E2 0 . . . 0 . . .
0 0 E3 0 0 ...
0 0 · · · En 0 . . .
Ĥ = (3.10)
.. .. .. .. . .
. . . . . ...
0 0 0 0 ...
.. .. .. .. .. . .
. . . . . .
If the measurement resulted the particle having the energy state Ej . The state
vector |ψi is then projected into the eigenstate |Ej i casing of what-so-called the
wavefunction collapse :
hEj |ψi = αj (3.11)
We may also calculate the expected-value for the energy:
3.5 Problems
1. Wilson’s chamber is a sealed environment containing a supersaturated vapour,
when a charged particle - like an electron or alpha particle- passes through
the chamber. It leaves a track of cloud behind it. Discuss why we can ’see’
24 CHAPTER 3. POSTULATES OF QUANTUM MECHANICS
Figure 3.2: Alpha particles from a Radium source in a cloud chamber, notice the
path of the particle
the quantum particle’s path in this case although we have stated there is no
path defined for quantum particles ?
2. If we let the position operator be a multiplicative one; i.e Q̂i = q i ; show
that -in order to satisfy the commutation relation discussed in the lecture
∂
the momentum operator needs to be P̂j = ~i j .
∂q
∂
3. From the previous problem, why we can’t use ~ . As a definition for the
∂qj
momentum operator, and −iq i as the position operator ?
4. Which is ‘bigger’ the phase space or the Hilbert space ? Provide a supporting
argument for your answer .
5. In a thought experiment , imagine having a quantum coin. A coin which
obeys the laws of quantum mechanics. Describe it mathematically
6. An electron can take 3 possible energy states E1 = 0.5eV , E2 = 1.2eV and
E3 = 1.6eV . With probabilitie: P1 = 0.8, P2 = 0.13 and P3 = 0.07.
(a) Write the Hamiltonian in matrix form.
(b) Write the normalised eigenbasis
(c) Find hEi and σ(E).
(d) What is the state ket after measuring the system and finding it taking
the second energy state ?
(e) Show that hψ|ψi = 1
3
Remember that |αn |2 = P (En )
Chapter 4
The Wavefunction
∂
Ĥ|ψ(t)i = i~ |ψ(t)i (4.1)
∂t
What we are interested in knowing for a free particle is its position, we therefore
project the state ket into the position space , and get the wavefunction:
25
26 CHAPTER 4. THE WAVEFUNCTION
p2
H(p) = (4.7)
2m
p̂2
The Hamiltonian operator that acts on the free-particle Hilbert space is Ĥ = ,
2m
since we have the position representation for the momentum operator p̂. We have
the Hamiltonian operator:
def ~2 ∂ 2
Ĥ = − (4.8)
2m ∂x2
Plunging it in the Schrödinger’s equation; to obtain:
~2 ∂ 2 ψ(x, t) ∂ψ(x, t)
− 2
= i~ (4.9)
2m ∂x ∂t
Observe the similarity between Schrödinger’s equation and the classical wave equa-
tion. However, we have derived this equation from an axiomatic approach. In order
to solve (4.9) we need to use a mathematical trick known as the separation of
variables. Assume that we can write the wavefunction as the product of two
functions:
ψ(x, t) = ϕ(x)h(t) (4.10)
Substituting in (4.9), and rearranging, we obtain:
~2 1 d2 ϕ(x) 1 dh(t)
− 2
= i~ (4.11)
2m ϕ(x) dx h(t) dt
4.3. THE FREE-PARTICLE SOLUTION 27
Each side of the equation (4.11) depends only on one variable, and since they equal
each other. This implies :
~2 1 d2 ϕ(x)
− = Const. (4.12a)
2m ϕ(x) dx2
1 dh(t)
i~ = Const. (4.12b)
h(t) dt
Evidently, the ’constant’ is indeed the eigen-energy of the particle E. We start
by solving the second equation (4.12b), the solution yields, what-so-called the
stationary states time evolution:
with ω = E~ . Observe this result can be obtained directly from the Heisenberg
picture ( Show how !) All systems of which we can separate their time dependence
in this way is called stationary states. They shall be the main focus in these notes.
~2 d2 ϕ(x)
− = Eϕ(x) (4.14)
2m dx2
This has a particular solution of the form
With k 2 = 2mE
~2
, having the dimension of inverse length; we recognise k being the
wavenumber. The solution (4.15) can be written in terms of the momentum - by
the relation p = ~j.-:
i
u(x) = Ce ~ px (4.16)
This solution is known as the plane wave solution. It represents a wave propagating
in the +x direction. This solution can be used to find ϕ(x) by the superposition
principle; the ’constant’ of integration is not constant in fact. Rather, it is a
function of p . In order to see this, recall that the eigen-energy of the free particle
2
p2
E = 2m = (~k)
2m
putting this in (4.9) we shall have a continuous spectrum of
eigen-energies. Then use the spectral theorem we shall have therefore :
Z +∞
1 i
ϕ(x) = √ ϕ̃(p)e ~ px dp (4.17)
2π~ −∞
28 CHAPTER 4. THE WAVEFUNCTION
Figure 4.1: Simulated time evolution of a Gaussian wavefunction for the free par-
ticle, observe the dispersion of the wave as time progresses
of the particle’s position- gets wider and wider as time progresses. The equation
(4.17) indicates that the position wavefunction is composed of infinite number of
momentum wavefunctions. The movement of the quantum particle is therefore
expressed in terms of a wavepacket resulting from infinite number of waves inter-
fering; we have run a computer code representing the wavepacket as seen in figure
4.3 :
∂j(x, t)
⇔ + ρ̇(x, t) = 0 (4.23)
∂x
This is the continuity equation in quantum mechanics , implying probability is
conserved .
1. he wave function must be single valued. This means that for any given values
ofx and ψ(x, t) must have a unique value. This is a way of guaranteeing that
there is only a single value for the probability of the system being in a given
state.
3. The wave function must be continuous everywhere. That is, there are no
sudden jumps in the probability density when moving through space. If a
function has a discontinuity such as a sharp step upwards or downwards,
this can be seen as a limiting case of a very rapid change in the function.
Such a rapid change would mean that the derivative of the function was very
large (either a very large positive or negative number). In the limit of a
step function, this would imply an infinite derivative. Since the momentum
4.6. PROBLEMS 31
of the system is found using the momentum operator, which is a first order
derivative, this would imply an infinite momentum, which is not possible in
a physically realistic system. Such an infinite derivative would also violate
condition 4.
4.6 Problems
1. Calculate hpi and hEi for the free particle
dω
vg =
dk
Show that this equals the ’classical velocity’ for the particle.
6. Show that the free particles could have both k > 0 and k < 0, i.e e−ikx is
also a solution to the free particle. What does that mean physically ?
σ(x)σ(p) ∼ ~ (5.2)
This is a form of the celebrated uncertainty relation for position and momentum,
which is directly derived from solving the free particle problem. The physical
meaning for this relation is that in order to have a well-defined position for the
wave-packet ( sharp width), we need to superimpose wide momentum wavefunc-
tions ; and vice versa ( figure 5.1). Implying the impossibility for having an
absolute accurate measurement for either momentum and position regardless of
the ’apparatus’ used. Moreover, the increase in the accuracy in one implies the
decrease in the other quantity.
1
This principle is what -alone- explains the stability of atoms, counter to what
the classical theory of electrodynamics predicts 2 . Because electrons are bound
within the atomic radius ∼ 1 ÅThey have to possess a velocity uncertainty of
1
The uncertainty principle is a fundamental property in nature unrelated to our methods of
measurement
2
In classical electrodynamics, an accelerating charge emits electromagnetic radiation
33
34 CHAPTER 5. THE UNCERTAINTY PRINCIPLE
Figure 5.1: The position wavefunction is made from interference of infinite number
of momentum wavefunctions
Figure 5.2: A real picture of Hydrogen atom; taken by special techniques. Indicat-
ing the probability of electron’s position around the nucleus, in complete agreement
with quantum theory. Reference:Stodolna, A. S., et al. Phys. Rev. lett. 110.21
(2013)
about ∼ 3 × 103 m/s or few eV c2 of energy. For the electron to fall into the
nucleus, i.e. being bound by the nuclear radius ∼ 10−5 Å; it requires an enormous
energy ∼ 22 MeV or more. Thus, electrons are forbidden from falling into the nu-
cleus. Figure 5.2 shows how the uncertainty principle, and schrödinger’s equation
agrees completely with nature, the picture illustrates a real hydrogen atom’s elec-
tron wavefunction indicating the probability of finding the electron in the vicinity
of the atom.
5.2. GENERAL UNCERTAINTY PRINCIPLE 35
In the position representation this equals to dxψ ∗ σ(A)2 ψ. We define the Schwartz
R
inequality :
|A|2 |B|2 ≥ |AB|2 (5.4)
For A = σ(X) and B = σ(P ); we have :
1 1
XP = [X, P ] + (X P + P X) (5.6)
2 2
Proof :
Start by Writing :
1 1
XP = XP − P X + XP
2 2
Moreover :
1 1 1 1
XP = P X + XP − P X + XP
2 2 2 2
Gathering terms, we obtain :
1 1
XP = [X, P ] + (XP + P X)
2 2
1 1
V ar(X)b V ar(P ) ≥ | i~|2 + |(hP Xi) + hXP i|2 (5.7)
2 4
Since both hP Xi and hXP i are equal, and the expression 41 |(hP Xi) + hXP i|2 is
non-negative . We may write:
3
We shal drop the ’hat’ whenever it is understood we are talking about an operator
36 CHAPTER 5. THE UNCERTAINTY PRINCIPLE
~2
V ar(X) V ar(P ) ≥ (5.8)
4
Taking the square root :
~
σ(X)σ(P ) ≥ (5.9)
2
In fact, this result can be generalised for any two non-compatible observables A
and B:
1
σ(A)σ(B) ≥ |h[A, B]i| (5.10)
2
This is the general form of uncertainty principle.
~
σ(E)σ(t) ≥ (5.12)
2
This principle holds for any quantum system, not just the harmonic oscillator.
However, it is easier to see in this case.
The time-energy uncertainty principle has rather deep implications, in particular
for the law of energy conservation. One can ’trade’ energy with time, by adding
energy -from nothing- to the system provided it is for short period of time. This is
an important phenomena observed a lot in particle physics. For example the weak
interaction which allow beta radiation ( neutrons decaying into protons or vice
versa). A particle called W- boson , having about 80 times the mass of proton is
created from nothing, but for a very short time ∼ 10−18 sec. Just enough to allow
the weak interaction to occur.
5.4. EHRENFEST THEOREM 37
Figure 5.3: A diagram showing the weak interaction mediated by the W boson
which is created from nothing by the uncertainty principle
d 1 ∂ Ω̂
hΩ̂i = h[Ω̂, Ĥ]i + h i (5.14)
dt i~ ∂t
38 CHAPTER 5. THE UNCERTAINTY PRINCIPLE
This is the exact mathematical formula for Ehrenfest theorem. Now, we turn to
applying this theorem to the operators p and X, in order to recover the classical
equations of motion:
˙ = 1 h[p, V (x)]i
hpi
i~
p2
since p is time-independent, and the Hamiltonian takes the form : H = 2m +V.
Working in the position representation, the above commutator is written explicitly
as : Z Z
∗ ∂ ∂
− dxψ (V ψ) + dxψ ∗ V (ψ)
∂x ∂x
Applying the product rule to the first expression :
Z Z Z
∗ ∂ ∗ ∂ ∂
− dxψ (V ) ψ − dxψ V (ψ) + dxψ ∗ V (ψ)
∂x ∂x ∂x
Thus we get:
˙ = −h ∂V i
hpi (5.15)
∂x
Which is Newton’s second law, or one of Hamilton’s equations ( recall that F =
− dV
dx
)
∂V
h[p, H]i = −h i (5.16)
∂x
Let’s now apply Ehrenfest theorem. to the position operator X:
2
˙ = 1 h[X, p ]i
hXi
i~ 2m
since X is also time-independent, and the potential is only a function of position.
We need to use the result from lecture (3), equation (22)
df (B)
[A, f (B)] = [A, B]
dB
with A = X, B = p and f (p) = p2 ; we obtain:
˙ = 1 d
hXi h[X, p] (p2 )i
2im~ dp
⇔
1
= (i~)(2hpi)
2im~
Finally :
˙ = 1 hpi
hXi (5.17)
m
The definition for the classical velocity, or the second Hamilton’s equation:
1
h[X, H]i = hpi (5.18)
m
5.5. PROBLEMS 39
5.5 Problems
1. We define the first ionisation energy , as the energy needed to free an
electron from the vicinity of its atom. For Carbon 14, it is found; by detailed
quantum mechanical calculations and experimental verification that the first
ionisation energy is Emin = 11.3 eV.
(a) Estimate using the uncertainty principle Emin , if you know the electron
in the C(14) atom is confined to a box of x = 0.182 nm. and E = p2 /2m
(b) Provide an explanation, from the uncertainty principle for not observing
’protons’ being confined to atoms; use the last calculations on C(14) as
a guide.
Figure 6.1: Sketch of the scattering problem in quantum mechanics. Notice that
we use the wave nature of the quantum particles
the probability for a scattering with a certain solid angle Ω. This is calculated from
41
42 CHAPTER 6. BASIC SCATTERING THEORY
With:
2mE
k2 =
~2
6.4. THE OPTICAL THEOREM IN 1-D 43
This equation resembles the scattering amplitude and S is known as the S-matrix,
and it plays an important rôle in scattering problems. As the S-matrix elements
characterises the full properties of the scattering process.
Recall the current density in (6.2) is conserved, i.e.:
jR = jL (6.5)
That implies:
|A|2 − |B|2 = |C|2 − |D|2 (6.6)
Since:
~k
|A|2 − |B|2
jL = (6.7a)
m
~k
|C|2 − |D|2
jR = (6.7b)
m
Since we have in our problem the incident flux is from the left only ⇒ D = 0. We
have thereby:
2 2
C 2
B
TL = = |S21 | RL = = |S11 |2
A A
taking the canonical form. When (V 6= 0), S should be written according to the
continuity conditions. Let :
dψ(x)L dψ(x)R
ψ(x = 0)L = ψ(x = 0)R |x=0 = |x=0 (6.11)
dx dx
We obtain the expression for the S-matrix:
2ir 1 + 2it
S= , (6.12)
1 + 2it 2ir∗ 1−2it
1+2it
∗
Or:
|r|2 + |t|2 = =(t) (6.13)
Which is the optical theorem in one dimension.
Chapter 7
Stationary States
for a large enough potential compared to the particle’s energy. This problem is
very important example to study discrete spectrum .
We start by a particle trapped in a potential well, of width a. The Schrödinger’s
equation for this particle is written as - in position representation-:
∂ ~2 ∂ 2
i~ ψ(x, t) = − ψ(x, t) + V (x)ψ(x, t) (7.1)
∂t 2m ∂x2
With : (
0, for 0 < x < a
V (x) = (7.2)
∞, otherwise
45
46 CHAPTER 7. STATIONARY STATES
Since the above equation is clearly separable, similar to the free particle. It resem-
bles a stationary state. We then write the eigenvalue problem:
~2 ∂ 2
u(x; E) + V (x)u(x; E) = Eu(x; E) (7.3)
2m ∂x2
Because there is a null probability of the particle being outside the box. And the
second condition on the wavefunction :
du(x) du(x)
|x=0 = |x=a (7.5)
dx dx
This condition comes from naturally from the analysis of the problem. The first
derivative of the wavefunction is proportional to the momentum of the particle,
we expect the particle will ’bump’ with both walls in the same manner. Although,
this violates the Born conditions discussed in lecture 5, but keep in mind that
the infinite well is an unphysical example!
Now, we rewrite Schrödinger’s equation as:
d2 u(x)
+ k 2 u(x) = 0 (7.6)
dx2
√
with k = 2mE~; this differential equation is solved by the substitution u(x) =
eRx Resulting:
u(x) = Aeikx + Be−ikx (7.7)
Where A and B are constants, we use the identity :
We observe that, using the boundary conditions we obtain :
Figure 7.2: First,second, and third lowest-energy eigenfunctions (red) and associ-
ated probability densities (blue) for the infinite square well potential
√
But since k = 2mE~, we conclude that energy takes discrete values :
~2 n2 π 2
En =
2ma2
=E0 n2 (7.10)
~ π2 2
With E0 = 2ma 2 , the ground energy. And n here denotes the quantum number
for the excited states of the particle in the box. Now, we mayqwrite the energy-
eigenfunctions, after calculating the normalisation factor C = a2 eiϕ 1
r
2 nπx i(ωt+φ)
ψn (x, t) = sin( ) e (7.11)
a a
2
Evaluation of this integral gives:
i
!
1 − (−1)n e− ~ pa
r
aπ
v(p) = n 2 (7.15)
~ n2 π 2 − a2 ~p2
Figure 7.3: Momentum probability density function for the ground state and three
more excitation states.
2
The following identity was used :
ax
ax
sin bx dx = a2e+b2 (a sin bx + b cos bx)
R
e
7.2. SIMPLE HARMONIC OSCILLATOR 49
[a, a† ] = I (7.21a)
[a, H 0 ] = a (7.21b)
[a† , H 0 ] = −a† (7.21c)
3
The operators a, a† and H 0 along with the commutator operation [·, ·] satisfy the su(1, 1)
algebra.
50 CHAPTER 7. STATIONARY STATES
We also define the number operator N ≡ a† a that acts on the eigenstates |ni
resulting an eigenvalue of n :
N |ni = n|ni
as a result we may conclude that
a|0i = 0 (7.22)
acting on the ’ground state’by the inhalation operator, kills it. Moreover
√
a|ni = n|n − 1i (7.23)
p
a† |ni = (n + 1)|n + 1i (7.24)
Hence, The Hamiltonian acting on these states will result (the energy eigen-
value)
1
Ĥ|ni = ~ω(n + )|ni (7.25)
2
Implying that the ’number states’ are the excitation states for the quantum har-
monic oscillator. The creation and inhalation operators excite or deceit it, and it
has a discrete energy spectrum of :
1
En = ~ω n + (7.26)
2
Even in the ground state, the quantum harmonic oscillator has a non-vanishing
energy. This is a direct result for the uncertainty principle in time and energy.
7.3 Problems
7.3.1 Particle in a box
1. Using the uncertainty principle for position and momentum, estimate the
ground state energy for an infinite well of width a, compare the obtained
result with the one found in the lecture.
5. What is the ground state, first excited and second excited states energies for
an electron trapped in an infinite well of width 1 Å.
7.3. PROBLEMS 53
5. Find the eigenfunction ψ1 (x), and show that it is orthogonal to ψ0 (x) seen
in the lecture.
6. From problem (3), verify the uncertainty relation for position and momentum
.
√
7. Verify that a† |ni = n + 1|n + 1i
Angular Momentum
With ∧ being the cross ( wedge) product between the position ~ri and linear mo-
mentum p~i of the ith degree of freedom in the system. For a single particle in 3 D
we give a precise definition for the angular momentum :
x̂ ŷ ẑ
~ = x y z
L (8.2)
px py p z
Figure 8.1: Illustration for the angular momentum of a classical rotating particle
55
56 CHAPTER 8. ANGULAR MOMENTUM
L± = Lx ± iLy (8.8)
The rising and lowering operators are expressed in the coordinate representa-
tion as :
L± = ±e±iϕ (∂θ ± i cot θ∂ϕ ) (8.10)
8.3. THE SPHERICAL HARMONICS 57
• Orthonormality :
Z
Y`m
1
1
(θ, ϕ)Y`∗m
2
2
(θ, ϕ)dΩ = δm1 ,m2 δ`1 ,`2 (8.16)
angles
58 CHAPTER 8. ANGULAR MOMENTUM
• Since the spherical harmonics form an orthonormal basis, the product of two
of them is again expressed in terms of spherical harmonics. Take the product
Y`m
1
1
(θ, ϕ) · Y`m2
2
(θ, ϕ), we can directly conclude that the resultant product is
a multiple of the spherical harmonics having M = m1 + m2 since the term
containing m is only an exponential, moreover, L taking the range |`1 − `2 | ≤
L ≤ ||`1 + `2 |. The general rule for multiplication is given by Wigner 3j-
symbols ( or Clebsh-Gordon coefficients) C(`1 , m1 ; `2 , m2 ; L, M )that shall be
studied later in the addition of angular momenta.
r
X (2`1 + 1)(2`2 + 1)(2L + 1)
Y`m 1
(θ, ϕ) · Y`m 2
(θ, ϕ) =
1 2
M,L
4π
`1 `2 L ∗M `1 `2 L
× YL (θ, ϕ)
m1 m2 M 0 0 0
(8.17)
These relations will prove useful as we discuss the addition of angular mo-
menta.
∞ X
` r
va·r
X 4π r` v ` λm
e = Y m. (8.19)
2` + 1 (` + m)!(` − m)! `
p
`=0 m=−`
with
λ 1
a = ẑ − (x̂ + iŷ) + (x̂ − iŷ) (8.20)
2 2λ
λ here is a real parameter
More properties are found in the textbooks. We list here some of the spherical
harmonics and their graphical representation:
r
1 1
Y00 (θ, ϕ) = (8.21)
2 π
8.5. ANGULAR MOMENTUM EIGENSTATES 59
r r
1 3 1 3 (x − iy)
Y1−1 (θ, ϕ) = · e−iϕ · sin θ = · (8.22)
2 2π 2 2π r
r r
1 3 1 3 z
Y10 (θ, ϕ) = · cos θ = · (8.23)
2 π 2 π r
r r
1 3 1 3 (x + iy)
Y11 (θ, ϕ) = − · eiϕ · sin θ = − · (8.24)
2 2π 2 2π r
Figure 8.2: Graphical representation for some of the spherical harmonics, the
colour coding represents the probability density calculated via |Ym` (θ, ϕ)|2
|β, mi (8.25)
such that :
since the angular momentum will be totally alight with the wither z or z− after
successive application of L+ or L− . If we let `~ be the total angular momentum
eigenvalue, then obviously mmax = ` and mmin = −`.
Now we analyse (5b) further :
We may now write a full description for the angular momentum spectrum:
5. There are other types of angular momenta, that shall be explored later, same
analysis will be applied to them.
8.7 Problems
1. Verify that the operator L2 commutes with all the angular momentum op-
erator components Lz , Lx andLy .
2. Verify the commutator algebra relations for L± and Lz using the commuta-
tor relations between angular momentum operator components commutation
relations.
4. Write the Hamiltonian operator for a system of two particles with reduced
mass µ orbiting each other, in the position representation.
8. What is the minimal length of the momentum vector observable that carries
the quantum number m = −4~ ?
In classical mechanics, the Hamiltonian H for two bodies interacting via a - time
independent- potential is given by :
2
X (pi )2
(H = + V (r) (9.1)
i=1
2µ
Where, µ = MM+m m
is the reduced mass, and r is the radial separation between the
bodies. pi is the canonical momentum, that we may decompose into two parts :
L
pr =linear momentum pt = angular momentum (9.2)
r
Hence, we may write (9.1) as :
p2 L2
H= + + V (r) (9.3)
2µ 2µr2
since we know that the moment of inertia I = µr2 we can therefore write: recall that L =
Iω
p2 L2
H= + + V (r) (9.4)
2µ 2I
The potential for the Hydrogen atom is the Coulomb potential, given by the for-
mula:
ke2
V (r) = − (9.5)
r
Hence we write the Hamiltonian function as :
p2 L2 ke2
H(p, r) = + − (9.6)
2µ 2µr2 r
63
64 CHAPTER 9. THE HYDROGEN ATOM
Due to the spherical symmetry, it is logical to project the state |Ψi into the
configuration space in spherical polar coordinates, hereby the the Hilbert space
is :
H : (L, dµ)
where dµ = dr d2 dφ sin φ2 dθ , the volume element in the spherical polar coordi-
nates and the wavefunction:
The H-atom is surely a stationary state Ψ(r, φ, θ; t) = ψ(r, φ, θ)e−iωt .Hence the
time-independent Schrödinger’s equation is written as :
~2 2 ke2
− ∇ ψ(r, φ, θ) − ψ(r, φ, θ) = Eψ(r, φ, θ) (9.8)
2µ r
It was found -mathematically- that the wavefunction can be separated into three
parts:
ψ(r, φ, θ) = R(r)P (φ)F (θ) (9.9)
Where R(r) is the radial function, and P (φ)F (θ) make up the spherical Harmonics
Y (φ, θ).
• Ym` (φ, θ), the associated spherical harmonics, the eigenfunction for the oper-
ators L̂ and L̂z .
We can find the energy spectrum for the Hydrogen atom from 9.11 :
µ e2
1 13.6 eV
En = − = − (9.12)
82o ~2 n2 n2
Figure 9.1: Energy levels of the idealised H-atom and the well-known spectral
series associated with electron transitions
and so on . However, the number of electrons that can occupy each energy level is
determined by Pauli exclusion principle: stating that no two electrons in the
atom can wave an overlapping wavefunctions. In other words each wavefunction
can describe one electron only. Meaning no electrons in the atom can have all of
their quantum numbers identical.
affect the energy spectrum. However, we shall see later that it player a röle in an
interaction inside the H-atom affecting the energy spectrum.
9.6 Problems
1. If you know that the ground state radial function of the H-atom is given by:
R10 = e−r/a0
(a) Normalize it .
(b) Construct ψ100 .
(c) Calculate hri , hr2 i, hxi and hx2 i, what do you observe?
2. Derive the Rydberg formula for the H-atom. Then obtain similar formulas
for D-atom and positronium (electron orbiting its anti particle) .
(a) If there is only one electron in the H-atom, how do you explain the
existence of multiple emission lines ?
(b) Explain the spectrum in detail, using the Rydberg formula.
(c) Spectroscopy is used to distinguish between the isotopes of elements (
H2 and D2 for example). Predict the D2 spectrum and draw it, compare
it the given H2 spectrum.
68 CHAPTER 9. THE HYDROGEN ATOM
5. Compare the energy spectrum of the H-atom, with the one in a 3-D box of
dimension 1 Å.
6. Show that:
∂U 1
h i = h 2i
∂r r
If U is the Coulomb potential.
(a) Start by using the wavefunction ψ100 , and show it is valid for all values
ofr up to r = 0.
(b) Assume that the wavefunction is constant over the small volume of the
‘spherical nucleus’ then calculate the probability P ≈ Vnuclus × |ψ100 |2 .
(c) The previous step can be done in more ‘sophisticate way’ by expanding
the wavefunction around the origin using the parameter = 2rnuclear /a
(d) substitute for a0 and rnuclear to estimate the probability.
Use these equations to generate the first two associated polynomials. Are
they orthonormal over the interval [0, +∞] and weighting function w(x) =
xp e−x ?
Chapter 10
Spin
ψ(x0 ) −→ ψ(x0 + )
Using the Taylor series expansion of the translated wavefuntion around the point
x0 we can write :
dψ
ψ(x0 + ) = ψ(x0 ) + + O(2 ). (10.1)
dx
since is an infinitesimal parameter , the order terms of 2 are considered vanishing,
hence:
dψ
ψ(x0 + ) ∼ ψ(x0 ) + (10.2)
dx
i
It won’t affect the expansion if we multiplied and divided by ~
:
i ~ dψ
ψ(x0 + ) ∼ ψ(x0 ) + , (10.3)
~ i dx
~ d
we hereby identify i dx
as the m operator, P̂ . Thus,
i
ψ(x0 + ) ∼ ψ(x0 ) + P̂ (ψ) . (10.4)
~
This equation basically tells us that the translation is generated by the momen-
tum operator , or the momentum operator is the generator of translation.
69
70 CHAPTER 10. SPIN
Same argument can be made for 3-D, in Cartesian coordinates ~r = (x, y, x) , being
~ That is,
translated by δr.
~
ψ(~r) −→ ψ(~r + δr)
By the same argument done before, the multivariate Tylor expansion :
~ i ~
ψ(~r + δr) ∼ δr ∇(ψ) . (10.5)
~ i
We know that the ’linear momentum operator is defined by P~ ~i ∇ , where the gra-
dient operator here is in the Cartesian coordinates .
Now, we can run the same argument for multi-particle system with f number of
degrees of freedom in generalised coordinates : (q 1 , q 2 , . . . q f ), the transformation
of generalised coordinates is not restricted translations, but also rotations ( if some
of the generalised coordinates correspond to angles for example. ). But such trans-
formation is written in the form : ∂q∂ i , that is, it depends on the derivatives of the
coordinates - for a general configuration space- .
freedom of the system are not affected by themselves. This is the main difference
between internal and external degrees of freedom of a system.
In fact, there is a theorem in mathematical physics called the Coleman Mandula
theorem, that separates space and time translation symmetries and internal ones
( one cannot be obtained from the other.) unless supersymmetry is used, this is
beyond the scope of our study.
The distinction between, scape and time (explicitly) dependent observables, and
internal degrees of freedom is crucial in the understand of elementary particle
physics in one hand or quantum information in the other. There are many ex-
amples for observables originating from internal degrees of freedom of quantum
particles. . Electric charge, magnetic dipole moment ( from free electron), quan- We should em-
tum numbers like: leptonic or baryonic quantum numbers, strangeness...etc phisise that they
are degrees of
freedom and
10.3 Mathematical description for internal de- not internal
structure , as
grees of freedom ( elementary)
As we mentioned above, the internal degrees of freedom for a quantum system particles are as-
has an independent Hilbert space that we dealt with earlier ( we can always go to far-as we know
L2 for external ones). Now, we need to study how such space can be defined, or dimentionless .
constructed.
We start by defining the state ket |ψ, decomposed into the basis for the Hilbert
space {|αi i}.
XH
dim
|ψi = αi |αi i (10.6)
i=1
Such that, the dimension of the Hilbert space is the same as the number of the
internal degrees of freedom described by that space. The coefficients αi are the
probability amplitudes of detecting the system having the state αi , i.e. usually the
quantum system is in a state of superposition, until it is measured. Identically to
what is studied before.
We can picture a state space for these degrees of freedom and define transforma-
tions just like we have done in the external ones 10.3 Since the magnitude of |ψi
should be 1, i.e.
||ψi| = 1,
any transformation made upon this ket is ought to a rotation in the state space.
If that space is quantised, the rotation is only possible in integer multiples , and
via a ladder ( raising and lowering) operators similar to what we have seen in
the (orbital) angular momentum. Hence all these internal degrees of freedom are
physical realisations of the same mathematical structure discussed earlier in the
72 CHAPTER 10. SPIN
Figure 10.1: The quantum state can be abstractly represented as a vector in the
state space
angular momentum.
These facts are very powerful, because it allows us to apply the same techniques,
and thought processes to many physical systems ( with modifications that appears
naturally ). Depending on the nature of the physical problem ( the symmetry, the
number of degrees of freedom ..etc). To illustrate this power, this technique ( which
is known formally as representation theory) is used extensively in Fundamental
physics. The whole standard model of particle physics is fundamentally built
upon the same idea of representation, quantum electrodynamics and quantum
chromodynamics ..etc
splits either in the positive or negative direction ( say the B field is aligned with
the z axis ). The split is due to a change of energy of the beam, because electron’s
have magnetic dipole moment, and it takes a ± values or 1/2 values only, otherwise
... If the electron has an integer multiples of some magnetic dipole moment, not
1/2 the splitting would be at least in three ways, for ± direction and renaming of
the beam passing through unaffected. Quantitatively we write the change of the
energy is given by :
∆H = µ ~
~ · B. (10.7)
~ = Bz only the projection of the magnetic dipole moment in the z direction
Since B
counts, we hence conclude that :
1
µz = ± · const. (10.8)
2
Same effect is observed if the B field was aligned with any axis not just the z
axis. Hence the observable of the dipole magnetic moment is associated with a
vector operator~ˆµ that has an eigenvalues of 1/2 of some constant. This constant
was found experimentally, and called the Born magnetron µB for the electron,
but generally it is noted by γ the gyromagnetic ratio, being more careful, Born
magnetron and the gyromagnetic ratio are not exactly equal, due to relativistic
effects, and Thomas precession γ = gs · µB where gs is called Landé g-factor .
In fact we write :
~ˆ
S
~ˆ = −gs µB
µ (10.9)
~
ˆ the spin operator .
We call the operator~S
1
hŜz i = ± ~ (10.10)
2
10.5. ELECTRON’S SPIN 75
We hall stick with z but any component gives as similar results as the Sz . And the
~ is just to reserve the dimension of the spin, since it is an angular momentum.
We let , for convince denote :
1 1
hŜz i = ms ms = − ~, + ~ (10.11)
2 2
Note how similarities in notation with the orbital angular momentum is appear-
ing. Now, we turn into studying the spin operators more, via repeating the S-G
experiment in the following way:
Start with S-G apparatus in the z direction, take the half of the initial beam that
is aligned in the +z direction ( having ms = +1/2~) passed though another S-G
apparatus but aligned in the x direction, it shall split the beam once again. This
time in the ±x directions. One would expect that by these two apparatuses; we
were able to measure two observables associated with Ŝz and Ŝx . Nevertheless,
experiments have shown that this is not true, as taking the 1/4 of the beam that
passed via the +z first and +x second, to a third S-G apparatus aligned in the z.
We expect only one beam to pass in the +z direction, but this doe not happen.
The beam splits into two a third time ! This does not happen if we passed it (
initially) into three consecutive S-G apparatuses aligned in the z direction. Hence,
we cannot measure the spin in to direction simultaneously. This is mathematically
written as :
[Ŝz , Ŝx ] 6= 0 (10.12)
In fact the commutation relation for the spin operators take the form for the indices
i, j, k taking the values x, y, z:
[Ŝi Ŝj ] = i~kij Ŝk (10.13)
1
Which is the same as for the L̂’s that we studied before. We can therefore,
adopting the philosophy of previous lectures define the following operators:
Ŝ± = Ŝx ± iŜy (10.14)
and:
Ŝ 2 = Ŝx2 + Ŝy2 + Ŝz2 (10.15)
With the eigenstates :
|s, ms i (10.16)
But the eigenvalue s take one value only s = 12 ~ and ms as we have seen takes the
values ms = ±1
2
~:
Ŝz |s, ms i p
= ms ~|s, ms i
(10.17)
Ŝ|s, ms i = ~ s(s + 1)|s, ms i
1
The symbol kij is called the Levi-Civita symbol and it is equal to 0 if i = j and +1 for even
permutation and −1 for odd permutations of the indices
76 CHAPTER 10. SPIN
Moreover, the pauli spin matrices have an important mathematical property called
the Clifford algebra relation:
{σ i , σ j } = σ i σ j + σ j σ i = 2δ ij (10.24)
We call the operation {·, ·} the anticommutator. From this property we can
easily prove that:
(σ 1 )2 = (σ 2 )2 = (σ 3 )2 = δ ij (10.25)
So far, we only dealt with abstract, mathematical entities, and their properties
(commutation relations), In order to use them in physical calculations, we need
to find a proper representation for them in order to be realised in the physical
world. We shall find that there are many possible representations for the Pauli
spin matrices, that have a direct physical application and meaning. In fact, spin
of the electron is only a simple application of the representations for the Pauli spin
matrices, others go as deep as the electroweak interaction in the standard model
of particle physics, and supersymmetry! However, there are other representations,
that seems to be of an interest of mathematicians mainly like the quaternions (a
higher form of complex numbers that has 3 imaginary units) .
Now, the action of the spin operators Ŝ1 , Ŝ2 , Ŝ3 and the ladder spin operators
Ŝ+ , Ŝ− on the kets is :
~ ~ ~
Ŝ1 |χ± i = |χ∓ i Ŝ2 |χ± i = ±i |χ∓ i Ŝ3 |χ± i = ± |χ± i
2 2 2
Ŝ± |χ∓ i = ~|χ± i Ŝ± |χ± i = 0 (10.28)
We can easily express the operators above as matrices, and with the help of the
identity:
~ = ~ ~σ
S (10.29)
2
We may write the explicit expression of the Pauli matrices:
0 1
σ1 =
1 0
0 −i
σ2 =
i 0
1 0
σ3 = .
0 −1
And:
0 1
σ+ =
0 0
0 0
σ− =
1 0
We shall only deal with the index-down matrices and drop the ket notion on the
eigenstate, calling them. For convenience and consistency with quantum mechanics
textbooks.
Sometimes, the states χ+ , χ− are denoted by α and β, respectively. The spinor χ
is defined as:
χ+
χ = (10.30)
χ−
Moreover, we can define the Hermitian conjugate of the spinor:
χ † = χ∗+ χ∗−
(10.31)
That satisfies:
χ †χ = 1 (10.32)
We can calculate the expected value for an operator Ω̂ acting on the spin Hilbert
space by:
hΩ̂i = χ † Ω̂ χ (10.33)
10.8. GEOMETRIC REPRESENTATION 79
Notice that, in order for the spinor to return to its original state, before rota-
tion, one needs not to make a 2φ rotation. Rather a rotation by 4π. This is the
main characteristic of spinors, that makes them ‘ very’ different from vectors, and
manifesting itself in terms of the phase factor in the geometrical representation.
80 CHAPTER 10. SPIN
Sometimes we denote this characteristic by saying that the ‘ group’ of spin trans-
formations double covers the ‘group’ of spatial transformations.
In order to picture this in a deeper way, one can think of the spinor’s internal space
as a Möbius band - illustrated in the figure 1. A vector on the Möbius band needs
to be transported along the band twice, in order to return to its initial state.
~ ·S
H = −γ B ~ (10.40)
such that γ = e/m, the ratio between the electron’s charge and its mass.And
~ ·S
B ~ = BSz .
It is clear that:
[H, Sz ] = 0 (10.41)
Implying that there exist eigenstates for H and S simultaneously. Since we already
know the eigenstates for Sz , and represented by the spinor χ . We then write :
Hχ = Eχ (10.42)
or:
−γB Sz χ = Eχ. (10.43)
Since, Sz χ = ± 21 ~χ The eigenenergies are:
E± = ∓µB B (10.44)
10.10. ELECTRON PARAMAGNETIC RESONANCE EPR 81
e~
The constant µB = 2m e
is Born magneton. It is necessary to add another constant
gs as we have seen earlier to this equation, known as the Landé g-factor, because
the electron precesses in the magnetic field. We then have:
E± = ∓gs µB B (10.45)
or: i 1 γBt
i 21 γBt 1 −i 12 γBt 0 Ae 2
χ(t) = Ae + Be = 1 (10.48)
0 1 Be−i 2 γBt
Ems = ms gs µB B (10.49)
hνr = gs µB B (10.50)
upper energy level nupper and the lower level nlower , using Maxwell-Boltzmann
statistics, under a thermodynamic temperature T :
nupper Eupper − Elower ∆E hνr
= exp − = exp − = exp − (10.51)
nlower kT kT kT
10.11 Problems
~ the vector product with itself is equal o
1. Show that for Spin vector S,
~ ∧S
S ~ = i~S
~
~ ∧S
hint: use the expression for the vector product S ~ = εk Si Sj ~ek
ij
4. Show that,
σ i σ j = δ ij + iij
kσ
k
. hint: use both the Clifford algebra and commutation relations the Pauli
matrices obey.
7. Use the definition of the spin vector in terms of the Pauli matrices to prove
that
3
hŜ 2 i = ~2
4
8. Given a vector ~a, find the dot product:
~a · ~σ
9. Prove that :
det(σ i ) = 1
Show that the Clifford algebra relation also holds for the gamma matrices :
{γ i , γ j } = 2δ ij I
84 CHAPTER 10. SPIN
Find their eigenvalues and eigenvectors They form the matrix representation
for the Pauli spin matrices, moreover, the results obtained will form the
explicit expression of the eigenstates |χ± i in column form.
12. Given
−1 −i
Ω̂ =
i 2
an operator acting on the spin Hilbert space:
13. Since the spin operator is a vector in the 3-D space. Show that if it is rotated
by an angle γ around the z-axis, the commutation relations algebra remains
invariant, for the rotated operator .
14. Calculate the EPR frequency for an experiment with B = 100 Gauß. where
µB = 9.274 × 10−24 Joule/Tesla. and the material used is Holmium with
gs = 1.97.
Appendix A
Given a system having α particles, with associated position ~rα and momenta p~α .
We define the virial function as:
X
ς= p~α · ~rα (A.1)
α
dς X ~
= ṗα · ~rα + p~α · ~ṙα (A.2)
dt α
Since we are dealing with many-particle system. We can take the time average for
the previous expression
R τ dς
dς 0
dt
h i = R τdt (A.3)
dt 0
dt
ς(τ ) − ς(0)
=
τ
Now, if the system has a periodic motion of a period τ . The time average for the
derivative of the virial function will vanish. even if the system does not admit a
periodic motion, the virial function ought to be bounded, hence one can integrate
dς
over a sufficiently large interval such that the time average h i will approach zero.
dt
Hence, we have ( at least as an approximation): recall that F~ = ṗ~
X X
h p~α · ~ṙα i = −h ṗ~α · ~rα i (A.4)
α α
85
86 APPENDIX A. THE VIRIAL THEOREM
We can now identify the LHS being twice the kinetic energy , the RHS is the force
dotted with the position :
1 X~
hT i = − h Fα · ~rα i (A.5)
2 α
This is the Virial theorem , the expected value for the kinetic energy for a system
is equal to its virial function.
It is interesting to look at forces that arise from central potential taking the form
:
V = krn+1 (A.6)
Hence, by the virial theorem eq (A.5):
1 d
hT i = hr · (krn+1 )i (A.7)
2 dr
1
= h(n + 1)krn+1 i
2
n+1
= hV i
2
For Columb and gravitational potentials, n = −2. Therefore we have :
1
hT i = − hV i (A.8)
2
Appendix B
ψ ∗ (x)ψ(x)dx = 1
R
|ψn |2 = 1
P
hψ|ψi = hψ|ψi = n
Note that we sometimes denote the basis |ei i by |ii. Here we consider the
vector |ψi to be normalised.
87
88 APPENDIX B. SUMMERY OF HILBERT SPACES
This is knows as the spectral decomposition , where ωi and |ωi i are the eigen-
values and eigenbasis respectively. The expected-value for Ω̂ is written as:
hΩ̂i = hψ|Ω̂|ψi
Proof:
Expanding the above expression :
! ! !
X X X
ψk∗ hωk | ωj |ωj ihωj | ψi |ωi i =
k j i
XX XX
⇔ ψk∗ ωj hωk |hωj i ψi hωj |hωi i =
k j
| {z } j i
| {z }
=δkj =δij
X X
⇔ ωj |ψj |2 = ωj P (j)
j
|{z} j
=ψj∗ ψj
That implies that the eigenvalue of the function of the operator is the function of
the eigenvalue itself.
We now attempt to find the matrix element for an operator Ω̂ not expressed in
the eigenbasis. Starting by :
|φi = Ω̂|ψi
expanding this expression:
X X
φk |ki = Ω̂ ψi |ii
k i
B.2. OPERATOR PROPERTIES 89
Taking the jth component of the LHS, by projecting on the basis |ji:
X
φj = ψi hĵ|Ω|ii
i
⇔ ωj δij
Hence the operator is diagonalised by the eigenbasis.
90 APPENDIX B. SUMMERY OF HILBERT SPACES
Bibliography
[1] James Binney and David Skinner. The physics of quantum mechanics. Oxford
University Press, 2013.
[2] Paul Adrien Maurice Dirac. The principles of quantum mechanics. Num-
ber 27. Oxford university press, 1981.
[4] Stephen Gasiorowicz. Quantum physics. John Wiley & Sons, 2007.
[7] Lev Davidovich Landau, Evgenii Mikhailovich Lifshitz, JB Sykes, John Stew-
art Bell, and ME Rose. Quantum mechanics, non-relativistic theory. Physics
Today, 11:56, 1958.
[9] Jun John Sakurai and Jim Napolitano. Modern quantum mechanics. Addison-
Wesley, 2011.
[11] Cohen Claude Tannoudji, Diu Bernard, and Laloë Franck. Mécanique quan-
tique. tome i. 1973.
91
92 BIBLIOGRAPHY
[12] Angus Ellis Taylor and David C Lay. Introduction to functional analysis,
volume 2. Wiley New York, 1958.