Thanks to visit codestin.com
Credit goes to www.scribd.com

0% found this document useful (0 votes)
9 views40 pages

Lecture Notes CAT1 Topics

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
9 views40 pages

Lecture Notes CAT1 Topics

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 40

Introduction to Quantum Mechanics and

Quantum Computing
(Lecture Notes for Engineering Physics : BAPHY105)

Vishnudath K N
Assistant Professor
Department of Physics, School of Advanced Sciences
VIT Vellore
2

Disclaimer
These lecture notes do not claim originality. They are an assimilation of existing materials
from various standard textbooks and lecture notes, compiled and reorganized to suit the
structure and requirements of the course syllabus.

Note
These lecture notes are continuously updated based on classroom discussions and ques-
tions asked by students. There may still be mistakes, and if you find any, please feel free
to write to me at [email protected].
Contents

1 Wave-Particle Duality 5
1.1 Electromagnetic Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.1.1 Young’s Double Slit Experiment . . . . . . . . . . . . . . . . . . . 7
1.2 Particle Nature of Radiation . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.2.1 Blackbody Radiation . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.2.2 The Ultraviolet Catastrophe . . . . . . . . . . . . . . . . . . . . . 10
1.2.3 Wien’s Radiation Law and Its Limitation . . . . . . . . . . . . . . 11
1.2.4 Planck’s Radiation Formula (1900) . . . . . . . . . . . . . . . . . 11
1.2.5 Einstein’s Photon Hypothesis (1905): Light consists of particles . 13
1.2.6 Photon Momentum . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.2.7 Compton Effect: Further Confirmation of the Photon Model . . . 14
1.3 de Broglie Hypothesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.3.1 Relativistic Case . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.4 The Wavefunction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.4.1 Normalization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
1.4.2 The Superposition Principle . . . . . . . . . . . . . . . . . . . . . 19
1.4.3 The Double Slit Experiment Revisited . . . . . . . . . . . . . . . 20
1.4.4 A First Look at the Schrödinger Equation . . . . . . . . . . . . . 22
1.5 The Stern–Gerlach Experiment . . . . . . . . . . . . . . . . . . . . . . . 23

2 Mathematical Foundations of Quantum Mechanics 27


2.1 Linear Vector Spaces and Basis . . . . . . . . . . . . . . . . . . . . . . . 27
2.2 Inner Products in Vector Space . . . . . . . . . . . . . . . . . . . . . . . 29
2.3 Hilbert Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.4 Geometric Meaning of the Inner Product . . . . . . . . . . . . . . . . . . 32
2.4.1 Vectors as Column Vectors . . . . . . . . . . . . . . . . . . . . . . 33
2.5 Dirac Notation: Bras and Kets . . . . . . . . . . . . . . . . . . . . . . . 33
2.5.1 From Geometry to Quantum Mechanics . . . . . . . . . . . . . . 34
2.6 Matrix Representations and Operators . . . . . . . . . . . . . . . . . . . 38
2.6.1 Operators as Matrices . . . . . . . . . . . . . . . . . . . . . . . . 38
2.6.2 Matrix operations: . . . . . . . . . . . . . . . . . . . . . . . . . . 39

3
Course Objectives and Outcomes

Course Objectives
1. Understand the origin and importance of quantum mechanics.

2. Study the principles of quantum mechanics.

3. Apply quantum mechanics for the understanding of quantum computing.

Course Outcomes
At the end of this course, students will be able to:

1. Identify limitations of classical physics through experimental observations.

2. Apply matrix algebra and Dirac notation to quantum problems.

3. Solve 1D and 3D potential box problems.

4. Understand quantum computing principles.

5. Demonstrate quantum phenomena via experiments and simulations.

4
Module 1

Wave-Particle Duality

Classical physics treats particles and waves as separate components. The only apparent
thing that a stone (particle) dropped into water and the ripples (waves) it creates have
in common is their ability to carry energy and momentum. The mechanics of particles
and the optics of waves are traditionally independent disciplines. Each has its own set of
principles based on distinct experiments.
Electrons possess charge and mass and behave according to the laws of particle mechanics
→ We think of electrons as particles.
Electromagnetic waves exhibit diffraction, interference, and polarization under suitable
circumstances → We think of them as waves.
We will see in this course that a moving electron can be interpreted as a wave manifes-
tation as well. Similarly, we shall see that under other circumstances electromagnetic
waves behave as though they consist of streams of particles.

1.1 Electromagnetic Waves


Wave: a disturbance that travels through a medium or space, transferring energy from
one point to another without causing permanent displacement of the medium itself.
Types of Waves:
(i) Mechanical Waves: These waves require a medium to travel. e.g., sound waves trav-
eling through air or water waves traveling through water.
(ii) Electromagnetic Waves: These waves, like light and radio waves, can travel through
a vacuum and do not require a medium.

Waves are described by properties like wavelength, amplitude, frequency, and speed.

Electromagnetic (EM) waves: Coupled oscillations of electric and magnetic fields


that move with the speed of light and exhibit typical wave behavior (see Fig. 1.1). The
speed c of EM waves in free space is given by

1
c= √ = 2.998 × 108 m/s (1.1)
µ0 ϵ0

where ϵ0 is the electric permittivity of free space and µ0 is its magnetic permeability.
James Clerk Maxwell unified the theories of electricity and magnetism into a single frame-
work called electromagnetism and proposed the notion of EM waves. The German physi-

5
6 MODULE 1. WAVE-PARTICLE DUALITY

cist Heinrich Hertz showed that EM waves indeed exist and behave exactly as Maxwell
had predicted.

Figure 1.1: Schematic representation of an EM wave. Note that the oscillating electric
field and the magnetic field are perpendicular to each other and also to the direction of
propagation. (Adapted from Modern Physics by Arthur Beiser, 6th edition, for instruc-
tional use only.)

A characteristic property of all waves is that they obey the principle of superpo-
sition: When two or more waves of the same nature travel past a point at
the same time, the instantaneous amplitude at that point is the vector sum
of the instantaneous amplitudes of the individual waves.
Instantaneous amplitude: the value of the quantity whose variations constitute the
wave at a certain place and time. For example:
(i) the instantaneous amplitude of a wave in a stretched string is the displacement of the
string from its normal position
(ii) the instantaneous amplitude of a water wave is the height of the water surface relative
to its normal level
(iii) the instantaneous amplitude of a sound wave is the change in pressure relative to
the normal pressure
(iv) For a plane EM wave in vacuum, the electric and magnetic fields are related by
E = cB. Hence, its instantaneous amplitude can be taken as either E or B, where E
and B denote the magnitudes of the electric and magnetic fields, respectively, and c is
the speed of the electromagnetic wave.

Returning to the idea of superposition: when two or more light waves meet at the
same place, they combine to form a new wave. The size (amplitude) of this new wave
depends on how the original waves add up.

• If the waves are in the same phase (they rise and fall together), they add up and
make a bigger wave - this is called constructive interference (left panel of Fig. 1.3).

• If the waves are out of phase (they rise and fall at opposite times), they cancel
each other, either partly or completely - this is called destructive interference (right
panel of Fig. 1.3).

If the original waves have different frequencies, the result will be a mix of both con-
structive and destructive interference.
1.1. ELECTROMAGNETIC WAVES 7

Figure 1.2: The EM spectrum. The visible light spectrum is the portion of the electro-
magnetic spectrum that is visible to the human eye, ranging from approximately 380 to
700 nanometers. This range includes the colors of the rainbow. (Adapted from Modern
Physics by Arthur Beiser, 6th edition, for instructional use only.)

1.1.1 Young’s Double Slit Experiment


The interference of light waves was first shown in 1801 by Thomas Young. He used two
narrow slits lit by monochromatic light from a single source (Fig. 1.4). Light from each
slit spread out like it was coming from the slit itself. This spreading is called diffraction,
which, like interference, is a property of waves.
Due to interference, the screen behind the slits does not appear uniformly illuminated.
Instead, we see alternating bright and dark fringes.

• Where the path difference between the two waves is an odd number of half wave-
lengths (λ/2, 3λ/2, 5λ/2, ...), destructive interference happens, and we get dark
fringes. (Here, path difference ∆x = ( 2n+1
2
)λ, n = 0, 1, 2, ...)

• Where the path difference is zero or a whole number of wavelengths (λ, 2λ, 3λ,
...), constructive interference happens, and we see bright fringes. (Here, path
8 MODULE 1. WAVE-PARTICLE DUALITY

Figure 1.3: The left panel shows constructive interference whereas the right panel shows
destructive interference. (Adapted from Modern Physics by Arthur Beiser, 6th edition,
for instructional use only.)

Figure 1.4: Young’s double slit experiment. (Adapted from Modern Physics by Arthur
Beiser, 6th edition, for instructional use only.)

difference ∆x = (nλ, n = 0, 1, 2, ...)

• In between, the interference is partial, so the brightness changes smoothly between


dark and bright.

According to what we know from classical physics, interference and diffrac-


tion happen only with waves. Classical particles do not show such patterns. If light
were just a stream of classical particles, we would expect only two bright spots behind
the slits, not an interference pattern.
So, Young’s experiment showed that light behaves like a wave. Later, Maxwell’s
theory explained that these waves are electromagnetic. Until the late 19th century,
this wave picture of light was fully accepted.

1.2 Particle Nature of Radiation


1.2.1 Blackbody Radiation
All objects emit energy continuously in the form of radiation, and the spectrum of this
emission depends mainly on temperature. At room temperature, most radiation is in the
infrared region and thus invisible to the human eye.
A body’s ability to emit radiation is closely linked to how well it absorbs it. At a
constant temperature, a body in thermal equilibrium must absorb and emit energy at
the same rate.
1.2. PARTICLE NATURE OF RADIATION 9

An ideal absorber that takes in all incident radiation, regardless of frequency, is


called a blackbody. In the lab, a close approximation is a hollow cavity with a small
hole. Radiation entering the hole reflects repeatedly and is almost entirely absorbed by
the inner walls, which also emit radiation characteristic of the temperature.
The radiation observed coming out of the hole is called blackbody radiation.

Figure 1.5: Blackbody spectra. The shape of the spectrum depends only on temperature.
Higher temperatures produce more radiation and shift the peak to higher frequencies. The
Spectral Energy Density u(ν) dν is the amount of energy per unit volume in the
frequency interval between ν and ν + dν. (Adapted from Modern Physics by Arthur
Beiser, 6th edition, for instructional use only.)

Key observations:
• Hotter objects emit more radiation overall. For instance, we can see from Fig. 1.5
that the spectrum for T = 1800 K lies entirely above that for 1200 K. This obser-
vation connects to a fundamental quantitative result:

– The area under each blackbody curve in Fig. 1.5 corresponds to the total
energy density, that is, the total energy per unit volume of the radiation
at that temperature. This total energy density increases rapidly with temper-
ature and is proportional to T 4 .
– Closely related to this is the Stefan–Boltzmann Law, which tells us how
much total energy is radiated per unit surface area per unit time by a
blackbody:
P = σT 4 , σ = 5.67 × 10−8 W/m2 · K4 (1.2)
where P is the total radiated power per unit surface area, and σ is the Ste-
fan–Boltzmann constant1 .
1
In general, P = ϵσT 4 , where ϵ is the emissivity of the surface, with 0 ≤ ϵ ≤ 1. A perfect blackbody
has ϵ = 1.
10 MODULE 1. WAVE-PARTICLE DUALITY

For example, a blackbody at 1800 K emits over five times as much energy per
unit area as one at 1200 K:
 4  4
1800 3
= = 5.06.
1200 2
This explains why the 1800 K curve lies entirely above the 1200 K one. The
hotter blackbody emits significantly more energy across all frequencies.

• The peak of the radiation spectrum shifts to higher frequencies (shorter wave-
lengths) as temperature increases. For example:

– A room-temperature object emits mostly in the infrared.


– A red-hot iron bar peaks in the visible red region.
– A white-hot object emits across the visible spectrum.

This shift is described by Wien’s Displacement Law. When the spectrum is


plotted as a function of frequency as in Fig.1.5, the peak occurs at a frequency νmax
such that
νmax
= a, with a ≈ 5.88 × 1010 Hz/K. (1.3)
T
You might be more familiar with the wavelength form, which states

λmax T = b, with b ≈ 2.898 × 10−3 m · K. (1.4)

It is important to note that νmax ̸= c/λmax because the relationship between fre-
quency and wavelength is nonlinear. The shapes of the blackbody curves plotted
against ν and λ are different, and so are the positions of their peaks.

1.2.2 The Ultraviolet Catastrophe


Rayleigh and Jeans tried to explain the blackbody spectrum using classical physics and
ideas from statistical mechanics.
They derived a formula (called the Rayleigh–Jeans law) for the energy distribution of
blackbody radiation:
8πkB T 2
u(ν) dν = ν dν (1.5)
c3
where:
• ν is the frequency of radiation,

• T is the absolute temperature,

• kB is Boltzmann’s constant,

• c is the speed of light.


This formula works well at low frequencies. However, it predicts that the energy
becomes infinite at high frequencies (ultraviolet region), which is not true in experiments.
This failure of classical physics is called the ultraviolet catastrophe.
It showed that classical physics could not explain blackbody radiation at high fre-
quencies, and a new theory was needed.
1.2. PARTICLE NATURE OF RADIATION 11

Figure 1.6: Comparison of the Rayleigh–Jeans formula with the observed blackbody
spectrum at 1500 K. The classical formula predicts too much energy at high frequencies.
This mismatch is called the ultraviolet catastrophe, and it was later seen as evidence
for Planck’s idea of quantized energy exchange. (Adapted from Modern Physics by Arthur
Beiser, 6th edition, for instructional use only.)

1.2.3 Wien’s Radiation Law and Its Limitation


Wilhelm Wien proposed an empirical formula to fit the blackbody spectrum at high
frequencies:
u(ν) dν = Aν 3 e−Bν/T dν (1.6)
where A and B are constants determined by fitting the experimental data.
Wien’s formula fits the experimental data well at high frequencies (short wavelengths),
but it fails at low frequencies. It underestimates the energy radiated in the infrared region.
Later, when Planck derived his radiation law, he showed that Wien’s constants are related
to fundamental constants:
8πh h
A= 3 , B= (1.7)
c kB
Thus, while Rayleigh–Jeans formula works at low frequencies and Wien’s law works
at high frequencies, neither gives a complete picture.

1.2.4 Planck’s Radiation Formula (1900)


In 1900, Max Planck proposed a new formula (which he later called a “lucky guess”) for
the spectral energy density of blackbody radiation:

8πhν 3 1
u(ν) dν = 3
· hν/k T dν (1.8)
c e B −1

where h = 6.626 × 10−34 J s is Planck’s constant.

This formula:
12 MODULE 1. WAVE-PARTICLE DUALITY

• Matches experimental results at all frequencies.

• Reduces to the Rayleigh–Jeans formula at low frequencies.

• Predicts zero energy at very high frequencies, solving the ultraviolet catastrophe.
Postulates of Planck’s Hypothesis:
• Energy exchange between matter and radiation occurs in discrete units (quanta).

• The energy of each quantum is proportional to the frequency: E = hν.

• Oscillators in the blackbody cavity walls can only have energies that are integer
multiples of hν.

• The probability of a given energy state is determined by Boltzmann statistics2 .


This was the first step toward the development of quantum theory3 . Planck received
the Nobel Prize in 1918 for his discovery of energy quanta.

• High-frequency limit (hν ≫ kB T ):


In this limit, the exponential term dominates:

ehν/kB T ≫ 1 ⇒ ehν/kB T − 1 ≈ ehν/kB T

Substituting into Planck’s formula:

8πhν 3 −hν/kB T
u(ν) dν ≈ e dν
c3
This is the same as Wien’s radiation law.

• Low-frequency limit (hν ≪ kB T ):


In this limit, use the Taylor series expansion:

x2 hν
ex ≈ 1 + x + + ··· so ehν/kB T − 1 ≈ ,
2! kB T
2
where we have ignored terms of order khν
B T
and higher. Substituting into Planck’s
formula:
8πhν 3 kB T 8πν 2 kB T
u(ν) dν ≈ · dν = dν
c3 hν c3
This is the Rayleigh-Jeans law.

Exercise: Compare the energy quanta of two oscillators:


(a) A 660 Hz tuning fork (a macroscopic harmonic oscillator),
(b) An atomic oscillator that emits or absorbs orange light of frequency 5.00 × 1014 Hz.
That is, the probability of an oscillator being in a state of energy E is proportional to e−E/kB T .
2

This weighting is crucial for calculating the average energy of each oscillator.
3
Both Wien’s displacement law and the Stefan–Boltzmann law can be derived from Planck’s radi-
ation law by appropriate mathematical manipulations: the former by finding the peak of the spectral
distribution, and the latter by integrating the total energy over all frequencies.
1.2. PARTICLE NATURE OF RADIATION 13

Model Valid At Limitation


Wien’s Law High frequencies (UV) Fails at low frequencies
Rayleigh-Jeans Law Low frequencies (IR) Diverges at high frequencies (UV catastrophe)
Planck’s Law All frequencies Matches experiment completely

Table 1.1: Comparison of radiation laws

Note: The tuning fork has a total vibrational energy of 0.04 J. Use this to comment on
whether quantization of energy is significant in the case of the tuning fork.
c
Exercise: Use the relation ν = to express the three radiation laws (Planck’s, Rayleigh
λ
c dν c
Jeans, and Wien’s) in terms of wavelength λ. (Hint : ν = =⇒ = − 2 =⇒ dν =
λ dλ λ
c
− 2 dλ, you can drop the negative sign since we are talking about energy density and it
λ
is always positive. )
Note: In the wavelength form, the Rayleigh Jeans law is valid at long wave-
lengths, while Wien’s law is accurate at short wavelengths.

Homework: Review the photoelectric effect and Einstein’s explanation of


it. (This is the work for which Einstein was awarded the Nobel Prize in
Physics in 1921.)

1.2.5 Einstein’s Photon Hypothesis (1905): Light consists of


particles
Einstein took Planck’s quantization seriously and extended it.
He proposed that light itself is made of particles, now called photons, and each
photon carries energy
E = hν (1.9)

This bold step explained the photoelectric effect, where light must deliver energy in whole
quanta to individual electrons.
Thus, Einstein interpreted Planck’s energy quantum as a real particle of light, the
photon.

1.2.6 Photon Momentum


Once Einstein proposed that light is made of photons with energy E = hν, it became
natural to ask: Do photons also carry momentum?
The answer comes from Einstein’s special theory of relativity, which relates the energy
and momentum of a particle through the following equation:

E 2 = p2 c2 + m2 c4 (1.10)

You do not need to worry about deriving this formula. Just know that it connects the
total energy E of a free particle to its momentum p and rest mass energy mc2 .
14 MODULE 1. WAVE-PARTICLE DUALITY

In the above equation, c is the speed of light in vacuum and m is the rest mass of the
particle. For a particle with zero rest mass (m = 0), such as the photon,4 this
simplifies to:
E = pc (1.11)
Solving for momentum gives:
E
p= (1.12)
c
Since each photon has energy E = hν, its momentum is:

p= (1.13)
c
Using the relation c = νλ, this can also be written as:
h h hc
p= , λ= = (1.14)
λ p E
This expression tells us that a photon, though massless, carries momentum that de-
pends on its wavelength.

1.2.7 Compton Effect: Further Confirmation of the Photon


Model
One of the most striking confirmations of the photon picture came from an experiment
by A.H. Compton in 1923.
In this experiment, X-rays were fired at a material containing loosely bound electrons.
Compton observed that after scattering, the X-rays had a slightly longer wavelength (i.e.,
lower energy), and electrons were ejected from the material, similar to how one particle
recoils when struck by another.
This observation could not be explained using the classical wave theory of light.
Instead, it made perfect sense if we assume that light is made up of photons, each
carrying energy E = hν and momentum.
In this view:
• A photon collides with an electron at rest.

• Some energy and momentum are transferred to the electron.

• The photon scatters with lower energy (longer wavelength), and the electron recoils
(see Fig. 1.7).
Using conservation of energy and momentum, Compton derived the formula for the
increase in wavelength of the scattered photon:
h
∆λ = λ′ − λ = (1 − cos θ) (1.15)
me c
Here,
4
Special theory of relativity also tells us that any particle traveling at the speed of light in vacuum
must be massless. Since the photon, which is the quantum of light, always moves at this speed, its mass
must be 0. At this stage, we just trust Einstein and his theory of special relativity; we do not dare to
question him. ;)
1.3. DE BROGLIE HYPOTHESIS 15

Figure 1.7: A photon-electron collision in which a photon is scattered by a stationary


electron through an angle θ. Since the electron recoils with momentum Pf , the photon’s
momentum decreases from pi to pf , and hence the photon wavelength increases. (Adapted
from Introduction to Quantum Mechanics by A. C. Phillips, for instructional use only.)

• λ and λ′ are the initial and scattered photon wavelengths,


• θ is the angle through which the photon is scattered,
• h
me c
≈ 2.43 × 10−12 m is the Compton wavelength of the electron.
This result showed that light must carry momentum, just like a particle. It provided
strong experimental support for the photon model, along with earlier evidence like the
photoelectric effect.

1.3 de Broglie Hypothesis


From Planck and Einstein, we have learned that electromagnetic waves have particle-like
properties. What about particles like electrons? Do they have wave-like properties?
That is exactly what the French physicist Louis de Broglie proposed in 1924. From
the previous section, we have seen that the wavelength of a photon with momentum p is
given by
h
λ= . (1.16)
p
De Broglie suggested that the above equation is completely general and applies to material
particles as well as to photons.
Now, that should come as a surprise to you. After all, from what we know so far,
particles are just particles - they have mass, energy, and momentum. Who would have
thought that they also have a wavelength?
Well, welcome to the next surprise of quantum mechanics.
The wavelength associated with a particle of momentum p given by Eq. 1.16 is known
as the de Broglie wavelength.

Exercise: Find the de Broglie wavelengths of


(a) a 46-g golf ball with a velocity of 30 m/s, and
(b) an electron with a velocity of 107 m/s.
Compute the ratio of the two wavelengths.
16 MODULE 1. WAVE-PARTICLE DUALITY

1.3.1 Relativistic Case


What is the de Broglie wavelength for a relativistic particle? To find this, remember that
for a relativistic particle,
E 2 = p2 c2 + m2 c4 (1.17)
From this, we can express the momentum as:
p
(E − mc2 )(E + mc2 )
p= .
c
Substituting into the de Broglie formula λ = hp , we get:

hc
λ= p . (1.18)
(E − mc2 )(E + mc2 )
Special Cases:
• Ultra-relativistic limit (very high energy): When the particle’s energy is much
greater than its rest energy, i.e., E ≫ mc2 , we can neglect the mass term. Then
the expression simplifies to:
hc
λ≈ . (1.19)
E
This is exactly the same as for a photon, which makes sense since photons are
massless and always travel at the speed of light (E ≫ mc2 is always true, since
mc2 = 0 for photons.).
• Non-relativistic limit (low speeds): When the particle is moving slowly com-
pared to the speed of light, that is, when the classical kinetic energy EK ≪ mc2 ,
the total energy can be approximated as5
p2
E ≈ mc2 + EK , where EK = (1.20)
2m
Thus in this limit, the momentum is given by
p
p = 2mEK (1.21)

and therefore, the de Broglie wavelength becomes


h h
λ= =√ . (1.22)
p 2mEK
5
Proof: We have E 2 = p2 c2 + m2 c4 , which can be written as
r
p
2 p2
E = m2 c4 + p2 c2 = mc 1 + 2 2 .
m c
Since we are in the non-relativistic limit where p ≪ mc, we can apply the binomial approximation
√ ϵ
1+ϵ≈1+ for small ϵ.
2
p2
Using this with ϵ = m2 c2 , we get

p2 p2
 
E ≈ mc2 1 + = mc2 + ,
2m2 c2 2m

where the second term is the classical kinetic energy EK = 12 mv 2 .


1.4. THE WAVEFUNCTION 17

In most real-life situations, particles of matter have very small de Broglie wavelengths,
making their wave nature hard to observe. However, since electrons have a tiny mass,
they have relatively larger wavelengths, making their wave-like behavior easier to detect
in experiments.

Exercise: Calculate the de Broglie wavelength of an electron that is accelerated through


an electric potential difference V . (Hint: Assume the electron is non-relativistic, and re-
member that an electron acquires a kinetic energy equal to eV when accelerated through
a potential V .)

De Broglie did not initially have direct experimental evidence to support his idea.
However, he showed that it could naturally explain why energy levels in atoms are re-
stricted to certain specific values. This was something Bohr had to postulate in his 1913
model of the hydrogen atom.
A few years later, de Broglie’s equation 1.16 was confirmed by experiments. These
experiments showed that electrons passing through crystals produced diffraction patterns,
similar to those produced by waves.
Before we look at one of the experiments that shows the wave nature of electrons, let
us try to understand what kind of wave is involved in de Broglie’s matter waves.

1.4 The Wavefunction


In water waves, what changes is the height of the water. In sound waves, it is the pressure
that varies. In light waves, the electric and magnetic fields change. So, what changes in
matter waves?

In matter waves, the quantity that varies is called the wave function. It is usually
written as the Greek letter Ψ (capital P si) or ψ (small psi). The wave function in
general is a complex-valued function and it depends on position and time.
Its value at a point ⃗r = (x, y, z) and time t is related to the chance of finding the
particle at that location and time.
In fact, in quantum mechanics, the wave function determines the state in which a
particle is. The state is the information that tells us all we need to know about the
particle at a fixed time, with the idea that the laws of physics will then dictate how
the state evolves at all later times. In the classical world, the state of the particle
is determined by its position ⃗r and its velocity ⃗v = d⃗ r
dt
. If one specifies both bits of
2
information at some time t0 , then we can use the equation of motion F⃗ = m ddt2⃗r to
determine ⃗r(t) and ⃗v (t) for all time.
However, as mentioned above, in the quantum world, the state of a particle is
determined by its wavefunction Ψ(⃗r, t). As we will see, if we know the wavefunction
at some time, say t0 , then we have all the information that we need to determine the
state at all other times.
The square of the wave function’s absolute value, written as |Ψ|2 , has a very nice
interpretation6 . This is called the probability density. It tells us how likely we are
to find the particle at a certain point and time. This idea was first introduced by Max
Born in 1926.
6
Just in case, |Ψ|2 = Ψ∗ Ψ, where Ψ∗ is the complex conjugate of Ψ.
18 MODULE 1. WAVE-PARTICLE DUALITY

From the probability density P (⃗r, t) = |Ψ(⃗r, t)|2 , one can compute actual probabilities
by multiplying by a volume: the probability that the particle sits in some small
volume dV centred around point ⃗r = (x, y, z) is P (⃗r, t) dV .

A large value of |Ψ|2 means a high chance of finding the particle there. A small value
means a low chance. But as long as |Ψ|2 ̸= 0, there is some chance, however small.

It’s important to remember:

• The wave function Ψ(x, t) describes the probability amplitude of finding a


particle at position x and time t.

• The square modulus |Ψ(x, t)|2 gives the probability density of finding the
particle at a specific location.

• Even if the wave function is spread out, the particle itself is not. When measured,
we always detect the whole particle at a single point.

• You will never detect a fraction of a particle (such as 20% of an electron), but
there can be a 20% chance of finding the entire particle at a particular location.

• The wave function must be single-valued. At each point (x, t), Ψ(x, t) must return
a unique complex number. This ensures well-defined probabilities.

• The wave function must be finite and continuous. It cannot blow up to infinity
or have sudden jumps. Otherwise, it would not represent a physical state.

• The first derivative dΨ


dx
must also be continuous, except possibly at points where
the potential energy has a sharp change. This ensures smooth behavior of the
probability flow and agrees with the Schrödinger equation.

• The wave function must vanish at infinity. That means Ψ(x, t) → 0 as x → ±∞.
Otherwise, the particle would have a nonzero chance of being found infinitely far
away.

• The wave function must be normalizable. The total probability of finding the
particle somewhere in space must be equal to 1:
Z ∞
|Ψ(x, t)|2 dx = 1
−∞

This is necessary to interpret |Ψ|2 as a probability density.

1.4.1 Normalization
You may recall from your mathematics lessons that the total probability of all possible
outcomes must add up to 1.
In quantum mechanics, |Ψ|2 represents the probability density. This means the
total probability of finding the particle anywhere in space is given by the integral of |Ψ|2
over all space.
1.4. THE WAVEFUNCTION 19

If Ψ is a valid wave function that describes a physical particle, then it must satisfy
the condition:7 Z
|Ψ(⃗r, t)|2 d3 r = 1 (1.23)
V

This is called the normalization condition. It ensures that the particle exists some-
where in space with 100% probability. Wave functions that satisfy this condition are said
to be normalized.

In practice, this is not a big issue. Suppose we have a wave function Ψ(⃗r, t) that is
not normalized but instead satisfies:
Z
d3 r |Ψ(⃗r, t)|2 = N < ∞ (1.24)

If N is a finite number, we say the wave function is normalisable8 . For a wave function
to be normalisable, it must go to zero fast enough as |⃗r| → ∞. In such cases,
√ we can

always construct a normalized wave function Ψ (⃗r, t) by dividing Ψ(⃗r, t) by N :

1
Ψ′ (⃗r, t) = √ Ψ(⃗r, t) (1.25)
N

This new wave function Ψ′ (⃗r, t) is now normalized, meaning:


Z
d3 r |Ψ′ (⃗r, t)|2 = 1 (1.26)

1.4.2 The Superposition Principle


Another key idea in quantum mechanics is the superposition principle.
It tells us that if a particle can be described by a wave function Ψ1 (⃗r, t) in one
situation, and by another wave function Ψ2 (⃗r, t) in a different situation, then when both
situations are equally possible, the total wave function is the sum:

Ψ(⃗r, t) = Ψ1 (⃗r, t) + Ψ2 (⃗r, t) (1.27)

This may sound strange at first, but it has been confirmed by many experi-
ments. It is one of the things that makes quantum mechanics very different from classical
physics.
In classical physics, if you have two possibilities, you add the probabilities.
But in quantum mechanics, we add the wave functions first, and then compute
the probability using:
P (⃗r, t) = |Ψ(⃗r, t)|2 (1.28)
This rule leads to interference effects - patterns that would be impossible if we added
probabilities directly.
We will see a striking example of this in the double slit experiment with electrons.
7
Where we denote the volume element dV = d3 r
8
Such functions are called square-integrable in mathematics, meaning their modulus squared has a
finite integral over all space.
20 MODULE 1. WAVE-PARTICLE DUALITY

1.4.3 The Double Slit Experiment Revisited


We repeat Young’s double slit experiment that we had discussed earlier, but this time,
instead of electromagnetic waves, we will use particles.

Classical Case
Let us first consider classical objects, like tennis balls. There is a wall with two openings.
These openings act as the double slits. Behind the wall is a detector screen.
We stand in front of the wall and throw tennis balls toward it. We throw them
randomly. Some balls hit the wall and bounce back. Others pass through one of the slits
and reach the detector.
We then observe how the balls are spread on the detector.

Figure 1.8: The distribution for classical particles. (Adapted


from lecture notes on Quantum Mechanics by David Tong,
https://www.damtp.cam.ac.uk/user/tong/quantum.html, for instructional use only.)

This situation is quite straightforward. Suppose we close the bottom slit and keep
only the top slit open. The detector then records a certain distribution of tennis balls.
This is shown as the yellow curve in Fig. 1.8. Let us call this distribution P1 .
Next, we close the top slit and open only the bottom slit. The balls again form a
similar distribution on the detector, just shifted downward. This is shown as the blue
curve, and we call it P2 .
Now we open both slits. In this case, the resulting pattern is simply the sum of the
two previous ones:
P12 = P1 + P2 (1.29)
This is shown as the green curve in the figure. Everything behaves just as we would
expect from classical particles.
1.4. THE WAVEFUNCTION 21

Had we been using electromagnetic waves instead of tennis balls, we would have seen
an interference pattern.

Quantum Case
With this classical picture in mind, we now turn to the quantum world. We perform
the double-slit experiment again, but this time using quantum particles - for example,
electrons. If we perform the experiment keeping only one of the two slits open, then the
result will not be very different from the classical case.
However, if we keep both the slits open and if the slits are small enough, the result is
very different from the classical case. When we detect an electron, it appears as a single
point on the detector screen. This is just like what we expect from a classical particle.
We never see a spread-out wave hitting the screen. Each electron arrives as a tiny dot.
But after sending many electrons, we see a pattern. The probability distribution that
builds up on the screen shows interference - like a wave. An example of this interference
pattern is shown in Fig. 1.9.

Figure 1.9: The double slit experiment, performed by Hitachi. The results show
the build-up of the interference pattern from 8 electrons, to 270 electrons, to 2000
electrons, and finally to 160,000 electrons where the interference fringes are clearly
visible. (Adapted from lecture notes on Quantum Mechanics by David Tong,
https://www.damtp.cam.ac.uk/user/tong/quantum.html, for instructional use only.
Original source: https://www.hitachi.com/rd/research/materials/quantum/doubleslit/index.html)

The most surprising part is this: There are some places on the screen where fewer
electrons are detected when both slits are open than when only one slit is open. This
cannot be explained using classical ideas. In classical physics, a particle must go through
either the first slit or the second slit. But here, it seems that each particle somehow
knows that both slits are open.

We can understand this using what we know about quantum mechanics.


If only the first slit is open, the electron is described by a wave function Ψ1 (⃗r). The
probability of detection at position ⃗r is:
P1 (⃗⃗r) = |Ψ1 (⃗r)|2 (1.30)
If only the second slit is open, we get another wave function Ψ2 (⃗r), and the probability
of detecting an electron at the point ⃗r is:
P2 (⃗r) = |Ψ2 (⃗r)|2 (1.31)
22 MODULE 1. WAVE-PARTICLE DUALITY

Now suppose that both slits are open (remember that both slits are identical and
equidistant from the electron source). Quantum mechanics tells us that the wave func-
tions add, not the probabilities. So we have:

Ψ12 (⃗r) = Ψ1 (⃗r) + Ψ2 (⃗r) (1.32)

Then the total probability of detecting an electron at the point ⃗r is:

P12 (⃗r) = |Ψ1 (⃗r) + Ψ2 (⃗r)|2 (1.33)

Expanding this gives:

P12 (⃗r) = |Ψ1 (⃗r)|2 + |Ψ2 (⃗r)|2 + Ψ∗1 (⃗r)Ψ2 (⃗r) + Ψ∗2 (⃗r)Ψ1 (⃗r) (1.34)

The last two terms are called interference terms. They show that:

P12 (⃗r) ̸= P1 (⃗r) + P2 (⃗r) (1.35)

These interference terms are what create the wave-like interference pattern on the
screen, even though we send the electrons one at a time.

1.4.4 A First Look at the Schrödinger Equation


As mentioned earlier, in classical physics, the motion of a particle is described using
Newton’s second law:
d2⃗r
F⃗ = m 2 , (1.36)
dt
where F⃗ is the force acting on the particle, m is its mass, and ⃗r(t) is its position at time
t.
In quantum mechanics, instead of a definite position, we describe the particle using a
wave function ψ(⃗r, t), which tells us the probability of finding the particle at position ⃗r
and time t.
The rule for how this wave function changes with time is given by the Schrödinger
equation. For a particle moving in one dimension (the position vector ⃗r = x, the y and
z components are 0), the equation is:

∂ψ(x, t) ℏ2 ∂ 2 ψ(x, t)
iℏ =− + V (x)ψ(x, t), (1.37)
∂t 2m ∂x2
where:

• ℏ= h

is the reduced Planck constant,

• m is the mass of the particle,

• V (x) is the potential energy at position x.

This is the time-dependent Schrödinger equation in one dimension.


In three dimensions, the equation becomes:

∂ψ(⃗r, t) ℏ2 2
iℏ =− ∇ ψ(⃗r, t) + V (⃗r)ψ(⃗r, t), (1.38)
∂t 2m
1.5. THE STERN–GERLACH EXPERIMENT 23

where ⃗r = (x, y, z) is the position in 3D, and


∂2 ∂2 ∂2
∇2 = + +
∂x2 ∂y 2 ∂z 2
is called the Laplacian operator.
We will mostly focus on the one-dimensional version in this course. It is perfectly
fine if you find the equation a bit intimidating at this stage. As we proceed
through the course, we will carefully understand what each term means and how to use
the Schrödinger equation to analyze simple physical systems. For now, just remember
that the Schrödinger equation plays a similar role in quantum mechanics as
Newton’s equation of motion does in classical mechanics, because it tells us
how the particle evolves with time.

Dimension of the Wave Function


The wave function ψ(⃗r) must be normalized so that the total probability is 1:
Z
|ψ(⃗r)|2 dn r = 1.

This means that |ψ(⃗r)|2 must have the dimensions of the inverse of the integration mea-
sure, dn r.

• In 1D (n = 1): The integration is over a length element dr = dx, so [ψ] = length−1/2 .


• In 2D (n = 2): The integration is over an area element d2 r = dA = dx dy, so
[ψ] = length−1 .
• In 3D (n = 3): The integration is over a volume element d3 r = dV = dx dy dz, so
[ψ] = length−3/2 .

This ensures that the probability element |ψ(⃗r)|2 dn r is dimensionless.

1.5 The Stern–Gerlach Experiment


After learning about wave–particle duality and the idea of a wavefunction, we now look
at an important experiment that shows another strange feature of quantum mechanics:
the idea of spin.

What is the Stern–Gerlach Experiment?


This experiment was performed by Otto Stern and Walther Gerlach in 1922.

• A beam of silver atoms is produced by heating silver in an oven.


• These atoms come out through a small hole and form a narrow beam.
• The beam passes through a region where there is a magnetic field that varies with
position.
• After that, the atoms hit a screen or detector.
24 MODULE 1. WAVE-PARTICLE DUALITY

Figure 1.10: The Stern–Gerlach Experiment setup. The magnetic field is non-uniform
and points mostly in the z-direction. (Adapted from Modern Quantum Mechanics by J.
J. Sakurai and Jim Napolitano, 2nd edition, for instructional use only.)

Magnetic Moment and Energy


You may remember from school that a current loop or a bar magnet has a property called
a magnetic moment, which tells us how strongly it interacts with a magnetic field.
In simple terms, the magnetic moment of a current loop is given by:

µ ⃗
⃗ = I · A, (1.39)

where:
• I is the current flowing in the loop,
• A
⃗ is the area vector of the loop (its direction is given by the right-hand rule).

This magnetic moment tries to align with an external magnetic field, just like a
compass needle aligns with Earth’s magnetic field.
In quantum physics, even particles like electrons - which are not little loops of wire
- behave as if they have a magnetic moment. This magnetic moment is related to an
intrinsic property called spin.
For an electron, the magnetic moment is related to its spin by:
 e 
⃗ = −g
µ ⃗
S, (1.40)
2m
where µ⃗ is the magnetic moment, e is the charge of the electron, m is the mass of the
⃗ is the spin of the particle, and g ≈ 2 is a constant called the g-factor.
electron, S
The energy of a magnetic moment in a magnetic field B ⃗ is given by:


E = −⃗µ · B. (1.41)

So, depending on whether the magnetic moment points along or against the magnetic
field, the particle will have different energies. This is why particles get deflected differently
1.5. THE STERN–GERLACH EXPERIMENT 25

in the Stern–Gerlach experiment. (Eq. 1.41 tells us how much potential energy the
magnetic moment has in the magnetic field. The force on a magnetic moment comes
⃗ with respect
from how the field changes with position - that is, the rate of change of B
to position.)

What do we expect classically?

Classically, if the electron behaves like a tiny spinning charged object (like a rotating
ball), it would have a magnetic moment. Because the atoms are randomly oriented,
we would expect to see a continuous spread on the detector - atoms deflected in many
directions depending on their magnetic moment.
So we might expect a broad smear on the detector screen.

What do we see in reality?

Surprisingly, the beam does not spread out smoothly.


Instead, the beam splits into exactly two spots - one slightly above and one slightly
below the central line. This is something that classical physics cannot explain!
This result suggests that the magnetic moment (and hence, the property called spin)
of the electron can only take two possible values along the direction of the magnetic
field - either “up” or “down”.
We call these:
ℏ ℏ
Sz = + and Sz = − , (1.42)
2 2
where ℏ is the reduced Planck constant.

What does this mean?

This was the first clear evidence that:

• Electrons (and other particles) have an intrinsic angular momentum called


spin.

• Spin is a quantum property - it does not come from the electron spinning like a
ball.

• The values of spin components are quantised. You cannot get in-between values.
You only get discrete outcomes like “up” or “down”.

• Measuring spin in one direction (say, z) gives either +ℏ/2 or −ℏ/2. If we instead
measured spin in the x-direction, we would again get two outcomes - now Sx = +ℏ/2
or −ℏ/2.

This idea - that quantum measurements only give certain allowed values - is called
quantisation.
26 MODULE 1. WAVE-PARTICLE DUALITY

Why is this important?


The Stern–Gerlach experiment is important because:

• It clearly shows the need for quantum mechanics.

• It gives us a real example of a two-state quantum system, which is a basic idea


behind qubits in quantum computing.

• It helps us understand how measurement works in quantum theory.

We will come back to the Stern–Gerlach experiment again later in the course.

Exercises
1. A wave has a wavelength of λ = 600 nm. Calculate its frequency. (Use the speed
of light c = 3 × 108 m/s.)

2. Calculate the momentum of a photon of wavelength λ = 700 nm. (Use Planck’s


constant h = 6.626 × 10−34 J·s.)

3. Calculate the momentum of a photon of frequency ν = 6 × 1014 Hz. (Use c =


3 × 108 m/s.)

4. Calculate the Compton wavelength shift when an X-ray photon is scattered by an


electron through an angle of 90◦ .
(Use me = 9.1 × 10−31 kg, h = 6.626 × 10−34 J·s, c = 3 × 108 m/s.)

5. The normalized wavefunction of a particle in one dimension is given by


2
Ψ(x) = Ae−x , for all x ∈ (−∞, ∞).
R∞ 2
Determine the normalization constant A. (Hint: Use the Gaussian integral e−x dx =
√ R∞ 2 √ −∞
π and then to find −∞ e−2x dx, use change of variable x → u/ 2.)

6. A wavefunction is defined in the interval x ∈ [0, 1] as

Ψ(x) = Ax(1 − x).

Determine the normalization constant A.


Module 2

Mathematical Foundations of
Quantum Mechanics

Why study this module?


Quantum mechanics is based on a mathematical language that’s quite different from
what we use in classical physics. Instead of working with positions, velocities and forces,
quantum mechanics uses states, operators and probabilities.
To actually do quantum mechanics (not just talk about it), we need to learn that
language. That’s what this module is about. It gives you the essential mathematical
tools of quantum theory. We start with linear algebra and build up to something called
Hilbert space.
You already know what a vector is, like in 3D space. In this module, we generalize
that idea to:

• Vectors with complex components (because quantum amplitudes are complex)

• Vector spaces that can have infinite dimensions (like spaces of functions)

• Operations like inner products, orthogonality and projections, which are key to
understanding how quantum systems evolve and how we measure them

If classical mechanics is geometry with vectors and forces, then quantum mechanics
is geometry in an abstract space of possibilities.
This module helps you get comfortable with that new space. It’s not just math in the
background, it’s the actual framework that quantum mechanics lives in.

2.1 Linear Vector Spaces and Basis


• A vector space is a collection of vectors (like |α⟩ , |β⟩ , |γ⟩, etc.) along with a set
of scalars (like a, b, c, . . .).

• The space is closed under two operations:

– Vector addition: Adding two vectors gives another vector in the same space:

|α⟩ + |β⟩ = |γ⟩ (2.1)

27
28 MODULE 2. MATHEMATICAL FOUNDATIONS OF QUANTUM MECHANICS

– Scalar multiplication: Multiplying a vector by a scalar gives another vector


in the space:
a |α⟩ = |γ⟩ (2.2)

• Vector addition satisfies the following rules:

– Commutative:
|α⟩ + |β⟩ = |β⟩ + |α⟩ (2.3)
– Associative:
|α⟩ + (|β⟩ + |γ⟩) = (|α⟩ + |β⟩) + |γ⟩ (2.4)
– There exists a zero vector |0⟩ such that:

|α⟩ + |0⟩ = |α⟩ (2.5)

– Every vector has an additive inverse (or negative) such that:

|α⟩ + (− |α⟩) = |0⟩ (2.6)

• Scalar multiplication satisfies:

– Distributivity over vector addition:

a(|α⟩ + |β⟩) = a |α⟩ + a |β⟩ (2.7)

– Distributivity over scalar addition:

(a + b) |α⟩ = a |α⟩ + b |α⟩ (2.8)

– Associativity of scalar multiplication:

a(b |α⟩) = (ab) |α⟩ (2.9)

– Multiplying by 1 and 0 gives:

1 |α⟩ = |α⟩ , 0 |α⟩ = |0⟩ (2.10)

• A linear combination means multiplying vectors by scalars and adding them:

a |α⟩ + b |β⟩ + c |γ⟩ + · · · (2.11)

• A vector is linearly independent if it cannot be written as a linear combination


of other vectors.

• A set of vectors is linearly independent if no vector in the set can be expressed


as a combination of the others.

• A set of vectors spans the space if every vector in the space can be written as a
linear combination of them.

• A basis is a set of vectors that are:


2.2. INNER PRODUCTS IN VECTOR SPACE 29

– Linearly independent, and


– Span the entire space.

• The dimension of a vector space is the number of vectors in any basis of that
space.

• Any vector |v⟩ can be written uniquely as a linear combination of basis vectors:
X
|v⟩ = ci |ei ⟩ (2.12)
i

where |ei ⟩ are the basis vectors and ci are the components of the vector in that
basis.

• This means a vector is represented by a list (or tuple) of its components:

|α⟩ ↔ (a1 , a2 , . . . , an ) (2.13)

• Operations on vectors translate to operations on their components:

|α⟩ + |β⟩ ↔ (a1 + b1 , a2 + b2 , . . .) (2.14)

c |α⟩ ↔ (ca1 , ca2 , . . .) (2.15)

• The zero vector has all components equal to zero:

|0⟩ ↔ (0, 0, . . . , 0) (2.16)

• The negative of a vector has all components with the opposite sign:

− |α⟩ ↔ (−a1 , −a2 , . . . , −an ) (2.17)

• Working with components is often easier, but it depends on your choice of basis.
The same vector will have different components in different bases.

2.2 Inner Products in Vector Space


• The dot product in three-dimensional cartesian space generalizes to the inner
product, which works in any dimension and allows complex vectors.

• The inner product between two vectors |α⟩ and |β⟩ is a complex number ⟨α|β⟩
satisfying:

1. ⟨β|α⟩ = ⟨α|β⟩∗
2. ⟨α|α⟩ ≥ 0, and ⟨α|α⟩ = 0 ⇐⇒ |α⟩ = |0⟩
3. ⟨α|bβ + cγ⟩ = b ⟨α|β⟩ + c ⟨α|γ⟩

• A vector space with an inner product is called an inner product space.


30 MODULE 2. MATHEMATICAL FOUNDATIONS OF QUANTUM MECHANICS

Vector Space 3D Cartesian Vec- Polynomials of De- Complex 2 × 2 Matrices


Concept tors gree ≤ 5 with real
coefficients
 
a b
Vectors ⃗v = (x, y, z) p(x) = a0 + a1 x + · · · + A =   , a, b, c, d ∈
a5 x 5 c d
C
Scalars R R C
(x1 , y1 , z1 ) + (x2 , y2 , z2 )
   
a1 a2 b1 b2
Addition = (x1 + x2 , y1 + y2 , p(x) + q(x)   +   =
z1 + z2 ) a3 a4 b3 b4
 
a1 + b 1 a2 + b 2
 
a3 + b 3 a4 + b 4
 
λa λb
Scalar Multi- a(x, y, z) = a · p(x) λA =  , λ ∈ C
plication (ax, ay, az) λc λd
 
0 0
Zero Vector (0, 0, 0) 0 + 0x + · · · + 0x5  
0 0
Basis î, ĵ, k̂ {1, x, x2 , . . . , x5 } Standard
  matrices:

1 0 0 1
 , ,
0 0 0 0
   
0 0 0 0
 , 
1 0 0 1
Dimension 3 6 4 (over C)
Linear Inde- Vectors not in same Monomials Basis matrices above are
pendence plane {1, x, x2 , . . .} linearly independent

Table 2.1: Examples of vector spaces across different contexts

• The norm (or length) of a vector is defined as:


p
∥α∥ ≡ ⟨α|α⟩ (2.18)

• A vector with norm 1 is called normalized or a unit vector.

• Two vectors are orthogonal if ⟨α|β⟩ = 0.

• A set of vectors {|αi ⟩} is orthonormal if1 :

⟨αi |αj ⟩ = δij (2.19)


1
δij , known as Kronecker delta, is defined as δij = 1 if i = j, and 0 otherwise.
2.3. HILBERT SPACE 31

• In an orthonormal basis, inner products and norms simplify (Explained later in


section 2.4):
⟨α|β⟩ = a∗1 b1 + a∗2 b2 + · · · + a∗n bn (2.20)
⟨α|α⟩ = |a1 |2 + |a2 |2 + · · · + |an |2 (2.21)
ai = ⟨ei |α⟩ , bi = ⟨ei |β⟩ (2.22)

• These generalize the familiar 3D formulas:

⃗a · ⃗b = ax bx + ay by + az bz , |⃗a|2 = a2x + a2y + a2z (2.23)

• The Schwarz inequality (also called the Cauchy–Schwarz inequality):

| ⟨α|β⟩ |2 ≤ ⟨α|α⟩ ⟨β|β⟩ (2.24)

2.3 Hilbert Space


We have now understood what it means to have a complex vector space, how to define an
inner product on it, and how this leads to notions of length, angles, orthogonality, and
projection, even when the vectors are functions.
Let us now see how this framework becomes the natural setting for quantum mechan-
ics.

What Is a Hilbert Space?


A Hilbert space is a complex vector space equipped with an inner product that satisfies
an important additional condition: it is complete with respect to the norm defined by
the inner product.
More precisely, it has the following properties:
• It is a complex vector space → we can add functions and multiply them by complex
numbers.
• It has an inner product ⟨ϕ|ψ⟩, which allows us to define notions like length, angle,
and orthogonality.
• It is complete, meaning that certain sequences of vectors (called Cauchy sequences)
always converge to a well-defined vector within the space.

Why Hilbert Space?


In quantum mechanics, the state of a particle is represented by a wavefunction ψ(x),
which is a complex-valued function of position. We want to treat these functions as
vectors in a space where we can do linear algebra, take inner products, expand in bases,
and define observables as operators (as we will see later).
However, not all functions are suitable. To interpret |ψ(x)|2 as a probability density,
we require2 : Z ∞
|ψ(x)|2 dx < ∞ (2.25)
−∞
2
This is just the condition of Normalizability that we have seen in Module 1.
32 MODULE 2. MATHEMATICAL FOUNDATIONS OF QUANTUM MECHANICS

Such functions are called square-integrable, and they form a space denoted L2 (R).
This space, together with the inner product
Z ∞
⟨ϕ|ψ⟩ = ϕ∗ (x)ψ(x)dx (2.26)
−∞

is an example of a Hilbert space.

Summary
A Hilbert space is:
• A complex vector space
• Equipped with an inner product
• Complete with respect to the norm defined by that inner product
It provides the correct mathematical setting to describe quantum states as vectors
(or more precisely, functions) with well-defined lengths, angles, and limits. From now on,
when we refer to “the space of quantum states,” we will mean a Hilbert space.

2.4 Geometric Meaning of the Inner Product


In quantum mechanics, the inner product ⟨ϕ|ψ⟩ measures the “overlap” between the
states |ϕ⟩ and |ψ⟩. To build some intuition, consider an analogy with 3D vectors.
Suppose |ψ⟩ is a 3-dimensional vector:
|ψ⟩ = a |i⟩ + b |j⟩ + c |k⟩ (2.27)
where |i⟩ , |j⟩ , |k⟩ are orthonormal basis vectors in Euclidean space. The coefficients are
obtained using inner products:
⟨i|ψ⟩ = a, ⟨j|ψ⟩ = b, ⟨k|ψ⟩ = c (2.28)
Each inner product tells us how much of |ψ⟩ lies in the direction of a given basis vector,
i.e., it is the projection (or shadow) of |ψ⟩ along that axis.
In this sense, the inner product is a measure of directional alignment between two
vectors. The full vector can be reconstructed from its projections:
|ψ⟩ = ⟨i|ψ⟩ |i⟩ + ⟨j|ψ⟩ |j⟩ + ⟨k|ψ⟩ |k⟩ (2.29)
We can generalize this to any discrete vector space. Suppose {|ei ⟩} denotes
an orthonormal basis for any vector space. Then any vector in that space, |ψ⟩, can be
represented as a linear combination of the basis vectors:
X X
|ψ⟩ = ci |ei ⟩ = ⟨ei |ψ⟩ |ei ⟩ (2.30)
i i

This can be seen easily by taking


X X X
⟨ej |ψ⟩ = ⟨ej | ci |ei ⟩⟩ = ci ⟨ej |ei ⟩ = ci δij = cj (2.31)
i i i

where, in the last step, we have used the orthonormality property ⟨ej |ei ⟩ = δij .
2.5. DIRAC NOTATION: BRAS AND KETS 33

2.4.1 Vectors as Column Vectors


P
Since any vector can be represented as |ψ⟩ = i ci |ei ⟩, we can also write |ψ⟩ simply in
terms of its projections along the basis vectors:
 
c1
c2 
|ψ⟩ =   (2.32)
..
.

2.5 Dirac Notation: Bras and Kets


Paul Dirac introduced a compact and powerful notation for quantum mechanics based
on abstract vector spaces. He split the inner product ⟨ϕ|ψ⟩ into two halves: a ket |ψ⟩
and a bra ⟨ϕ|. In this formalism:

• A ket |ψ⟩ represents a vector (state) in a Hilbert space.

• A bra ⟨ϕ| is the Hermitian adjoint (or dual) of a ket. Thus, ⟨ϕ| = (|ϕ⟩)† . The
adjoint operation is defined as

(|ϕ⟩)† = (|ϕ⟩∗ )T (2.33)

that is, the transpose of the complex conjugate. For every vector in the ket space,
there exists a corresponding bra vector (or dual vector) in the dual space. Thus,
for the ket vector in Eq. 2.32,

⟨ψ| = (|ψ⟩)† = c∗1 c∗2 · · ·



(2.34)

• The quantity ⟨ϕ|ψ⟩  denotes


 the overlap
 between
 the two vectors |ϕ⟩ and |ψ⟩. Thus
c1 b1
in general, if |ψ⟩ = c2  and |ϕ⟩ = b2 , then,
c3 b3

⟨ϕ|ψ⟩ = (|ϕ⟩)† |ψ⟩ = b∗1 c1 + b∗2 c2 + b∗3 c3 . (2.35)

Kets as Vectors, Bras as Linear Functionals


Kets live in a complex vector space, while bras live in the dual space. The bra ⟨ϕ| acts
on a ket |ψ⟩ to produce a complex number:

⟨ϕ|ψ⟩ ∈ C (2.36)

In infinite-dimensional spaces (such as the space of square-integrable functions), we


interpret: Z ∞
⟨ϕ|ψ⟩ = ϕ∗ (x)ψ(x) dx (2.37)
−∞

This generalizes the notion of the dot product and expresses the overlap (or interference)
between two states.
34 MODULE 2. MATHEMATICAL FOUNDATIONS OF QUANTUM MECHANICS

Summary
• |ψ⟩ (ket vector): abstract state vector
• ⟨ϕ| (bra vector): Hermitian adjoint of a ket
• ⟨ϕ|ψ⟩: complex number (measure of overlap)

2.5.1 From Geometry to Quantum Mechanics


In mathematics, the inner product between two vectors measures their “directional over-
lap” in an abstract vector space. In quantum mechanics, the inner product ⟨ei |ψ⟩ between
a basis vector |ei ⟩ and a state vector |ψ⟩ has a more specific interpretation: it is the prob-
ability amplitude for obtaining the outcome corresponding to |ei ⟩ when the system is
measured in that basis. The modulus squared | ⟨ei |ψ⟩ |2 gives the probability of finding
the system in the state |ei ⟩ upon measurement.
Suppose we have a quantum state |ψ⟩ in a space spanned by some orthonormal basis
{|ei ⟩}. Then we can express the state as:
X
|ψ⟩ = ⟨ei |ψ⟩ |ei ⟩ (2.38)
i

Thus, just as in geometry the inner product measures “directional overlap,” in quan-
tum mechanics it simultaneously encodes projection and the probability of obtaining a
particular measurement outcome.

Example: Spin- 21 System in the {|e1 ⟩ , |e2 ⟩} Basis


Consider a spin- 12 particle (such as an electron). In the standard basis, we can take
   
1 0
|e1 ⟩ = , |e2 ⟩ = .
0 1
These could represent, for example, spin-up and spin-down along the z-axis.
A general state |ψ⟩ in this basis can be written as:
 
c
|ψ⟩ = c1 |e1 ⟩ + c2 |e2 ⟩ = 1 ,
c2
where c1 and c2 are complex numbers called probability amplitudes.

Probability Amplitudes and Probabilities


The inner products with the basis vectors give:
⟨e1 |ψ⟩ = c1 , ⟨e2 |ψ⟩ = c2 .
• c1 is the probability amplitude for finding the particle in the state |e1 ⟩ when mea-
sured in this basis.
• c2 is the probability amplitude for |e2 ⟩.
The probabilities are:
P (e1 ) = |c1 |2 , P (e2 ) = |c2 |2 .
2.5. DIRAC NOTATION: BRAS AND KETS 35

Normalization Condition
Since the system must be found in one of these two basis states upon measurement, the
probabilities must sum to 1:
|c1 |2 + |c2 |2 = 1.
This condition is equivalent to:
⟨ψ|ψ⟩ = 1.
From the column-vector form:
 
c1
⟨ψ|ψ⟩ = c∗1 c∗2 = |c1 |2 + |c2 |2 .

c2

What If the State is Not Normalized?


Suppose we have
⟨ψ|ψ⟩ = N = |c1 |2 + |c2 |2 ,
where N > 1 but finite. To normalize the state, we define:

|ψ⟩ c1 c2
|ψ ′ ⟩ = √ = √ |e1 ⟩ + √ |e2 ⟩ .
N N N
Now:
|c1 |2 + |c2 |2
⟨ψ ′ |ψ ′ ⟩ = = 1,
N
and the probabilities become:

|c1 |2 |c2 |2
P (e1 ) = , P (e2 ) = .
N N

Worked Example
Let  
2+i
|ψ⟩ = .
1
First compute:

|c1 |2 = |2 + i|2 = (2)2 + (1)2 = 5, |c2 |2 = |1|2 = 1.

Thus:
N = |c1 |2 + |c2 |2 = 5 + 1 = 6.
The normalized state is:  
′ 1 2+i
|ψ ⟩ = √ .
6 1
Probabilities are:
5 1
P (e1 ) = , P (e2 ) = .
6 6
These now sum to 1, as required.
36 MODULE 2. MATHEMATICAL FOUNDATIONS OF QUANTUM MECHANICS

Practice Problems
1. Linear Combination and Basis
Let |e1 ⟩ = (1, 0) and |e2 ⟩ = (0, 1) be an orthonormal basis for a 2D complex vector
space. Let
|ψ⟩ = 3 |e1 ⟩ + (1 + i) |e2 ⟩

(a) Write |ψ⟩ as a column vector.


(b) Is this a linear combination of {|e1 ⟩ , |e2 ⟩}?
(c) What are the components c1 = ⟨e1 |ψ⟩ and c2 = ⟨e2 |ψ⟩?

2. Linear Independence in a 2D Vector Space


Check if the following vectors are linearly independent :
" # " #
2 4
|v1 ⟩ = , |v2 ⟩ = .
3 6

Hint.

• Two vectors in R2 (or C2 ) are linearly dependent iff one is a scalar multiple of
the other.
• To test this, try to find a scalar c such that |v2 ⟩ = c |v1 ⟩. Equate components
and check consistency:

(v2 )1 (v2 )2
c= , c= ,
(v1 )1 (v1 )2

provided the denominators are nonzero. If both equal the same c, the vectors
are dependent. If the components give a contradiction, they are independent.
• Handle zero components carefully: if a component of |v1 ⟩ is zero, use the other
component to determine c, and then check the zero-component equation for
consistency.

Solution.
(v2 )1 4 (v2 )2 6
= = 2, = = 2.
(v1 )1 2 (v1 )2 3
Both components give the same scalar c = 2. Hence |v2 ⟩ = 2 |v1 ⟩, so {|v1 ⟩ , |v2 ⟩} is
linearly dependent.

3. Linear Independence in 3D
Check if the following vectors are linearly independent :
     
1 0 2
|v1 ⟩ = 0 , |v2 ⟩ = 1 , |v3 ⟩ = 1 .
     

2 3 7
2.5. DIRAC NOTATION: BRAS AND KETS 37

Hint.
Direct-combination approach: try to express one column as a linear combination of
the others. For example attempt

|v3 ⟩ = a |v1 ⟩ + b |v2 ⟩ .

This gives a small linear system for a, b obtained by equating components:




 a · 1 + b · 0 = 2,

a · 0 + b · 1 = 1,


a · 2 + b · 3 = 7.

Solve the first two equations for a, b and then check the third for consistency. If
consistent, the columns are dependent and you have an explicit relation.3
Solution From the first two component equations:

a = 2, b = 1.

Check the third component:

2 · a + 3 · b = 2 · 2 + 3 · 1 = 4 + 3 = 7,

which matches the third component of |v3 ⟩. Therefore

|v3 ⟩ = 2 |v1 ⟩ + 1 |v2 ⟩ ,


4
so the three kets are linearly dependent.

4. Norm and Normalization


Given a vector |α⟩ = (2, i) in a 2D complex vector space with inner product

⟨α|β⟩ = α1∗ β1 + α2∗ β2

(a) Compute ⟨α|α⟩ and hence find ∥α∥.


(b) Normalize |α⟩ to get a unit vector |α̃⟩.

5. Inner Product and Orthogonality


Let |ϕ⟩ = (1, −1), |χ⟩ = (1, 1). Consider the standard complex inner product.

(a) Compute ⟨ϕ|χ⟩.


(b) Are |ϕ⟩ and |χ⟩ orthogonal?

3
An equivalent shortcut in R3 is to form the matrix M = [|v1 ⟩ |v2 ⟩ |v3 ⟩] and compute det M : if
det M = 0, the vectors are dependent; if det M ̸= 0, they are independent.
4
Equivalently, det M = 0.
38 MODULE 2. MATHEMATICAL FOUNDATIONS OF QUANTUM MECHANICS

6. Orthonormal Basis Expansion


Let {|e1 ⟩ , |e2 ⟩} be an orthonormal basis and a state is given by

1 i
|ψ⟩ = √ |e1 ⟩ + √ |e2 ⟩
2 2

(a) Find the norm of |ψ⟩.


(b) What is the probability of measuring the state |e2 ⟩?

7. Hilbert Space and Square-Integrability


R∞
Which of the following functions ψ(x) are square-integrable on R? (meaning −∞
|ψ(x)|2 dx <
∞?)

1
(a) ψ(x) = 1+x2
(b) ψ(x) = sin(x)
2
(c) ψ(x) = e−x

Justify your answers qualitatively.

2.6 Matrix Representations and Operators


Once we fix a basis, we can represent all the objects in Dirac notation concretely using
matrices and vectors.
We have already seen that in a finite-dimensional Hilbert space, we can write a ket
|ψ⟩ as a column vector:  
ψ1
 ψ2 
|ψ⟩ =  ..  (2.39)
 
 . 
ψn
and the corresponding bra ⟨ψ| is the Hermitian adjoint (conjugate transpose):

⟨ψ| = ψ1∗ ψ2∗ · · · ψn∗



(2.40)

Thus the inner product becomes a matrix multiplication:


 
ψ1 n
∗ ∗
  .
.
⟨ϕ|ψ⟩ = ϕ1 · · · ϕn  .  =
 X ∗
ϕi ψi (2.41)
ψn i=1

2.6.1 Operators as Matrices


Linear operators as matrices
A linear operator  is just a mathematical rule that takes one vector and gives you
another vector in the same vector space. If our space has dimension n, we can write
2.6. MATRIX REPRESENTATIONS AND OPERATORS 39

both the input and output vectors as columns with n numbers. In that case, Â can be
represented by an n × n square matrix.
In symbols:   
A11 · · · A1n ψ1
 .. ... .
..   ... 
 |ψ⟩ =  .  = |ϕ⟩ (2.42)
 
An1 · · · Ann ψn
This means: apply the matrix to the column representing |ψ⟩ to get a new column
representing |ϕ⟩.
Example: 2D case
Think of a 2D vector |ψ⟩ like  
2
|ψ⟩ =
3
This might represent a point in the x-y plane.
Now let the operator  be represented by the matrix
 
1 2
 =
0 1

When  acts on |ψ⟩:


      
1 2 2 1×2+2×3 8
 |ψ⟩ = = = = |ϕ⟩
0 1 3 0×2+1×3 3

Here, Â has transformed the vector |ψ⟩ into a new vector |ϕ⟩ in the same space.
In short:

• The vector is the “input”

• The matrix is the “machine” (operator)

• The output is another vector in the same space

2.6.2 Matrix operations:


• Transpose: Denoted T T , obtained by swapping rows and columns.

• Symmetric matrix: T T = T .

• Antisymmetric matrix: T T = −T .

• Complex conjugate: Denoted T ∗ , obtained by taking the complex conjugate of


each entry.

• Real matrix: T ∗ = T , Purely imaginary matrix: T ∗ = −T .

• Hermitian conjugate (adjoint): Denoted T † , defined as T † = (T ∗ )T .

• Hermitian matrix: T † = T .

• Skew-Hermitian matrix: T † = −T .
40 MODULE 2. MATHEMATICAL FOUNDATIONS OF QUANTUM MECHANICS

• Inverse: Denoted T −1 , defined by T T −1 = T −1 T = I, where I is the identity


matrix. A matrix has an inverse only if it is square and its determinant is nonzero.

• The commutator of two matrices A and B is denoted by [A, B] and is defined as


[A, B] = AB − BA.

Useful properties:

• (AT )T = A

• (A∗ )∗ = A

• (A† )† = A

• (AB)T = B T AT

• (AB)∗ = A∗ B ∗

• (AB)† = B † A†

• (A−1 )T = (AT )−1 (if A−1 exists)

• (A−1 )∗ = (A∗ )−1 (if A−1 exists)

• (A−1 )† = (A† )−1 (if A−1 exists)

• (AB)−1 = B −1 A−1 (if A−1 and B −1 exist)

You might also like