ISAMP QC Notes Lecture 2
ISAMP QC Notes Lecture 2
Lecture Notes
May 2, 2025
Contents
1
1.7.2 Dirac Notation – Bra-Ket Formalism . . . . . . . . . . . . . . . . . . . . . . . 17
1.8 Postulates of Quantum Mechanics – A Conceptual and Experimental Foundation . . 17
1.9 Measurement in Quantum Mechanics and Collapse of the Wave Function . . . . . . 19
1.9.1 The Measurement Problem: A Departure from Classical Thinking . . . . . . 19
1.9.2 States and Measurements in Dirac Notation . . . . . . . . . . . . . . . . . . . 20
1.9.3 Collapse of the Wave Function . . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.9.4 Projection Operators and State Update Rule . . . . . . . . . . . . . . . . . . 20
1.9.5 Expectation Values and Repeated Measurements . . . . . . . . . . . . . . . . 21
1.9.6 Commuting Observables and Simultaneous Measurement . . . . . . . . . . . 21
1.9.7 Conceptual and Philosophical Implications . . . . . . . . . . . . . . . . . . . 21
1.10 Interpretations of Quantum Mechanics – Beyond the Copenhagen View . . . . . . . 22
1.10.1 Why Do We Need Interpretations? . . . . . . . . . . . . . . . . . . . . . . . . 22
1.10.2 The Copenhagen Interpretation (Standard View) . . . . . . . . . . . . . . . . 22
1.10.3 Many-Worlds Interpretation (Everett, 1957) . . . . . . . . . . . . . . . . . . . 22
1.10.4 Pilot-Wave Theory (de Broglie–Bohm) . . . . . . . . . . . . . . . . . . . . . . 23
1.10.5 Objective Collapse Theories (GRW, Penrose) . . . . . . . . . . . . . . . . . . 23
1.10.6 Quantum Bayesianism (QBism) . . . . . . . . . . . . . . . . . . . . . . . . . . 23
1.10.7 Comparative Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
Classical Worldview: The universe is predictable, continuous, and deterministic. Energy and
matter behave in well-defined, continuous ways.
3
These efforts were based on the assumption that energy exchange between matter and electromag-
netic radiation is continuous, in line with the classical worldview.
Among these attempts, the most systematic and influential was by Lord Rayleigh and later
refined by Sir James Jeans. They applied the equipartition theorem from classical statistical me-
chanics — which states that each mode of a vibrating system should, on average, have an energy
of 1/2kB T .
Assuming the cavity walls contained standing electromagnetic waves, they derived what became
known as the Rayleigh-Jeans Law and using equipartition theorem and classical electromagnetism,
Lord Rayleigh (1900) and Sir James Jeans (1905) derived:
8πν 2 kT
I(ν, T ) =
c3
This equation correctly predicted the radiation intensity at low frequencies, but disastrously
diverged at high frequencies — suggesting that black bodies should emit infinite energy in the
ultraviolet range.
This non-physical result became infamous as the ultraviolet catastrophe, exposing a fundamental
failure of classical physics in describing microscopic phenomena.
It was this failure — despite the sincere and mathematically rigorous efforts of Rayleigh and
Jeans — that set the stage for a radical new approach: the birth of quantum theory.
E = hν
Here, h is a new fundamental constant of nature—now known as Planck’s constant—with a
value of approximately 6.626 × 10−34 Js.
With this assumption, Planck derived what is now called Planck’s Radiation Law, which
perfectly matched the black body spectrum across all frequencies:
Planck’s Law
8πhν 3 1
I(ν, T ) = 3
· hν/kT
c e −1
This formula smoothly transitions between the low-frequency behavior predicted by the Rayleigh-
Jeans law and the high-frequency suppression observed in experiments—effectively curing the ul-
traviolet catastrophe.
E = hν
He applied this idea to explain the long-standing puzzle of the photoelectric effect—a phe-
nomenon in which electrons are ejected from a metal surface when light shines on it.
Classical wave theory predicted that:
• Increasing light intensity should increase the energy of emitted electrons.
• A time delay should occur before emission, especially at low light intensities.
However, experiments showed:
• Electrons are emitted instantly, even with very faint light.
• No electrons are ejected if the light’s frequency is below a certain threshold, regardless of
intensity.
• The kinetic energy of the emitted electrons depends on the frequency, not intensity.
Einstein resolved these contradictions by proposing:
A single electron absorbs a single photon. If the photon’s energy exceeds the metal’s
work function ϕ, the electron is ejected with kinetic energy K = hν − ϕ.
This explanation not only matched all observations but also provided the first compelling
evidence for the particle-like behavior of light, marking the emergence of wave-particle
duality.
Einstein’s work on the photoelectric effect earned him the Nobel Prize in Physics in
1921—not for relativity, but for introducing the concept of photons and advancing quantum theory.
h
λ=
p
Here, p = mv for a non-relativistic particle, and h is Planck’s constant.
This relation is now known as the De Broglie wavelength, and it applies to all mat-
ter—electrons, atoms, and even large objects (though wavelengths become negligible at macroscopic
scales).
De Broglie extended this idea within his doctoral thesis, which was initially met with skepticism.
However, it earned strong support from Einstein and was soon confirmed through experiments that
changed the course of quantum mechanics.
—
If particles like electrons exhibit wave-like behavior, then what exactly is “waving”?
What does a matter wave physically represent?
For light, the wave was clearly defined in terms of oscillating electric and magnetic fields. But
for particles, there was no analogous physical field. Despite much philosophical debate, scientists
struggled to ascribe a direct physical meaning to the matter wave.
Eventually, physicists began to shift their focus—not on what the wave was—but on how it
mathematically behaved. This led to the concept of the wave function, a central object in
quantum theory.
ψ(x, t) = Aei(kx−ωt)
could represent a free particle, but not one localized in space. This is a plane wave—it is
delocalized, extending infinitely in both directions. Such a wave describes a particle with a definite
momentum (since p = h̄k) but completely uncertain position. This contradicts our intuition and
the Heisenberg uncertainty principle.
To describe a particle that is localized in space, like a moving electron, physicists proposed
a new idea: the particle’s wave function must be a superposition of many plane waves, each
with slightly different momentum. The result is a wave packet.
The result is a wave function that is localized in space (i.e., it forms a “packet”), but it is made
up of a range of momenta. This beautifully illustrates the position-momentum uncertainty
principle:
h̄
∆x · ∆p
2
Here:
• |ψ⟩ is the quantum state (wave function).
• |un ⟩ or |p⟩ are the eigenstates of some Hermitian operator (e.g., momentum or Hamiltonian).
Conclusion
From De Broglie’s initial hypothesis to the wave packet formulation, and finally to the realization of
wave functions as vectors in Hilbert space, the evolution of the “matter wave” concept represents
one of the most intellectually satisfying transitions in physics. It unifies classical and quantum
insights under a precise and powerful mathematical structure, laying the foundation for everything
from atomic physics to quantum computing.
If a particle is described by a wave function, then what equation governs its evolution
in space and time?
This question led to one of the greatest breakthroughs in theoretical physics. In 1925–1926,
Austrian physicist Erwin Schrödinger sought an equation that would describe how the wave
function of a matter wave evolves with time.
Schrödinger was inspired by:
• De Broglie’s relation λ = hp ,
• Hamiltonian mechanics — especially the analogy between energy in classical systems and
wave behavior.
His goal was to construct a wave equation for matter that would:
p2
E= + V (x)
2m
p2 −h̄2 d2
→
2m 2m dx2
Thus, Schrödinger proposed the following time-independent wave equation:
h̄2 d2 ψ(x)
− + V (x)ψ(x) = Eψ(x)
2m dx2
h̄2 ∂ 2
∂ψ(x, t)
ih̄ = − + V (x) ψ(x, t)
∂t 2m ∂x2
This is the time-dependent Schrödinger equation, and it is the fundamental equation of
motion in non-relativistic quantum mechanics. It plays a role analogous to Newton’s second law
F = ma in classical mechanics.
2. Linearity: If ψ1 and ψ2 are solutions, then any linear combination αψ1 + βψ2 is also a
solution.
d
Ĥ|ψ(t)⟩ = ih̄ |ψ(t)⟩
dt
where Ĥ is the Hamiltonian operator, which contains both kinetic and potential energy:
p̂2
Ĥ = + V (x̂)
2m
Here, |ψ(t)⟩ is the quantum state vector in Hilbert space.
• It forms the basis for modern quantum technologies, including lasers, transistors, quantum
dots, and quantum computing.
Schrödinger’s equation is not merely a differential equation—it’s a new law of nature. It defines
a new language in which nature communicates at the microscopic scale.
Final Remark
Schrödinger’s wave equation unified the earlier ideas of De Broglie’s matter waves, Planck’s quanti-
zation, and Born’s probabilistic interpretation into a coherent mathematical framework. Its beauty
lies in its simplicity, its power in its predictive success. It remains a cornerstone of quantum
theory—a triumph of 20th-century physics.
He discarded the classical notion of electron trajectories in favor of focusing purely on quantities
that could be measured—such as the frequency and intensity of light emitted by atomic transitions.
where:
• ωmn = Em −En
h̄ is the transition frequency.
xp ̸= px
This became a cornerstone of quantum theory and replaced the classical Poisson brackets of
Hamiltonian mechanics.
The time evolution of observables followed from the quantum analogue of Hamilton’s equations:
!
d i ∂ Â
= [Ĥ, Â] +
dt h̄ ∂t
Here, Â is any observable operator and Ĥ is the Hamiltonian operator. This is the **Heisenberg
equation of motion**, which governs the evolution of operators rather than states.
• In Schrödinger picture, the states evolve with time, while operators remain fixed.
• In Heisenberg picture, the operators evolve with time, while states remain fixed.
They are connected via a unitary transformation, and both yield the same physical predictions.
The choice between them is often a matter of convenience or context.
2. It introduced non-commuting quantities and paved the way for operator algebra.
3. It revealed the deep role of symmetries, spectra, and algebraic structures in quantum systems.
4. It inspired the later development of quantum field theory and the use of operator formalism
in all modern quantum formulations.
Conclusion
Heisenberg’s matrix mechanics, though less intuitive than wave mechanics, was a foundational
advance in quantum theory. It demonstrated that the heart of quantum mechanics lies not in
picturable models but in mathematical structure. It complemented Schrödinger’s formulation and
marked the beginning of modern operator-based quantum physics. Together, these parallel formula-
tions culminated in the unified, abstract framework of Hilbert space quantum mechanics developed
by Dirac and von Neumann.
Â|a⟩ = a|a⟩
This is not a mathematical curiosity—it lies at the heart of quantum indeterminacy. From this,
the Heisenberg uncertainty principle follows:
h̄
∆x · ∆p ≥
2
This principle tells us that there are fundamental limits to the precision with which we can
simultaneously know certain pairs of observables.
—
⟨Â⟩ = ⟨ψ|Â|ψ⟩
This expression reflects the statistical nature of quantum mechanics.
—
In this framework, the entire machinery of quantum mechanics becomes a generalized linear
algebra on infinite-dimensional complex vector spaces—a stunning synthesis of physical intuition
and mathematical rigor.
—
Conclusion
The operator formalism provides the language in which modern quantum mechanics is written.
Every physical measurement is tied to an operator; every evolution is generated by an operator
(the Hamiltonian); and the probabilistic structure of quantum theory emerges from how states
decompose in terms of eigenvectors of these operators.
This powerful abstraction is what makes quantum mechanics both difficult and beautiful—it
replaces trajectories with transformations, and measurements with projections in a space richer
than anything in classical physics.
• Vectors: Quantum states |ψ⟩ are vectors in an infinite-dimensional complex vector space.
• Basis Vectors: Like î, ĵ, k̂ in 3D, quantum states can be expanded as:
X
|ψ⟩ = cn |n⟩
n
⟨ϕ|ψ⟩ ∈ C
• Bra ⟨ϕ|: Represents the complex conjugate row vector (dual state).
This notation not only simplifies calculations but also emphasizes the abstract nature of quan-
tum states, independent of any particular basis (e.g., position or momentum).
—
Conclusion
With these postulates and the tools of linear algebra and operator theory, quantum mechanics
becomes a complete and predictive framework. The abstractness of Hilbert space and the precision
of bra-ket notation reflect the elegance and power of the quantum world—where physical reality
is encoded not in trajectories, but in complex-valued probability amplitudes and their transforma-
tions.
This probabilistic interpretation emerged from experiments like the double-slit experiment,
where particles interfere like waves but are detected as localized impacts. The wave function is not
directly observable, but its square modulus predicts experimental outcomes.
—
d h̄2 2
Position: x̂ = x, Momentum: p̂ = −ih̄ , Energy: Ĥ = − ∇ + V (x)
dx 2m
This postulate was driven by the failure of classical ideas to explain atomic spectra and the
success of operator-based matrix mechanics introduced by Heisenberg.
—
Âϕ = aϕ
If the system is not initially in an eigenstate of Â, the act of measurement causes the wave
function to “collapse” into one of its eigenstates ϕ, and the corresponding eigenvalue a is recorded.
This postulate accounts for the inherently probabilistic nature of quantum outcomes. The
probability of obtaining eigenvalue an when measuring  in state |ψ⟩ is:
⟨Â⟩ = ⟨ψ|Â|ψ⟩
This rule is fundamental in linking theory to experiment. It allows us to compute physical
quantities from wave functions, even when exact eigenstates are not involved.
—
∂
ĤΨ(⃗r, t) = ih̄ Ψ(⃗r, t)
∂t
Here, Ĥ is the Hamiltonian operator that includes kinetic and potential energy. This postulate
provides the dynamical law of quantum mechanics and replaces Newton’s second law.
—
• The failure of classical physics to explain atomic stability, black body radiation, and the
photoelectric effect.
• Experimental breakthroughs like electron diffraction, spectral lines, and the Stern-Gerlach
experiment.
Each postulate reflects a direct attempt to reconcile observations with theory, forming a re-
markably consistent and predictive structure.
Together, these five postulates define the architecture of quantum mechanics—balancing ab-
stract mathematical rigor with direct physical applicability.
Â|an ⟩ = an |an ⟩
The eigenstates {|an ⟩} form a complete orthonormal basis of the Hilbert space. Thus, the state
|ψ⟩ can be expanded as:
X
|ψ⟩ = cn |an ⟩, where cn = ⟨an |ψ⟩
n
Here, |cn |2 is the probability of obtaining outcome an when measuring Â. This is the Born
rule:
|ψ⟩ −→ |ak ⟩
This is known as the collapse postulate or state reduction. It marks a discontinuous,
non-unitary change in the state vector.
Importantly: - Before measurement, the system is in a superposition of many eigenstates. -
After measurement, the system “jumps” into one definite eigenstate.
This collapse is instantaneous and non-deterministic—it introduces fundamental randomness
into the evolution of the system.
—
This formalism confirms: - The state is projected onto the subspace corresponding to the
eigenvalue. - The denominator ensures the state is re-normalized after projection.
—
P̂k2 = P̂k
It also aligns with the idea that eigenstates are stationary states with respect to their own
observable.
—
[Â, B̂] = 0
In this case, there exists a common basis of eigenstates |an , bm ⟩, and the state can be expanded
accordingly:
X
|ψ⟩ = cnm |an , bm ⟩
n,m
Measuring  and B̂ yields definite outcomes (an , bm ), and the state collapses into the common
eigenstate |an , bm ⟩.
If  and B̂ do not commute, measuring one will disturb the other. This is directly connected
to the uncertainty principle.
—
Conclusion
In quantum mechanics, measurement is not just a passive inquiry—it is an active intervention
that reshapes the system’s state. The collapse of the wave function, though not described by the
Schrödinger equation, is essential to reconciling theory with experiment. The projection postulate,
expectation values, and Born probabilities together form a powerful, predictive framework for
extracting physical meaning from the abstract wave function.
While reproducing all predictions of standard quantum mechanics, it introduces hidden variables
and a physically real guiding wave.
—
These models aim to solve the measurement problem dynamically, but often require new physical
constants and lack experimental confirmation so far.
—
QBism offers a radical shift in viewpoint, reframing quantum theory as a user’s guide for making
predictions.
—
What if the electron itself has an intrinsic angular momentum—an internal “spin”—responsible
for the observed spectral splitting?
He imagined the electron spinning about its own axis and hypothesized that this could produce a
magnetic moment that interacts with external magnetic fields, thereby explaining the fine structure
and anomalous Zeeman effect.
Eager and excited, Kronig discussed this idea with Pauli, hoping to gain support.
• A spinning electron would require surface speeds faster than the speed of light to match
observed magnetic moments.
25
• The proposal lacked a solid theoretical foundation at the time.
eh̄
µs = g s = g s µB
2me
eh̄
where: - µB = 2m e
is the Bohr magneton, - gs ≈ 2 is the spin gyromagnetic ratio (later explained
by Dirac’s relativistic theory).
Their model naturally accounted for the doublet structure in alkali spectra and the anomalous
Zeeman effect.
“Well, that is a nice idea, though it may be wrong. But you don’t yet have a reputation,
so you have nothing to lose.”
Despite Uhlenbeck’s later hesitations—especially after discussing the idea with the eminent
physicist Hendrik Lorentz, who raised strong objections based on classical electrodynamics—Ehrenfest
had already taken decisive action.
Without waiting for further approval, he submitted their manuscript to Die Naturwissenschaften.
When the students worried whether they had made a mistake, Ehrenfest reassured them:
Their short communication was published in late 1925, and the concept of electron spin formally
entered the scientific literature. It was a bold leap—initiated by two graduate students, backed by
a wise and courageous mentor.
• Spin explained the Pauli exclusion principle and atomic structure more deeply.
Ralph Kronig, despite being first, missed historical credit—but would later make key contribu-
tions in solid-state physics. Pauli, to his credit, later acknowledged the importance of spin and
incorporated it into his own work.
• It is described by its own set of algebraic rules and operators, distinct from orbital angular
momentum.
Conclusion
The discovery of spin was one of the boldest conceptual leaps in 20th-century physics. Despite initial
rejection, persistence and open-minded mentorship brought it into the theoretical fold. Today,
spin is a cornerstone of quantum mechanics—underpinning everything from quantum statistics to
quantum computing.
S = Iω
where I is the moment of inertia and ω is the angular velocity. A rotating charged object should
produce a magnetic dipole moment:
S h̄
vsurface = ωa = a∝
I ma
To match the observed spin angular momentum of the electron using this classical model, one
finds that the required surface speed exceeds the speed of light unless the electron has a non-zero
radius greater than ∼ 10−17 m.
But high-energy scattering experiments show that the electron behaves as a point-like particle,
with no measurable size. Thus, the classical picture of spin is untenable, as it would violate special
relativity.
Conclusion
Spin is not a literal spinning motion of a charged ball—it is a purely quantum mechanical degree
of freedom, emergent from combining quantum theory with special relativity (as seen in Dirac’s
equation). It is responsible for magnetic properties, atomic structure, and forms the foundation of
modern quantum information science.
Understanding this distinction is crucial before interpreting spin-based experiments like Stern–Gerlach.
Experimental Setup
The key components of the experiment included:
• Silver Atom Source: Silver atoms were vaporized in a high-temperature oven to produce
a beam.
• Collimation Slits: The atomic beam was collimated using narrow slits to ensure a well-
defined trajectory.
• Inhomogeneous Magnetic Field: The beam passed through a region with a non-uniform
magnetic field, created by specially shaped magnetic poles.
• Detection Screen: After traversing the magnetic field, atoms struck a detection screen,
where their positions were recorded.
F = ∇(µ · B)
Assuming the magnetic field varies along the z-axis and the magnetic moment is aligned along
z, the force simplifies to:
∂Bz
Fz = µz
∂z
For a silver atom with a single unpaired electron (spin- 12 ), the magnetic moment is:
µz = −gs µB ms
where:
• µB = eh̄
2me is the Bohr magneton,
∂Bz
Fz = −gs µB ms
∂z
Deflection Calculation
The deflection z of the atom on the detector screen can be estimated using classical mechanics.
The atom experiences the force Fz over the interaction length L of the magnetic field, acquiring a
transverse velocity vz :
Fz L
vz =
mv
where v is the longitudinal velocity of the atom and m is its mass.
After exiting the magnetic field, the atom travels a distance D to the detector, during which it
moves transversely by:
D Fz LD
z = vz · =
v mv 2
Substituting Fz :
(−gs µB ms ) ∂B
∂z LD
z
z=
mv 2
This equation shows that the deflection is directly proportional to the magnetic field gradient
and the magnetic moment, and inversely proportional to the mass and square of the velocity of the
atom.
h̄ h̄
Ŝz ↑ = + ↑, Ŝz ↓ = − ↓
2 2
The magnetic moment operator is proportional to the spin operator:
Further Developments
Subsequent experiments built upon the Stern–Gerlach setup to explore more complex quantum
phenomena, including entanglement and the behavior of multi-particle systems. The principles
demonstrated remain foundational in modern physics research and applications.