0% found this document useful (0 votes)
4 views

ISAMP QC Notes Lecture 2

The ISAMP Workshop on Quantum Computation, held from May 1-12, 2025, features lecture notes by Dr. Sugam Kumar from the Inter-University Accelerator Centre in New Delhi. The document covers fundamental concepts of quantum mechanics, including historical developments, key theories, and significant experiments like the Stern-Gerlach experiment. It emphasizes the transition from classical physics to quantum theory, highlighting critical contributions from physicists such as Max Planck and Albert Einstein.

Uploaded by

Siddharth Sharma
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)
4 views

ISAMP QC Notes Lecture 2

The ISAMP Workshop on Quantum Computation, held from May 1-12, 2025, features lecture notes by Dr. Sugam Kumar from the Inter-University Accelerator Centre in New Delhi. The document covers fundamental concepts of quantum mechanics, including historical developments, key theories, and significant experiments like the Stern-Gerlach experiment. It emphasizes the transition from classical physics to quantum theory, highlighting critical contributions from physicists such as Max Planck and Albert Einstein.

Uploaded by

Siddharth Sharma
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/ 32

ISAMP Workshop on Quantum Computation

1–12 May 2025

Lecture Notes

Dr. Sugam Kumar


Inter-University Accelerator Centre (IUAC), New Delhi, India

May 2, 2025
Contents

1 Quantum Mechanics Fundamentals 3


1.1 Classical Physics Before 1900 – The Age of Determinism . . . . . . . . . . . . . . . . 3
1.1.1 The First Crisis: Black Body Radiation . . . . . . . . . . . . . . . . . . . . . 3
1.1.2 Classical Prediction: Rayleigh-Jeans Law (1900–1905) . . . . . . . . . . . . . 3
1.1.3 Max Planck’s Quantum Hypothesis (1900) . . . . . . . . . . . . . . . . . . . . 4
1.1.4 Einstein’s Extension – Photoelectric Effect (1905) . . . . . . . . . . . . . . . 5
1.2 De Broglie’s Hypothesis – Dual Nature of Matter (1923) . . . . . . . . . . . . . . . . 5
1.2.1 Wave-Particle Symmetry – De Broglie’s Insight . . . . . . . . . . . . . . . . . 6
1.2.2 Experimental Confirmation – Electron Diffraction . . . . . . . . . . . . . . . 6
1.3 Matter Waves and the Emergence of the Wave Function . . . . . . . . . . . . . . . . 7
1.3.1 From Monochromatic Waves to Wave Packets . . . . . . . . . . . . . . . . . . 7
1.3.2 Physical Interpretation of the Wave Function . . . . . . . . . . . . . . . . . . 8
1.3.3 Wave Function as a Superposition of Eigenstates . . . . . . . . . . . . . . . . 8
1.3.4 Wave Functions as Vectors in Hilbert Space . . . . . . . . . . . . . . . . . . . 8
1.4 The Schrödinger Equation – A Cornerstone of Quantum Mechanics . . . . . . . . . . 9
1.4.1 Historical Motivation and the Birth of Quantum Dynamics . . . . . . . . . . 9
1.4.2 Mathematical Derivation of the Time-Independent Schrödinger Equation . . 9
1.4.3 Time-Dependent Schrödinger Equation . . . . . . . . . . . . . . . . . . . . . 10
1.4.4 Physical Significance of the Schrödinger Equation . . . . . . . . . . . . . . . . 10
1.4.5 Mathematical Structure and Operator Form . . . . . . . . . . . . . . . . . . . 11
1.4.6 Importance of the Schrödinger Equation in Quantum Mechanics . . . . . . . 11
1.5 Heisenberg’s Matrix Mechanics – A Parallel Formulation of Quantum Theory . . . . 12
1.5.1 Historical Context: A Different Path to Quantum Mechanics . . . . . . . . . 12
1.5.2 Heisenberg’s Quantum Rule – Transition Amplitudes . . . . . . . . . . . . . . 12
1.5.3 Mathematical Structure of Matrix Mechanics . . . . . . . . . . . . . . . . . . 13
1.5.4 Comparison with Schrödinger’s Formulation . . . . . . . . . . . . . . . . . . . 13
1.5.5 Logical Power and Legacy of Matrix Mechanics . . . . . . . . . . . . . . . . . 13
1.6 Observables and Operators in Quantum Mechanics . . . . . . . . . . . . . . . . . . . 14
1.6.1 From Classical Quantities to Quantum Observables . . . . . . . . . . . . . . . 14
1.6.2 Historical Development of Operators . . . . . . . . . . . . . . . . . . . . . . . 14
1.6.3 Operators Corresponding to Common Observables . . . . . . . . . . . . . . . 14
1.6.4 Mathematical Structure and Properties of Operators . . . . . . . . . . . . . . 15
1.6.5 Canonical Commutation Relations and the Uncertainty Principle . . . . . . . 15
1.6.6 Measurement and Expectation Values . . . . . . . . . . . . . . . . . . . . . . 15
1.6.7 The Bigger Picture: Observables in the Hilbert Space Framework . . . . . . . 16
1.7 Linear Algebra, Dirac Notation, and the Postulates of Quantum Mechanics . . . . . 16
1.7.1 Linear Algebra – The Language of Quantum Theory . . . . . . . . . . . . . . 16

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

2 Stern–Gerlach Experiment and Spin 25


2.1 Historical Background: The Mystery of Spectral Lines and a New Degree of Freedom 25
2.2 Ralph Kronig’s Insight – A Rejected Proposal . . . . . . . . . . . . . . . . . . . . . . 25
2.3 Uhlenbeck and Goudsmit – The Bold Leap to Spin . . . . . . . . . . . . . . . . . . . 26
2.4 Aftermath and Legacy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.5 Conceptual Leap: Spin is Not Classical Rotation . . . . . . . . . . . . . . . . . . . . 27
2.5.1 Electron Does Not Spin — The Failure of Classical Analogy . . . . . . . . . . 27
2.6 The Stern–Gerlach Experiment: Quantization of Angular Momentum . . . . . . . . 28

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi


Chapter 1

Quantum Mechanics Fundamentals

Lecture 2: Quantum Mechanics Fundamentals – Origins and the Quantum


Paradigm Shift

1.1 Classical Physics Before 1900 – The Age of Determinism


Before the 20th century, physics was dominated by classical mechanics and electromagnetism:

• Newtonian Mechanics (1687): Deterministic laws of motion; particles follow well-defined


trajectories under forces.

• Maxwell’s Electromagnetism (1860s): Unified electricity and magnetism; light under-


stood as a transverse electromagnetic wave.

• Thermodynamics (1820–1870s): Laws governing energy, heat, and entropy.

• Statistical Mechanics (Boltzmann, Maxwell): Explained thermodynamic properties


from atomic motion.

Classical Worldview: The universe is predictable, continuous, and deterministic. Energy and
matter behave in well-defined, continuous ways.

1.1.1 The First Crisis: Black Body Radiation


In the late 19th century, classical physics was remarkably successful in explaining most physical
phenomena. However, certain experiments exposed cracks in its foundations. A key example was
the black body radiation problem.
Black Body: A black body is an idealized physical object that absorbs all incident electro-
magnetic radiation and emits radiation with a characteristic spectrum that depends only on its
temperature.
Experimental observations showed that the intensity of radiation emitted at various wavelengths
had a well-defined peak and fell off at shorter and longer wavelengths, depending on temperature

1.1.2 Classical Prediction: Rayleigh-Jeans Law (1900–1905)


In the late 19th century, several prominent physicists attempted to derive the spectrum of black
body radiation using the well-established principles of classical mechanics and electromagnetism.

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.

1.1.3 Max Planck’s Quantum Hypothesis (1900)


Faced with the ultraviolet catastrophe and the inability of classical physics to correctly describe
black body radiation, German physicist Max Planck sought a resolution—not by abandoning
classical ideas entirely, but by carefully modifying their foundation.
In December 1900, Planck made a profound and unprecedented assumption to match experi-
mental data:
The energy exchanged between matter and electromagnetic radiation is not continuous,
but quantized—it occurs in discrete packets or “quanta”.
This was a radical departure from the prevailing belief that energy could vary smoothly and
continuously.
Planck proposed that the energy E associated with each oscillator (or electromagnetic mode)
in a cavity is proportional to the frequency ν of that radiation:

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.

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi


Initially, Planck regarded his assumption as a mathematical trick —a temporary step to reconcile
theory with experiment. He did not fully grasp its deeper physical implications. He was, in fact,
reluctant to abandon classical continuity, and introduced quantization more out of necessity than
philosophical conviction.
Ironically, what Planck considered a minor adjustment turned out to be the foundation of
a scientific revolution. His hypothesis introduced the concept of **discreteness in nature**—a
cornerstone of what would soon become quantum theory.

1.1.4 Einstein’s Extension – Photoelectric Effect (1905)


While Planck introduced the concept of energy quantization to explain black body radiation, it was
Albert Einstein who gave it a bold and physical interpretation. In 1905, Einstein proposed that
light itself is quantized —it behaves not just as a wave, but also as a stream of discrete particles,
later called photons.
According to Einstein, each photon carries an energy proportional to its frequency:

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.

1.2 De Broglie’s Hypothesis – Dual Nature of Matter (1923)


The early quantum developments introduced a profound duality: radiation (light), long regarded
as a wave, was shown to also exhibit particle-like behavior. This led to a radical idea proposed
in 1923 by the French physicist Louis de Broglie—what if this duality is not exclusive to light?
Could matter also display wave-like properties?

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi


1.2.1 Wave-Particle Symmetry – De Broglie’s Insight
Inspired by the symmetry of nature and drawing from Planck and Einstein’s work, De Broglie
argued that if waves like light can behave like particles (photons), then perhaps particles like
electrons can also behave like waves. This principle, which he called the “wave-particle symmetry,”
was not derived from experiment, but from elegant reasoning.
De Broglie hypothesized that a material particle of momentum p is associated with a wave of
wavelength λ given by:

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.

1.2.2 Experimental Confirmation – Electron Diffraction


The wave nature of light had long been established via diffraction and interference. De Broglie’s
hypothesis suggested that electrons should also show similar behavior if directed through a crystal
or slit system.

Davisson–Germer Experiment (1927)


American physicists Clinton Davisson and Lester Germer performed a groundbreaking ex-
periment at Bell Labs. They directed a beam of low-energy electrons onto a nickel crystal and
observed:
- The scattered electrons produced an interference pattern. - The angles of intensity peaks
matched precisely with the predictions from Bragg’s law, assuming the electrons had a wavelength
given by λ = h/p.
This was a direct and stunning confirmation of the wave nature of electrons—validating De
Broglie’s prediction.

G.P. Thomson’s Experiment (1927)


Simultaneously and independently, George Paget Thomson (son of J.J. Thomson) conducted an
experiment in England. He accelerated electrons through a thin metal foil and observed diffraction
rings on a photographic plate—very similar to X-ray diffraction patterns.
- His results showed that electrons interfered like waves. - Ironically, while J.J. Thomson had
discovered the electron as a particle, his son G.P. Thomson confirmed its wave nature.
These two experiments—one in the United States and one in England—cemented the reality of
matter waves. Electrons, once the prototypical particles, could behave as waves when interacting
with periodic structures.

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi


Conclusion: De Broglie’s bold idea, supported by Planck’s quantization and Einstein’s photons,
revealed a deep truth: Wave-particle duality is a universal property of nature. Radiation can behave
like

1.3 Matter Waves and the Emergence of the Wave Function


Following the confirmation of De Broglie’s hypothesis through the experiments of Davisson-Germer
and G.P. Thomson, the scientific community was faced with a profound question:

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.

1.3.1 From Monochromatic Waves to Wave Packets


A key realization was that a single-frequency wave like

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.

Wave Packet Representation


A wave packet is constructed by adding together (integrating) many monochromatic waves with
different k-values:
Z ∞
Ψ(x, t) = A(k)ei(kx−ω(k)t) dk
−∞
Here:
• A(k) is the amplitude distribution of the plane waves.
h̄k2
• ω(k) is the dispersion relation (for a free particle: ω = 2m ).

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:

∆x · ∆p
2

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi


1.3.2 Physical Interpretation of the Wave Function
This Ψ(x, t) is the so-called wave function of the particle. Although it does not have direct
physical meaning itself, its squared magnitude gives the probability density of finding the particle
at position x and time t:
|Ψ(x, t)|2 = Probability density
This probabilistic interpretation—introduced by Max Born—was revolutionary. The wave func-
tion encapsulates all measurable information about the particle. The particle no longer “has” a
trajectory in the classical sense; instead, it is described by a cloud of probabilities governed by
Ψ(x, t).

1.3.3 Wave Function as a Superposition of Eigenstates


Just as a wave packet is a superposition of many waves with different momenta, in general quantum
mechanics, a wave function is expressed as a superposition of eigenstates of an observable (like
energy, momentum, or position).
Mathematically, this is written as:
X
|ψ⟩ = cn |un ⟩
n

or for continuous spectra: Z


|ψ⟩ = c(p)|p⟩ dp

Here:
• |ψ⟩ is the quantum state (wave function).

• |un ⟩ or |p⟩ are the eigenstates of some Hermitian operator (e.g., momentum or Hamiltonian).

• cn or c(p) are the complex amplitudes (coefficients) in the basis.


This form reflects a profound principle: the state of a quantum system can always be decom-
posed in terms of basis eigenstates of an observable. Measurement projects the state onto one of
these basis states.

1.3.4 Wave Functions as Vectors in Hilbert Space


The abstraction of wave functions leads us to a deeper mathematical structure: the Hilbert space.
Just like vectors in 3D space can be written as a linear combination of basis vectors (e.g.,
⃗v = aî + bĵ + ck̂), a quantum state |ψ⟩ is a “vector” in an abstract, infinite-dimensional vector
space known as Hilbert space.
X
|ψ⟩ = cn |n⟩
n
Here:
• |n⟩ are orthonormal basis vectors (eigenstates of an operator).

• The inner product ⟨n|m⟩ = δnm defines orthonormality.

• The norm ⟨ψ|ψ⟩ = 1 ensures total probability is conserved.

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi


This formalism allows physicists to treat quantum states using tools from linear algebra and
functional analysis. Operators like momentum or Hamiltonian act on these vectors, and measure-
ments correspond to projections onto eigenvectors.

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.

1.4 The Schrödinger Equation – A Cornerstone of Quantum Me-


chanics
1.4.1 Historical Motivation and the Birth of Quantum Dynamics
After the acceptance of the wave nature of matter, confirmed by the experiments of Davisson and
Germer, and the success of De Broglie’s matter wave hypothesis, it became clear that particles like
electrons are not merely point-like entities but are described by wave functions.
But a crucial question remained:

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 ,

• Classical wave equations (like that of a vibrating string),

• 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:

• Reduce to Newtonian mechanics in the classical limit,

• Be consistent with energy conservation,

• Yield results that match experimental data for atomic spectra.

1.4.2 Mathematical Derivation of the Time-Independent Schrödinger Equation


Let us now follow Schrödinger’s reasoning mathematically.
From classical mechanics, the total energy E of a particle is:

p2
E= + V (x)
2m

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi


De Broglie had proposed that particles exhibit wave-like behavior. A free particle’s wave func-
tion can be written as:
ψ(x) = Aei(kx−ωt)
Now use the relations:
p = h̄k, E = h̄ω
Therefore, for a time-independent wave function, the kinetic energy operator becomes:

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

This is known as the time-independent Schrödinger equation. It is a differential equation


that determines the spatial form of the wave function ψ(x) for a particle in a potential V (x).

1.4.3 Time-Dependent Schrödinger Equation


To describe the full evolution of a quantum state over time, Schrödinger extended his idea to
propose the time-dependent equation.
Starting from the energy relation:
∂ ∂
E → ih̄ , p → −ih̄
∂t ∂x
He generalized the wave equation to:

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.

1.4.4 Physical Significance of the Schrödinger Equation


The Schrödinger equation governs the entire dynamics of a quantum system. It tells us how the
wave function evolves with time. From the wave function, we can derive all physical quantities:

• Probability distributions |ψ(x, t)|2 ,

• Expectation values of observables,

• Energy eigenvalues and eigenstates,

• Tunneling probabilities and bound states,

• Time evolution and interference patterns.

It embodies the following essential features:

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi


1. Determinism of evolution: Given the wave function at time t0 , its future evolution is fully
determined by the equation.

2. Linearity: If ψ1 and ψ2 are solutions, then any linear combination αψ1 + βψ2 is also a
solution.

3. Superposition principle: The wave function naturally accommodates quantum interference


and entanglement.

1.4.5 Mathematical Structure and Operator Form


The Schrödinger equation can also be written in operator form:

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.

1.4.6 Importance of the Schrödinger Equation in Quantum Mechanics


• It is the foundation of wave mechanics—the first complete formulation of quantum me-
chanics.

• It accurately describes atomic, molecular, and solid-state systems.

• It explains phenomena like energy quantization, tunneling, and superposition.

• It enables precise prediction of measurement outcomes via the Born rule.

• 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.

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi


1.5 Heisenberg’s Matrix Mechanics – A Parallel Formulation of
Quantum Theory
1.5.1 Historical Context: A Different Path to Quantum Mechanics
While Erwin Schrödinger was developing his wave mechanics in 1925–1926, another groundbreak-
ing approach to quantum mechanics was being independently formulated by the young German
physicist Werner Heisenberg.
Heisenberg took a very different route—one that was rooted not in visualizable wave pictures,
but in abstract mathematics and observable quantities. His formulation came to be known as
matrix mechanics.

The Motivation Behind Matrix Mechanics


In the early 1920s, atomic spectroscopy was producing increasingly accurate measurements of spec-
tral lines emitted by atoms, especially hydrogen. These spectral lines followed empirical rules such
as the Rydberg formula, but classical physics could not explain their origin.
Heisenberg’s key insight was:

Physics should be based only on observable quantities, not on speculative, unobservable


concepts like electron orbits.

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.

1.5.2 Heisenberg’s Quantum Rule – Transition Amplitudes


Heisenberg proposed that instead of describing particles’ positions and momenta as functions of
time, one should describe transitions between energy levels using arrays of numbers that encode
frequencies and amplitudes of emitted radiation.
These arrays of transition quantities could be organized as a set of matrices:

Position: xmn (t) = xmn eiωmn t

where:

• m and n refer to discrete energy levels,

• xmn is the transition amplitude,

• ωmn = Em −En
h̄ is the transition frequency.

The key realization was that these matrices do not commute:

xp ̸= px

This non-commutativity captured the essence of quantum behavior.

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi


1.5.3 Mathematical Structure of Matrix Mechanics
In collaboration with Max Born and Pascual Jordan, Heisenberg developed a consistent formalism
of quantum mechanics based on matrix algebra.
They introduced canonical commutation relations:

[x̂, p̂] = x̂p̂ − p̂x̂ = ih̄

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.

1.5.4 Comparison with Schrödinger’s Formulation


Heisenberg’s matrix mechanics and Schrödinger’s wave mechanics appear very different at first
glance. Yet, they are mathematically equivalent.

• 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.

1.5.5 Logical Power and Legacy of Matrix Mechanics


Heisenberg’s approach was revolutionary in several ways:

1. It abandoned classical intuition and emphasized observable phenomena.

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.

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi


Summary Timeline
Year Contributor Development
1860s–1890s Maxwell, Boltzmann Classical theories of EM and thermodynamics
1900 Rayleigh & Jeans Classical black body radiation law (failed at UV)
1900 Max Planck Quantization of energy (E = hν)
1905 Einstein Photon theory of light (photoelectric effect)
1923 De Broglie Matter-wave duality
1925–26 Heisenberg, Schrödinger Quantum mechanics formalism

1.6 Observables and Operators in Quantum Mechanics


1.6.1 From Classical Quantities to Quantum Observables
In classical mechanics, physical quantities—like position, momentum, and energy—are represented
by real-valued functions of time and space. They have well-defined values at every instant and
evolve deterministically according to Newton’s or Hamilton’s equations.
However, in quantum mechanics, particles do not have definite positions or momenta until a
measurement is made. Instead, the system is described by a wave function ψ(x, t), which encodes
the probabilities of all possible outcomes.
This shift in perspective required a radical reinterpretation of how physical quantities (or ob-
servables) are handled.
In quantum mechanics, every observable is associated with a linear Hermitian operator
that acts on the system’s wave function or state vector.

1.6.2 Historical Development of Operators


The operator formalism was developed through the parallel efforts of:
• Heisenberg, who introduced matrix representations of observables.
• Schrödinger, who encoded observables as differential operators acting on wave functions.
• Dirac, who unified the formalism with bra-ket notation and abstract vector spaces.
These efforts culminated in the modern view where quantum systems reside in a Hilbert space,
and observables are represented by self-adjoint (Hermitian) operators on that space.

1.6.3 Operators Corresponding to Common Observables


Below is a summary of key classical observables and their quantum operator counterparts.
Observable Classical Expression Quantum Operator
Position x x̂ = x (multiplication operator)
d
Momentum p p̂ = −ih̄ dx
p2 h̄2 d2
Kinetic Energy 2m T̂ = − 2m dx2
Potential Energy V (x) V̂ = V (x) (acts multiplicatively)
Total Energy (Hamiltonian) H =T +V Ĥ = T̂ + V̂
d

Angular Momentum (1D) L = xp L̂ = −ih̄ x dx
Time (parameter) — Not an operator in standard QM

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi


1.6.4 Mathematical Structure and Properties of Operators


Quantum operators are not just placeholders—they possess deep mathematical properties:

• Linearity: Â(c1 ψ1 + c2 ψ2 ) = c1 Âψ1 + c2 Âψ2

• Hermitian:  = † — ensures that eigenvalues (measurable quantities) are real.

• Eigenvalue Equations: Observables satisfy:

Â|a⟩ = a|a⟩

where a is a measurable value (eigenvalue) and |a⟩ is the corresponding eigenstate.

• Commutators: The commutator of two operators  and B̂ is defined as:

[Â, B̂] = ÂB̂ − B̂ Â

If [Â, B̂] ̸= 0, the observables are said to be incompatible.

1.6.5 Canonical Commutation Relations and the Uncertainty Principle


The most famous commutator in quantum mechanics is:

[x̂, p̂] = ih̄

This is not a mathematical curiosity—it lies at the heart of quantum indeterminacy. From this,
the Heisenberg uncertainty principle follows:

∆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.

1.6.6 Measurement and Expectation Values


When we measure an observable Â, we do not obtain a deterministic value, but a probabilistic
outcome. The average value (expectation value) in state |ψ⟩ is:

⟨Â⟩ = ⟨ψ|Â|ψ⟩
This expression reflects the statistical nature of quantum mechanics.

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi


1.6.7 The Bigger Picture: Observables in the Hilbert Space Framework
From an abstract perspective:

• The wave function |ψ⟩ is a vector in a complex Hilbert space.

• Observables are Hermitian linear operators acting on that space.

• Eigenstates of an operator form a basis of the Hilbert space.

• Measurement projects |ψ⟩ onto one of the operator’s eigenstates.

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.

1.7 Linear Algebra, Dirac Notation, and the Postulates of Quan-


tum Mechanics
1.7.1 Linear Algebra – The Language of Quantum Theory
Quantum mechanics is fundamentally a theory of vectors and operators in an abstract vector
space called a Hilbert space. Many of the ideas introduced earlier—wave functions, eigenstates,
operators—are deeply rooted in linear algebra.
Key concepts include:

• 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

where |n⟩ are orthonormal basis vectors.

• Inner Product: Defines the overlap between two states:

⟨ϕ|ψ⟩ ∈ C

• Norm: ⟨ψ|ψ⟩ = 1 ensures total probability is conserved.

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi


• Linear Operators: Observables act as operators on vectors. For example, momentum
operator:
d
p̂|ψ⟩ = −ih̄ |ψ⟩
dx
This mathematical structure allows us to apply powerful theorems (orthogonality, diagonaliza-
tion, spectral decomposition) to quantum systems.

1.7.2 Dirac Notation – Bra-Ket Formalism


Introduced by Paul Dirac, this compact notation elegantly expresses the structure of quantum
mechanics:

• Ket |ψ⟩: Represents a column vector (state).

• Bra ⟨ϕ|: Represents the complex conjugate row vector (dual state).

• Inner Product: ⟨ϕ|ψ⟩ is a complex number.

• Outer Product: |ψ⟩⟨ϕ| is an operator.

• Expectation Value: ⟨ψ|Â|ψ⟩ is the average measured value of observable Â.

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.

1.8 Postulates of Quantum Mechanics – A Conceptual and Ex-


perimental Foundation
The formal structure of quantum mechanics is based on a set of five postulates. These postulates
were not invented arbitrarily; they were distilled from decades of experimental puzzles, theoretical
insights, and logical consistency.
Each postulate reflects a profound shift from classical ideas—where systems have definite tra-
jectories—to a quantum worldview governed by probabilities, wave functions, and operators.

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi


Postulate 1: The Wave Function and the State of a System
A quantum system is completely described by a wave function Ψ(⃗r, t), which depends on spatial
coordinates ⃗r and time t. This wave function encodes all measurable properties of the system.
The probability of finding a particle at position ⃗r at time t is given by:

P (⃗r, t) = |Ψ(⃗r, t)|2


That is, the probability density is Ψ∗ (⃗r, t)Ψ(⃗r, t), and the total probability over all space must
be one:
Z
|Ψ(⃗r, t)|2 d3 r = 1

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.

Postulate 2: Observables Correspond to Hermitian Operators


Every classical observable (like position, momentum, energy) corresponds to a Hermitian operator
 in quantum mechanics. These operators act on the wave function or state vector.
Hermitian operators have real eigenvalues and orthonormal eigenfunctions. Thus, measurable
quantities in quantum theory are the eigenvalues of these operators.
Examples include:

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.

Postulate 3: Measurement and the Eigenvalue Equation


In any measurement of an observable Â, the only values that can be observed are its eigenvalues
a, obtained from the eigenvalue equation:

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:

P (an ) = |⟨an |ψ⟩|2


Even though the pre-measurement wave function may not be an eigenfunction, it can always
be expressed as a superposition of the eigenfunctions of Â. The superposition principle, tested in
countless interference and tunneling experiments, confirms this structure.

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi


Postulate 4: Expectation Value of an Observable
If a system is in a state Ψ, the expectation value (i.e., the average outcome over many repeated
measurements) of the observable  is given by:
Z
⟨Â⟩ = Ψ∗ (⃗r, t)ÂΨ(⃗r, t) d3 r

or equivalently, in Dirac notation:

⟨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.

Postulate 5: Time Evolution via the Schrödinger Equation


The time evolution of the wave function is governed by the time-dependent Schrödinger equation:


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.

Where the Postulates Come From – A Brief Perspective


These postulates were not deduced from pure logic—they arose as a coherent response to:

• 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.

• The need for a mathematically consistent framework, developed independently by Schrödinger,


Heisenberg, and Dirac.

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.

1.9 Measurement in Quantum Mechanics and Collapse of the Wave


Function
1.9.1 The Measurement Problem: A Departure from Classical Thinking
In classical mechanics, measurement is passive—measuring a system merely reveals its pre-existing
state. But in quantum mechanics, measurement plays a fundamentally active role: it not only
reveals an outcome but also changes the state of the system.

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi


This non-classical behavior is encoded in the measurement postulate and is most dramati-
cally illustrated by the phenomenon known as the collapse of the wave function.

1.9.2 States and Measurements in Dirac Notation


Let a quantum system be in a state |ψ⟩, a unit vector in Hilbert space. Let  be a Hermitian
operator corresponding to some observable (e.g., position, momentum, energy). The eigenvalue
equation for  is:

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:

P (an ) = |⟨an |ψ⟩|2


1.9.3 Collapse of the Wave Function


If a measurement yields the outcome ak , then immediately after the measurement, the system is
found in the corresponding eigenstate |ak ⟩:

|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.

1.9.4 Projection Operators and State Update Rule


We can formalize collapse using the projection operator:

P̂k = |ak ⟩⟨ak |


The state of the system after a measurement yielding ak is:

P̂k |ψ⟩ |ak ⟩⟨ak |ψ⟩


|ψ ′ ⟩ = q =p = |ak ⟩
⟨ψ|P̂k |ψ⟩ |⟨ak |ψ⟩|2

This formalism confirms: - The state is projected onto the subspace corresponding to the
eigenvalue. - The denominator ensures the state is re-normalized after projection.

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi


1.9.5 Expectation Values and Repeated Measurements
If no measurement is performed, the state continues to evolve unitarily via the Schrödinger equation.
But if we measure  repeatedly: - The first measurement collapses |ψ⟩ to some eigenstate |ak ⟩.
- Subsequent measurements of  yield the same value ak with probability 1.
This reflects the idempotency of projection:

P̂k2 = P̂k
It also aligns with the idea that eigenstates are stationary states with respect to their own
observable.

1.9.6 Commuting Observables and Simultaneous Measurement


Two observables  and B̂ can be simultaneously measured only if they commute:

[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.

1.9.7 Conceptual and Philosophical Implications


The measurement postulate introduces a deep tension in quantum mechanics: - Unitary evolu-
tion via Schrödinger equation is deterministic and reversible. - Measurement is probabilistic and
irreversible.
This dual character is central to many interpretations of quantum mechanics. It also underlies
paradoxes such as: - Schrödinger’s cat, - The EPR paradox and quantum entanglement, - Delayed
choice and quantum eraser experiments.
Nevertheless, the measurement framework is experimentally verified to exquisite precision. Ev-
ery quantum prediction—interference, tunneling, entanglement—is ultimately validated through
measurement outcomes interpreted via this postulate.

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.

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi


1.10 Interpretations of Quantum Mechanics – Beyond the Copen-
hagen View
1.10.1 Why Do We Need Interpretations?
As we have seen, the formalism of quantum mechanics is mathematically complete and experimen-
tally successful. The wave function evolves smoothly via the Schrödinger equation, and the act of
measurement collapses it into a definite eigenstate with a probabilistically determined outcome.
This picture—embodied in the Copenhagen interpretation—has guided the development of
quantum theory since its inception. But it raises deep philosophical questions:
• What does the wave function really represent?
• Is the collapse real or just a computational tool?
• Does the observer play a fundamental role in physical reality?
• Can we describe the quantum world objectively, or only through measurement?
These questions have inspired multiple interpretations of quantum mechanics, each offering a
different resolution of the measurement problem and a different vision of quantum reality.

1.10.2 The Copenhagen Interpretation (Standard View)


• Developed by Niels Bohr, Werner Heisenberg, and others in the 1920s.
• The wave function ψ is a tool for computing probabilities—not a physical object.
• Measurement causes an instantaneous, non-unitary collapse of the wave function.
• The observer plays a special role—what is real depends on what is observed.
• The boundary between quantum system and classical apparatus is ambiguous but necessary.
This view is operational and pragmatic: “Shut up and calculate.” It remains the most widely
taught interpretation and underlies the standard measurement postulates we presented earlier.

1.10.3 Many-Worlds Interpretation (Everett, 1957)


• Proposed by Hugh Everett III as a way to avoid collapse.
• The wave function ψ is real and universal—it never collapses.
• Every possible outcome of a quantum measurement actually occurs, in a separate “branch”
of the universe.
• The observer becomes entangled with the system; different branches contain different observer
outcomes.
• All evolution is unitary; there is no need for special measurement postulates.
This interpretation is mathematically clean but conceptually extravagant—it implies a vast
multiverse of ever-splitting realities.

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi


1.10.4 Pilot-Wave Theory (de Broglie–Bohm)
• Developed by Louis de Broglie (1927) and refined by David Bohm (1952).
• Particles have definite positions at all times, guided by a hidden “pilot wave” (the wave
function).
• The wave function evolves by the Schrödinger equation and influences particle trajectories.
• No collapse occurs—probabilities arise from ignorance of initial conditions.
• Provides a deterministic but non-local account of quantum phenomena.

While reproducing all predictions of standard quantum mechanics, it introduces hidden variables
and a physically real guiding wave.

1.10.5 Objective Collapse Theories (GRW, Penrose)


• Modify the Schrödinger equation to include spontaneous, random collapses.
• Collapse is a real physical process, not linked to observation.
• Ghirardi–Rimini–Weber (GRW) model introduces collapse events at random times.
• Roger Penrose proposed that gravity may trigger collapse of superpositions.

These models aim to solve the measurement problem dynamically, but often require new physical
constants and lack experimental confirmation so far.

1.10.6 Quantum Bayesianism (QBism)


• The wave function represents the observer’s personal belief about the system.
• Measurement updates this belief using Bayesian probability.
• Emphasizes subjectivity and information—quantum theory is about an agent’s expectations.
• All probabilities are epistemic (knowledge-based), not ontic (reality-based).

QBism offers a radical shift in viewpoint, reframing quantum theory as a user’s guide for making
predictions.

1.10.7 Comparative Overview


Interpretation Collapse? Deterministic? Wave Function is Real?
Copenhagen Yes No No (epistemic)
Many-Worlds No Yes Yes
Pilot-Wave No Yes Yes
Objective Collapse Yes (spontaneous) No Yes
QBism No N/A No (subjective)

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi


Conclusion
All interpretations make the same experimental predictions—so far. They differ not in results, but
in the metaphysical picture they paint of reality. While the Copenhagen interpretation remains the
standard, it is not the only lens through which quantum mechanics can be understood.
As you explore quantum theory more deeply, these interpretations offer rich perspectives on
what the mathematics might really mean—and invite you to reflect not only on particles and fields,
but on the nature of measurement, reality, and knowledge itself.

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi


Chapter 2

Stern–Gerlach Experiment and Spin

2.1 Historical Background: The Mystery of Spectral Lines and a


New Degree of Freedom
In the early 1920s, atomic physics was in a state of rapid development. The Bohr-Sommerfeld
model had explained many aspects of hydrogen spectra, but some puzzling features remained—most
notably, the anomalous Zeeman effect.
The Zeeman effect refers to the splitting of spectral lines in a magnetic field. While the normal
Zeeman effect could be explained by orbital angular momentum and classical electromagnetism,
the anomalous version—which showed additional fine structure splitting—remained mysterious.
Physicists began to suspect that atoms contained an additional internal degree of freedom. This
eventually led to the idea of spin, but the path was neither simple nor immediate.

2.2 Ralph Kronig’s Insight – A Rejected Proposal


In early 1925, a young German physicist named Ralph Kronig, who had recently earned his PhD,
visited Wolfgang Pauli in Hamburg. Kronig had just attended a talk by Niels Bohr, where the
limitations of the Bohr-Sommerfeld model and the unexplained fine-structure lines were discussed.
Inspired by these issues, Kronig proposed a bold idea:

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.

Pauli’s Sharp Dismissal


To Kronig’s surprise, Pauli harshly criticized the idea, calling it ”nonsense.” He argued:

• A spinning electron would require surface speeds faster than the speed of light to match
observed magnetic moments.

• There was no classical justification for such internal degrees of freedom.

25
• The proposal lacked a solid theoretical foundation at the time.

Disheartened, Kronig withdrew the idea and never published it.


Ironically, just months later, two Dutch graduate students would propose the exact same
idea—this time with better luck.

2.3 Uhlenbeck and Goudsmit – The Bold Leap to Spin


Later that same year (1925), two physics students at the University of Leiden—George Uhlen-
beck and Samuel Goudsmit—were studying spectral fine structure under the guidance of Paul
Ehrenfest, a prominent theoretical physicist.
Like Kronig, they were intrigued by the unexplained patterns in atomic spectra. Through
private discussions and growing frustration with existing models, they independently proposed:

Electrons possess an intrinsic angular momentum, or “spin”, of fixed magnitude.



They assigned the spin an angular momentum of S = 2 and an associated magnetic moment
of:

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.

Ehrenfest’s Crucial Intervention


Uhlenbeck and Goudsmit approached their advisor Paul Ehrenfest with their idea that electrons
possess intrinsic spin. At first, Ehrenfest was skeptical—after all, the idea of a point particle like
the electron “spinning” contradicted classical expectations, especially from relativity.
However, after discussing it with them, Ehrenfest recognized their bold thinking and the poten-
tial explanatory power behind the idea. He encouraged them to write up their result quickly and
told them:

“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:

“You are both young enough to be able to afford a stupidity.”

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.

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi


2.4 Aftermath and Legacy
Though initially controversial, their idea gained support when:

• Dirac’s relativistic wave equation (1928) derived the electron’s spin 1


2 and its magnetic moment
from first principles.

• Stern–Gerlach experiments confirmed quantized angular momentum projection values.

• 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.

2.5 Conceptual Leap: Spin is Not Classical Rotation


Though it was motivated by imagining the electron “spinning,” we now understand that:

• Spin is an intrinsic quantum property, not arising from spatial rotation.

• It is described by its own set of algebraic rules and operators, distinct from orbital angular
momentum.

• Spin operators obey the SU(2) algebra:

[Ŝx , Ŝy ] = ih̄Ŝz , etc.

• Eigenvalues of Ŝ 2 are s(s + 1)h̄2 , and for electrons, s = 21 .

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.

2.5.1 Electron Does Not Spin — The Failure of Classical Analogy


At first glance, the idea of electron spin evokes an image of a tiny charged sphere rotating about
its axis—much like a spinning planet or a gyroscope. This analogy, however, quickly breaks down
when examined through the lens of classical physics.

The Classical Argument


In classical mechanics, the angular momentum S of a rigid rotating body is given by:

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:

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi


q
µS = S
2m
Now consider a uniformly charged solid sphere of radius a and mass m, rotating fast enough to
produce the known spin angular momentum S = h̄2 . Its surface speed would need to satisfy:

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.

The Quantum View of Spin


Quantum mechanics provides a consistent framework in which:
• Spin is not associated with spatial motion.
• The spin angular momentum Ŝ is an intrinsic, non-classical property of the particle.
• For an electron, the magnitude and projection values are:
p 1 h̄
S = s(s + 1)h̄, s = , Sz = ±
2 2
This quantization has no classical analogue. Furthermore, spin cannot be described by a wave
function alone; instead, it requires more general mathematical objects such as spinors and operators
acting in Hilbert space.
Despite this, spin still leads to observable physical effects. For example, an electron’s magnetic
moment due to spin is:
e
µS = −g S, with g ≈ 2
2me

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.

2.6 The Stern–Gerlach Experiment: Quantization of Angular Mo-


mentum
Historical Background
In 1922, Otto Stern and Walther Gerlach conducted an experiment to test the hypothesis of space
quantization—the idea that atomic-scale angular momentum components are quantized. They di-
rected a beam of silver atoms through a non-uniform magnetic field and observed the deflection

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi


Figure 2.1: Schematic of the Stern–Gerlach experiment setup: A beam of silver atoms passes
through an inhomogeneous magnetic field, resulting in discrete deflections corresponding to quan-
tized spin states.

pattern on a detector screen. Contrary to classical expectations of a continuous distribution, the


beam split into discrete spots, providing direct evidence of quantized angular momentum projec-
tions.

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.

Classical vs. Quantum Expectations


Classically, one would expect the magnetic moments of atoms to be oriented randomly, leading
to a continuous distribution of deflections on the detection screen. However, the experiment re-
vealed discrete spots, indicating that the magnetic moment—and thus the angular momentum—is
quantized.

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi


Mathematical Framework
The force on a magnetic moment µ in a magnetic field gradient is given by:

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:

• gs ≈ 2 is the electron g-factor,

• µB = eh̄
2me is the Bohr magneton,

• ms = ± 21 is the spin magnetic quantum number.

Substituting, the force becomes:

∂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.

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi


Quantum Mechanical Interpretation
The discrete splitting observed in the Stern–Gerlach experiment cannot be explained by classical
physics. Quantum mechanically, the spin angular momentum operator Ŝz has eigenvalues:

h̄ h̄
Ŝz ↑ = + ↑, Ŝz ↓ = − ↓
2 2
The magnetic moment operator is proportional to the spin operator:

µ̂z = −gs µB Ŝz /h̄


Therefore, the measurement of the atom’s position on the detector corresponds to a measure-
ment of its spin state along the z-axis, collapsing the spin state into one of the eigenstates ↑ or
↓.

Significance and Impact


The Stern–Gerlach experiment provided the first direct evidence of quantum spin and the quan-
tization of angular momentum. It demonstrated that particles such as electrons possess intrinsic
angular momentum that can take on only discrete values. This experiment was pivotal in the de-
velopment of quantum mechanics and has profound implications in fields like quantum computing
and magnetic resonance imaging.

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.

ISAMP Workshop on Quantum Computation: Sugam Kumar, IUAC, New Delhi

You might also like