You are on page 1of 182

Quantum Mechanics

An Introductory Framework

PDF generated using the open source mwlib toolkit. See http://code.pediapress.com/ for more information.
PDF generated at: Tue, 12 Apr 2011 09:51:09 UTC

Contents
Articles
1. Introductory Principles
History of Quantum Mechanics

1
1

Basic Concepts of Quantum Mechanics

10

Introduction to Quantum Mechanics

17

2. The Quantum Theories

34

Old Quantum Theory

34

Quantum Mechanics after 1925

41

3. The Interpretation of Quantum Mechanics

58

Interpretations of Quantum Mechanics

58

The Copenhagen Interpretation

70

4. Einstein's Objections

76

Principle of Locality

76

EPR Paradox

79

Bell's Theorem

88

5. Schrdinger's Objections
Schrdinger's Cat

6. Measurement Problems

101
101
107

The Measurement Problem

107

Measurement in Quantum Mechanics

109

7. Advanced Concepts

117

Quantum Number

117

Quantum Information

121

Quantum Statistical Mechanics

123

8. Advanced Topics

126

Quantum Field Theory

126

String Theory

140

Quantum Gravity

157

Appendix

166

Quantum

166

Quantum state

168

References
Article Sources and Contributors

174

Image Sources, Licenses and Contributors

178

Article Licenses
License

179

1. Introductory Principles
History of Quantum Mechanics
The history of quantum mechanics, as it
interlaces with the history of quantum
chemistry, began essentially with a number
of different scientific discoveries: the 1838
discovery of cathode rays by Michael
Faraday; the 1859-1860 winter statement of
the black body radiation problem by Gustav
Kirchhoff; the 1877 suggestion by Ludwig
Boltzmann that the energy states of a
physical system could be discrete; and the
1900 quantum hypothesis by Max Planck
that any energy-radiating atomic system can
theoretically be divided into a number of
discrete "energy elements" (epsilon) such
that each of these energy elements is
proportional to the frequency with which
each of them individually radiate energy, as
defined by the following formula:

Niels Bohr's 1913 quantum model of the atom, which incorporated an


explanation of Johannes Rydberg's 1888 formula, Max Planck's 1900 quantum
hypothesis, i.e. that atomic energy radiators have discrete energy values ( = h),
J.J.Thomson's 1904 plum pudding model, Albert Einstein's 1905 light quanta
postulate, and Ernest Rutherford's 1907 discovery of the atomic nucleus.

where h is a numerical value called Planck's constant. Then, in 1905, to explain the photoelectric effect (1839),
which demonstrated that shining light of particular wavelengths on certain materials causes electrons to be ejected
from those materials, Albert Einstein postulated, as based on Planck's quantum hypothesis, that light itself consists of
individual quantum particles, which later came to be called photons (1926). The phrase "quantum mechanics" was
first used in Max Born's 1924 paper "Zur Quantenmechanik". In the years to follow, this theoretical basis slowly
began to be applied to chemical structure, reactivity, and bonding.

Overview
In short, in 1900, German physicist Max Planck introduced the idea that energy is quantized, to derive a formula for
the observed frequency dependence of the energy emitted by a black body. In 1905, Einstein explained the
photoelectric effect by postulating that light, or more generally all electromagnetic radiation, can be divided into a
finite number of "energy quanta" that are localized points in space. From the introduction section of his March 1905
quantum paper, On a heuristic viewpoint concerning the emission and transformation of light, Einstein states:
According to the assumption to be contemplated here, when a light ray is spreading from a point, the energy is
not distributed continuously over ever-increasing spaces, but consists of a finite number of energy quanta that

History of Quantum Mechanics


are localized in points in space, move without dividing, and can be absorbed or generated only as a whole.
This statement has been called the most revolutionary sentence written by a physicist of the twentieth century.[1]
These energy quanta later came to be called "photons", a term introduced by Gilbert N. Lewis in 1926. The idea that
each photon had to consist of energy in terms of quanta was a remarkable achievement; it effectively solved the
problem of black body radiation attaining infinite energy, which occurred in theory if light were to be explained only
in terms of waves. In 1913, Bohr explained the spectral lines of the hydrogen atom, again by using quantization, in
his paper of July 1913 On the Constitution of Atoms and Molecules.
These theories, though successful, were strictly phenomenological: during this time, there was no rigorous
justification for quantization, aside, perhaps, from Henri Poincar's discussion of Planck's theory in his 1912 paper
Sur la thorie des quanta.[2] [3] They are collectively known as the old quantum theory.
The phrase "quantum physics" was first used in Johnston's Planck's Universe in Light of Modern Physics (1931).
In 1924, the French physicist Louis de Broglie put forward his theory of matter waves by stating that particles can
exhibit wave characteristics and vice versa. This theory was for a single particle and derived from special relativity
theory. Building on de Broglie's approach, modern quantum mechanics was born in 1925, when the German
physicists Werner Heisenberg and Max Born developed matrix mechanics and the Austrian physicist Erwin
Schrdinger invented wave mechanics and the non-relativistic Schrdinger equation as an approximation to the
generalised case of de Broglie's theory.[4] Schrdinger subsequently showed that the two approaches were
equivalent.
Heisenberg formulated his uncertainty principle in 1927, and the Copenhagen interpretation started to take shape at
about the same time. Starting around 1927, Paul Dirac began the process of unifying quantum mechanics with
special relativity by proposing the Dirac equation for the electron. The Dirac equation achieves the relativistic
description of the wavefunction of an electron that Schrdinger failed to obtain. It predicts electron spin and led
Dirac to predict the existence of the positron. He also pioneered the use of operator theory, including the influential
bra-ket notation, as described in his famous 1930 textbook. During the same period, Hungarian polymath John von
Neumann formulated the rigorous mathematical basis for quantum mechanics as the theory of linear operators on
Hilbert spaces, as described in his likewise famous 1932 textbook. These, like many other works from the founding
period still stand, and remain widely used.
The field of quantum chemistry was pioneered by physicists Walter Heitler and Fritz London, who published a study
of the covalent bond of the hydrogen molecule in 1927. Quantum chemistry was subsequently developed by a large
number of workers, including the American theoretical chemist Linus Pauling at Caltech, and John C. Slater into
various theories such as Molecular Orbital Theory or Valence Theory.
Beginning in 1927, attempts were made to apply quantum mechanics to fields rather than single particles, resulting
in what are known as quantum field theories. Early workers in this area included P.A.M. Dirac, W. Pauli, V.
Weisskopf, and P. Jordan. This area of research culminated in the formulation of quantum electrodynamics by R.P.
Feynman, F. Dyson, J. Schwinger, and S.I. Tomonaga during the 1940s. Quantum electrodynamics is a quantum
theory of electrons, positrons, and the electromagnetic field, and served as a role model for subsequent quantum field
theories. The theory of quantum chromodynamics was formulated beginning in the early 1960s. The theory as we
know it today was formulated by Politzer, Gross and Wilczek in 1975. Building on pioneering work by Schwinger,
Higgs and Goldstone, the physicists Glashow, Weinberg and Salam independently showed how the weak nuclear
force and quantum electrodynamics could be merged into a single electroweak force, for which they received the
1979 Nobel Prize in Physics.

History of Quantum Mechanics

Timeline
The following timeline shows the key steps and contributors in the precursory development of quantum mechanics
and quantum chemistry:
Date

Person

Contribution

1900

Max Planck

To explain black body radiation (1862), he suggested that electromagnetic energy could only be emitted in
quantized form, i.e. the energy could only be a multiple of an elementary unit E = h, where h is Planck's constant
and is the frequency of the radiation.

1901

Frederick Soddy and


Ernest Rutherford

Discovered nuclear transmutation when they found that radioactive thorium was converting itself into radium
through a process of nuclear decay.

1902

Gilbert N. Lewis

To explain the octet rule (1893), he developed the cubical atom theory in which electrons in the form of dots
were positioned at the corner of a cube and suggested that single, double, or triple bonds result when two atoms
are held together by multiple pairs of electrons (one pair for each bond) located between the two atoms (1916).

1904

Richard Abegg

Noted the pattern that the numerical difference between the maximum positive valence, such as +6 for H2SO4, and
the maximum negative valence, such as -2 for H2S, of an element tends to be eight (Abegg's rule).

1905

Albert Einstein

Explained the photoelectric effect (1839), i.e. that shining light on certain materials can function to eject electrons
from the material, he postulated, as based on Plancks quantum hypothesis (1900), that light itself consists of
individual quantum particles (photons).

1905

Albert Einstein

First to explain the effects of Brownian motion as caused by the kinetic energy (i.e., movement) of atoms, which
was subsequently, experimentally verified by Jean Baptiste Perrin, thereby settling the century-long dispute about
the validity of John Dalton's atomic theory.

1905

Albert Einstein

Publishes his Special Theory of Relativity.

1905

Albert Einstein

Determines the equivalence of matter and energy.

1907

Ernest Rutherford

To test the plum pudding model (1904), he fired, positively-charged, alpha particles at gold foil and noticed that
some bounced back thus showing that an atom has a small-sized positively charged atomic nucleus at its center.

1909

Geoffrey Ingram Taylor Demonstrated that interference patterns of light were generated even when the light energy introduced consisted
of only one photon. This discovery of the wave-particle duality of matter and energy was fundamental to the later
development of quantum field theory.

1909
and
1916

Albert Einstein

Showed that, if Planck's law of black-body radiation is accepted, the energy quanta must also carry momentum p
= h / , making them full-fledged particles.

1911

Lise Meitner and Otto


Hahn

Performed an experiment that showed that the energies of electrons emitted by beta decay had a continuous rather
than discrete spectrum. This was in apparent contradiction to the law of conservation of energy, as it appeared that
energy was lost in the beta decay process. A second problem was that the spin of the Nitrogen-14 atom was 1, in
contradiction to the Rutherford prediction of . These anomalies were later explained by the discoveries of the
neutrino and the neutron.

1912

Victor Hess

Discovers the existence of cosmic radiation.

1912

Henri Poincar

Published an influential mathematical argument in support of the essential nature of energy quanta.

1913

Robert Andrews
Millikan

Publishes the results of his "oil drop" experiment, in which he precisely determines the electric charge of the
electron. Determination of the fundamental unit of electric charge made it possible to calculate the Avogadro
constant (which is the number of atoms or molecules in one mole of any substance) and thereby to determine the
atomic weight of the atoms of each element.

1913

Johannes Stark and


Antonino Lo Surdo

Independently discovered the shifting and splitting of the spectral lines of atoms and molecules due to the
presence of the light source in an external static electric field.

1913

Niels Bohr

To explain the Rydberg formula (1888), which correctly modeled the light emission spectra of atomic hydrogen,
Bohr hypothesized that negatively charged electrons revolve around a positively charged nucleus at certain fixed
quantum distances and that each of these spherical orbits has a specific energy associated with it such that
electron movements between orbits requires quantum emissions or absorptions of energy.

[2] [3]

History of Quantum Mechanics

1915

Albert Einstein

First presents to the Prussian Academy of Science what are now known as the Einstein field equations. These
equations specify how the geometry of space and time is influenced by whatever matter is present, and form the
core of Einstein's General Theory of Relativity. Although this theory is not directly applicable to quantum
mechanics, theorists of quantum gravity seek to reconcile them.

1916

Arnold Sommerfeld

To account for the Zeeman effect (1896), i.e. that atomic absorption or emission spectral lines change when the
light source is subjected to a magnetic field, he suggested there might be elliptical orbits in atoms in addition to
spherical orbits.

1918

Ernest Rutherford

Noticed that, when alpha particles were shot into nitrogen gas, his scintillation detectors showed the signatures of
hydrogen nuclei. Rutherford determined that the only place this hydrogen could have come from was the nitrogen,
and therefore nitrogen must contain hydrogen nuclei. He thus suggested that the hydrogen nucleus, which was
known to have an atomic number of 1, was an elementary particle, which he decided must be the protons
hypothesized by Eugen Goldstein.

1919

Irving Langmuir

Building on the work of Lewis (1916), he coined the term "covalence" and postulated that coordinate covalent
bonds occur when two electrons of a pair of atoms come from both atoms and are equally shared by them, thus
explaining the fundamental nature of chemical bonding and molecular chemistry.

1922

Arthur Compton

Found that X-ray wavelengths increase due to scattering of the radiant energy by "free electrons". The scattered
quanta have less energy than the quanta of the original ray. This discovery, known as the "Compton effect," or
"Compton scattering" demonstrates the "particle" concept of electromagnetic radiation.

1922

Otto Stern and Walther


Gerlach

Stern-Gerlach experiment detects discrete values of angular momentum for atoms in the ground state passing
through an inhomogeneous magnetic field leading to the discovery of the spin of the electron.

1923

Louis De Broglie

Postulated that electrons in motion are associated with waves the lengths of which are given by Plancks constant
h divided by the momentum of the mv = p of the electron: = h / mv = h / p.

1924

Satyendra Nath Bose

His work on quantum mechanics provides the foundation for Bose-Einstein statistics, the theory of the
Bose-Einstein condensate, and the discovery of the boson.

1925

Friedrich Hund

Outlined the rule of maximum multiplicity which states that when electrons are added successively to an atom
as many levels or orbits are singly occupied as possible before any pairing of electrons with opposite spin occurs
and made the distinction that the inner electrons in molecules remained in atomic orbitals and only the valence
electrons needed to be in molecular orbitals involving both nuclei.

1925

Werner Heisenberg

Developed the matrix mechanics formulation of Quantum Mechanics.

1925

Wolfgang Pauli

Outlined the Pauli exclusion principle which states that no two identical fermions may occupy the same
quantum state simultaneously.

1926

Gilbert Lewis

Coined the term photon, which he derived from the Greek word for light, (transliterated phs).

1926

Erwin Schrdinger

Used De Broglies electron wave postulate (1924) to develop a wave equation that represents mathematically the
distribution of a charge of an electron distributed through space, being spherically symmetric or prominent in
certain directions, i.e. directed valence bonds, which gave the correct values for spectral lines of the hydrogen
atom.

1927

Charles Drummond
Ellis (along with James
Chadwick and
colleagues)

Finally established clearly that the beta decay spectrum is in fact continuous and not discrete, posing a problem
that would later by solved by theorizing (and later discovering) the existence of the neutrino.

1927

Walter Heitler

Used Schrdingers wave equation (1926) to show how two hydrogen atom wavefunctions join together, with
plus, minus, and exchange terms, to form a covalent bond.

1927

Robert Mulliken

In 1927 Mulliken worked, in coordination with Hund, to develop a molecular orbital theory where electrons are
assigned to states that extend over an entire molecule and, in 1932, introduced many new molecular orbital
terminologies, such as bond, bond, and bond.

1928

Paul Dirac

In the Dirac equations, Paul Dirac integrated the principle of special relativity with quantum electrodynamics and
hypothesized the existence of the positron.

History of Quantum Mechanics

1928

Linus Pauling

Outlined the nature of the chemical bond in which he used Heitlers quantum mechanical covalent bond model
(1927) to outline the quantum mechanical basis for all types of molecular structure and bonding and suggested
that different types of bonds in molecules can become equalized by rapid shifting of electrons, a process called
resonance (1931), such that resonance hybrids contain contributions from the different possible electronic
configurations.

1929

John Lennard-Jones

Introduced the linear combination of atomic orbitals approximation for the calculation of molecular orbitals.

1930

Wolfgang Pauli

In a famous letter, Pauli suggested that, in addition to electrons and protons, atoms also contained an extremely
light neutral particle which he called the "neutron." He suggested that this "neutron" was also emitted during beta
decay and had simply not yet been observed. Later it was determined that this particle was actually the almost
massless neutrino.

1931

Walther Bothe and


Herbert Becker

Found that if the very energetic alpha particles emitted from polonium fell on certain light elements, specifically
beryllium, boron, or lithium, an unusually penetrating radiation was produced. At first this radiation was thought
to be gamma radiation, although it was more penetrating than any gamma rays known, and the details of
experimental results were very difficult to interpret on this basis. Some scientists began to hypothesize the
possible existence of another fundamental, atomic particle.

1931

Enrico Fermi

Renamed Pauli's "neutron" to neutrino to distinguish it from the then-hypothetical possibility of a much more
massive neutron.

1932

Irne Joliot-Curie and


Frdric Joliot

Showed that if the unknown radiation generated by alpha particles fell on paraffin or any other
hydrogen-containing compound, it ejected protons of very high energy. This was not in itself inconsistent with the
proposed gamma ray nature of the new radiation, but detailed quantitative analysis of the data became
increasingly difficult to reconcile with such a hypothesis.

1932

James Chadwick

Performed a series of experiments showing that the gamma ray hypothesis for the unknown radiation produced by
alpha particles was untenable, and that the new particles must be the neutrons hypothesized by Enrico Fermi.
Chadwick suggested that, in fact, the new radiation consisted of uncharged particles of approximately the same
mass as the proton, and he performed a series of experiments verifying his suggestion.

1932

Werner Heisenberg

Applied perturbation theory to the two-electron problem and showed how resonance arising from electron
exchange could explain exchange forces.

1932

Mark Oliphant

Building upon the nuclear transmutation experiments of Ernest Rutherford done a few years earlier, fusion of light
nuclei (hydrogen isotopes) was first observed by Oliphant in 1932. The steps of the main cycle of nuclear fusion
in stars were subsequently worked out by Hans Bethe throughout the remainder of that decade.

1932

Carl D. Anderson

Experimentally proves the existence of the positron.

1933

Le Szilrd

First theorized the concept of a nuclear chain reaction. He filed a patent for his idea of a simple nuclear reactor the
following year.

1934

Enrico Fermi

Published a very successful model of beta decay in which neutrinos were produced.

1934

Enrico Fermi

Studies the effects of bombarding uranium isotopes with neutrons.

1934

N.N.Semyonov

Develops the total quantitative chain chemical reaction theory. The idea of the chain reaction, developed by
Semyonov, is the basis of various high technologies using the incineration of gas mixtures. The idea was also used
for the description of the nuclear reaction.

1935

Hideki Yukawa

Published his hypothesis of the Yukawa Potential and predicted the existence of the pion, stating that such a
potential arises from the exchange of a massive scalar field, such as would be found in the field of the pion. Prior
to Yukawa's paper, it was believed that the scalar fields of the fundamental forces necessitated massless particles.

1936

Carl D. Anderson

Discovered muons while he studied cosmic radiation.

1937

Carl Anderson

Experimentally proves the existence of the pion.

1938

Charles Coulson

Made the first accurate calculation of a molecular orbital wavefunction with the hydrogen molecule.

1938

Otto Hahn, Fritz


Strassmann, Lise
Meitner, and Otto
Robert Frisch

Hahn and Strassmann sent a manuscript to Naturwissenschaften reporting they had detected the element barium
after bombarding uranium with neutrons. Simultaneously, they communicated these results to Meitner. Meitner,
and her nephew Frisch, correctly interpreted these results as being nuclear fission. Frisch confirmed this
experimentally on 13 January 1939.

History of Quantum Mechanics

1939

Le Szilrd and Enrico


Fermi

Discovered neutron multiplication in uranium, proving that a chain reaction was indeed possible.

1942

Kan-Chang Wang

First proposed the use of beta capture to experimentally detect neutrinos.

1942

Enrico Fermi

Created the first artificial self-sustaining nuclear chain reaction, called Chicago Pile-1 (CP-1), in a racquets court
below the bleachers of Stagg Field at the University of Chicago on December 2, 1942.

1945

Manhattan Project

First nuclear fission explosion.

1947

G. D. Rochester and C.
C. Butler

Published two cloud chamber photographs of cosmic ray-induced events, one showing what appeared to be a
neutral particle decaying into two charged pions, and one which appeared to be a charged particle decaying into a
charged pion and something neutral. The estimated mass of the new particles was very rough, about half a
proton's mass. More examples of these "V-particles" were slow in coming, and they were soon given the name
kaons.

1948

Sin-Itiro Tomonaga and Independently introduced perturbative renormalization as a method of correcting the original Lagrangian of a
Julian Schwinger
quantum field theory so as to eliminate an infinite series of counterterms that would otherwise result.

1948

Richard Feynman

Stated the path integral formulation of quantum mechanics.

1949

Freeman Dyson

Determined the equivalence of the formulations of quantum electrodynamics that existed by that time Richard
Feynman's diagrammatic path integral formulation and the operator method developed by Julian Schwinger and
[5]
Sin-Itiro Tomonaga. A by-product of that demonstration was the invention of the Dyson series.

1951

Clemens C. J. Roothaan Derived the Roothaan-Hall equations, putting rigorous molecular orbital methods on a firm basis.
and George G. Hall

1952

Manhattan Project

First explosion of a thermonuclear bomb.

1954

Chen Ning Yang and


Robert Mills

Derived a gauge theory for nonabelian groups, leading to the successful formulation of both electroweak
unification and quantum chromodynamics.

1955
and
1956

Murray Gell-Mann and


Kazuhiko Nishijima

Independently derived the Gell-MannNishijima formula, which relates the baryon number B, the strangeness S,
and the isospin Iz of hadrons to the charge Q, eventually leading to the systematic categorization of hadrons and,
ultimately, the Quark Model of hadron composition.

1956

P. Kuroda

Predicted that self-sustaining nuclear chain reactions should occur in natural uranium deposits.

1956

Clyde L. Cowan and


Frederick Reines

Experimentally proved the existence of the neutrino.

1957

William Alfred Fowler, In their 1957 paper Synthesis of the Elements in Stars, they explained how the abundances of essentially all but
Margaret Burbidge,
the lightest chemical elements could be explained by the process of nucleosynthesis in stars.
Geoffrey Burbidge, and
Fred Hoyle

1961

Clauss Jnsson

Performed Young's double-slit experiment (1909) for the first time with particles other than photons by using
electrons and with similar results, confirming that massive particles also behaved according to the wave-particle
duality that is a fundamental principle of quantum field theory.

1961

Sheldon Lee Glashow

Extended the electroweak unification models developed by Julian Schwinger by including a short range neutral
current, the Z0. The resulting symmetry structure that Glashow proposed, SU(2) X U(1), formed the basis of the
accepted theory of the electroweak interactions.

1962

Leon M. Lederman,
Melvin Schwartz and
Jack Steinberger

Showed that more than one type of neutrino exists by detecting interactions of the muon neutrino (already
hypothesised with the name "neutretto")

1962

Murray Gell-Mann and


Yuval Ne'eman

Independently classified the hadrons according to a system that Gell-Mann called the "Eightfold Way," and which
ultimately led to the quark model (1964) of hadron composition.

1962

Jeffrey Goldstone,
Yoichiro Nambu,
Abdus Salam, and
Steven Weinberg

Developed what is now known as Goldstone's Theorem, in which it was proved that, if there is continuous
symmetry transformation under which the Lagrangian is invariant, then either the vacuum state is also invariant
under the transformation, or there must exist spinless particles of zero mass, thereafter called Nambu-Goldstone
bosons.

History of Quantum Mechanics

1963

Nicola Cabibbo

Developed the mathematical matrix by which the first two (and ultimately three) generations of quarks could be
predicted.

1964

Murray Gell-Mann and


George Zweig

Independently proposed the quark model of hadrons, predicting the arbitrarily named up, down, and strange
quarks. Gell-Mann is credited with coining the term "quark," which he found in James Joyce's book Finnegans
Wake.

1964

Franois Englert,
Robert Brout, Peter
Higgs, Gerald Guralnik,
C. R. Hagen, and Tom
Kibble

Postulated that a fundamental quantum field, now called the Higgs field, permeates space and, by way of the
Higgs mechanism, provides mass to all the elementary subatomic particles that interact with it. While the Higgs
field is postulated to confer mass on quarks and leptons, it represents only a tiny portion of the masses of other
subatomic particles, such as protons and neutrons. In these, gluons that bind quarks together confer most of the
particle mass. The Higgs mechanism, which gives mass to vector bosons, was theorized in 1964 by Franois
Englert and Robert Brout. In October of the same year, Peter Higgs, working from the ideas of Philip Anderson
reached the same conclusions; and, independently, by Gerald Guralnik, C. R. Hagen, and Tom Kibble, who
worked out the results by the spring of 1963.

1964

Sheldon Lee Glashow


and James Bjorken

Predicted the existence of the charm quark. The addition was proposed because it allowed for a better description
of the weak interaction (the mechanism that allows quarks and other particles to decay), equalized the number of
known quarks with the number of known leptons, and implied a mass formula that correctly reproduced the
masses of the known mesons.

1967

Steven Weinberg and


Abdus Salam

Published a paper in which he described Yang-Mills Theory using the SU(2) X U(1) supersymmetry group,
thereby yielding a mass for the W particle of the Weak Interaction via spontaneous symmetry breaking.

1968

Stanford University

Deep inelastic scattering experiments at the Stanford Linear Accelerator Center (SLAC) showed that the proton
contained much smaller, point-like objects and was therefore not an elementary particle. Physicists at the time
were reluctant to identify these objects with quarks, instead calling them "partons" a term coined by Richard
Feynman. The objects that were observed at SLAC would later be identified as up and down quarks. Nevertheless,
"parton" remains in use as a collective term for the constituents of hadrons (quarks, antiquarks, and gluons). The
strange quark's existence was indirectly validated by the SLAC's scattering experiments: not only was it a
necessary component of Gell-Mann and Zweig's three-quark model, but it provided an explanation for the kaon
(K) and pion () hadrons discovered in cosmic rays in 1947.

1970

Sheldon Lee Glashow,


John Iliopoulos and
Luciano Maiani

Presented further reasoning for the existence of the as-yet undiscovered charm quark.

1971

Martinus J. G. Veltman
and Gerardus 't Hooft

Showed that, if the symmetries of Yang-Mills Theory were to be broken according to the method suggested by
Peter Higgs, then Yang-Mills theory can be renormalized. The renormalization of Yang-Mills Theory predicted
the existence of a massless particle, called the gluon, which could explain the nuclear Strong Force. It also
explained how the particles of the Weak Interaction, the W and Z bosons, obtained their mass via spontaneous
symmetry breaking and the Yukawa interaction.

1972

Francis Perrin

Discovered the existence of "natural nuclear fission reactors" in uranium deposits in Oklo, Gabon, where analysis
of isotope ratios demonstrated that self-sustaining, nuclear chain reactions had occurred. The conditions under
which a natural nuclear reactor could exist were predicted in 1956 by P. Kuroda.

1973

Makoto Kobayashi and


Toshihide Maskawa

Noted that the experimental observation of CP violation could be explained if an additional pair of quarks existed.
The two new quarks were eventually named top and bottom.

1974

Pier Giorgio Merli

Performed Young's double-slit experiment (1909) using a single electron with similar results, confirming the
existence of quantum fields for massive particles.

1974

Burton Richter and


Samuel Ting

Charm quarks were produced almost simultaneously by two teams in November 1974 (see November Revolution)
one at SLAC under Burton Richter, and one at Brookhaven National Laboratory under Samuel Ting. The
charm quarks were observed bound with charm antiquarks in mesons. The two discovering parties had
independently assigned the discovered meson two different symbols, J and ; thus, it became formally known as
the J/ meson. The discovery finally convinced the physics community of the quark model's validity.

1975

Martin Lewis Perl

With his colleagues at the SLACLBL group, he detected the tau in a series of experiments between 1974 and
1977.

1977

Leon Lederman

Observed the bottom quark with his team at Fermilab. This discovery was a strong indicator of the top quark's
existence: without the top quark, the bottom quark would have been without a partner that was required by the
mathematics of the theory.

History of Quantum Mechanics

1983

Carlo Rubbia and


Simon van der Meer

Unambiguous signals of W particles were seen in January 1983 during a series of experiments conducted by Carlo
Rubbia and Simon van der Meer at the Super Proton Synchrotron. The actual experiments were called UA1 (led
by Rubbia) and UA2 (led by Peter Jenni), and were the collaborative effort of many people. Simon van der Meer
was the driving force on the use of the accelerator. UA1 and UA2 found the Z particle a few months later, in May
1983.

1995

Fermilab

The top quark was finally observed by a team at Fermilab. It had a mass much greater than had been previously
expected almost as great as a gold atom.

1995

Eric Cornell, Carl


Wieman and Wolfgang
Ketterle

The first "pure" BoseEinstein condensate was created by Eric Cornell, Carl Wieman, and co-workers at JILA.
They did this by cooling a dilute vapor consisting of approximately two thousand rubidium-87 atoms to below
170 nK using a combination of laser cooling and magnetic evaporative cooling. About four months later, an
independent effort led by Wolfgang Ketterle at MIT created a condensate made of sodium-23. Ketterle's
condensate had about a hundred times more atoms, allowing him to obtain several important results such as the
observation of quantum mechanical interference between two different condensates.

2000

CERN

CERN scientists publish experimental results in which they claim to have observed indirect evidence of the
existence of a quark-gluon plasma, which they call a "new state of matter."

Founding experiments

Thomas Young's double-slit experiment demonstrating the wave nature of light (c1805)
Henri Becquerel discovers radioactivity (1896)
J. J. Thomson's cathode ray tube experiments (discovers the electron and its negative charge) (1897)
The study of black body radiation between 1850 and 1900, which could not be explained without quantum
concepts.
The photoelectric effect: Einstein explained this in 1905 (and later received a Nobel prize for it) using the concept
of photons, particles of light with quantized energy
Robert Millikan's oil-drop experiment, which showed that electric charge occurs as quanta (whole units), (1909)
Ernest Rutherford's gold foil experiment disproved the plum pudding model of the atom which suggested that the
mass and positive charge of the atom are almost uniformly distributed. (1911)
Otto Stern and Walther Gerlach conduct the Stern-Gerlach experiment, which demonstrates the quantized nature
of particle spin (1920)

Clinton Davisson and Lester Germer demonstrate the wave nature of the electron[6] in the Electron diffraction
experiment (1927)
Clyde L. Cowan and Frederick Reines confirm the existence of the neutrino in the neutrino experiment (1955)
Claus Jnsson`s double-slit experiment with electrons (1961)
The Quantum Hall effect, discovered in 1980 by Klaus von Klitzing. The quantized version of the Hall effect has
allowed for the definition of a new practical standard for electrical resistance and for an extremely precise
independent determination of the fine structure constant.
The experimental verification of quantum entanglement by Alain Aspect in 1982.

History of Quantum Mechanics

References
[1] Folsing, Albrecht (1997), Albert Einstein: A Biography, trans. Ewald Osers, Viking
[2] McCormmach, Russell (Spring, 1967), "Henri Poincar and the Quantum Theory", Isis 58 (1): 3755, doi:10.1086/350182
[3] Irons, F. E. (August, 2001), "Poincar's 191112 proof of quantum discontinuity interpreted as applying to atoms", American Journal of
Physics 69 (8): 879884, doi:10.1119/1.1356056
[4] Hanle, P.A. (December 1977), "Erwin Schrodinger's Reaction to Louis de Broglie's Thesis on the Quantum Theory.", Isis 68 (4): 606609,
doi:10.1086/351880
[5] F. J. Dyson, Phys. Rev. 75, 486, 1736 (1949)
[6] The Davisson-Germer experiment, which demonstrates the wave nature of the electron (http:/ / hyperphysics. phy-astr. gsu. edu/ hbase/
quantum/ davger2. html)

Further reading
Bacciagaluppi, Guido; Valentini (2009), Quantum theory at the crossroads: reconsidering the 1927 Solvay
conference, Cambridge, UK: Cambridge University Press, arXiv:quant-ph/0609184, ISBN9780521814218,
OCLC227191829
Bernstein, Jeremy (2009), Quantum Leaps (http://books.google.com/?id=j0Me3brYOL0C&
printsec=frontcover), Cambridge, Massachusetts: Belknap Press of Harvard University Press,
ISBN9780674035416
Jammer, Max (1966), The conceptual development of quantum mechanics, New York: McGraw-Hill,
OCLC534562
Jammer, Max (1974), The philosophy of quantum mechanics: The interpretations of quantum mechanics in
historical perspective, New York: Wiley, ISBN0471439584, OCLC969760

External links
A History of Quantum Mechanics (http://www-groups.dcs.st-and.ac.uk/~history/HistTopics/
The_Quantum_age_begins.html)
A Brief History of Quantum Mechanics (http://www.oberlin.edu/physics/dstyer/StrangeQM/history.html)
Homepage of the Quantum History Project (http://quantum-history.mpiwg-berlin.mpg.de/)

Basic Concepts of Quantum Mechanics

10

Basic Concepts of Quantum Mechanics


This article is intended as an accessible, non-technical introduction to the subject. For the main encyclopedia
article, see Quantum mechanics. For a somewhat more technical introduction to the subject that requires
some algebra, see Introduction to quantum mechanics.
Quantum mechanics explains the behaviour of matter and energy on the scale of atoms and subatomic particles.
Classical physics explains matter and energy at the macroscopic level of the scale familiar to human experience,
including the behavior of astronomical bodies. It remains the key to measurement for much of modern science and
technology; but at the end of the 19th Century observers discovered phenomena in both the large (macro) and the
small (micro) worlds which classical physics could not explain.[1]
This article describes the limitations of classical physics, and explains the main concepts of the quantum theory
which supplanted it in the early decades of the 20th Century. These concepts are described in roughly the order they
were first discovered.

Origins in black body radiation


Thermal radiation is electromagnetic
radiation emitted from the surface of an
object due to the object's temperature. If the
object is heated sufficiently, it starts to emit
light at the red end of the spectrum: it is red
hot. Heating it further causes the colour to
change, as light of shorter wavelengths
(higher frequencies) becomes stronger. A
good emitter is also a good absorber. When
it is cold, such an object looks black, as it
emits practically no visible light, absorbing
all light that falls on it. Consequently, an
ideal emitter is known as a black body, and
the radiation it emits is called black body
radiation.

Hot metalwork from a blacksmith. The yellow-orange glow is the visible part of
the thermal radiation emitted due to the high temperature. Everything else in the
picture is glowing with thermal radiation as well, but less brightly and at longer
wavelengths that the human eye cannot see. A far-infrared camera records this
radiation.

In the late 19th Century, thermal radiation


had been fairly well characterised
experimentally. The wavelength at which
the radiation is strongest is given by Wien's displacement law, and the overall power emitted per unit area is given by
the StefanBoltzmann law. As temperature increases, the glow colour changes from red to yellow to white to blue.
Even as the peak wavelength moves into the ultra-violet, enough radiation continues to be emitted in the blue
wavelengths that the body continues to appear blue. It never becomes invisibleindeed, the radiation of visible light
increases monotonically with temperature.[2] Physicists were searching for a theoretical explanation of these
experimental results.
The answer from classical physics is called the RayleighJeans law. This law agrees with experimental results at
long wavelengths. At short wavelengths, however, it predicts that energy is emitted by a hot body at an infinite rate.
This result, which is clearly wrong, is known as the ultraviolet catastrophe.

Basic Concepts of Quantum Mechanics

11

The first model which was able to explain the full spectrum of thermal radiation
was put forward by Max Planck in 1900.[3] He modelled the thermal radiation as
being in equilibrium by analogy with a set of harmonic oscillators. But to
reproduce the experimental results, each oscillator had to have an exact number of
energy units, rather than being able to have any arbitrary amount of energy. In
other words, the energy of each oscillator was quantised.[4] He determined that the
energy of each oscillator was proportional to the frequency, i.e. that it was always
an exact multiple of a constant now known as the Planck constant.
Max Planck on a German postage

Planck's law was the first quantum theory in physics, and Planck won the Nobel
stamp
Prize in 1918 "in recognition of the services he rendered to the advancement of
Physics by his discovery of energy quanta".[5] At the time, however, Planck's own
view was that quantisation was purely a mathematical trick to explain the unexpected experimental results, rather
than (as we now know) a fundamental change in our understanding of the world.[6]

Photons: the quantisation of light


In 1905, Albert Einstein took an extra step. He suggested that quantisation wasn't
just a mathematical trick, but that the energy in a beam of light occurs in individual
packets now called photons.[7] The energy of a single photon is Planck's constant
multiplied by the photon's frequency.
Einstein's proposal was able to explain several puzzling properties of the
photoelectric effect, which is the way certain metals give off electrons when light
falls on them.[8] :24 For centuries, scientists had debated between two possible
theories of light: was it a wave or did it instead consist of a stream of tiny particles?
By the 19th Century, the debate was generally considered to have been settled in
favour of the wave theory, because it was able to explain observed effects such as
Einstein's portrait by Harm
refraction, diffraction and polarisation. Because of this, Einstein's proposal was met
Kamerlingh Onnes at the
by great scepticism. Eventually, however, his particle analogy became accepted, as
University of Leiden in 1920
it helped explain how light delivers energy in multiples of certain set values, called
quanta of energy. Nevertheless, the wave analogy remained indispensable for
helping to explain other phenomena of light, such as diffraction.
Thus, for the first time, was an object demonstrating both wave-like and particle-like characteristics modelled as
being based upon discrete energy levels.

Basic Concepts of Quantum Mechanics

12

Bohr model of the atom


By the early 20th century, it was known that atoms consisted of a diffuse cloud of
negatively-charged electrons surrounding a small, dense, positively-charged
nucleus. This suggested a model in which the electrons circled around the nucleus
like planets orbiting the sun.[9] However, it was also known that the atom in this
model would be unstable: the orbiting electrons should give off electromagnetic
radiation, causing them to lose energy and spiral towards the nucleus, colliding
with it in a fraction of a second.
A second, related, puzzle was the emission spectrum of atoms. When a gas is
heated, it gives off light at certain discrete frequencies; for example, the visible
light given off by hydrogen consists of four different colours, as shown in the
picture below. In contrast, white light contains light at the whole range of visible
frequencies.

Sketch by Christian Gori from a


photograph of Niels Bohr

Emission spectrum of hydrogen. When excited,


hydrogen gas gives off light in four distinct
colours (spectral lines) in the visible spectrum, as
well as a number of lines in the infra-red and
ultra-violet.

In 1913 Niels Bohr proposed a new model of the atom that included quantised electron orbits. This solution became
known as the Bohr model of the atom. In Bohr's model, electrons could inhabit only particular orbits around the
atomic nucleus. When an atom emits or absorbs energy, the electron does not move in a continuous trajectory from
one orbit around the nucleus to another, as might be expected in classical theory. Instead, the electron jumps
instantaneously from one orbit to another, giving off the difference in energy as light in the form of a photon.[10] :6
The possible energies of the photons given off by each element in the periodic table are determined by the difference
in energy between the orbits, so the emission spectrum for each element will contain a number of lines.[11] The Bohr
model was able to explain the emission spectrum of hydrogen, but wasn't able to make accurate predictions for
multi-electron atoms, or to explain why some spectral lines are brighter than others.

Basic Concepts of Quantum Mechanics

13

Particle-wave duality
Quantum mechanics is based upon the concept that subatomic particles
can have both wave-like and particle-like properties. This phenomenon
is known as waveparticle duality. The explanation stems from a
theory proposed by French physicist Louis de Broglie in 1924, that
subatomic particles such as electrons are associated with waves.
Experiments later showed that he was correct: electrons can bend
around objects and can display wave shapes.[10] :6
Consequently, neither wave nor particle is an entirely satisfactory
model to use in understanding light. Indeed, astrophysicist A.S.
Eddington proposed in 1927 that "We can scarcely describe such an
entity as a wave or as a particle; perhaps as a compromise we had
better call it a 'wavicle' ".[12] This term was later popularised by
mathematician Banesh Hoffmann.[13] :172

Light streaming through windows at Chicago's


Union Station in 1943. Quantum mechanics
shows that light acts both as waves and as
particles.

The concept of waves and particles, and the analogies which use them,
are mechanisms of classical physics. Quantum mechanics, which seeks
to explain nature at a level underlying that of the atoms which comprise matter, cannot be understood in such terms.
The classical concepts presuppose an artificial division of matter (as particles) and energy (as waves) that has no
objective validity on the sub-atomic level. If the distinction no longer holds true, it is not surprising if classes of
object can exhibit the characteristics of either.

Uncertainty principle
Suppose that we want to measure the position and speed of an object -- for example
a car going through a radar speed trap. Naively, we assume that (at a particular
moment in time) the car has a definite position and speed, and how accurately we
can measure these values depends on the quality of our measuring equipment -- if
we improve the precision of our measuring equipment, we will get a result that is
closer to the true value. In particular, we would assume that how precisely we
measure the speed of the car does not affect its position, and vice versa.
In 1927 German physicist Werner Heisenberg proved that in the sub-atomic world
such assumptions are not correct.[15] Quantum mechanics shows that certain pairs
of physical properties, such as position and speed, cannot both be known to
arbitrary precision. He showed that the more precisely one of them is known, the
less precisely the other can be known. This statement is known as the uncertainty
Werner Heisenberg at the age of
principle (or Heisenberg's uncertainty principle). It is not a statement about the
26. Heisenberg won the Nobel
Prize in Physics in 1932 for the
accuracy of our measuring equipment, but about the nature of the system itself -work that he did at around this
our naive assumption that an object has a definite position and speed is incorrect.
[14]
time.
On a scale of cars and people, these uncertainties are still present but are too small
to be noticed; yet they are large enough that when dealing with individual atoms
and electrons they become critical.[16]
Heisenberg gave, as an illustration, the measurement of the position and momentum of an electron using a photon of
light. In measuring the electron's position, the higher the frequency of the photon the more accurate is the
measurement of the position of the impact, but the greater is the disturbance of the electron, which absorbs a random
amount of energy, rendering the measurement obtained of its momentum increasingly uncertain (momentum is

Basic Concepts of Quantum Mechanics

14

velocity multiplied by mass), for one is necessarily measuring its post-impact disturbed momentum, from the
collision products, not its original momentum. With a photon of lower frequency the disturbance - hence uncertainty
- in the momentum is less, but so is the accuracy of the measurement of the position of the impact.
The uncertainty principle shows mathematically that the product of the uncertainty in the position and momentum of
a particle can never be less than a certain value, and that this value is related to Planck's constant. It is, in point of
fact, up to a small numerical factor equal to Planck's constant.

Schrdinger's wave equation


Although Heisenberg had no problem with the existence of discontinuous quantum
jumps, Austrian physicist Erwin Schrdinger hoped that a theory based on
continuous wave-like properties[17] could avoid what he called (in the reported
words of Wilhelm Wien[18] ) "this nonsense about quantum jumps."
Building on De Broglie's theoretical model of particles as waves, Schrdinger
accordingly brought forth in 1926 what has been called "the fundamental equation"
of quantum mechanics.[19]
In point of fact, Schrdinger's wave equation is as central to quantum mechanics as
Einstein's equation
is to Relativity.
The equation describes the probability waves which govern the motion of
sub-atomic particles, "and it specifies how these waves are altered by external
influences. Schrdinger established the correctness of the equation by applying it to
the hydrogen atom, predicting many of its properties with remarkable accuracy.
The equation is used extensively in atomic, nuclear, and solid-state physics."[19]

Erwin Schrdinger, about 1933,


age 46

Atomic orbital model


Bohr's model of the atom was essentially two-dimensional an electron orbiting in a plane around its nuclear
"sun." In modern theory, orbital has replaced the earlier word orbit concerning the position of an electron in relation
to the nucleus of an atom. It is often depicted as a three-dimensional region within which there is a 95 percent
probability of finding the electron.[20]
The uncertainty principle states that an electron cannot be viewed as having an exact location at any given time. The
concepts of exact position and exact velocity (distance traveled per unit of time) taken together really have no
meaning in nature.[21] An orbital, then, is a "cloud" of possible locations in which an electron might be found, a
distribution of probabilities rather than a precise location.[22]

Basic Concepts of Quantum Mechanics

15

Quantum field theory


The idea of quantum field theory began in the late 1920s with British physicist Paul
Dirac, when he attempted to quantise the electromagnetic field a procedure for
constructing a quantum theory starting from a classical theory.
A field in physics is "a region or space in which a given effect (such as magnetism)
exists."[23] Other effects that manifest themselves as fields are gravitation and static
electricity.[24] In 2008, physicist Richard Hammond wrote that
Sometimes we distinguish between quantum mechanics (QM) and
quantum field theory (QFT). QM refers to a system in which the
number of particles is fixed, and the fields (such as the
electromechanical field) are continuous classical entities. QFT . . . goes
a step further and allows for the creation and annihilation of particles .
...
He added, however, that quantum mechanics is often used to refer to "the entire
notion of quantum view."[25] :108
In 1931, Dirac proposed the existence of particles that later became known as
anti-matter.[26] Dirac shared the Nobel Prize in physics for 1933 with Schrdinger,
"for the discovery of new productive forms of atomic theory."[27]

Practical use

This sculpture in Bristol, England


a series of clustering cones
presents the idea of small worlds
which Paul Dirac studied to reach
his discovery of anti-matter.

The main value of the quantum mechanics theory is its practical applications.
Examples include the laser, the transistor, the electron microscope, and magnetic resonance imaging. The study of
semiconductors led to the invention of the diode and the transistor, which are indispensable for modern electronics.
In even the simple light switch, quantum tunnelling is vital, as otherwise the electrons in the electric current could
not penetrate the potential barrier made up of a layer of oxide. Flash memory chips found in USB drives also use
quantum tunnelling, to erase their memory cells.[28]

Notes
[1] Quantum Mechanics from [[National Public Radio (http:/ / www. pbs. org/ trasnsistor/ science/ info/ quantum. html)]]
[2] Landau, L. D.; E. M. Lifshitz (1996). Statistical Physics (3rd Edition Part 1 ed.). Oxford: Butterworth-Heinemann.
[3] This was published (in German) as Planck, Max (1901). "Ueber das Gesetz der Energieverteilung im Normalspectrum" (http:/ / www. physik.
uni-augsburg. de/ annalen/ history/ historic-papers/ 1901_309_553-563. pdf). Ann. Phys. 309 (3): 55363. doi:10.1002/andp.19013090310. ..
English translation: " On the Law of Distribution of Energy in the Normal Spectrum (http:/ / dbhs. wvusd. k12. ca. us/ webdocs/
Chem-History/ Planck-1901/ Planck-1901. html)".
[4] The word "quantum" comes from the Latin word for "how much" (as does "quantity"). Something that is "quantised", like the energy of
Planck's harmonic oscillators, can only have specific values. For example, in most countries money is effectively quantised: the "quantum of
money" being the lowest-value coin in circulation. "Mechanics" is the branch of science that deals with the action of forces on objects, and so
"quantum mechanics" is the form of mechanics that deals with objects for which particular properties are quantised.
[5] "The Nobel Prize in Physics 1918" (http:/ / nobelprize. org/ nobel_prizes/ physics/ laureates/ 1918/ ). The Nobel Foundation. . Retrieved
2009-08-01.
[6] Kragh, Helge (1 December 2000). "Max Planck: the reluctant revolutionary" (http:/ / physicsworld. com/ cws/ article/ print/ 373).
PhysicsWorld.com. .
[7] Einstein, Albert (1905). "ber einen die Erzeugung und Verwandlung des Lichtes betreffenden heuristischen Gesichtspunkt" (http:/ / www.
zbp. univie. ac. at/ dokumente/ einstein1. pdf). Annalen der Physik 17: 132148. ., translated into English as On a Heuristic Viewpoint
Concerning the Production and Transformation of Light (http:/ / lorentz. phl. jhu. edu/ AnnusMirabilis/ AeReserveArticles/ eins_lq. pdf). The
term "photon" was introduced in 1926.
[8] Stephen Hawking, The Universe in a Nutshell, Bantam, 2001.

Basic Concepts of Quantum Mechanics


[9] The classical model of the atom is called the planetary model or the Rutherford model, after Ernest Rutherford who proposed it in 1911, based
on the Geiger-Marsden gold foil experiment which first demonstrated the existence of the nucleus.
[10] World Book Encyclopedia, 2007.
[11] Dicke and Wittke, Introduction to Quantum Mechanics, p. 10f.
[12] A.S. Eddington, The Nature of the Physical World, the course of Gifford Lectures that Eddington delivered in the University of Edinburgh in
January to March 1927, Kessinger Publishing, 2005, p. 201. (http:/ / books. google. com/ books?id=PGOTKcxSqMUC& pg=PA201&
lpg=PA201& dq=We+ can+ scarcely+ describe+ such+ an+ entity+ as+ a+ wave+ or+ as+ a+ particle;+ perhaps+ as+ a+ compromise+ we+
had+ better+ call+ it+ a+ `wavicle& source=bl& ots=K0IfGzaXli& sig=zgrQiBJbHRLuUzVBT-yy8jZhC1Y& hl=en&
ei=i8g1SpOHC4PgtgOu_4jVDg& sa=X& oi=book_result& ct=result& resnum=1)
[13] Banesh Hoffman, The Strange Story of the Quantum, Dover, 1959
[14] Heisenberg's Nobel Prize citation (http:/ / nobelprize. org/ nobel_prizes/ physics/ laureates/ 1932/ )
[15] Heisenberg first published his work on the uncertainty principle in the leading German physics journal Zeitschrift fr Physik: Heisenberg,
W. (1927). "ber den anschaulichen Inhalt der quantentheoretischen Kinematik und Mechanik". Z. Phys. 43 (3-4): 172198.
doi:10.1007/BF01397280.
[16] Nobel Prize in Physics presentation speech, 1932 (http:/ / nobelprize. org/ nobel_prizes/ physics/ laureates/ 1932/ press. html)
[17] Schrdinger's formulation of quantum mechanics based on waves is sometimes referred to as "wave mechanics", to distinguish it from the
matrix mechanics formulation of Heisenberg, Max Born and Pascual Jordan.
[18] W. Moore, Schrdinger: Life and Thought, Cambridge University Press (1989), p. 222 (http:/ / www. amazon. com/ dp/ 0521437679)
[19] "Schrodinger Equation (Physics)," Encyclopedia Britannica (http:/ / www. britannica. com/ EBchecked/ topic/ 528298/
Schrodinger-equation)
[20] "Orbital (chemistry and physics)," Encyclopedia Britannica (http:/ / www. britannica. com/ EBchecked/ topic/ 431159/ orbital)
[21] "Uncertainty principle," Encyclopedia Britannica (http:/ / www. britannica. com/ EBchecked/ topic/ 614029/ uncertainty-principle)
[22] "What paths do electronics actually take in an orbital..." Physics Forums, January 4, 2005 (http:/ / www. physicsforums. com/ archive/
index. php/ t-58605. html)
[23] "Mechanics," Merriam-Webster Online Dictionary (http:/ / www. merriam-webster. com/ dictionary/ field)
[24] "Field," Encyclopedia Britannica (http:/ / www. britannica. com/ EBchecked/ topic/ 206162/ field)
[25] Richard Hammond, The Unknown Universe, New Page Books, 2008. ISBN 978-1-60163-003-2
[26] The Physical World website (http:/ / www. physicalworld. org/ restless_universe/ html/ ru_dira. html)
[27] "The Nobel Prize in Physics 1933" (http:/ / nobelprize. org/ nobel_prizes/ physics/ laureates/ 1933/ ). The Nobel Foundation. . Retrieved
2007-11-24.
[28] Durrani, Z. A. K.; Ahmed, H. (2008). Vijay Kumar. ed. Nanosilicon. Elsevier. p.345. ISBN9780080445281.

References
Further reading
Richard Feynman, 1985. QED: The Strange Theory of Light and Matter, Princeton University Press. ISBN
0-691-08388-6

16

Introduction to Quantum Mechanics

17

Introduction to Quantum Mechanics


Quantum mechanics is the body of scientific principles which attempts to explain the
behavior of matter and its interactions with energy on the scale of atoms and atomic
particles.
Just before 1900, it became clear that classical physics was unable to model certain
phenomena. Coming to terms with these limitations led to the development of quantum
mechanics, a major revolution in physics. This article describes how the limitations of
classical physics were discovered, and the main concepts of the quantum theories
which replaced them in the early decades of the 20th Century.[1] These concepts are
described in roughly the order they were first discovered; for a more complete history
of the subject, see History of quantum mechanics.
Some aspects of quantum mechanics can seem counter-intuitive, because they describe
behavior quite different than that seen at larger length scales, where classical physics is
an excellent approximation. In the words of Richard Feynman, quantum mechanics
deals with "nature as she is absurd."[2]
Many types of energy, such as photons (discrete units of light), behave in some respects
like particles and in other respects like waves. Radiators of photons (such as neon
lights) have emission spectra which are discontinuous, in that only certain frequencies
of light are present. Quantum mechanics predicts the energies, the colors, and the
spectral intensities of all forms of electromagnetic radiation.
But quantum mechanics theory ordains that the more closely one pins down one
measure (such as the position of a particle), the less precise another measurement
pertaining to the same particle (such as its momentum) must become. Put another way,
measuring position first and then measuring momentum does not have the same
outcome as measuring momentum first and then measuring position; the act of
measuring the first property necessarily introduces additional energy into the
micro-system being studied, thereby perturbing that system.

Left to right: Max Planck,


Albert Einstein, Niels Bohr,
Louis de Broglie, Max Born,
Paul Dirac, Werner
Heisenberg, Wolfgang Pauli,
Erwin Schrdinger, Richard
Feynman.

Even more disconcerting, pairs of particles can be created as entangled twins which
means that a measurement which pins down one property of one of the particles will
instantaneously pin down the same or another property of its entangled twin, regardless of the distance separating
them though this may be regarded as merely a mathematical, rather than a real, anomaly.

Introduction to Quantum Mechanics

18

The first quantum theory: Max Planck and black body radiation.

Hot metalwork from a blacksmith. The yellow-orange glow is the visible part of
the thermal radiation emitted due to the high temperature. Everything else in the
picture is glowing with thermal radiation as well, but less brightly and at longer
wavelengths that the human eye cannot see. A far-infrared camera will show this
radiation.

Thermal radiation is electromagnetic


radiation emitted from the surface of an
object due to the object's temperature. If an
object is heated sufficiently, it starts to emit
light at the red end of the spectrum it is
"red hot". Heating it further causes the
colour to change, as light at shorter
wavelengths (higher frequencies) begins to
be emitted. It turns out that a perfect emitter
is also a perfect absorber. When it is cold,
such an object looks perfectly black, as it
emits practically no visible light, because it
absorbs all the light that falls on it.
Consequently, an ideal thermal emitter is
known as a black body, and the radiation it
emits is called black body radiation.

In the late 19th century, thermal radiation


had been fairly well characterized
experimentally. The wavelength at which the radiation is strongest is given by Wien's displacement law, and the
overall power emitted per unit area is given by the StefanBoltzmann law. So, as temperature increases, the glow
colour changes from red to yellow to white to blue. Even as the peak wavelength moves into the ultra-violet, enough
radiation continues to be emitted in the blue wavelengths that the body continues to appear blue. It never becomes
invisibleindeed, the radiation of visible light increases monotonically with temperature.[3] Physicists were
searching for a theoretical explanation for these experimental results.
The "answer" found using classical
physics is the RayleighJeans law.
This law agrees with experimental
results at long wavelengths. At short
wavelengths,
however,
classical
physics predicts that energy will be
emitted by a hot body at an infinite
rate. This result, which is clearly
wrong, is known as the ultraviolet
catastrophe.
The first model which was able to
explain the full spectrum of thermal
radiation was put forward by Max
Planck in 1900.[4] He modelled the
thermal radiation as being in
equilibrium, using a set of harmonic
oscillators.
To
reproduce
the
experimental results he had to assume

The peak wavelength and total power radiated by a black body vary with temperature.
Classical electromagnetism drastically overestimates these intensities, particularly at short
wavelengths.

Introduction to Quantum Mechanics

19

that each oscillator produced an integral number of units of energy at its one characteristic frequency, rather than
being able to emit any arbitrary amount of energy. In other words, the energy of each oscillator was "quantized".[5]
The quantum of energy for each oscillator, according to Planck, was proportional to the frequency of the oscillator;
the constant of proportionality is now known as the Planck constant. The Planck constant, usually written as h, has
the value 6.631034J s, and so the energy, E, of an oscillator of frequency f is given by
where

[6]

Planck's law was the first quantum theory in physics, and Planck won the Nobel Prize in 1918 "in recognition of the
services he rendered to the advancement of Physics by his discovery of energy quanta".[7] At the time, however,
Planck's view was that quantization was purely a mathematical trick, rather than (as we now know) a fundamental
change in our understanding of the world.[8]

Photons: the quantisation of light


In 1905, Albert Einstein took an extra step. He suggested that quantisation wasn't
just a mathematical trick: the energy in a beam of light occurs in individual
packets, which are now called photons.[9] The energy of a single photon is given by
its frequency multiplied by Planck's constant:

For centuries, scientists had debated between two possible theories of light: was it a
wave or did it instead consist of a stream of tiny particles? By the 19th century, the
debate was generally considered to have been settled in favour of the wave theory,
as it was able to explain observed effects such as refraction, diffraction and
polarization. James Clerk Maxwell had shown that electricity, magnetism and light
Einstein's portrait by Harm
are all manifestations of the same phenomenon: the electromagnetic field.
Kamerlingh Onnes at the
Maxwell's equations, which are the complete set of laws of classical
University of Leiden in 1920
electromagnetism, describe light as waves: a combination of oscillating electric and
magnetic fields. Because of the preponderance of evidence in favour of the wave theory, Einstein's ideas were met
initially by great scepticism. Eventually, however, the photon model became favoured; one of the most significant
pieces of evidence in its favour was its ability to explain several puzzling properties of the photoelectric effect,
described in the following section. Nonetheless, the wave analogy remained indispensable for helping to understand
other characteristics of light, such as diffraction.

The photoelectric effect


In 1887 Heinrich Hertz observed that light can eject electrons from
metal.[10] In 1902 Philipp Lenard discovered that the maximum
possible energy of an ejected electron is related to the frequency of
the light, not to its intensity; if the frequency is too low, no
electrons are ejected regardless of the intensity. The lowest
frequency of light which causes electrons to be emitted, called the
threshold frequency, is different for every metal. This is at odds
with classical electromagnetism, which predicts that the electron's
energy should be proportional to the intensity of the radiation.
Light (red arrows, left) is shone upon a metal. If the
light is of sufficient frequency (i.e. sufficient energy),
electrons are ejected (blue arrows, right).

Introduction to Quantum Mechanics


Einstein explained the effect by postulating that a beam of light is a stream of particles (photons), and that if the
beam is of frequency f each photon has an energy equal to hf (i.e. an integer multiple of Planck's constant).[10] An
electron is likely to be struck only by a single photon, which imparts at most an energy hf to the electron[10] (in point
of fact, it logically cannot be struck by more than one photon, since the first it absorbs will cause it to eject).
Therefore, the intensity of the beam has no effect;[11] only its frequency determines the maximum energy that can be
imparted to the electron.[10]
To explain the threshold effect, Einstein argued that it takes a certain amount of energy, called the work function,
denoted by , to remove an electron from the metal.[10] This amount of energy is different for each metal. If the
energy of the photon is less than the work function then it does not carry sufficient energy to remove the electron
from the metal. The threshold frequency, f0, is the frequency of a photon whose energy is equal to the work function:
If f is greater than f0, the energy hf is enough to remove an electron. The ejected electron has a kinetic energy EK
which is, at most, equal to the photon's energy minus the energy needed to dislodge the electron from the metal:

Einstein's description of light as being composed of particles extended Planck's notion of quantised energy: a single
photon of a given frequency f delivers an invariant amount of energy hf. In other words, individual photons can
deliver more or less energy, but only depending on their frequencies. However, although the photon is a particle it
was still being described as having the wave-like property of frequency. Once again, the particle account of light was
being "compromised".[12] [13]
The relationship between the frequency of electromagnetic radiation and the energy of each individual photon is why
ultraviolet light can cause sunburn, but visible or infrared light cannot. A photon of ultraviolet light will deliver a
high amount of energyenough to contribute to cellular damage such as a sunburn. A photon of infrared light will
deliver a lower amount of energyonly enough to warm one's skin. So an infrared lamp can warm a large surface,
perhaps large enough to keep people comfortable in a cold room, but it cannot give anyone a sunburn.
If each individual photon had identical energy, it would not be correct to talk of a "high energy" photon. Light of
high frequency would carry more energy only because of a wave effect, i.e. because there were more photons
arriving per second. If you doubled the frequency, you would double the number of energy units arriving each
second: an argument based on intensity (i.e. on the number of photons per second). Einstein rejected that
wave-dependent classical approach, in favour of a particle-based analysis where the energy of the particle must be
absolute (since it is measured from a single impact only), and varies with frequency in discrete steps (i.e. is
quantised). Hence he arrived at the concept of quantised energy levels.
In nature, single photons are not encountered. The sun emits photons continuously at all electromagnetic frequencies,
so they appear to propagate as a continuous wave, not as discrete units. The emission sources available to Hertz and
Lennard in the 19th Century also had that characteristic. And the traditional mechanisms for generating photons are
classical devices, in which the energy output is regulated by varying the frequency. But Einstein proposed that
although a particular frequency is tied to a specific energy level, the frequency is dependent on the energy level, not
vice versa (contrary to the tenets of classical physics). This formed his solution to the photoelectric effect, even
though it was counter-intuitive.
And although the energy imparted by the photon is invariant at any given frequency, the initial energy-state of the
electron prior to absorption is not. Therefore anomalous results may occur for individual electrons, but statistically a
process of averaging will smooth out the results if a large enough number of electrons are emitted. This point is
helpful in comprehending the distinction between the study of individual particles in quantum dynamics and the
study of massed particles in classical physics.

20

Introduction to Quantum Mechanics

The quantisation of matter: the Bohr model of the atom


By the dawn of the 20th Century, it was known that atoms consist of a diffuse cloud of negatively-charged electrons
surrounding a small, dense, positively-charged nucleus. This suggested a model in which the electrons circle around
the nucleus like planets orbiting a sun.[14] However, it was also known that the atom in this model would be
unstable: according to classical theory the orbiting electrons should give off electromagnetic radiation, causing them
to lose energy and spiral toward the nucleus, colliding with it in a fraction of a second.
A second, related, puzzle was the emission spectrum of atoms. When a gas is heated, it gives off light only at
discrete frequencies. For example, the visible light given off by hydrogen consists of four different colours, as shown
in the picture below. By contrast, white light consists of a continuous emission across the whole range of visible
frequencies.

Emission spectrum of hydrogen. When excited,


hydrogen gas gives off light in four distinct
colours (spectral lines) in the visible spectrum, as
well as a number of lines in the infra-red and
ultra-violet.

In 1885 the Swiss mathematician Johann Balmer discovered that each wavelength (lambda) in the visible spectrum
of hydrogen is related to some integer n by the equation

where B is a constant which Balmer determined to be equal to 364.56nm. Thus Balmer's constant was the basis of a
system of discrete, i.e. quantised, integers.
In 1888 Johannes Rydberg generalized and greatly increased the explanatory utility of Balmer's formula. He
predicted that hydrogen will emit light of wavelength (lambda) where is related to two integers n and m
according to what is now known as the Rydberg formula:[15]

where R is the Rydberg constant, equal to 0.0110nm1, and n must be greater than m.
Rydberg's formula accounts for the four visible wavelengths of hydrogen by setting m = 2 and n = 3,4,5,6. It also
predicts additional wavelengths in the emission spectrum: for m = 1 and for n > 1, the emission spectrum should
contain certain ultraviolet wavelengths, and for m = 3 and n > 3, it should also contain certain infrared wavelengths.
Experimental observation of these wavelengths came several decades later: in 1908 Louis Paschen found some of the
predicted infrared wavelengths, and in 1914 Theodore Lyman found some of the predicted ultraviolet
wavelengths.[15]

21

Introduction to Quantum Mechanics

22

Bohr's model
In 1913 Niels Bohr proposed a new model of the atom
that included quantized electron orbits.[16] In Bohr's
model, electrons could inhabit only certain orbits
around the atomic nucleus. When an atom emitted (or
absorbed) energy, the electron did not move in a
continuous trajectory from one orbit around the nucleus
to another, as might be expected classically. Instead,
the electron would jump instantaneously from one orbit
to another, giving off the emitted light in the form of a
photon.[17] The possible energies of photons given off
by each element were determined by the differences in
energy between the orbits, and so the emission
spectrum for each element would contain a number of
lines.[18]
Bohr theorised that the angular momentum, L, of an
electron is quantised:

The Bohr model of the atom, showing an electron quantum jumping


to ground state n = 1.

where n is an integer and h is the Planck constant. Starting from this assumption, Coulomb's law and the equations of
circular motion show that an electron with n units of angular momentum will orbit a proton at a distance r given by
,
where ke is the Coulomb constant, m is the mass of an electron, and e is the charge on an electron. For simplicity this
is written as

where a0, called the Bohr radius, is equal to 0.0529nm. The Bohr radius is the radius of the smallest allowed orbit.
The energy of the electron[19] can also be calculated, and is given by
.
Thus Bohr's assumption that angular momentum is quantised means that an electron can only inhabit certain orbits
around the nucleus, and that it can have only certain energies. A consequence of these constraints is that the electron
will not crash into the nucleus: it cannot continuously emit energy, and it cannot come closer to the nucleus than a0
(the Bohr radius).
An electron loses energy by jumping instantaneously from its original orbit to a lower orbit; the extra energy is
emitted in the form of a photon. Conversely, an electron that absorbs a photon gains energy, hence it jumps to an
orbit that is farther from the nucleus.
Each photon from glowing atomic hydrogen is due to an electron moving from a higher orbit, with radius rn, to a
lower orbit, rm. The energy E of this photon is the difference in the energies En and Em of the electron:

Introduction to Quantum Mechanics

23

Since Planck's equation shows that the photon's energy is related to its wavelength by E = hc/, the wavelengths of
light which can be emitted are given by

This equation has the same form as the Rydberg formula, and predicts that the constant R should be given by

Therefore the Bohr model of the atom can predict the emission spectrum of hydrogen in terms of fundamental
constants.[20] However, it was not able to make accurate predictions for multi-electron atoms, or to explain why
some spectral lines are brighter than others.

Wave-particle duality
In 1924, Louis de Broglie proposed the idea that just as light has both wave-like and particle-like properties, matter
also has wave-like properties.[21] The wavelength, , associated with a particle is related to its momentum, p:[22] [23]

The relationship, called the de Broglie hypothesis, holds for all types of matter. Thus all matter exhibits properties of
both particles and waves.
Three years later, the wave-like nature of electrons was demonstrated by showing that a beam of electrons could
exhibit diffraction, just like a beam of light. At the University of Aberdeen, George Thomson passed a beam of
electrons through a thin metal film and observed the predicted interference patterns. At Bell Labs, Davisson and
Germer guided their beam through a crystalline grid. Similar wave-like phenomena were later shown for atoms and
even small molecules. De Broglie was awarded the Nobel Prize for Physics in 1929 for his hypothesis; Thomson and
Davisson shared the Nobel Prize for Physics in 1937 for their experimental work.
This is the concept of wave-particle duality: neither the classical concepts of "particle" or "wave" can fully describe
the behavior of quantum-scale objects, either photons or matter. Wave-particle duality is an example of the principle
of complementarity in quantum physics. An elegant example of wave-particle duality, the double slit experiment, is
discussed in the section below.
De Broglie's treatment of quantum events served as a jumping off point for Schrdinger when he set about to
construct a wave equation to describe quantum theoretical events.

The double-slit experiment

Light from one slit interferes with light from the


other, producing an interference pattern (the 3
fringes shown at the right).

In the double-slit experiment as originally performed by Thomas


Young and Augustin Fresnel in 1827, a beam of light is directed
through two narrow, closely spaced slits, producing an interference
pattern of light and dark bands on a screen. If one of the slits is covered
up, one might naively expect that the intensity of the fringes due to
interference would be halved everywhere. In fact, a much simpler
pattern is seen, a simple diffraction pattern. Closing one slit results in a
much simpler pattern diametrically opposite the open slit. Exactly the
same behaviour can be demonstrated in water waves, and so the
double-slit experiment was seen as a demonstration of the wave nature
of light.

Introduction to Quantum Mechanics

24

The double-slit experiment has also been


performed using electrons, atoms, and even
molecules, and the same type of interference
pattern is seen. Thus all matter possesses
both particle and wave characteristics.
Even if the source intensity is turned down
so that only one particle (e.g. photon or
electron) is passing through the apparatus at
a time, the same interference pattern
develops over time. The quantum particle
acts as a wave when passing through the
double slits, but as a particle when it is
detected. This is a typical feature of
quantum complementarity: a quantum
The diffraction pattern produced when light is shone through one slit (top) and the
interference pattern produced by two slits (bottom). The interference pattern from
particle will act as a wave when we do an
two slits is much more complex, demonstrating the wave-like propagation of light.
experiment to measure its wave-like
properties, and like a particle when we do an
experiment to measure its particle-like properties. Where on the detector screen any individual particle shows up will
be the result of an entirely random process.

Application to the Bohr model


De Broglie expanded the Bohr model of the atom by showing that an electron in orbit around a nucleus could be
thought of as having wave-like properties. In particular, an electron will be observed only in situations that permit a
standing wave around a nucleus. An example of a standing wave is a violin string, which is fixed at both ends and
can be made to vibrate. The waves created by a stringed instrument appear to oscillate in place, moving from crest to
trough in an up-and-down motion. The wavelength of a standing wave is related to the length of the vibrating object
and the boundary conditions. For example, because the violin string is fixed at both ends, it can carry standing waves
of wavelengths 2l/n, where l is the length and n is a positive integer. De Broglie suggested that the allowed electron
orbits were those for which the circumference of the orbit would be an integer number of wavelengths.

Development of modern quantum mechanics


In 1925, building on de Broglie's hypothesis, Erwin Schrdinger developed the
equation that describes the behaviour of a quantum mechanical wave. The
equation, called the Schrdinger equation after its creator, is central to quantum
mechanics, and defines the permitted stationary states of a quantum system, and
describes how the quantum state of a physical system changes in time.
Schrdinger was able to calculate the energy levels of hydrogen by treating a
hydrogen atom's electron as a classical wave, moving in a well of electrical
potential created by the proton. This calculation accurately reproduced the energy
levels of the Bohr model.

Erwin Schrdinger, about 1933,


age 46

Introduction to Quantum Mechanics

25
At around the same time, Werner Heisenberg was trying to find an explanation for
the intensities of the different lines in the hydrogen emission spectrum. By means
of a series of mathematical analogies, Heisenberg wrote out the quantum
mechanical analogue for the classical computation of intensities. Shortly
afterwards, Heisenberg's colleague Max Born realised that Heisenberg's method of
calculating the probabilities for transitions between the different energy levels
could best be expressed by using the mathematical concept of matrices.[25]

Werner Heisenberg at the age of


26. Heisenberg won the Nobel
Prize in Physics in 1932 for the
work that he did at around this
[24]
time.

In May 1926, Schrdinger proved that Heisenberg's matrix mechanics and his own
wave mechanics made the same predictions about the properties and behaviour of
the electron; mathematically, the two theories were identical. Yet the two men
disagreed on the interpretation of their mutual theory. Heisenberg saw no problem
in the existence of discontinuous quantum jumps, but Schrdinger hoped that a
theory based on continuous wave-like properties could avoid what he called (in the
words of Wilhelm Wien[26] ) "this nonsense about quantum jumps."

Copenhagen interpretation
Bohr, Heisenberg and others tried to explain what these experimental results and mathematical models really mean.
Their description, known as the Copenhagen interpretation of quantum mechanics, aimed to describe the nature of
reality that was being probed by the measurements and described by the mathematical formulations of quantum
mechanics.
The main principles of the Copenhagen interpretation are:
1. A system is completely described by a wave function, . (Heisenberg)
2. How changes over time is given by the Schrdinger equation.
3. The description of nature is essentially probabilistic. The probability of an event for example, where on the
screen a particle will show up in the two slit experiment is related to the square of the amplitude of its wave
function. (Born rule, due to Max Born, which gives a physical meaning to the wavefunction in the Copenhagen
interpretation: the probability amplitude)
4. It is not possible to know the values of all of the properties of the system at the same time; those properties that
are not known with precision must be described by probabilities. (Heisenberg's uncertainty principle)
5. Matter, like energy, exhibits a wave-particle duality. An experiment can demonstrate the particle-like properties
of matter, or its wave-like properties; but not both at the same time. (Complementarity principle due to Bohr)
6. Measuring devices are essentially classical devices, and measure classical properties such as position and
momentum.
7. The quantum mechanical description of large systems should closely approximate the classical description.
(Correspondence principle of Bohr and Heisenberg)
Various consequences of these principles are discussed in more detail in the following subsections.

Introduction to Quantum Mechanics

Uncertainty principle
Suppose that we want to measure the position and speed of an object for example a car going through a radar
speed trap. Naively, we assume that the car has a definite position and speed at a particular moment in time, and how
accurately we can measure these values depends on the quality of our measuring equipment if we improve the
precision of our measuring equipment, we will get a result that is closer to the true value. In particular, we would
assume that how precisely we measure the speed of the car does not affect its position, and vice versa.
In 1927, Heisenberg proved that these assumptions are not correct.[27] Quantum mechanics shows that certain pairs
of physical properties, like position and speed, cannot both be known to arbitrary precision: the more precisely one
property is known, the less precisely the other can be known. This statement is known as the uncertainty principle.
The uncertainty principle isn't a statement about the accuracy of our measuring equipment, but about the nature of
the system itself our naive assumption that the car had a definite position and speed was incorrect. On a scale of
cars and people, these uncertainties are too small to notice, but when dealing with atoms and electrons they become
critical.[28]
The uncertainty principle shows mathematically that the product of the uncertainty in the position and momentum of
a particle (momentum is velocity multiplied by mass) could never be less than a certain value, and that this value is
related to Planck's constant.

Wave function collapse


Wavefunction collapse is a forced term for whatever happened when it becomes appropriate to replace the
description of an uncertain state of a system by a description of the system in a definite state. Explanations for the
nature of the process of becoming certain are controversial. At any time before a photon "shows up" on a detection
screen it can only be described by a set of probabilities for where it might show up. When it does show up, for
instance in the CCD of an electronic camera, the time and the space where it interacted with the device are known
within very tight limits. However, the photon has disappeared, and the wave function has disappeared with it. In its
place some physical change in the detection screen has appeared, e.g., an exposed spot in a sheet of photographic
film.

Eigenstates and eigenvalues


For a more detailed introduction to this subject, see: Introduction to eigenstates
Because of the uncertainty principle, statements about both the position and momentum of particles can only assign a
probability that the position or momentum will have some numerical value. Therefore it is necessary to formulate
clearly the difference between the state of something that is indeterminate, such as an electron in a probability cloud,
and the state of something having a definite value. When an object can definitely be "pinned-down" in some respect,
it is said to possess an eigenstate.

The Pauli exclusion principle


Wolfgang Pauli proposed the following concise statement of his 1925 principle: "There cannot exist an atom in such
a quantum state that two electrons within [it] have the same set of quantum numbers."[29]
He developed the exclusion principle from what he called a "two-valued quantum degree of freedom" to account for
the observation of a doublet, meaning a pair of lines differing by a small amount (e.g. on the order of 0.15 ), in the
spectrum of atomic hydrogen. The existence of these closely spaced lines in the bright-line spectrum meant that there
was more energy in the electron orbital from magnetic moments than had previously been described.
In early 1925, Uhlenbeck and Goudsmit proposed that electrons rotate about an axis in the same way that the earth
rotates on its axis. They proposed to call this property spin. Spin would account for the missing magnetic moment,
and allow two electrons in the same orbital to occupy distinct quantum states if they, "spun" in opposite directions,

26

Introduction to Quantum Mechanics


thus satisfying the Exclusion Principle. A new quantum number was then needed, one to represent the momentum
embodied in the rotation of each electron.

Application to the hydrogen atom


Schrdinger was able to calculate the energy levels of hydrogen by treating a hydrogen atom's electron as a wave,
represented by the "wave function" , in a electric potential well, V, created by the proton. In Bohr's theory,
electrons orbited the nucleus much as planets orbit the sun. Schrdinger's model gives the probability of the electron
being located at any particular location following the uncertainty principle, the electron cannot be viewed as
having an exact location at any given time. The solutions to Schrdinger's equation are called atomic orbitals,
"clouds" of possible locations in which an electron might be found, a distribution of probabilities rather than a
precise location. Orbitals have a range of different shapes in three dimensions. The energies of the different orbitals
can be calculated, and they accurately reproduce the energy levels of the Bohr model.
Within Schrdinger's picture, each electron has four properties:
1. An "orbital" designation, indicating whether the particle wave is one that is closer to the nucleus with less energy
or one that is farther from the nucleus with more energy;
2. The "shape" of the orbital, spherical or otherwise;
3. The "inclination" of the orbital, determining the magnetic moment of the orbital around the z-axis.
4. The "spin" of the electron.
The collective name for these properties is the quantum state of the electron. The quantum state can be described by
giving a number to each these properties; these are known as the electron's quantum numbers. The quantum state of
the electron is described by its wavefunction. The Pauli exclusion principle demands that no two electrons within an
atom may have the same values of all four numbers.

The shapes of the first five atomic orbitals: 1s, 2s, 2px,2py, and 2pz. The colors show the
phase of the wavefunction.

The first property describing the orbital is the principal quantum number, n, which is the same as in Bohr's model. n
denotes the energy level of each orbital. The possible values for n are integers:
The next quantum number, the azimuthal quantum number, denoted l, describes the shape of the orbital. The shape is
a consequence of the angular momentum of the orbital. The angular momentum represents the resistance of a
spinning object to speeding up or slowing down under the influence of external force. The azimuthal quantum
number represents the orbital angular momentum of an electron around its nucleus. The possible values for l are
integers from 0 to n - 1:
. The shape of each orbital has its own letter as well. The first shape
is denoted by the letter s (for "spherical"). The next shape is denoted by the letter p and has the form of a dumbbell.
The other orbitals have more complicated shapes (see atomic orbital), and are denoted by the letters d, f, and g.
The third quantum number, the magnetic quantum number, describes the magnetic moment of the electron, and is
denoted by ml (or simply m). The possible values for ml are integers from -l to l:
. The magnetic quantum number measures the component of the angular
momentum in a particular direction. The choice of direction is arbitrary, conventionally the z-direction is chosen.
The fourth quantum number, the spin quantum number (pertaining to the "orientation" of the electron's spin) is
denoted ms, with values +12 or -12.

27

Introduction to Quantum Mechanics


The chemist Linus Pauling wrote, by way of example:
In the case of a helium atom with two electrons in the 1 s orbital, the Pauli Exclusion Principle requires that
the two electrons differ in the value of one quantum number. Their values of n, l, and ml are the same;
moreover, they have the same spin, s = 12. Accordingly they must differ in the value of ms, which can have the
value of +12 for one electron and 12 for the other."[29]
It is the underlying structure and symmetry of atomic orbitals, and the way that electrons fill them, that determines
the organisation of the periodic table and the structure and strength of chemical bonds between atoms.

Dirac wave equation


In 1928, Paul Dirac extended the Pauli equation, which described spinning electrons, to account for special relativity.
The result was a theory that dealt properly with events, such as the speed at which an electron orbits the nucleus,
occurring at a substantial fraction of the speed of light. By using the simplest electromagnetic interaction, Dirac was
able to predict the value of the magnetic moment associated with the electron's spin, and found the experimentally
observed value, which was too large to be that of a spinning charged sphere governed by classical physics. He was
able to solve for the spectral lines of the hydrogen atom, and to reproduce from physical first principles
Sommerfeld's successful formula for the fine structure of the hydrogen spectrum.
Dirac's equations sometimes yielded a negative value for energy, for which he proposed a novel solution: he posited
the existence of an antielectron and of a dynamical vacuum. This led to many-particle quantum field theory.

Quantum entanglement
The Pauli exclusion principle says that
two electrons in one system cannot be
in the same state. Nature leaves open
the possibility, however, that two
electrons can have both states
"superimposed" over them. Recall that
the wave functions that emerge
Superposition of two quantum characteristics, and two resolution possibilities.
simultaneously from the double slits
arrive at the detection screen in a state
of superposition. Nothing is certain until the superimposed waveforms "collapse," At that instant an electron shows
up somewhere in accordance with the probabilities that are the squares of the amplitudes of the two superimposed
waveforms. The situation there is already very abstract. A concrete way of thinking about entangled photons,
photons in which two contrary states are superimposed on each of them in the same event, is as follows:
Imagine that the superposition of a state that can be mentally labeled as blue and another state that can be mentally
labeled as red will then appear (in imagination, of course) as a purple state. Two photons are produced as the result
of the same atomic event. Perhaps they are produced by the excitation of a crystal that characteristically absorbs a
photon of a certain frequency and emits two photons of half the original frequency. So the two photons come out
"purple." If the experimenter now performs some experiment that will determine whether one of the photons is either
blue or red, then that experiment changes the photon involved from one having a superposition of "blue" and "red"
characteristics to a photon that has only one of those characteristics. The problem that Einstein had with such an
imagined situation was that if one of these photons had been kept bouncing between mirrors in a laboratory on earth,
and the other one had traveled halfway to the nearest star, when its twin was made to reveal itself as either blue or
red, that meant that the distant photon now had to lose its "purple" status too. So whenever it might be investigated, it
would necessarily show up, instantaneously, in the opposite state to whatever its twin had revealed.

28

Introduction to Quantum Mechanics


Suppose that some species of animal life carries both male and female characteristics in its genetic potential. It will
become either male or female depending on some environmental change. Perhaps it will remain indeterminate until
the weather either turns very hot or very cold. Then it will show one set of sexual characteristics and will be locked
into that sexual status by epigenetic changes, the presence in its system of high levels of androgen or estrogen, etc.
There are actually situations in nature that are similar to this scenario, but now imagine that if twins are born, then
they are forbidden by nature to both manifest the same sex. So if one twin goes to Antarctica and changes to become
a female, then the other twin will turn into a male despite the fact that local weather has done nothing special to it.
Such a world would be very hard to explain. How can something that happens to one animal in Antarctica affect its
twin in Redwood, California? Is it mental telepathy? What? How can the change be instantaneous? Even a radio
message from Antarctica would take a certain amount of time.
In trying to show that quantum mechanics was not a complete theory, Einstein started with the theory's prediction
that two or more particles that have interacted in the past can appear strongly correlated when their various
properties are later measured. He sought to explain this seeming interaction in a classical way, through their common
past, and preferably not by some "spooky action at a distance." The argument is worked out in a famous paper,
Einstein, Podolsky, and Rosen (1935; abbreviated EPR), setting out what is now called the EPR paradox. Assuming
what is now usually called local realism, EPR attempted to show from quantum theory that a particle has both
position and momentum simultaneously, while according to the Copenhagen interpretation, only one of those two
properties actually exists and only at the moment that it is being measured. EPR concluded that quantum theory is
incomplete in that it refuses to consider physical properties which objectively exist in nature. (Einstein, Podolsky, &
Rosen 1935 is currently Einstein's most cited publication in physics journals.) In the same year, Erwin Schrdinger
used the word "entanglement" and declared: "I would not call that one but rather the characteristic trait of quantum
mechanics." [30] The question of whether entanglement is a real condition is still in dispute.[31] The Bell inequalities
are the most powerful challenge to Einstein's claims.

Quantum electrodynamics
Quantum electrodynamics (QED) is the name of the quantum theory of the electromagnetic force. Understanding
QED begins with understanding electromagnetism. Electromagnetism can be called "electrodynamics" because it is a
dynamic interaction between electrical and magnetic forces. Electromagnetism begins with the electric charge.
Electric charges are the sources of, and create, electric fields. An electric field is a field which exerts a force on any
particles that carry electric charges, at any point in space. This includes the electron, proton, and even quarks, among
others. As a force is exerted, electric charges move, a current flows and a magnetic field is produced. The magnetic
field, in turn causes electric current (moving electrons). The interacting electric and magnetic field is called an
electromagnetic field.
The physical description of interacting charged particles, electrical currents, electrical fields, and magnetic fields is
called electromagnetism.
In 1928 Paul Dirac produced a relativistic quantum theory of electromagnetism. This was the progenitor to modern
quantum electrodynamics, in that it had essential ingredients of the modern theory. However, the problem of
unsolvable infinities developed in this relativistic quantum theory. Years later, renormalization solved this problem.
Initially viewed as a suspect, provisional procedure by some of its originators, renormalization eventually was
embraced as an important and self-consistent tool in QED and other fields of physics. Also, in the late 1940s
Feynman's diagrams showed all possible interactions of a given event. The diagrams showed that the
electromagnetic force is the interactions of photons between interacting particles.
An example of a prediction of quantum electrodynamics which has been verified experimentally is the Lamb shift.
This refers to an effect whereby the quantum nature of the electromagnetic field causes the energy levels in an atom
or ion to deviate slightly from what they would otherwise be. As a result, spectral lines may shift or split.

29

Introduction to Quantum Mechanics


In the 1960s physicists realized that QED broke down at extremely high energies. From this inconsistency the
Standard Model of particle physics was discovered, which remedied the higher energy breakdown in theory. The
Standard Model unifies the electromagnetic and weak interactions into one theory. This is called the electroweak
theory.

Interpretations
The physical measurements, equations, and predictions pertinent to quantum mechanics are all consistent and hold a
very high level of confirmation. However, the question of what these abstract models say about the underlying nature
of the real world has received competing answers.

Notes
[1]
[2]
[3]
[4]

Much of the universe on the largest scale also does not conform to classical physics, because of general relativity.
Richard P. Feynman, QED, p. 10
Landau, L. D.; E. M. Lifshitz (1996). Statistical Physics (3rd Edition Part 1 ed.). Oxford: Butterworth-Heinemann. ISBN0521653142.
This was published (in German) as Planck, Max (1901). "Ueber das Gesetz der Energieverteilung im Normalspectrum" (http:/ / www. physik.
uni-augsburg. de/ annalen/ history/ historic-papers/ 1901_309_553-563. pdf). Ann. Phys. 309 (3): 55363. doi:10.1002/andp.19013090310. .
English translation: " On the Law of Distribution of Energy in the Normal Spectrum (http:/ / dbhs. wvusd. k12. ca. us/ webdocs/
Chem-History/ Planck-1901/ Planck-1901. html)".

[5] The word "quantum" comes from the Latin word for "how much" (as does "quantity"). Something which is "quantized", like the energy of
Planck's harmonic oscillators, can only take specific values. For example, in most countries money is effectively quantized, with the "quantum
of money" being the lowest-value coin in circulation. "Mechanics" is the branch of science that deals with the action of forces on objects, and
so "quantum mechanics" is the part of mechanics that deals with objects for which particular properties are quantized.
[6] Francis Weston Sears (1958). Mechanics, Wave Motion, and Heat (http:/ / books. google. com/ books?hl=en& q="Mechanics,+ Wave+
Motion,+ and+ Heat"+ "where+ n+ =+ 1,"& btnG=Search+ Books). Addison-Wesley. p.537. .
[7] "The Nobel Prize in Physics 1918" (http:/ / nobelprize. org/ nobel_prizes/ physics/ laureates/ 1918/ ). The Nobel Foundation. . Retrieved
2009-08-01.
[8] Kragh, Helge (1 December 2000). "Max Planck: the reluctant revolutionary" (http:/ / physicsworld. com/ cws/ article/ print/ 373).
PhysicsWorld.com.
[9] Einstein, Albert (1905). "ber einen die Erzeugung und Verwandlung des Lichtes betreffenden heuristischen Gesichtspunkt" (http:/ / www.
zbp. univie. ac. at/ dokumente/ einstein1. pdf). Annalen der Physik 17: 132148. ., translated into English as On a Heuristic Viewpoint
Concerning the Production and Transformation of Light (http:/ / lorentz. phl. jhu. edu/ AnnusMirabilis/ AeReserveArticles/ eins_lq. pdf). The
term "photon" was introduced in 1926.
[10] Taylor, J. R.; Zafiratos, C. D.; Dubson, M. A. (2004). Modern Physics for Scientists and Engineers. Prentice Hall. pp.1279.
ISBN0135897890.
[11] Actually there can be intensity-dependent effects, but at intensities achievable with non-laser sources these effects are unobservable.
[12] Dicke and Wittke, Introduction to Quantum Mechanics, p. 12
[13] Einstein's photoelectric effect equation can be derived and explained without requiring the concept of "photons". That is, the electromagnetic
radiation can be treated as a classical electromagnetic wave, as long as the electrons in the material are treated by the laws of quantum
mechanics. The results are quantitatively correct for thermal light sources (the sun, incandescent lamps, etc) both for the rate of electron
emission as well as their angular distribution. For more on this point, see http:/ / ntrs. nasa. gov/ archive/ nasa/ casi. ntrs. nasa. gov/
19680009569_1968009569. pdf
[14] The classical model of the atom is called the planetary model, or sometimes the Rutherford model after Ernest Rutherford who proposed it in
1911, based on the Geiger-Marsden gold foil experiment which first demonstrated the existence of the nucleus.
[15] Taylor, J. R.; Zafiratos, C. D.; Dubson, M. A. (2004). Modern Physics for Scientists and Engineers. Prentice Hall. pp.1478.
ISBN0135897890.
[16] McEvoy, J. P.; Zarate, O. (2004). Introducing Quantum Theory. Totem Books. pp.7089, especially p. 89. ISBN1840465778.
[17] World Book Encyclopedia, page 6, 2007.
[18] Dicke and Wittke, Introduction to Quantum Mechanics, p. 10f.
[19] In this case, the energy of the electron is the sum of its kinetic and potential energies. The electron has kinetic energy by virtue of its actual
motion around the nucleus, and potential energy because of its electromagnetic interaction with the nucleus.
[20] The model can be easily modified to account of the emission spectrum of any system consisting of a nucleus and a single electron (that is,
ions such as He+ or O7+ which contain only one electron).
[21] J. P. McEvoy and Oscar Zarate (2004). Introducing Quantum Theory. Totem Books. p.110f. ISBN1-84046-577-8.
[22] Aezel, Amir D., Entanglrment, p. 51f. (Penguin, 2003) ISBN 0-452-28457
[23] J. P. McEvoy and Oscar Zarate (2004). Introducing Quantum Theory. Totem Books. p.114. ISBN1-84046-577-8.

30

Introduction to Quantum Mechanics


[24] Heisenberg's Nobel Prize citation (http:/ / nobelprize. org/ nobel_prizes/ physics/ laureates/ 1932/ )
[25] For a somewhat more sophisticated look at how Heisenberg transitioned from the old quantum theory and classical physics to the new
quantum mechanics, see Heisenberg's entryway to matrix mechanics.
[26] W. Moore, Schrdinger: Life and Thought, Cambridge University Press (1989), p. 222.
[27] Heisenberg first published his work on the uncertainty principle in the leading German physics journal Zeitschrift fr Physik: Heisenberg,
W. (1927). "ber den anschaulichen Inhalt der quantentheoretischen Kinematik und Mechanik". Z. Phys. 43 (34): 172198.
doi:10.1007/BF01397280.
[28] Nobel Prize in Physics presentation speech, 1932 (http:/ / nobelprize. org/ nobel_prizes/ physics/ laureates/ 1932/ press. html)
[29] Linus Pauling, The Nature of the Chemical Bond, p. 47
[30] E. Schrdinger, Proceedings of the Cambridge Philosophical Society, 31 (1935), p. 555says: "When two systems, of which we know the
states by their respective representation, enter into a temporary physical interaction due to known forces between them and when after a time
of mutual influence the systems separate again, then they can no longer be described as before, viz., by endowing each of them with a
representative of its own. I would not call that one but rather the characteristic trait of quantum mechanics."
[31] "Quantum Nonlocality and the Possibility of Superluminal Effects", John G. Cramer, http:/ / www. npl. washington. edu/ npl/ int_rep/
qm_nl. html

References
Bernstein, Jeremy (2005). "Max Born and the quantum theory". American Journal of Physics 73 (11).
Beller, Mara (2001). Quantum Dialogue: The Making of a Revolution. University of Chicago Press.
Bohr, Niels (1958). Atomic Physics and Human Knowledge. John Wiley & Sons. ASINB00005VGVF.
ISBN0486479285. OCLC530611.
de Broglie, Louis (1953). The Revolution in Physics. Noonday Press. LCCN53010401.
Einstein, Albert (1934). Essays in Science. Philosophical Library. ISBN0486470113. LCCN55003947.
Feigl, Herbert; Brodbeck, May (1953). Readings in the Philosophy of Science. Appleton-Century-Crofts.
ISBN0390304883. LCCN53006438.
Feynman, Richard P. (1949). "Space-Time Approach to Quantum Electrodynamics" (http://www.physics.
princeton.edu/~mcdonald/examples/QED/feynman_pr_76_769_49.pdf). Physical Review 76 (6): 769789.
doi:10.1103/PhysRev.76.769.
Fowler, Michael (1999). The Bohr Atom. University of Virginia.
Heisenberg, Werner (1958). Physics and Philosophy. Harper and Brothers. ISBN0061305499. LCCN99010404.
Lakshmibala, S. (2004). "Heisenberg, Matrix Mechanics and the Uncertainty Principle". Resonance, Journal of
Science Education 9 (8).
Liboff, Richard L. (1992). Introductory Quantum Mechanics (2nd ed.).
Lindsay, Robert Bruce; Margenau, Henry (1957). Foundations of Physics. Dover. ISBN0918024188.
LCCN57014416.
McEvoy, J. P.; Zarate, Oscar. Introducing Quantum Theory. ISBN1-874166-37-4.
Nave, Carl Rod (2005). "Quantum Physics" (http://hyperphysics.phy-astr.gsu.edu/hbase/quacon.
html#quacon). HyperPhysics. Georgia State University.
Peat, F. David (2002). From Certainty to Uncertainty: The Story of Science and Ideas in the Twenty-First
Century. Joseph Henry Press.
Reichenbach, Hans (1944). Philosophic Foundations of Quantum Mechanics. University of California Press.
ISBN0486404595. LCCNa44004471.
Schlipp, Paul Arthur (1949). Albert Einstein: Philosopher-Scientist. Tudor Publishing Company.
LCCN50005340.
Scientific American Reader, 1953.
Sears, Francis Weston (1949). Optics (3rd ed.). Addison-Wesley. ISBN0195046013. LCCN51001018.
Shimony, A. (1983). "(title not given in citation)". Foundations of Quantum Mechanics in the Light of New
Technology (S. Kamefuchi et al., eds.). Tokyo: Japan Physical Society. pp.225.; cited in: Popescu, Sandu; Daniel
Rohrlich (1996). "Action and Passion at a Distance: An Essay in Honor of Professor Abner Shimony".
arXiv:quant-ph/9605004[quant-ph].

31

Introduction to Quantum Mechanics


Tavel, Morton; Tavel, Judith (illustrations) (2002). Contemporary physics and the limits of knowledge (http://
books.google.com/?id=SELS0HbIhjYC&pg=PA200&dq=Wave+function+collapse). Rutgers University
Press. ISBN9780813530772.
Van Vleck, J. H.,1928, "The Correspondence Principle in the Statistical Interpretation of Quantum Mechanics,"
Proc. Nat. Acad. Sci. 14: 179.
Wheeler, John Archibald; Feynman, Richard P. (1949). "Classical Electrodynamics in Terms of Direct
Interparticle Action". Reviews of Modern Physics 21 (3): 425433. doi:10.1103/RevModPhys.21.425.
Wieman, Carl; Perkins, Katherine (2005). "Transforming Physics Education". Physics Today.
Westmoreland; Benjamin Schumacher (1998). "Quantum Entanglement and the Nonexistence of Superluminal
Signals". arXiv:quant-ph/9801014[quant-ph].
Bronner, Patrick; Strunz, Andreas; Silberhorn, Christine; Meyn, Jan-Peter (2009). "Demonstrating quantum
random with single photons". European Journal of Physics 30 (5): 11891200. doi:10.1088/0143-0807/30/5/026.

Further reading
The following titles, all by working physicists, attempt to communicate quantum theory to lay people, using a
minimum of technical apparatus.
Jim Al-Khalili (2003) Quantum: A Guide for the Perplexed. Weidenfield & Nicholson.
Richard Feynman (1985) QED: The Strange Theory of Light and Matter. Princeton University Press. ISBN
0-691-08388-6
Ford, Kenneth (2005) The Quantum World. Harvard Univ. Press. Includes elementary particle physics.
Ghirardi, GianCarlo (2004) Sneaking a Look at God's Cards, Gerald Malsbary, trans. Princeton Univ. Press. The
most technical of the works cited here. Passages using algebra, trigonometry, and bra-ket notation can be passed
over on a first reading.
Tony Hey and Walters, Patrick (2003) The New Quantum Universe. Cambridge Univ. Press. Includes much about
the technologies quantum theory has made possible.
N. David Mermin (1990) Spooky actions at a distance: mysteries of the QT in his Boojums all the way through.
Cambridge Univ. Press: 110-176. The author is a rare physicist who tries to communicate to philosophers and
humanists.
Roland Omnes (1999) Understanding Quantum Mechanics. Princeton Univ. Press.
Victor Stenger (2000) Timeless Reality: Symmetry, Simplicity, and Multiple Universes. Buffalo NY: Prometheus
Books. Chpts. 5-8.
Martinus Veltman (2003) Facts and Mysteries in Elementary Particle Physics. World Scientific Publishing
Company.
A website with good introduction to Quantum mechanics can be found here. (http://www.chem1.com/acad/
webtext/atoms/atpt-4.html)

External links
Takada, Kenjiro, Emeritus professor at Kyushu University, " Microscopic World -- Introduction to Quantum
Mechanics. (http://www2.kutl.kyushu-u.ac.jp/seminar/MicroWorld1_E/MicroWorld_1_E.html)"
Quantum Theory. (http://www.encyclopedia.com/doc/1E1-quantumt.html)
Quantum Mechanics. (http://www.aip.org/history/heisenberg/p07.htm)
Planck's original paper (http://dbhs.wvusd.k12.ca.us/webdocs/Chem-History/Planck-1901/Planck-1901.
html) on Planck's constant.
Everything you wanted to know about the quantum world. (http://www.newscientist.com/channel/
fundamentals/quantum-world) From the New Scientist.
This Quantum World. (http://thisquantumworld.com/ht/index.php)

32

Introduction to Quantum Mechanics


The Quantum Exchange (http://www.compadre.org/quantum) (tutorials and open source learning software).
Theoretical Physics wiki (http://theoreticalphysics.wetpaint.com)
" Uncertainty Principle, (http://www.thebigview.com/spacetime/index.html)" a recording of Werner
Heisenberg's voice.
Single and double slit interference (http://class.phys.psu.edu/251Labs/10_Interference_&_Diffraction/
Single_and_Double-Slit_Interference.pdf)
Time-Evolution f a Wavepacket in a Square Well (http://demonstrations.wolfram.com/
TimeEvolutionOfAWavepacketInASquareWell/) An animated demonstration of a wave packet dispersion over
time.
Experiments with single photons (http://www.didaktik.physik.uni-erlangen.de/quantumlab/english/) An
introduction into quantum physics with interactive experiments
Hitachi video recording of double-slit experiment done with electrons. You can see the interference pattern build
up over time. (http://www.youtube.com/watch?v=oxknfn97vFE)

33

34

2. The Quantum Theories


Old Quantum Theory
The old quantum theory was a collection of results from the years 19001925 which predate modern quantum
mechanics. The theory was never complete or self-consistent, but was a collection of heuristic prescriptions which
are now understood to be the first quantum corrections to classical mechanics.[1] The Bohr model was the focus of
study, and Arnold Sommerfeld[2] made a crucial contribution by quantizing the z-component of the angular
momentum, which in the old quantum era was inappropriately called space quantization (Richtungsquantelung).
This allowed the orbits of the electron to be ellipses instead of circles, and introduced the concept of quantum
degeneracy. The theory would have correctly explained the Zeeman effect, except for the issue of electron spin.
The main tool was Bohr Sommerfeld quantization, a procedure for selecting out certain discrete set of states of a
classical integrable motion as allowed states. These are like the allowed orbits of the Bohr model of the atom; the
system can only be in one of these states and not in any states in between. The theory did not extend to chaotic
motions, because it required a full multiply periodic trajectory of the classical system for all time in order to pose the
quantum conditions.
The old quantum theory lives on as an approximation technique in quantum mechanics, called the WKB method.
Semi-classical approximations were a popular research subject in the 1970s and 1980s, after Gutzwiller discovered a
semi-classical description for systems which are classically chaotic (see quantum chaos).

Basic principles
The basic idea of the old quantum theory is that the motion in an atomic system is quantized, or discrete. The system
obeys classical mechanics except that not every motion is allowed, only those motions which obey the old quantum
condition:

where the

are the momenta of the system and the

are the corresponding coordinates. The quantum numbers

are integers and the integral is taken over one period of the motion. The integral is an area in phase space, which
is a quantity called the action, which is quantized in units of Planck's constant. For this reason, Planck's constant was
often called the quantum of action.
In order for the old quantum condition to make sense, the classical motion must be separable, meaning that there are
separate coordinates in terms of which the motion is periodic. The periods of the different motions do not have to
be the same, they can even be incommensurate, but there must be a set of coordinates where the motion decomposes
in a multi-periodic way.
The motivation for the old quantum condition was the correspondence principle, complemented by the physical
observation that the quantities which are quantized must be adiabatic invariants. Given Planck's quantization rule for
the harmonic oscillator, either condition determines the correct classical quantity to quantize in a general system up
to an additive constant.

Old Quantum Theory

35

Examples
Harmonic oscillator
The simplest system in the old quantum theory is the Harmonic oscillator, whose Hamiltonian is:

The level sets of H are the orbits, and the quantum condition is that the area enclosed by an orbit in phase space is an
integer. It follows that the energy is quantized according to the Planck rule:

a result which was known well before, and used to formulate the old quantum condition.
The thermal properties of a quantized oscillator may be found by averaging the energy in each of the discrete states
assuming that they are occupied with a Boltzmann weight:

kT is Boltzmann constant times the absolute temperature, which is the temperature as measured in more natural units
of energy. The quantity
is more fundamental in thermodynamics than the temperature, because it is the
thermodynamic potential associated to the energy.
From this expression, it is easy to see that for large values of

, for very low temperatures, the average energy U in

the Harmonic oscillator approaches zero very quickly, exponentially fast. The reason is that kT is the typical energy
of random motion at temperature T, and when this is smaller than
, there is not enough energy to give the
oscillator even one quantum of energy. So the oscillator stays in its ground state, storing next to no energy at all.
This means that at very cold temperatures, the change in energy with respect to beta, or equivalently the change in
energy with respect to temperature, is also exponentially small. The change in energy with respect to temperature is
the specific heat, so the specific heat is exponentially small at low temperatures, going to zero like

At small values of

, at high temperatures, the average energy U is equal to

. This reproduces the

equipartition theorem of classical thermodynamics--- every harmonic oscillator at temperature T has energy kT on
average. This means that the specific heat of an oscillator is constant in classical mechanics and equal to k. For a
collection of atoms connected by springs, a reasonable model of a solid, the total specific heat is equal to the total
number of oscillators times k. There are overall three oscillators for each atom, corresponding to the three possible
directions of independent oscillations in three dimensions. So the specific heat of a classical solid is always 3k per
atom, or in chemistry units, 3R per mole of atoms.
Monatomic solids at room temperatures have approximately the same specific heat of 3k per atom, but at low
temperatures they don't. The specific heat is smaller at colder temperatures, and it goes to zero at absolute zero. This
is true for all material systems, and this observation is called the third law of thermodynamics. Classical mechanics
cannot explain the third law, because in classical mechanics the specific heat is independent of the temperature.
This contradiction between classical mechanics and the specific heat of cold materials was noted by James Clerk
Maxwell in the 19th century, and remained a deep puzzle for those who advocated an atomic theory of matter.
Einstein resolved this problem in 1906 by proposing that atomic motion is quantized. This was the first application
of quantum theory to a mechanical systems. A short while later, Debye gave a quantitative theory of solid specific
heats in terms of quantized oscillators with various frequencies (see Einstein solid and Debye model).

Old Quantum Theory

36

One dimensional potential


One dimensional problems are easy to solve. At any energy E, the value of the momentum p is found from the
conservation equation:

which is integrated over all values of q between the classical turning points, the places where the momentum
vanishes. The integral is easiest for a particle in a box of length L, where the quantum condition is:

which gives the allowed momenta:

and the energy levels

Another easy case to solve with the old quantum theory is a linear potential on the positive halfline, the constant
confining force F binding a particle to an impenetrable wall. This case is much more difficult in the full quantum
mechanical treatment, and unlike the other examples, the semiclassical answer here is not exact but approximate,
becoming more accurate at large quantum numbers.

so that the quantum condition is:

Which determines the energy levels.

Rotator
Another simple system is the rotator. A rotator consists of a mass M at the end of a massless rigid rod of length R and
in two dimensions has the Lagrangian:

which determines that the momentum J conjugate to


requires that J multiplied by the period of

, the polar angle,

. The old quantum condition

is an integer multiple of Planck's constant:

the angular momentum to be an integer multiple of


was enough to determine the energy levels.

. In the Bohr model, this restriction imposed on circular orbits

In three dimensions, a rigid rotator can be described by two angles and , where is the inclination relative
to an arbitrarily chosen z-axis while is the rotator angle in the projection to the xy plane. The kinetic energy is
again the only contribution to the Lagrangian:

And the conjugate momenta are

and

. The equation of motion for

is trivial:

is a constant:

Old Quantum Theory


which is the z-component of the angular momentum. The quantum condition demands that the integral of the
constant as varies from 0 to
is an integer multiple of h:

And m is called the magnetic quantum number, because the z component of the angular momentum is the magnetic
moment of the rotator along the z direction in the case where the particle at the end of the rotator is charged.
Since the three dimensional rotator is rotating about an axis, the total angular momentum should be restricted in the
same way as the two-dimensional rotator. The two quantum conditions restrict the total angular momentum and the
z-component of the angular momentum to be the integers l,m. This condition is reproduced in modern quantum
mechanics, but in the era of the old quantum theory it led to a paradox: how can the orientation of the angular
momentum relative to the arbitrarily chosen z-axis be quantized? This seems to pick out a direction in space.
This phenomenon, the quantization of angular momentum about an axis, was given the name space quantization,
because it seemed incompatible with rotational invariance. In modern quantum mechanics, the angular momentum is
quantized the same way, but the discrete states of definite angular momentum in any one orientation are quantum
superpositions of the states in other orientations, so that the process of quantization does not pick out a preferred
axis. For this reason, the name "space quantization" fell out of favor, and the same phenomenon is now called the
quantization of angular momentum.

Hydrogen atom
The angular part of the Hydrogen atom is just the rotator, and gives the quantum numbers l and m. The only
remaining variable is the radial coordinate, which executes a periodic one dimensional potential motion, which can
be solved.
For a fixed value of the total angular momentum L, the Hamiltonian for a classical Kepler problem is (the unit of
mass and unit of energy redefined to absorb two constants):

Fixing the energy to be constant and solving for the radial momentum p, the quantum condition integral is:

which is elementary, and gives a new quantum number k which determines the energy in combination with l. The
energy is:

and it only depends on the sum of k and l, which is the principal quantum number n. Since k is positive, the allowed
values of l for any given n are no bigger than n. The energies reproduce those in the Bohr model, except with the
correct quantum mechanical multiplicities, with some ambiguity at the extreme values.
The semiclassical hydrogen atom is called the Sommerfeld model, and its orbits are ellipses of various sizes at
discrete inclinations. The Sommerfeld model predicted that the magnetic moment of an atom measured along an axis
will only take on discrete values, a result which seems to contradict rotational invariance but which was confirmed
by the SternGerlach experiment. BohrSommerfeld theory is a part of the development of quantum mechanics
and describes the possibility of atomic energy levels being split by a magnetic field.

37

Old Quantum Theory

38

Relativistic orbit
Arnold Sommerfeld derived the relativistic solution of atomic energy levels.[3] We will start this derivation with the
relativistic equation for energy in the electric potential

After substitution

we get

For momentum

and their ratio

the equation of motion is (see Binet

equation)

with solution

The angular shift of periapsis per revolution is given by

With the quantum conditions

and

we will obtain energies

where

is the fine-structure constant. This solution is same as the solution of the Dirac equation[4].

De Broglie waves
In 1905, Einstein noted that the entropy of the quantized electromagnetic field oscillators in a box is, for short
wavelength, equal to the entropy of a gas of point particles in the same box. The number of point particles is equal to
the number of quanta. Einstein concluded that the quanta were localizable objects, particles of light, and named them
photons.
Einstein's theoretical argument was based on thermodynamics, on counting the number of states, and so was not
completely convincing. Nevertheless, he concluded that light had attributes of both waves and particles, more
precisely that an electromagnetic standing wave with frequency with the quantized energy:

Old Quantum Theory

39

should be thought of as consisting of n photons each with an energy


photons were related to the wave.

. Einstein could not describe how the

The photons have momentum as well as energy, and the momentum had to be

where

is the wavenumber of

the electromagnetic wave. This is required by relativity, because the momentum and energy form a four-vector, as do
the frequency and wave-number.
In 1924, as a PhD candidate, Louis de Broglie proposed a new interpretation of the quantum condition. He suggested
that all matter, electrons as well as photons, are described by waves obeying the relations.

or, expressed in terms of wavelength

instead,

He then noted that the quantum condition:

counts the change in phase for the wave as it travels along the classical orbit, and requires that it be an integer
multiple of
. Expressed in wavelengths, the number of wavelengths along a classical orbit must be an integer.
This is the condition for constructive interference, and it explained the reason for quantized orbitsthe matter waves
make standing waves only at discrete frequencies, at discrete energies.
For example, for a particle confined in a box, a standing wave must fit an integer number of wavelengths between
twice the distance between the walls. The condition becomes:

so that the quantized momenta are:

reproducing the old quantum energy levels.


This development was given a more mathematical form by Einstein, who noted that the phase function for the
waves:
in a mechanical system should be identified with the solution to the HamiltonJacobi equation, an
equation which even Hamilton considered to be a short-wavelength limit of a wave mechanics.
These ideas led to the development of Schrdinger equation.

Kramers transition matrix


The old quantum theory was formulated only for special mechanical systems which could be separated into action
angle variables which were periodic. It did not deal with the emission and absorption of radiation. Nevertheless,
Hendrik Kramers was able to find heuristics for describing how emission and absorption should be calculated.
Kramers suggested that the orbits of a quantum system should be Fourier analyzed, decomposed into harmonics at
multiples of the orbit frequency:

The index n describes the quantum numbers of the orbit, it would be nlm in the Sommerfeld model. The frequency
is the angular frequency of the orbit
while k is an index for the Fourier mode. Bohr had suggested that the
k-th harmonic of the classical motion correspond to the transition from level n to level nk.
Kramers proposed that the transition between states were analogous to classical emission of radiation, which
happens at frequencies at multiples of the orbit frequencies. The rate of emission of radiation is proportional to
, as it would be in classical mechanics. The description was approximate, since the Fourier components did

Old Quantum Theory


not have frequencies that exactly match the energy spacings between levels.
This idea led to the development of matrix mechanics.

History
The old quantum theory was sparked by the work of Max Planck on the emission and absorption of light, and began
in earnest after the work of Albert Einstein on the specific heats of solids. Einstein, followed by Debye, applied
quantum principles to the motion of atoms, explaining the specific heat anomaly.
In 1913, Niels Bohr identified the correspondence principle and used it to formulate a model of the Hydrogen atom
which explained the line spectrum. In the next few years Arnold Sommerfeld extended the quantum rule to arbitrary
integrable systems making use of the principle of adiabatic invariance of the quantum numbers introduced by
Lorentz and Einstein. Sommerfeld's model was much closer to the modern quantum mechanical picture than Bohr's.
Throughout the 1910s and well into the 1920s, many problems were attacked using the old quantum theory with
mixed results. Molecular rotation and vibration spectra were understood and the electron's spin was discovered,
leading to the confusion of half-integer quantum numbers. Max Planck introduced the zero point energy and Arnold
Sommerfeld semiclassically quantized the relativistic hydrogen atom. Hendrik Kramers explained the Stark effect.
Bose and Einstein gave the correct quantum statistics for photons.
Kramers gave a prescription for calculating transition probabilities between quantum states in terms of Fourier
components of the motion, ideas which were extended in collaboration with Werner Heisenberg to a semiclassical
matrix-like description of atomic transition probabilities. Heisenberg went on to reformulate all of quantum theory in
terms of a version of these transition matrices, creating Matrix mechanics.
In 1924, Louis de Broglie introduced the wave theory of matter, which was extended to a semiclassical equation for
matter waves by Albert Einstein a short time later. In 1926 Erwin Schrdinger found a completely quantum
mechanical wave-equation, which reproduced all the successes of the old quantum theory without ambiguities and
inconsistencies. Schrdinger's wave mechanics developed separately from matrix mechanics until Schrdinger and
others proved that the two methods predicted the same experimental consequences. Paul Dirac later proved in 1926
that both methods can be obtained from a more general method called transformation theory.
Matrix mechanics and wave mechanics put an end to the era of the old-quantum theory.

References
[1]
[2]
[3]
[4]

ter Haar, D. (1967). The Old Quantum Theory. Pergamon Press. pp.206. ISBN0080121012.
Sommerfeld, Arnold (1919). Atombau und Spektrallinien'. Braunschweig: Friedrich Vieweg und Sohn. ISBN3871444847.
Arnold Sommerfeld (1924). Atombau und Spektrallinien. Braunschweig. ISBN3871444847.
http:/ / www. iop. org/ EJ/ article/ 1063-7869/ 47/ 5/ L06/ PHU_47_5_L06. pdf

Further reading
Thewlis, J., ed (1962). Encyclopaedic Dictionary of Physics.

40

Quantum Mechanics after 1925

Quantum Mechanics after 1925


Quantum mechanics, also known as
quantum physics or quantum theory, is a
branch of physics providing a mathematical
description of the dual particle-like and
wave-like behaviour and interaction of matter
and energy. Quantum mechanics describes
the time evolution of physical systems via a
mathematical structure called the wave
function. The wave function encapsulates the
probability that the system is to be found in a
given state at a given time. Quantum
mechanics also allows one to calculate the
effect on the system of making measurements
of properties of the system by defining the
effect of those measurements on the wave
function. This leads to the well-known
uncertainty principle as well as enduring
debate over the role of the experimenter,
epitomised in the Schrodinger's Cat thought
experiment.
Quantum mechanics differs significantly
from classical mechanics in its predictions
Some trajectories of a harmonic oscillator (a ball attached to a spring) in classical
when the scale of observations becomes
mechanics (A-B) and quantum mechanics (C-H). In quantum mechanics, the
comparable to the atomic and sub-atomic
position of the ball is represented by a wave (called the wavefunction), with real
scale, the so-called quantum realm. However,
part shown in blue and imaginary part in red. Some of the trajectories, such as
C,D,E,F, are standing waves (or "stationary states"). Each standing-wave
many macroscopic properties of systems can
frequency
is proportional to a possible energy level of the oscillator. This "energy
only be fully understood and explained with
quantization" does not occur in classical physics, where the oscillator can have
the use of quantum mechanics. Phenomena
any energy.
such as superconductivity, the properties of
materials such as semiconductors and nuclear
and chemical reaction mechanisms observed as macroscopic behaviour, cannot be explained using classical
mechanics.
The term was coined by Max Planck, and derives from the observation that some physical quantities can be changed
only by discrete amounts, or quanta, as multiples of the Planck constant, rather than being capable of varying
continuously or by any arbitrary amount. For example, the angular momentum, or more generally the action, of an
electron bound into an atom or molecule is quantized. Although an unbound electron does not exhibit quantized
energy levels, one which is bound in an atomic orbital has quantized values of angular momentum. In the context of
quantum mechanics, the waveparticle duality of energy and matter and the uncertainty principle provide a unified
view of the behavior of photons, electrons and other atomic-scale objects.
The mathematical formulations of quantum mechanics are abstract. Similarly, the implications are often
counter-intuitive in terms of classical physics. The centerpiece of the mathematical formulation is the wavefunction
(defined by Schrdinger's wave equation), which describes the probability amplitude of the position and momentum
of a particle. Mathematical manipulations of the wavefunction usually involve the bra-ket notation, which requires

41

Quantum Mechanics after 1925


an understanding of complex numbers and linear functionals. The wavefunction treats the object as a quantum
harmonic oscillator and the mathematics is akin to that of acoustic resonance.
Many of the results of quantum mechanics do not have models that are easily visualized in terms of classical
mechanics; for instance, the ground state in the quantum mechanical model is a non-zero energy state that is the
lowest permitted energy state of a system, rather than a traditional classical system that is thought of as simply being
at rest with zero kinetic energy.
Fundamentally, it attempts to explain the peculiar behaviour of matter and energy at the subatomic levelan attempt
which has produced more accurate results than classical physics in predicting how individual particles behave. But
many unexplained anomalies remain.
Historically, the earliest versions of quantum mechanics were formulated in the first decade of the 20th Century,
around the time that atomic theory and the corpuscular theory of light as interpreted by Einstein first came to be
widely accepted as scientific fact; these later theories can be viewed as quantum theories of matter and
electromagnetic radiation.
Following Schrdinger's breakthrough in deriving his wave equation in the mid-1920s, quantum theory was
significantly reformulated away from the old quantum theory, towards the quantum mechanics of Werner
Heisenberg, Max Born, Wolfgang Pauli and their associates, becoming a science of probabilities based upon the
Copenhagen interpretation of Niels Bohr. By 1930, the reformulated theory had been further unified and formalized
by the work of Paul Dirac and John von Neumann, with a greater emphasis placed on measurement, the statistical
nature of our knowledge of reality, and philosophical speculations about the role of the observer.
The Copenhagen interpretation quickly became (and remains) the orthodox interpretation. However, due to the
absence of conclusive experimental evidence there are also many competing interpretations.
Quantum mechanics has since branched out into almost every aspect of physics, and into other disciplines such as
quantum chemistry, quantum electronics, quantum optics and quantum information science. Much 19th Century
physics has been re-evaluated as the classical limit of quantum mechanics and its more advanced developments in
terms of quantum field theory, string theory, and speculative quantum gravity theories.

History
The history of quantum mechanics dates back to the 1838 discovery of cathode rays by Michael Faraday. This was
followed by the 1859 statement of the black body radiation problem by Gustav Kirchhoff, the 1877 suggestion by
Ludwig Boltzmann that the energy states of a physical system can be discrete, and the 1900 quantum hypothesis of
Max Planck.[1] Planck's hypothesis that energy is radiated and absorbed in discrete "quanta", or "energy elements",
precisely matched the observed patterns of black body radiation. According to Planck, each energy element E is
proportional to its frequency :

where h is Planck's constant. Planck cautiously insisted that this was simply an aspect of the processes of absorption
and emission of radiation and had nothing to do with the physical reality of the radiation itself.[2] However, in 1905
Albert Einstein interpreted Planck's quantum hypothesis realistically and used it to explain the photoelectric effect, in
which shining light on certain materials can eject electrons from the material. Einstein postulated that light itself
consists of individual quanta of energy, later called photons.[3]
The foundations of quantum mechanics were established during the first half of the twentieth century by Niels Bohr,
Werner Heisenberg, Max Planck, Louis de Broglie, Albert Einstein, Erwin Schrdinger, Max Born, John von
Neumann, Paul Dirac, Wolfgang Pauli, David Hilbert, and others. In the mid-1920s, developments in quantum
mechanics led to its becoming the standard formulation for atomic physics. In the summer of 1925, Bohr and
Heisenberg published results that closed the "Old Quantum Theory". Out of deference to their dual state as particles,
light quanta came to be called photons (1926). From Einstein's simple postulation was born a flurry of debating,

42

Quantum Mechanics after 1925


theorizing and testing. Thus the entire field of quantum physics emerged, leading to its wider acceptance at the Fifth
Solvay Conference in 1927.
The other exemplar that led to quantum mechanics was the study of electromagnetic waves such as light. When it
was found in 1900 by Max Planck that the energy of waves could be described as consisting of small packets or
quanta, Albert Einstein further developed this idea to show that an electromagnetic wave such as light could be
described as a particle - later called the photon - with a discrete energy that was dependent on its frequency. This led
to a theory of unity between subatomic particles and electromagnetic waves called waveparticle duality in which
particles and waves were neither one nor the other, but had certain properties of both.
While quantum mechanics traditionally described the world of the very small, it is also needed to explain certain
recently investigated macroscopic systems such as superconductors and superfluids.
The word quantum derives from Latin, meaning "how great" or "how much".[4] In quantum mechanics, it refers to a
discrete unit that quantum theory assigns to certain physical quantities, such as the energy of an atom at rest (see
Figure 1). The discovery that particles are discrete packets of energy with wave-like properties led to the branch of
physics dealing with atomic and sub-atomic systems which is today called quantum mechanics. It is the underlying
mathematical framework of many fields of physics and chemistry, including condensed matter physics, solid-state
physics, atomic physics, molecular physics, computational physics, computational chemistry, quantum chemistry,
particle physics, nuclear chemistry, and nuclear physics.[5] Some fundamental aspects of the theory are still actively
studied.[6]
Quantum mechanics is essential to understand the behavior of systems at atomic length scales and smaller. For
example, if classical mechanics governed the workings of an atom, electrons would rapidly travel towards and
collide with the nucleus, making stable atoms impossible. However, in the natural world the electrons normally
remain in an uncertain, non-deterministic "smeared" (waveparticle wave function) orbital path around or through
the nucleus, defying classical electromagnetism.[7]
Quantum mechanics was initially developed to provide a better explanation of the atom, especially the differences in
the spectra of light emitted by different isotopes of the same element. The quantum theory of the atom was
developed as an explanation for the electron remaining in its orbit, which could not be explained by Newton's laws
of motion and Maxwell's laws of classical electromagnetism.
Broadly speaking, quantum mechanics incorporates four classes of phenomena for which classical physics cannot
account:

The quantization of certain physical properties


Waveparticle duality
The uncertainty principle
Quantum entanglement

Mathematical formulations
In the mathematically rigorous formulation of quantum mechanics developed by Paul Dirac[8] and John von
Neumann,[9] the possible states of a quantum mechanical system are represented by unit vectors (called "state
vectors"). Formally, these reside in a complex separable Hilbert space (variously called the "state space" or the
"associated Hilbert space" of the system) well defined up to a complex number of norm 1 (the phase factor). In other
words, the possible states are points in the projective space of a Hilbert space, usually called the complex projective
space. The exact nature of this Hilbert space is dependent on the system; for example, the state space for position and
momentum states is the space of square-integrable functions, while the state space for the spin of a single proton is
just the product of two complex planes. Each observable is represented by a maximally Hermitian (precisely: by a
self-adjoint) linear operator acting on the state space. Each eigenstate of an observable corresponds to an eigenvector
of the operator, and the associated eigenvalue corresponds to the value of the observable in that eigenstate. If the

43

Quantum Mechanics after 1925


operator's spectrum is discrete, the observable can only attain those discrete eigenvalues.
In the formalism of quantum mechanics, the state of a system at a given time is described by a complex wave
function, also referred to as state vector in a complex vector space.[10] This abstract mathematical object allows for
the calculation of probabilities of outcomes of concrete experiments. For example, it allows one to compute the
probability of finding an electron in a particular region around the nucleus at a particular time. Contrary to classical
mechanics, one can never make simultaneous predictions of conjugate variables, such as position and momentum,
with accuracy. For instance, electrons may be considered to be located somewhere within a region of space, but with
their exact positions being unknown. Contours of constant probability, often referred to as "clouds", may be drawn
around the nucleus of an atom to conceptualize where the electron might be located with the most probability.
Heisenberg's uncertainty principle quantifies the inability to precisely locate the particle given its conjugate
momentum.[11]
According to one interpretation, as the result of a measurement the wave function containing the probability
information for a system collapses from a given initial state to a particular eigenstate. The possible results of a
measurement are the eigenvalues of the operator representing the observable which explains the choice of
Hermitian operators, for which all the eigenvalues are real. We can find the probability distribution of an observable
in a given state by computing the spectral decomposition of the corresponding operator. Heisenberg's uncertainty
principle is represented by the statement that the operators corresponding to certain observables do not commute.
The probabilistic nature of quantum mechanics thus stems from the act of measurement. This is one of the most
difficult aspects of quantum systems to understand. It was the central topic in the famous Bohr-Einstein debates, in
which the two scientists attempted to clarify these fundamental principles by way of thought experiments. In the
decades after the formulation of quantum mechanics, the question of what constitutes a "measurement" has been
extensively studied. Newer interpretations of quantum mechanics have been formulated that do away with the
concept of "wavefunction collapse"; see, for example, the relative state interpretation. The basic idea is that when a
quantum system interacts with a measuring apparatus, their respective wavefunctions become entangled, so that the
original quantum system ceases to exist as an independent entity. For details, see the article on measurement in
quantum mechanics.[12] Generally, quantum mechanics does not assign definite values. Instead, it makes predictions
using probability distributions; that is, it describes the probability of obtaining possible outcomes from measuring an
observable. Often these results are skewed by many causes, such as dense probability clouds[13] or quantum state
nuclear attraction.[14] [15] Naturally, these probabilities will depend on the quantum state at the "instant" of the
measurement. Hence, uncertainty is involved in the value. There are, however, certain states that are associated with
a definite value of a particular observable. These are known as eigenstates of the observable ("eigen" can be
translated from German as meaning inherent or characteristic).[16]
In the everyday world, it is natural and intuitive to think of everything (every observable) as being in an eigenstate.
Everything appears to have a definite position, a definite momentum, a definite energy, and a definite time of
occurrence. However, quantum mechanics does not pinpoint the exact values of a particle's position and momentum
(since they are conjugate pairs) or its energy and time (since they too are conjugate pairs); rather, it only provides a
range of probabilities of where that particle might be given its momentum and momentum probability. Therefore, it
is helpful to use different words to describe states having uncertain values and states having definite values
(eigenstate). Usually, a system will not be in an eigenstate of the observable (particle) we are interested in. However,
if one measures the observable, the wavefunction will instantaneously be an eigenstate (or generalised eigenstate) of
that observable. This process is known as wavefunction collapse, a controversial and much debated process.[17] It
involves expanding the system under study to include the measurement device. If one knows the corresponding wave
function at the instant before the measurement, one will be able to compute the probability of collapsing into each of
the possible eigenstates. For example, the free particle in the previous example will usually have a wavefunction that
is a wave packet centered around some mean position x0, neither an eigenstate of position nor of momentum. When
one measures the position of the particle, it is impossible to predict with certainty the result.[12] It is probable, but not

44

Quantum Mechanics after 1925


certain, that it will be near x0, where the amplitude of the wave function is large. After the measurement is
performed, having obtained some result x, the wave function collapses into a position eigenstate centered at x.[18]
The time evolution of a quantum state is described by the Schrdinger equation, in which the Hamiltonian (the
operator corresponding to the total energy of the system) generates time evolution. The time evolution of wave
functions is deterministic in the sense that, given a wavefunction at an initial time, it makes a definite prediction of
what the wavefunction will be at any later time.[19]
During a measurement, on the other hand, the change of the wavefunction into another one is not deterministic; it is
unpredictable, i.e. random. A time-evolution simulation can be seen here.[20] [21] Wave functions can change as time
progresses. An equation known as the Schrdinger equation describes how wavefunctions change in time, a role
similar to Newton's second law in classical mechanics. The Schrdinger equation, applied to the aforementioned
example of the free particle, predicts that the center of a wave packet will move through space at a constant velocity,
like a classical particle with no forces acting on it. However, the wave packet will also spread out as time progresses,
which means that the position becomes more uncertain. This also has the effect of turning position eigenstates
(which can be thought of as infinitely sharp wave packets) into broadened wave packets that are no longer position
eigenstates.[22]
Some wave functions produce probability
distributions that are constant, or independent of
time, such as when in a stationary state of
constant energy, time drops out of the absolute
square of the wave function. Many systems that
are treated dynamically in classical mechanics are
described by such "static" wave functions. For
example, a single electron in an unexcited atom is
pictured classically as a particle moving in a
circular trajectory around the atomic nucleus,
whereas in quantum mechanics it is described by
a static, spherically symmetric wavefunction
surrounding the nucleus (Fig. 1). (Note that only
the lowest angular momentum states, labeled s,
are spherically symmetric).[23]
The Schrdinger equation acts on the entire
probability amplitude, not merely its absolute
Fig. 1: Probability densities corresponding to the wavefunctions of an
value. Whereas the absolute value of the
electron in a hydrogen atom possessing definite energy levels (increasing
from the top of the image to the bottom: n = 1, 2, 3, ...) and angular
probability amplitude encodes information about
momentum
(increasing across from left to right: s, p, d, ...). Brighter areas
probabilities, its phase encodes information about
correspond to higher probability density in a position measurement.
the interference between quantum states. This
Wavefunctions like these are directly comparable to Chladni's figures of
gives rise to the wave-like behavior of quantum
acoustic modes of vibration in classical physics and are indeed modes of
oscillation as well: they possess a sharp energy and thus a keen frequency.
states. It turns out that analytic solutions of
The
angular momentum and energy are quantized, and only take on discrete
Schrdinger's equation are only available for a
values like those shown (as is the case for resonant frequencies in
small number of model Hamiltonians, of which
acoustics).
the quantum harmonic oscillator, the particle in a
box, the hydrogen molecular ion and the
hydrogen atom are the most important representatives. Even the helium atom, which contains just one more electron
than hydrogen, defies all attempts at a fully analytic treatment. There exist several techniques for generating
approximate solutions. For instance, in the method known as perturbation theory one uses the analytic results for a
simple quantum mechanical model to generate results for a more complicated model related to the simple model by,

45

Quantum Mechanics after 1925


for example, the addition of a weak potential energy. Another method is the "semi-classical equation of motion"
approach, which applies to systems for which quantum mechanics produces weak deviations from classical behavior.
The deviations can be calculated based on the classical motion. This approach is important for the field of quantum
chaos.
There are numerous mathematically equivalent formulations of quantum mechanics. One of the oldest and most
commonly used formulations is the transformation theory proposed by Cambridge theoretical physicist Paul Dirac,
which unifies and generalizes the two earliest formulations of quantum mechanics, matrix mechanics (invented by
Werner Heisenberg)[24] [25] and wave mechanics (invented by Erwin Schrdinger).[26] In this formulation, the
instantaneous state of a quantum system encodes the probabilities of its measurable properties, or "observables".
Examples of observables include energy, position, momentum, and angular momentum. Observables can be either
continuous (e.g., the position of a particle) or discrete (e.g., the energy of an electron bound to a hydrogen atom).[27]
An alternative formulation of quantum mechanics is Feynman's path integral formulation, in which a
quantum-mechanical amplitude is considered as a sum over histories between initial and final states; this is the
quantum-mechanical counterpart of action principles in classical mechanics.

Interactions with other scientific theories


The rules of quantum mechanics are fundamental; they assert that the state space of a system is a Hilbert space and
that observables of that system are Hermitian operators acting on that space; they do not tell us which Hilbert space
or which operators. These can be chosen appropriately in order to obtain a quantitative description of a quantum
system. An important guide for making these choices is the correspondence principle, which states that the
predictions of quantum mechanics reduce to those of classical physics when a system moves to higher energies or,
equivalently, larger quantum numbers (i.e. whereas a single particle exhibits a degree of randomness, in systems
incorporating millions of particles averaging takes over and, at the high energy limit, the statistical probability of
random behaviour approaches zero). In other words, classical mechanics is simply a quantum mechanics of large
systems. This "high energy" limit is known as the classical or correspondence limit. One can even start from an
established classical model of a particular system, and attempt to guess the underlying quantum model that would
give rise to the classical model in the correspondence limit.
When quantum mechanics was originally formulated, it was applied to models whose correspondence limit was
non-relativistic classical mechanics. For instance, the well-known model of the quantum harmonic oscillator uses an
explicitly non-relativistic expression for the kinetic energy of the oscillator, and is thus a quantum version of the
classical harmonic oscillator.
Early attempts to merge quantum mechanics with special relativity involved the replacement of the Schrdinger
equation with a covariant equation such as the Klein-Gordon equation or the Dirac equation. While these theories
were successful in explaining many experimental results, they had certain unsatisfactory qualities stemming from
their neglect of the relativistic creation and annihilation of particles. A fully relativistic quantum theory required the
development of quantum field theory, which applies quantization to a field rather than a fixed set of particles. The
first complete quantum field theory, quantum electrodynamics, provides a fully quantum description of the
electromagnetic interaction. The full apparatus of quantum field theory is often unnecessary for describing
electrodynamic systems. A simpler approach, one employed since the inception of quantum mechanics, is to treat
charged particles as quantum mechanical objects being acted on by a classical electromagnetic field. For example,
the elementary quantum model of the hydrogen atom describes the electric field of the hydrogen atom using a
classical
Coulomb potential. This "semi-classical" approach fails if quantum fluctuations in the
electromagnetic field play an important role, such as in the emission of photons by charged particles.
Quantum field theories for the strong nuclear force and the weak nuclear force have been developed. The quantum
field theory of the strong nuclear force is called quantum chromodynamics, and describes the interactions of
subnuclear particles: quarks and gluons. The weak nuclear force and the electromagnetic force were unified, in their

46

Quantum Mechanics after 1925


quantized forms, into a single quantum field theory known as electroweak theory, by the physicists Abdus Salam,
Sheldon Glashow and Steven Weinberg. These three men shared the Nobel Prize in Physics in 1979 for this work.[28]
It has proven difficult to construct quantum models of gravity, the remaining fundamental force. Semi-classical
approximations are workable, and have led to predictions such as Hawking radiation. However, the formulation of a
complete theory of quantum gravity is hindered by apparent incompatibilities between general relativity, the most
accurate theory of gravity currently known, and some of the fundamental assumptions of quantum theory. The
resolution of these incompatibilities is an area of active research, and theories such as string theory are among the
possible candidates for a future theory of quantum gravity.
Classical mechanics has been extended into the complex domain, and complex classical mechanics exhibits
behaviours similar to quantum mechanics.[29]

Quantum mechanics and classical physics


Predictions of quantum mechanics have been verified experimentally to a extremely high degree of accuracy.
According to the correspondence principle between classical and quantum mechanics, all objects obey the laws of
quantum mechanics, and classical mechanics is just an approximation for large systems (or a statistical quantum
mechanics of a large collection of particles). The laws of classical mechanics thus follow from the laws of quantum
mechanics as a statistical average at the limit of large systems or large quantum numbers.[30] However, chaotic
systems do not have good quantum numbers, and quantum chaos studies the relationship between classical and
quantum descriptions in these systems.
Quantum coherence is an essential difference between classical and quantum theories, and is illustrated by the
Einstein-Podolsky-Rosen paradox. Quantum interference involves adding together probability amplitudes, whereas
when classical waves interfere there is an adding together of intensities. For microscopic bodies, the extension of the
system is much smaller than the coherence length, which gives rise to long-range entanglement and other nonlocal
phenomena characteristic of quantum systems.[31] Quantum coherence is not typically evident at macroscopic scales,
although an exception to this rule can occur at extremely low temperatures, when quantum behavior can manifest
itself on more macroscopic scales (see Bose-Einstein condensate and Quantum machine). This is in accordance with
the following observations:
Many macroscopic properties of a classical system are a direct consequences of the quantum behavior of its parts.
For example, the stability of bulk matter (which consists of atoms and molecules which would quickly collapse
under electric forces alone), the rigidity of solids, and the mechanical, thermal, chemical, optical and magnetic
properties of matter are all results of the interaction of electric charges under the rules of quantum mechanics.[32]
While the seemingly exotic behavior of matter posited by quantum mechanics and relativity theory become more
apparent when dealing with extremely fast-moving or extremely tiny particles, the laws of classical Newtonian
physics remain accurate in predicting the behavior of the vast majority of large objectsof the order of the size of
large molecules and biggerat velocities much smaller than the velocity of light.[33]

47

Quantum Mechanics after 1925

Relativity and quantum mechanics


Main articles: Quantum gravity and Theory of everything
Even with the defining postulates of both Einstein's theory of general relativity and quantum theory being
indisputably supported by rigorous and repeated empirical evidence and while they do not directly contradict each
other theoretically (at least with regard to primary claims), they are resistant to being incorporated within one
cohesive model.[34]
Einstein himself is well known for rejecting some of the claims of quantum mechanics. While clearly contributing to
the field, he did not accept the more philosophical consequences and interpretations of quantum mechanics, such as
the lack of deterministic causality and the assertion that a single subatomic particle can occupy numerous areas of
space at one time. He also was the first to notice some of the apparently exotic consequences of entanglement and
used them to formulate the Einstein-Podolsky-Rosen paradox, in the hope of showing that quantum mechanics had
unacceptable implications. This was 1935, but in 1964 it was shown by John Bell (see Bell inequality) that, although
Einstein was correct in identifying seemingly paradoxical implications of quantum mechanical nonlocality, these
implications could be experimentally tested. Alain Aspect's initial experiments in 1982, and many subsequent
experiments since, have verified quantum entanglement.
According to the paper of J. Bell and the Copenhagen interpretation (the common interpretation of quantum
mechanics by physicists since 1927), and contrary to Einstein's ideas, quantum mechanics was not at the same time
a "realistic" theory
and a local theory.
The Einstein-Podolsky-Rosen paradox shows in any case that there exist experiments by which one can measure the
state of one particle and instantaneously change the state of its entangled partner, although the two particles can be
an arbitrary distance apart; however, this effect does not violate causality, since no transfer of information happens.
Quantum entanglement is at the basis of quantum cryptography, with high-security commercial applications in
banking and government.
Gravity is negligible in many areas of particle physics, so that unification between general relativity and quantum
mechanics is not an urgent issue in those applications. However, the lack of a correct theory of quantum gravity is an
important issue in cosmology and physicists' search for an elegant "theory of everything". Thus, resolving the
inconsistencies between both theories has been a major goal of twentieth- and twenty-first-century physics. Many
prominent physicists, including Stephen Hawking, have labored in the attempt to discover a theory underlying
everything, combining not only different models of subatomic physics, but also deriving the universe's four
forcesthe strong force, electromagnetism, weak force, and gravity from a single force or phenomenon. One of
the leaders in this field is Edward Witten, a theoretical physicist who formulated the groundbreaking M-theory,
which is an attempt at describing the supersymmetrical based string theory.

Attempts at a unified field theory


As of 2011 the quest for unifying the fundamental forces through quantum mechanics is still ongoing. Quantum
electrodynamics (or "quantum electromagnetism"), which is currently (in the perturbative regime at least) the most
accurately tested physical theory,[35] has been successfully merged with the weak nuclear force into the electroweak
force and work is currently being done to merge the electroweak and strong force into the electrostrong force.
Current predictions state that at around 1014 GeV the three aforementioned forces are fused into a single unified
field,[36] Beyond this "grand unification," it is speculated that it may be possible to merge gravity with the other
three gauge symmetries, expected to occur at roughly 1019 GeV. However and while special relativity is
parsimoniously incorporated into quantum electrodynamics the expanded general relativity, currently the best
theory describing the gravitation force, has not been fully incorporated into quantum theory.

48

Quantum Mechanics after 1925

Philosophical implications
Since its inception, the many counter-intuitive results of quantum mechanics have provoked strong philosophical
debate and many interpretations. Even fundamental issues such as Max Born's basic rules concerning probability
amplitudes and probability distributions took decades to be appreciated.
Richard Feynman said, "I think I can safely say that nobody understands quantum mechanics."[37]
The Copenhagen interpretation, due largely to the Danish theoretical physicist Niels Bohr, is the interpretation of the
quantum mechanical formalism most widely accepted amongst physicists. According to it, the probabilistic nature of
quantum mechanics is not a temporary feature which will eventually be replaced by a deterministic theory, but
instead must be considered to be a final renunciation of the classical ideal of causality. In this interpretation, it is
believed that any well-defined application of the quantum mechanical formalism must always make reference to the
experimental arrangement, due to the complementarity nature of evidence obtained under different experimental
situations.
Albert Einstein, himself one of the founders of quantum theory, disliked this loss of determinism in measurement. (A
view paraphrased as "God does not play dice with the universe.") Einstein held that there should be a local hidden
variable theory underlying quantum mechanics and that, consequently, the present theory was incomplete. He
produced a series of objections to the theory, the most famous of which has become known as the
Einstein-Podolsky-Rosen paradox. John Bell showed that the EPR paradox led to experimentally testable differences
between quantum mechanics and local realistic theories. Experiments have been performed confirming the accuracy
of quantum mechanics, thus demonstrating that the physical world cannot be described by local realistic theories.[38]
The Bohr-Einstein debates provide a vibrant critique of the Copenhagen Interpretation from an epistemological point
of view.
The Everett many-worlds interpretation, formulated in 1956, holds that all the possibilities described by quantum
theory simultaneously occur in a multiverse composed of mostly independent parallel universes.[39] This is not
accomplished by introducing some new axiom to quantum mechanics, but on the contrary by removing the axiom of
the collapse of the wave packet: All the possible consistent states of the measured system and the measuring
apparatus (including the observer) are present in a real physical (not just formally mathematical, as in other
interpretations) quantum superposition. Such a superposition of consistent state combinations of different systems is
called an entangled state. While the multiverse is deterministic, we perceive non-deterministic behavior governed by
probabilities, because we can observe only the universe, i.e. the consistent state contribution to the mentioned
superposition, we inhabit. Everett's interpretation is perfectly consistent with John Bell's experiments and makes
them intuitively understandable. However, according to the theory of quantum decoherence, the parallel universes
will never be accessible to us. This inaccessibility can be understood as follows: Once a measurement is done, the
measured system becomes entangled with both the physicist who measured it and a huge number of other particles,
some of which are photons flying away towards the other end of the universe; in order to prove that the wave
function did not collapse one would have to bring all these particles back and measure them again, together with the
system that was measured originally. This is completely impractical, but even if one could theoretically do this, it
would destroy any evidence that the original measurement took place (including the physicist's memory).

49

Quantum Mechanics after 1925

Applications
Quantum mechanics had enormous success in explaining many of the features of our world. The individual
behaviour of the subatomic particles that make up all forms of matterelectrons, protons, neutrons, photons and
otherscan often only be satisfactorily described using quantum mechanics. Quantum mechanics has strongly
influenced string theory, a candidate for a theory of everything (see reductionism) and the multiverse hypothesis.
Quantum mechanics is important for understanding how individual atoms combine covalently to form chemicals or
molecules. The application of quantum mechanics to chemistry is known as quantum chemistry. (Relativistic)
quantum mechanics can in principle mathematically describe most of chemistry. Quantum mechanics can provide
quantitative insight into ionic and covalent bonding processes by explicitly showing which molecules are
energetically favorable to which others, and by approximately how much.[40] Most of the calculations performed in
computational chemistry rely on quantum mechanics.[41]
Much of modern technology operates
at a scale where quantum effects are
significant. Examples include the laser,
the transistor (and thus the microchip),
the electron microscope, and magnetic
resonance imaging. The study of
semiconductors led to the invention of
the diode and the transistor, which are
indispensable for modern electronics.
Researchers are currently seeking
robust
methods
of
directly
manipulating quantum states. Efforts
are being made to develop quantum
cryptography, which will allow
guaranteed secure transmission of
information. A more distant goal is the
A working mechanism of a resonant tunneling diode device, based on the phenomenon of
development of quantum computers,
quantum tunneling through the potential barriers.
which are expected to perform certain
computational tasks exponentially
faster than classical computers. Another active research topic is quantum teleportation, which deals with techniques
to transmit quantum information over arbitrary distances.
Quantum tunneling is vital in many devices, even in the simple light switch, as otherwise the electrons in the electric
current could not penetrate the potential barrier made up of a layer of oxide. Flash memory chips found in USB
drives use quantum tunneling to erase their memory cells.
Quantum mechanics primarily applies to the atomic regimes of matter and energy, but some systems exhibit
quantum mechanical effects on a large scale; superfluidity (the frictionless flow of a liquid at temperatures near
absolute zero) is one well-known example. Quantum theory also provides accurate descriptions for many previously
unexplained phenomena such as black body radiation and the stability of electron orbitals. It has also given insight
into the workings of many different biological systems, including smell receptors and protein structures.[42] Recent
work on photosynthesis has provided evidence that quantum correlations play an essential role in this most
fundamental process of the plant kingdom.[43] Even so, classical physics often can be a good approximation to
results otherwise obtained by quantum physics, typically in circumstances with large numbers of particles or large
quantum numbers. (However, some open questions remain in the field of quantum chaos.)

50

Quantum Mechanics after 1925

51

Examples
Free particle
For example, consider a free particle. In quantum mechanics, there is wave-particle duality so the properties of the
particle can be described as the properties of a wave. Therefore, its quantum state can be represented as a wave of
arbitrary shape and extending over space as a wave function. The position and momentum of the particle are
observables. The Uncertainty Principle states that both the position and the momentum cannot simultaneously be
measured with full precision at the same time. However, one can measure the position alone of a moving free
particle creating an eigenstate of position with a wavefunction that is very large (a Dirac delta) at a particular
position x and zero everywhere else. If one performs a position measurement on such a wavefunction, the result x
will be obtained with 100% probability (full certainty). This is called an eigenstate of position (mathematically more
precise: a generalized position eigenstate (eigendistribution)). If the particle is in an eigenstate of position then its
momentum is completely unknown. On the other hand, if the particle is in an eigenstate of momentum then its
position is completely unknown.[44] In an eigenstate of momentum having a plane wave form, it can be shown that
the wavelength is equal to h/p, where h is Planck's constant and p is the momentum of the eigenstate.[45]

3D confined electron wave functions for each eigenstate in a Quantum Dot. Here, rectangular and triangular-shaped quantum dots are shown. Energy
states in rectangular dots are more s-type and p-type. However, in a triangular dot the wave functions are mixed due to confinement symmetry.

Step potential
The potential in this case is given by:

The step potential with incident and exiting


waves shown.

Quantum Mechanics after 1925

52

The solutions are superpositions of left and right moving waves:


,

where the wave vectors are related to the energy via


, and

and the coefficients A and B are determined from the boundary conditions and by imposing a continuous derivative
to the solution.
Each term of the solution can be interpreted as an incident, reflected of transmitted component of the wave, allowing
the calculation of transmission and reflection coefficients. In contrast to classical mechanics, incident particles with
energies higher than the size of the potential step are still partially reflected.

Rectangular potential barrier


This is a model for the quantum tunneling effect, which has important applications to modern devices such as flash
memory and the scanning tunneling microscope.

Particle in a box
The particle in a 1-dimensional potential energy box is the most simple
example where restraints lead to the quantization of energy levels. The
box is defined as having zero potential energy inside a certain region
and infinite potential energy everywhere outside that region. For the
1-dimensional case in the
direction, the time-independent
Schrdinger equation can be written as:[46]

1-dimensional potential energy box (or infinite


potential well)

Writing the differential operator

the previous equation can be seen to be evocative of the classic analogue

with

as the energy for the state

, in this case coinciding with the kinetic energy of the particle.

The general solutions of the Schrdinger equation for the particle in a box are:

Quantum Mechanics after 1925

53

or, from Euler's formula,

The presence of the walls of the box determines the values of C, D, and k. At each wall (x = 0 and x = L), = 0.
Thus when x = 0,

and so D = 0. When x = L,

C cannot be zero, since this would conflict with the Born interpretation. Therefore sin kL = 0, and so it must be that
kL is an integer multiple of . Therefore,

The quantization of energy levels follows from this constraint on k, since

Finite potential well


This is generalization of the infinite potential well problem to potential wells of finite depth.

Harmonic oscillator
As in the classical case, the potential for the quantum harmonic oscillator is given by:

This problem can be solved either by directly solving the Schrdinger equation directly, which is not trivial, or by
using the more elegant ladder method, first proposed by Paul Dirac. The eigenstates are given by:

where Hn are the Hermite polynomials:

and the corresponding energy levels are


.
This is another example which illustrates the quantification of energy for bound states.

Quantum Mechanics after 1925

Notes
[1] J. Mehra and H. Rechenberg, The historical development of quantum theory, Springer-Verlag, 1982.
[2] T.S. Kuhn, Black-body theory and the quantum discontinuity 1894-1912, Clarendon Press, Oxford, 1978.
[3] A. Einstein, ber einen die Erzeugung und Verwandlung des Lichtes betreffenden heuristischen Gesichtspunkt (On a heuristic point of view
concerning the production and transformation of light), Annalen der Physik 17 (1905) 132-148 (reprinted in The collected papers of Albert
Einstein, John Stachel, editor, Princeton University Press, 1989, Vol. 2, pp. 149-166, in German; see also Einstein's early work on the
quantum hypothesis, ibid. pp. 134-148).
[4] "Merriam-Webster.com" (http:/ / www. merriam-webster. com/ dictionary/ quantum). Merriam-Webster.com. 2010-08-13. . Retrieved
2010-10-15.
[5] Edwin Thall. "FCCJ.org" (http:/ / mooni. fccj. org/ ~ethall/ quantum/ quant. htm). Mooni.fccj.org. . Retrieved 2010-10-15.
[6] Compare the list of conferences presented here (http:/ / ysfine. com/ ).
[7] Oocities.com (http:/ / web. archive. org/ web/ 20091026095410/ http:/ / geocities. com/ mik_malm/ quantmech. html)
[8] P.A.M. Dirac, The Principles of Quantum Mechanics, Clarendon Press, Oxford, 1930.
[9] J. von Neumann, Mathematische Grundlagen der Quantenmechanik, Springer, Berlin, 1932 (English translation: Mathematical Foundations
of Quantum Mechanics, Princeton University Press, 1955).
[10] Greiner, Walter; Mller, Berndt (1994). Quantum Mechanics Symmetries, Second edition (http:/ / books. google. com/
books?id=gCfvWx6vuzUC& pg=PA52). Springer-Verlag. p.52. ISBN3-540-58080-8. .,
[11] "AIP.org" (http:/ / www. aip. org/ history/ heisenberg/ p08a. htm). AIP.org. . Retrieved 2010-10-15.
[12] Greenstein, George; Zajonc, Arthur (2006). The Quantum Challenge: Modern Research on the Foundations of Quantum Mechanics, Second
edition (http:/ / books. google. com/ books?id=5t0tm0FB1CsC& pg=PA215). Jones and Bartlett Publishers, Inc. p.215. ISBN0-7637-2470-X.
.,
[13] probability clouds are approximate, but better than the Bohr model, whereby electron location is given by a probability function, the wave
function eigenvalue, such that the probability is the squared modulus of the complex amplitude
[14] "Actapress.com" (http:/ / www. actapress. com/ PaperInfo. aspx?PaperID=25988& reason=500). Actapress.com. . Retrieved 2010-10-15.
[15] Hirshleifer, Jack (2001). The Dark Side of the Force: Economic Foundations of Conflict Theory (http:/ / books. google. com/
books?id=W2J2IXgiZVgC& pg=PA265). Campbridge University Press. p.265. ISBN0-521-80412-4. .,
[16] Dict.cc (http:/ / www. dict. cc/ german-english/ eigen. html)
De.pons.eu (http:/ / de. pons. eu/ deutsch-englisch/ eigen)
[17] "PHY.olemiss.edu" (http:/ / www. phy. olemiss. edu/ ~luca/ Topics/ qm/ collapse. html). PHY.olemiss.edu. 2010-08-16. . Retrieved
2010-10-15.
[18] "Farside.ph.utexas.edu" (http:/ / farside. ph. utexas. edu/ teaching/ qmech/ lectures/ node28. html). Farside.ph.utexas.edu. . Retrieved
2010-10-15.
[19] "Reddit.com" (http:/ / www. reddit. com/ r/ philosophy/ comments/ 8p2qv/ determinism_and_naive_realism/ ). Reddit.com. 2009-06-01. .
Retrieved 2010-10-15.
[20] Michael Trott. "Time-Evolution of a Wavepacket in a Square Well Wolfram Demonstrations Project" (http:/ / demonstrations. wolfram.
com/ TimeEvolutionOfAWavepacketInASquareWell/ ). Demonstrations.wolfram.com. . Retrieved 2010-10-15.
[21] Michael Trott. "Time Evolution of a Wavepacket In a Square Well" (http:/ / demonstrations. wolfram. com/
TimeEvolutionOfAWavepacketInASquareWell/ ). Demonstrations.wolfram.com. . Retrieved 2010-10-15.
[22] Mathews, Piravonu Mathews; Venkatesan, K. (1976). A Textbook of Quantum Mechanics (http:/ / books. google. com/
books?id=_qzs1DD3TcsC& pg=PA36). Tata McGraw-Hill. p.36. ISBN0-07-096510-2. .,
[23] "Wave Functions and the Schrdinger Equation" (http:/ / physics. ukzn. ac. za/ ~petruccione/ Phys120/ Wave Functions and the Schrdinger
Equation. pdf) (PDF). . Retrieved 2010-10-15.
[24] "Spaceandmotion.com" (http:/ / www. spaceandmotion. com/ physics-quantum-mechanics-werner-heisenberg. htm). Spaceandmotion.com. .
Retrieved 2010-10-15.
[25] Especially since Werner Heisenberg was awarded the Nobel Prize in Physics in 1932 for the creation of quantum mechanics, the role of Max
Born has been obfuscated. A 2005 biography of Born details his role as the creator of the matrix formulation of quantum mechanics. This was
recognized in a paper by Heisenberg, in 1940, honoring Max Planck. See: Nancy Thorndike Greenspan, "The End of the Certain World: The
Life and Science of Max Born" (Basic Books, 2005), pp. 124 - 128, and 285 - 286.
[26] "IF.uj.edu.pl" (http:/ / th-www. if. uj. edu. pl/ acta/ vol19/ pdf/ v19p0683. pdf) (PDF). . Retrieved 2010-10-15.
[27] "OCW.ssu.edu" (http:/ / ocw. usu. edu/ physics/ classical-mechanics/ pdf_lectures/ 06. pdf) (PDF). . Retrieved 2010-10-15.
[28] "The Nobel Prize in Physics 1979" (http:/ / nobelprize. org/ nobel_prizes/ physics/ laureates/ 1979/ index. html). Nobel Foundation. .
Retrieved 2010-02-16.
[29] Complex Elliptic Pendulum (http:/ / arxiv. org/ abs/ 1001. 0131), Carl M. Bender, Daniel W. Hook, Karta Kooner
[30] "Scribd.com" (http:/ / www. scribd. com/ doc/ 5998949/ Quantum-mechanics-course-iwhatisquantummechanics). Scribd.com. 2008-09-14. .
Retrieved 2010-10-15.
[31] Philsci-archive.pitt.edu (http:/ / philsci-archive. pitt. edu/ archive/ 00002328/ 01/ handbook. pdf)
[32] "Academic.brooklyn.cuny.edu" (http:/ / academic. brooklyn. cuny. edu/ physics/ sobel/ Nucphys/ atomprop. html).
Academic.brooklyn.cuny.edu. . Retrieved 2010-10-15.

54

Quantum Mechanics after 1925


[33] "Cambridge.org" (http:/ / assets. cambridge. org/ 97805218/ 29526/ excerpt/ 9780521829526_excerpt. pdf) (PDF). . Retrieved 2010-10-15.
[34] "There is as yet no logically consistent and complete relativistic quantum field theory.", p. 4. V. B. Berestetskii, E. M. Lifshitz, L P
Pitaevskii (1971). J. B. Sykes, J. S. Bell (translators). Relativistic Quantum Theory 4, part I. Course of Theoretical Physics (Landau and
Lifshitz) ISBN 0080160255
[35] "Life on the lattice: The most accurate theory we have" (http:/ / latticeqcd. blogspot. com/ 2005/ 06/ most-accurate-theory-we-have. html).
Latticeqcd.blogspot.com. 2005-06-03. . Retrieved 2010-10-15.
[36] Parker, B. (1993). Overcoming some of the problems. pp.259279.
[37] The Character of Physical Law (1965) Ch. 6; also quoted in The New Quantum Universe (2003) by Tony Hey and Patrick Walters
[38] "Plato.stanford.edu" (http:/ / plato. stanford. edu/ entries/ qm-action-distance/ ). Plato.stanford.edu. 2007-01-26. . Retrieved 2010-10-15.
[39] "Plato.stanford.edu" (http:/ / plato. stanford. edu/ entries/ qm-everett/ ). Plato.stanford.edu. . Retrieved 2010-10-15.
[40] Books.google.com (http:/ / books. google. com/ books?id=vdXU6SD4_UYC). Books.google.com. . Retrieved 2010-10-23.
[41] "en.wikiboos.org" (http:/ / en. wikibooks. org/ wiki/ Computational_chemistry/ Applications_of_molecular_quantum_mechanics).
En.wikibooks.org. . Retrieved 2010-10-23.
[42] Anderson, Mark (2009-01-13). "Discovermagazine.com" (http:/ / discovermagazine. com/ 2009/ feb/
13-is-quantum-mechanics-controlling-your-thoughts/ article_view?b_start:int=1& -C). Discovermagazine.com. . Retrieved 2010-10-23.
[43] "Quantum mechanics boosts photosynthesis" (http:/ / physicsworld. com/ cws/ article/ news/ 41632). physicsworld.com. . Retrieved
2010-10-23.
[44] Davies, P. C. W.; Betts, David S. (1984). Quantum Mechanics, Second edition (http:/ / books. google. com/ books?id=XRyHCrGNstoC&
pg=PA79). Chapman and Hall. p.79. ISBN0-7487-4446-0. .,
[45] Books.Google.com (http:/ / books. google. com/ books?id=tKm-Ekwke_UC). Books.Google.com. 2007-08-30. . Retrieved 2010-10-23.
[46] Derivation of particle in a box, chemistry.tidalswan.com (http:/ / chemistry. tidalswan. com/ index. php?title=Quantum_Mechanics)

References
The following titles, all by working physicists, attempt to communicate quantum theory to lay people, using a
minimum of technical apparatus.
Chester, Marvin (1987) Primer of Quantum Mechanics. John Wiley. ISBN 0-486-42878-8
Richard Feynman, 1985. QED: The Strange Theory of Light and Matter, Princeton University Press. ISBN
0-691-08388-6. Four elementary lectures on quantum electrodynamics and quantum field theory, yet containing
many insights for the expert.
Ghirardi, GianCarlo, 2004. Sneaking a Look at God's Cards, Gerald Malsbary, trans. Princeton Univ. Press. The
most technical of the works cited here. Passages using algebra, trigonometry, and bra-ket notation can be passed
over on a first reading.
N. David Mermin, 1990, "Spooky actions at a distance: mysteries of the QT" in his Boojums all the way through.
Cambridge University Press: 110-76.
Victor Stenger, 2000. Timeless Reality: Symmetry, Simplicity, and Multiple Universes. Buffalo NY: Prometheus
Books. Chpts. 5-8. Includes cosmological and philosophical considerations.
More technical:
Bryce DeWitt, R. Neill Graham, eds., 1973. The Many-Worlds Interpretation of Quantum Mechanics, Princeton
Series in Physics, Princeton University Press. ISBN 0-691-08131-X
Dirac, P. A. M. (1930). The Principles of Quantum Mechanics. ISBN0198520115. The beginning chapters make
up a very clear and comprehensible introduction.
Hugh Everett, 1957, "Relative State Formulation of Quantum Mechanics," Reviews of Modern Physics 29:
454-62.
Feynman, Richard P.; Leighton, Robert B.; Sands, Matthew (1965). The Feynman Lectures on Physics. 1-3.
Addison-Wesley. ISBN0738200085.
Griffiths, David J. (2004). Introduction to Quantum Mechanics (2nd ed.). Prentice Hall. ISBN0-13-111892-7.
OCLC40251748. A standard undergraduate text.
Max Jammer, 1966. The Conceptual Development of Quantum Mechanics. McGraw Hill.
Hagen Kleinert, 2004. Path Integrals in Quantum Mechanics, Statistics, Polymer Physics, and Financial Markets,
3rd ed. Singapore: World Scientific. Draft of 4th edition. (http://www.physik.fu-berlin.de/~kleinert/b5)

55

Quantum Mechanics after 1925


Gunther Ludwig, 1968. Wave Mechanics. London: Pergamon Press. ISBN 0-08-203204-1
George Mackey (2004). The mathematical foundations of quantum mechanics. Dover Publications. ISBN
0-486-43517-2.
Albert Messiah, 1966. Quantum Mechanics (Vol. I), English translation from French by G. M. Temmer. North
Holland, John Wiley & Sons. Cf. chpt. IV, section III.
Omns, Roland (1999). Understanding Quantum Mechanics. Princeton University Press. ISBN0-691-00435-8.
OCLC39849482.
Scerri, Eric R., 2006. The Periodic Table: Its Story and Its Significance. Oxford University Press. Considers the
extent to which chemistry and the periodic system have been reduced to quantum mechanics. ISBN
0-19-530573-6
Transnational College of Lex (1996). What is Quantum Mechanics? A Physics Adventure. Language Research
Foundation, Boston. ISBN0-9643504-1-6. OCLC34661512.
von Neumann, John (1955). Mathematical Foundations of Quantum Mechanics. Princeton University Press.
ISBN0691028931.
Hermann Weyl, 1950. The Theory of Groups and Quantum Mechanics, Dover Publications.
D. Greenberger, K. Hentschel, F. Weinert, eds., 2009. Compendium of quantum physics, Concepts, experiments,
history and philosophy, Springer-Verlag, Berlin, Heidelberg.

Further reading
Bernstein, Jeremy (2009). Quantum Leaps (http://books.google.com/books?id=j0Me3brYOL0C&
printsec=frontcover). Cambridge, Massachusetts: Belknap Press of Harvard University Press.
ISBN9780674035416.
Bohm, David (1989). Quantum Theory. Dover Publications. ISBN0-486-65969-0.
Eisberg, Robert; Resnick, Robert (1985). Quantum Physics of Atoms, Molecules, Solids, Nuclei, and Particles
(2nd ed.). Wiley. ISBN0-471-87373-X.
Liboff, Richard L. (2002). Introductory Quantum Mechanics. Addison-Wesley. ISBN0-8053-8714-5.
Merzbacher, Eugen (1998). Quantum Mechanics. Wiley, John & Sons, Inc. ISBN0-471-88702-1.
Sakurai, J. J. (1994). Modern Quantum Mechanics. Addison Wesley. ISBN0-201-53929-2.
Shankar, R. (1994). Principles of Quantum Mechanics. Springer. ISBN0-306-44790-8.

External links
A foundation approach to quantum Theory that does not rely on wave-particle duality. (http://www.mesacc.
edu/~kevinlg/i256/QM_basics.pdf)
The Modern Revolution in Physics (http://www.lightandmatter.com/html_books/6mr/ch01/ch01.html) - an
online textbook.
J. O'Connor and E. F. Robertson: A history of quantum mechanics. (http://www-history.mcs.st-andrews.ac.uk/
history/HistTopics/The_Quantum_age_begins.html)
Introduction to Quantum Theory at Quantiki. (http://www.quantiki.org/wiki/index.php/
Introduction_to_Quantum_Theory)
Quantum Physics Made Relatively Simple (http://bethe.cornell.edu/): three video lectures by Hans Bethe
H is for h-bar. (http://www.nonlocal.com/hbar/)
Quantum Mechanics Books Collection (http://www.freebookcentre.net/Physics/Quantum-Mechanics-Books.
html): Collection of free books
Course material
Doron Cohen: Lecture notes in Quantum Mechanics (comprehensive, with advanced topics). (http://arxiv.org/
abs/quant-ph/0605180)

56

Quantum Mechanics after 1925


MIT OpenCourseWare: Chemistry (http://ocw.mit.edu/OcwWeb/Chemistry/index.htm).
MIT OpenCourseWare: Physics (http://ocw.mit.edu/OcwWeb/Physics/index.htm). See 8.04 (http://ocw.
mit.edu/OcwWeb/Physics/8-04Spring-2006/CourseHome/index.htm)
Stanford Continuing Education PHY 25: Quantum Mechanics (http://www.youtube.com/stanford#g/c/
84C10A9CB1D13841) by Leonard Susskind, see course description (http://continuingstudies.stanford.edu/
courses/course.php?cid=20072_PHY 25) Fall 2007
5 Examples in Quantum Mechanics (http://www.physics.csbsju.edu/QM/)
Imperial College Quantum Mechanics Course. (http://www.imperial.ac.uk/quantuminformation/qi/tutorials)
Spark Notes - Quantum Physics. (http://www.sparknotes.com/testprep/books/sat2/physics/
chapter19section3.rhtml)
Quantum Physics Online : interactive introduction to quantum mechanics (RS applets). (http://www.
quantum-physics.polytechnique.fr/)
Experiments to the foundations of quantum physics with single photons. (http://www.didaktik.physik.
uni-erlangen.de/quantumlab/english/index.html)
Motion Mountain, Volume IV (http://www.motionmountain.net/download.html) - A modern introduction to
quantum theory, with several animations.
AQME (http://www.nanohub.org/topics/AQME) : Advancing Quantum Mechanics for Engineers by
T.Barzso, D.Vasileska and G.Klimeck online learning resource with simulation tools on nanohub
Quantum Mechanics (http://www.lsr.ph.ic.ac.uk/~plenio/lecture.pdf) by Martin Plenio
Quantum Mechanics (http://farside.ph.utexas.edu/teaching/qm/389.pdf) by Richard Fitzpatrick
Online course on Quantum Transport (http://nanohub.org/resources/2039)
FAQs
Many-worlds or relative-state interpretation. (http://www.hedweb.com/manworld.htm)
Measurement in Quantum mechanics. (http://www.mtnmath.com/faq/meas-qm.html)
Media
Lectures on Quantum Mechanics by Leonard Susskind (http://www.youtube.com/
view_play_list?p=84C10A9CB1D13841)
Everything you wanted to know about the quantum world (http://www.newscientist.com/channel/
fundamentals/quantum-world) archive of articles from New Scientist.
Quantum Physics Research (http://www.sciencedaily.com/news/matter_energy/quantum_physics/) from
Science Daily
Overbye, Dennis (December 27, 2005). "Quantum Trickery: Testing Einstein's Strangest Theory" (http://www.
nytimes.com/2005/12/27/science/27eins.html?ex=1293339600&en=caf5d835203c3500&ei=5090). The
New York Times. Retrieved April 12, 2010.
Audio: Astronomy Cast (http://www.astronomycast.com/physics/ep-138-quantum-mechanics/) Quantum
Mechanics June 2009. Fraser Cain interviews Pamela L. Gay.
Philosophy
"Quantum Mechanics" (http://plato.stanford.edu/entries/qm) entry by Jenann Ismael. in the Stanford
Encyclopedia of Philosophy
"Measurement in Quantum Theory" (http://plato.stanford.edu/entries/qm) entry by Henry Krips. in the
Stanford Encyclopedia of Philosophy

57

58

3. The Interpretation of Quantum Mechanics


Interpretations of Quantum Mechanics
An interpretation of quantum mechanics is a set of statements which attempt to explain how quantum mechanics
informs our understanding of nature. Although quantum mechanics has received thorough experimental testing,
many of these experiments are open to different interpretations. There exist a number of contending schools of
thought, differing over whether quantum mechanics can be understood to be deterministic, which elements of
quantum mechanics can be considered "real", and other matters.
This question is of special interest to philosophers of physics, as physicists continue to show a strong interest in the
subject. They usually consider an interpretation of quantum mechanics as an interpretation of the mathematical
formalism of quantum mechanics, specifying the physical meaning of the mathematical entities of the theory.

Historical background
The definition of terms used by researchers in quantum theory (such as wavefunctions and matrix mechanics)
progressed through many stages. For instance, Schrdinger originally viewed the wavefunction associated with the
electron as corresponding to the charge density of an object smeared out over an extended, possibly infinite, volume
of space. Max Born interpreted it as simply corresponding to a probability distribution. These are two different
interpretations of the wavefunction. In one it corresponds to a material field; in the other it corresponds to a
probability distribution specifically, the probability that the quantum of charge is located at any particular point
within spatial dimensions.
The Copenhagen interpretation was traditionally the most popular among physicists, next to a purely instrumentalist
position that denies any need for explanation (a view expressed in David Mermin's famous quote "shut up and
calculate", often misattributed to Richard Feynman.[1] ) However, the many-worlds interpretation has been gaining
acceptance;[2] a poll mentioned in "The Physics of Immortality" (published in 1994), of 72 "leading cosmologists
and other quantum field theorists" found that 58% supported the many-worlds interpretation, including Stephen
Hawking and Nobel laureates Murray Gell-Mann and Richard Feynman.[3] Moreover, the instrumentalist position
has been challenged by recent proposals for falsifiable experiments that might one day distinguish interpretations,
e.g. by measuring an AI consciousness[4] or via quantum computing.[5]

The nature of interpretation


What interpretations are interpretations of is a formalism a set of equations and formulae for generating results
and predictions and a phenomenology, a set of observations, including both those obtained by empirical research,
and more informal subjective ones (the fact that humans invariably observe an unequivocal world is important in the
interpretation of quantum mechanics) . These are the more-or-less fixed ingredients of an interpretation. The
ingredients that vary between interpretations are the ontology and the epistomology, which are concerned with what,
if anything, the interpreted theory is "really about". The same phenomenon may be given an ontological reading
under one interpretation, and an epistemological one under another. For instance, indeterminism may be attributed to
the real existence of a "maybe" in the universe (ontology) or to limitations of an observer's information and
predictive abilities (epistemology). Interpretations may be broadly classed as leaning more towards ontology, i.e.
realism, or towards anti-realism.

Interpretations of Quantum Mechanics


Some approaches tend to avoid giving any interpretation of phenomena or formalism. These can be described as
instrumentalist. Other approaches suggest modifications to the formalism, and are therefore, strictly speaking,
alternative theories rather than interpretations. In some cases, for instance Bohmian mechanics, it is open to debate
as to whether an approach is equivalent to the standard formalism.

Problems of Interpretation
The difficulties of interpretation reflect a number of points about the orthodox description of quantum mechanics,
including:
1. The abstract, mathematical nature of that description.
2. The existence of what appear to be non-deterministic and irreversible processes.
3. The phenomenon of entanglement, and in particular the correlations between remote events that are not expected
in classical theory.
4. The complementarity of the proffered descriptions of reality.
5. The role played by observers and the process of measurement.
6. The rapid rate at which quantum descriptions become more complicated as the size of a system increases.
Firstly, the accepted mathematical structure of quantum mechanics is based on fairly abstract mathematics, such as
Hilbert spaces and operators on those spaces. In classical mechanics and electromagnetism, on the other hand,
properties of a point mass or properties of a field are described by real numbers or functions defined on two or three
dimensional sets. These have direct, spatial meaning, and in these theories there seems to be less need to provide
special interpretation for those numbers or functions.
Furthermore, the process of measurement may play an essential role in quantum theory - a hotly contested point. The
world around us seems to be in a specific state, but quantum mechanics describes it by wave functions that govern
the probability of all values. In general, the wave-function assigns non-zero probabilities to all possible values of any
given physical quantity, such as position. How, then, do we see a particle in a specific position when its wave
function is spread across all space? In order to describe how specific outcomes arise from the probabilities, the direct
interpretation introduced the concept of measurement. According to the theory, wave functions interact with each
other and evolve in time in accordance with the laws of quantum mechanics until a measurement is performed, at
which point the system takes on one of its possible values, with a probability that's governed by the wave-function.
Measurement can interact with the system state in somewhat peculiar ways, as is illustrated by the double-slit
experiment.
Thus the mathematical formalism used to describe the time evolution of a non-relativistic system proposes two
opposed kinds of transformation:
Reversible transformations described by unitary operators on the state space. These transformations are
determined by solutions to the Schrdinger equation.
Non-reversible and unpredictable transformations described by mathematically more complicated transformations
(see quantum operations). Examples include the transformations undergone by a system as a result of
measurement.
A solution to the problem of interpretation consists in providing some form of plausible picture, by resolving the
second kind of transformation. This can be achieved by purely mathematical solutions, as offered by the
many-worlds or the consistent histories interpretations.
In addition to the unpredictable and irreversible character of measurement processes, there are other elements of
quantum physics that distinguish it sharply from classical physics and which are not present in any classical theory.
One of these is the phenomenon of entanglement, as illustrated in the EPR paradox, which seemingly violates
principles of local causality.[6]

59

Interpretations of Quantum Mechanics


Another obstruction to interpretation is the phenomenon of complementarity, which seems to violate basic principles
of propositional logic. Complementarity says there is no logical picture (one obeying classical propositional logic)
that can simultaneously describe and be used to reason about all properties of a quantum system S. This is often
phrased by saying that there are "complementary" propositions A and B that can each describe S, but not at the same
time. Examples of A and B are propositions using a wave description of S and a corpuscular description of S. The
latter statement is one part of Niels Bohr's original formulation, which is often equated to the principle of
complementarity itself.
Complementarity does not usually imply that it is classical logic which is at fault (although Hilary Putnam did take
that view in his paper "Is logic empirical?"). Rather, complementarity means that the composition of physical
properties for S (such as position and momentum both having values within certain ranges), using propositional
connectives, does not obey the rules of classical propositional logic (see also Quantum logic). As is now well-known
(Omns, 1999) the "origin of complementarity lies in the non-commutativity of [the] operators" that describe
observables (i.e. particles) in quantum mechanics.
Because the complexity of a quantum system is exponential in its number of degrees of freedom, it is difficult to
overlap the quantum and classical descriptions to see how the classical approximations are being made.

Problematic status of interpretations


As classical physics and non-mathematical language cannot match the precision of quantum mechanics mathematics,
anything said outside the mathematical formulation is necessarily limited in accuracy.
Also, the precise ontological status of each interpretation remains a matter of philosophical argument. In other
words, if we interpret the formal structure X of quantum mechanics by means of a structure Y (via a mathematical
equivalence of the two structures), what is the status of Y? This is the old question of saving the phenomena, in a
new guise.
Some physicists, for example Asher Peres and Chris Fuchs, argue that an interpretation is nothing more than a
formal equivalence between sets of rules for operating on experimental data, thereby implying that the whole
exercise of interpretation is unnecessary.

Instrumentalist interpretation
Any modern scientific theory requires at the very least an instrumentalist description that relates the mathematical
formalism to experimental practice and prediction. In the case of quantum mechanics, the most common
instrumentalist description is an assertion of statistical regularity between state preparation processes and
measurement processes. That is, if a measurement of a real-value quantity is performed many times, each time
starting with the same initial conditions, the outcome is a well-defined probability distribution agreeing with the real
numbers; moreover, quantum mechanics provides a computational instrument to determine statistical properties of
this distribution, such as its expectation value.
Calculations for measurements performed on a system S postulate a Hilbert space H over the complex numbers.
When the system S is prepared in a pure state, it is associated with a vector in H. Measurable quantities are
associated with Hermitian operators acting on H: these are referred to as observables.
Repeated measurement of an observable A where S is prepared in state yields a distribution of values. The
expectation value of this distribution is given by the expression

This mathematical machinery gives a simple, direct way to compute a statistical property of the outcome of an
experiment, once it is understood how to associate the initial state with a Hilbert space vector, and the measured
quantity with an observable (that is, a specific Hermitian operator).

60

Interpretations of Quantum Mechanics


As an example of such a computation, the probability of finding the system in a given state

61
is given by

computing the expectation value of a (rank-1) projection operator


The probability is then the non-negative real number given by

By abuse of language, a bare instrumentalist description could be referred to as an interpretation, although this usage
is somewhat misleading since instrumentalism explicitly avoids any explanatory role; that is, it does not attempt to
answer the question why?

Summary of common interpretations of QM


Classification adopted by Einstein
An interpretation (i.e. a semantic explanation of the formal mathematics of quantum mechanics) can be characterized
by its treatment of certain matters addressed by Einstein, such as:
Realism
Completeness
Local realism
Determinism
To explain these properties, we need to be more explicit about the kind of picture an interpretation provides. To that
end we will regard an interpretation as a correspondence between the elements of the mathematical formalism M and
the elements of an interpreting structure I, where:
The mathematical formalism M consists of the Hilbert space machinery of ket-vectors, self-adjoint operators
acting on the space of ket-vectors, unitary time dependence of the ket-vectors, and measurement operations. In
this context a measurement operation is a transformation which turns a ket-vector into a probability distribution
(for a formalization of this concept see quantum operations).
The interpreting structure I includes states, transitions between states, measurement operations, and possibly
information about spatial extension of these elements. A measurement operation refers to an operation which
returns a value and might result in a system state change. Spatial information would be exhibited by states
represented as functions on configuration space. The transitions may be non-deterministic or probabilistic or there
may be infinitely many states.
The crucial aspect of an interpretation is whether the elements of I are regarded as physically real. Hence the bare
instrumentalist view of quantum mechanics outlined in the previous section is not an interpretation at all, for it
makes no claims about elements of physical reality.
The current usage of "realism" and "completeness" originated in the 1935 paper in which Einstein and others
proposed the EPR paradox.[7] In that paper the authors proposed the concepts "element of reality" and the
"completeness" of a physical theory. They characterised "element of reality" as a quantity whose value can be
predicted with certainty before measuring or otherwise disturbing it, and defined a "complete physical theory" as one
in which every element of physical reality is accounted for by the theory. In a semantic view of interpretation, an
interpretation is complete if every element of the interpreting structure is present in the mathematics. Realism is also
a property of each of the elements of the maths; an element is real if it corresponds to something in the interpreting
structure. For example, in some interpretations of quantum mechanics (such as the many-worlds interpretation) the
ket vector associated to the system state is said to correspond to an element of physical reality, while in other
interpretations it is not.
Determinism is a property characterizing state changes due to the passage of time, namely that the state at a future
instant is a function of the state in the present (see time evolution). It may not always be clear whether a particular

Interpretations of Quantum Mechanics


interpretation is deterministic or not, as there may not be a clear choice of a time parameter. Moreover, a given
theory may have two interpretations, one of which is deterministic and the other not.
Local realism has two aspects:
The value returned by a measurement corresponds to the value of some function in the state space. In other words,
that value is an element of reality;
The effects of measurement have a propagation speed not exceeding some universal limit (e.g. the speed of light).
In order for this to make sense, measurement operations in the interpreting structure must be localized.
A precise formulation of local realism in terms of a local hidden variable theory was proposed by John Bell.
Bell's theorem, combined with experimental testing, restricts the kinds of properties a quantum theory can have. For
instance, Bell's theorem implies that quantum mechanics cannot satisfy local realism.

The Copenhagen interpretation


The Copenhagen interpretation is the "standard" interpretation of quantum mechanics formulated by Niels Bohr and
Werner Heisenberg while collaborating in Copenhagen around 1927. Bohr and Heisenberg extended the probabilistic
interpretation of the wavefunction proposed originally by Max Born. The Copenhagen interpretation rejects
questions like "where was the particle before I measured its position?" as meaningless. The measurement process
randomly picks out exactly one of the many possibilities allowed for by the state's wave function in a manner
consistent with the well-defined probabilities that are assigned to each possible state.

Many worlds
The many-worlds interpretation is an interpretation of quantum mechanics in which a universal wavefunction obeys
the same deterministic, reversible laws at all times; in particular there is no (indeterministic and irreversible)
wavefunction collapse associated with measurement. The phenomena associated with measurement are claimed to be
explained by decoherence, which occurs when states interact with the environment producing entanglement,
repeatedly splitting the universe into mutually unobservable alternate historiesdistinct universes within a greater
multiverse.

Consistent histories
The consistent histories interpretation generalizes the conventional Copenhagen interpretation and attempts to
provide a natural interpretation of quantum cosmology. The theory is based on a consistency criterion that allows the
history of a system to be described so that the probabilities for each history obey the additive rules of classical
probability. It is claimed to be consistent with the Schrdinger equation.
According to this interpretation, the purpose of a quantum-mechanical theory is to predict the relative probabilities of
various alternative histories (for example, of a particle).

Ensemble interpretation, or statistical interpretation


The Ensemble interpretation, also called the statistical interpretation, can be viewed as a minimalist interpretation.
That is, it claims to make the fewest assumptions associated with the standard mathematics. It takes the statistical
interpretation of Born to the fullest extent. The interpretation states that the wave function does not apply to an
individual system - for example, a single particle - but is an abstract statistical quantity that only applies to an
ensemble (a vast multitude) of similarly prepared systems or particles. Probably the most notable supporter of such
an interpretation was Einstein:
The attempt to conceive the quantum-theoretical description as the complete description of the individual
systems leads to unnatural theoretical interpretations, which become immediately unnecessary if one accepts
the interpretation that the description refers to ensembles of systems and not to individual systems.

62

Interpretations of Quantum Mechanics


Einstein in Albert Einstein: Philosopher-Scientist, ed. P.A. Schilpp (Harper & Row, New York)
The most prominent current advocate of the ensemble interpretation is Leslie E. Ballentine, Professor at Simon
Fraser University, author of the graduate level text book Quantum Mechanics, A Modern Development. An
experiment illustrating the ensemble interpretation is provided in Akira Tonomura's Video clip 1 .[8] It is evident
from this double-slit experiment with an ensemble of individual electrons that, since the quantum mechanical wave
function (absolutely squared) describes the completed interference pattern, it must describe an ensemble.

de BroglieBohm theory
The de BroglieBohm theory of quantum mechanics is a theory by Louis de Broglie and extended later by David
Bohm to include measurements. Particles, which always have positions, are guided by the wavefunction. The
wavefunction evolves according to the Schrdinger wave equation, and the wavefunction never collapses. The
theory takes place in a single space-time, is non-local, and is deterministic. The simultaneous determination of a
particle's position and velocity is subject to the usual uncertainty principle constraint. The theory is considered to be
a hidden variable theory, and by embracing non-locality it satisfies Bell's inequality. It has been shown to be
empirically equivalent to the Copenhagen interpretation, which retains locality but gives up counter factual
definiteness. The measurement problem is resolved, since the particles have definite positions at all times.[9]
Collapse is explained as phenomenological.[10]

Relational quantum mechanics


The essential idea behind relational quantum mechanics, following the precedent of special relativity, is that different
observers may give different accounts of the same series of events: for example, to one observer at a given point in
time, a system may be in a single, "collapsed" eigenstate, while to another observer at the same time, it may be in a
superposition of two or more states. Consequently, if quantum mechanics is to be a complete theory, relational
quantum mechanics argues that the notion of "state" describes not the observed system itself, but the relationship, or
correlation, between the system and its observer(s). The state vector of conventional quantum mechanics becomes a
description of the correlation of some degrees of freedom in the observer, with respect to the observed system.
However, it is held by relational quantum mechanics that this applies to all physical objects, whether or not they are
conscious or macroscopic. Any "measurement event" is seen simply as an ordinary physical interaction, an
establishment of the sort of correlation discussed above. Thus the physical content of the theory has to do not with
objects themselves, but the relations between them.[11] [12]
An independent relational approach to quantum mechanics was developed in analogy with David Bohm's elucidation
of special relativity,[13] in which a detection event is regarded as establishing a relationship between the quantized
field and the detector. The inherent ambiguity associated with applying Heisenberg's uncertainty principle is
subsequently avoided.[14]

Transactional interpretation
The transactional interpretation of quantum mechanics (TIQM) by John G. Cramer is an interpretation of quantum
mechanics inspired by the WheelerFeynman absorber theory.[15] It describes quantum interactions in terms of a
standing wave formed by retarded (forward-in-time) and advanced (backward-in-time) waves. The author argues that
it avoids the philosophical problems with the Copenhagen interpretation and the role of the observer, and resolves
various quantum paradoxes.

63

Interpretations of Quantum Mechanics

Stochastic mechanics
An entirely classical derivation and interpretation of Schrdinger's wave equation by analogy with Brownian motion
was suggested by Princeton University professor Edward Nelson in 1966.[16] Similar considerations had previously
been published, for example by R. Frth (1933), I. Fnyes (1952), and Walter Weizel (1953), and are referenced in
Nelson's paper. More recent work on the stochastic interpretation has been done by M. Pavon.[17] An alternative
stochastic interpretation was developed by Roumen Tsekov.[18]

Objective collapse theories


Objective collapse theories differ from the Copenhagen interpretation in regarding both the wavefunction and the
process of collapse as ontologically objective. In objective theories, collapse occurs randomly ("spontaneous
localization"), or when some physical threshold is reached, with observers having no special role. Thus, they are
realistic, indeterministic, no-hidden-variables theories. The mechanism of collapse is not specified by standard
quantum mechanics, which needs to be extended if this approach is correct, meaning that Objective Collapse is more
of a theory than an interpretation. Examples include the Ghirardi-Rimini-Weber theory[19] and the Penrose
interpretation.[20]

von Neumann/Wigner interpretation: consciousness causes the collapse


In his treatise The Mathematical Foundations of Quantum Mechanics, John von Neumann deeply analyzed the
so-called measurement problem. He concluded that the entire physical universe could be made subject to the
Schrdinger equation (the universal wave function). Since something "outside the calculation" was needed to
collapse the wave function, von Neumann concluded that the collapse was caused by the consciousness of the
experimenter.[21] This point of view was later more prominently expanded on by Eugene Wigner, but remains a view
held by very few physicists.[22]
Variations of the von Neumann interpretation include:
Subjective reduction research
This principle, that consciousness causes the collapse, is the point of intersection between quantum
mechanics and the mind/body problem; and researchers are working to detect conscious events
correlated with physical events that, according to quantum theory, should involve a wave function
collapse; but, thus far, results are inconclusive.[23] [24]
Participatory anthropic principle (PAP)
John Archibald Wheeler's participatory anthropic principle claims that consciousness plays some role in
bringing the universe into existence.[25]
Other physicists have elaborated their own variations of the von Neumann interpretation; including:
Henry P. Stapp (Mindful Universe: Quantum Mechanics and the Participating Observer)
Bruce Rosenblum and Fred Kuttner (Quantum Enigma: Physics Encounters Consciousness)

Many minds
The many-minds interpretation of quantum mechanics extends the many-worlds interpretation by proposing that the
distinction between worlds should be made at the level of the mind of an individual observer.

Quantum logic
Quantum logic can be regarded as a kind of propositional logic suitable for understanding the apparent anomalies
regarding quantum measurement, most notably those concerning composition of measurement operations of
complementary variables. This research area and its name originated in the 1936 paper by Garrett Birkhoff and John

64

Interpretations of Quantum Mechanics


von Neumann, who attempted to reconcile some of the apparent inconsistencies of classical boolean logic with the
facts related to measurement and observation in quantum mechanics.

Modal interpretations of quantum theory


Modal interpretations of quantum mechanics were first conceived of in 1972 by B. van Fraassen, in his paper A
formal approach to the philosophy of science. However, this term now is used to describe a larger set of models that
grew out of this approach. The Stanford Encyclopedia of Philosophy describes several versions:[26]
The Copenhagen variant
Kochen-Dieks-Healey Interpretations
Motivating Early Modal Interpretations, based on the work of R. Clifton, M. Dickson and J. Bub.

Time-symmetric theories
Several theories have been proposed which modify the equations of quantum mechanics to be symmetric with
respect to time reversal.[27] [28] [29] This creates retrocausality: events in the future can affect ones in the past, exactly
as events in the past can affect ones in the future. In these theories, a single measurement cannot fully determine the
state of a system (making them a type of hidden variables theory), but given two measurements performed at
different times, it is possible to calculate the exact state of the system at all intermediate times. The collapse of the
wavefunction is therefore not a physical change to the system, just a change in our knowledge of it due to the second
measurement. Similarly, they explain entanglement as not being a true physical state but just an illusion created by
ignoring retrocausality. The point where two particles appear to "become entangled" is simply a point where each
particle is being influenced by events that occur to the other particle in the future.

Other interpretations
As well as the mainstream interpretations discussed above, a number of other interpretations have been proposed
which have not made a significant scientific impact. These range from proposals by mainstream physicists to the
more occult ideas of quantum mysticism.

Comparison
The most common interpretations are summarized in the table below. The values shown in the cells of the table are
not without controversy, for the precise meanings of some of the concepts involved are unclear and, in fact, are
themselves at the center of the controversy surrounding the given interpretation.
No experimental evidence exists that distinguishes among these interpretations. To that extent, the physical theory
stands, and is consistent with itself and with reality; difficulties arise only when one attempts to "interpret" the
theory. Nevertheless, designing experiments which would test the various interpretations is the subject of active
research.
Most of these interpretations have variants. For example, it is difficult to get a precise definition of the Copenhagen
interpretation as it was developed and argued about by many people.

65

Interpretations of Quantum Mechanics

66

Interpretation

Author(s)

Ensemble
interpretation

Max Born, 1926

Agnostic

No

Yes

Agnostic

No

None

Copenhagen
interpretation

Niels Bohr, Werner


Heisenberg, 1927

No

No1

Yes

No

Yes2

None

de Broglie-Bohm
theory

Louis de Broglie, 1927,


David Bohm, 1952

Yes

Yes3

Yes4

Yes

No

None

von Neumann
interpretation

von Neumann, 1932,


Wheeler, Wigner

No

Yes

Yes

No

Yes

Causal

Quantum logic

Garrett Birkhoff, 1936

Agnostic

Agnostic

Yes5

No

No

Interpretational6

Many-worlds
interpretation

Hugh Everett, 1957

Yes

Yes

No

No

No

None

Time-symmetric
theories

Yakir Aharonov, 1964

Yes

Yes

Yes

Yes

No

No

Stochastic
interpretation

Edward Nelson, 1966

No

No

Yes

No

No

None

Many-minds
interpretation

H. Dieter Zeh, 1970

Yes

Yes

No

No

No

Interpretational7

Consistent histories

Robert B. Griffiths, 1984

Agnostic8

Agnostic8

No

No

No

Interpretational6

Objective collapse
theories

Ghirardi-Rimini-Weber, 1986

No

Yes

Yes

No

Yes

None

Transactional
interpretation

John G. Cramer, 1986

No

Yes

Yes

No

Yes9

None

Relational
interpretation

Carlo Rovelli, 1994

No

Yes

Agnostic10

No

Yes11

Intrinsic12

Deterministic? Wavefunction
real?

Unique
history?

Hidden
Collapsing
variables? wavefunctions?

Observer
role?

According to Bohr, the concept of a physical state independent of the conditions of its experimental observation
does not have a well-defined meaning. According to Heisenberg the wavefunction represents a probability, but
not an objective reality itself in space and time.
2
According to the Copenhagen interpretation, the wavefunction collapses when a measurement is performed.
3
Both particle AND guiding wavefunction are real.
4
Unique particle history, but multiple wave histories.
5
But quantum logic is more limited in applicability than Coherent Histories.
6
Quantum mechanics is regarded as a way of predicting observations, or a theory of measurement.
7
Observers separate the universal wavefunction into orthogonal sets of experiences.
8
If wavefunction is real then this becomes the many-worlds interpretation. If wavefunction less than real, but
more than just information, then Zurek calls this the "existential interpretation".
9
In the TI the collapse of the state vector is interpreted as the completion of the transaction between emitter and
absorber.
10
Comparing histories between systems in this interpretation has no well-defined meaning.
11
Any physical interaction is treated as a collapse event relative to the systems involved, not just macroscopic or
conscious observers.
12
The state of the system is observer-dependent, i.e., the state is specific to the reference frame of the observer.

Interpretations of Quantum Mechanics

Sources
Bub, J. and Clifton, R. 1996. A uniqueness theorem for interpretations of quantum mechanics, Studies in
History and Philosophy of Modern Physics 27B: 181-219
Rudolf Carnap, 1939, "The interpretation of physics," in Foundations of Logic and Mathematics of the
International Encyclopedia of Unified Science. University of Chicago Press.
Dickson, M., 1994, "Wavefunction tails in the modal interpretation" in Hull, D., Forbes, M., and Burian, R., eds.,
Proceedings of the PSA 1" 36676. East Lansing, Michigan: Philosophy of Science Association.
--------, and Clifton, R., 1998, "Lorentz-invariance in modal interpretations" in Dieks, D. and Vermaas, P., eds.,
The Modal Interpretation of Quantum Mechanics. Dordrecht: Kluwer Academic Publishers: 948.
Fuchs, Christopher, 2002, "Quantum Mechanics as Quantum Information (and only a little more)."
arXiv:quant-ph/0205039
-------- and A. Peres, 2000, "Quantum theory needs no interpretation," Physics Today.
Herbert, N., 1985. Quantum Reality: Beyond the New Physics. New York: Doubleday. ISBN 0-385-23569-0.
Hey, Anthony, and Walters, P., 2003. The New Quantum Universe, 2nd ed. Cambridge Univ. Press. ISBN
0-5215-6457-3.
Roman Jackiw and D. Kleppner, 2000, "One Hundred Years of Quantum Physics," Science 289(5481): 893.
Max Jammer, 1966. The Conceptual Development of Quantum Mechanics. McGraw-Hill.
--------, 1974. The Philosophy of Quantum Mechanics. Wiley & Sons.
Al-Khalili, 2003. Quantum: A Guide for the Perplexed. London: Weidenfeld & Nicholson.
de Muynck, W. M., 2002. Foundations of quantum mechanics, an empiricist approach. Dordrecht: Kluwer
Academic Publishers. ISBN 1-4020-0932-1.[30]
Roland Omns, 1999. Understanding Quantum Mechanics. Princeton Univ. Press.
Karl Popper, 1963. Conjectures and Refutations. London: Routledge and Kegan Paul. The chapter "Three views
Concerning Human Knowledge" addresses, among other things, instrumentalism in the physical sciences.
Hans Reichenbach, 1944. Philosophic Foundations of Quantum Mechanics. Univ. of California Press.
Max Tegmark and J. A. Wheeler, 2001, "100 Years of Quantum Mysteries," Scientific American 284: 68.
Bas van Fraassen, 1972, "A formal approach to the philosophy of science," in R. Colodny, ed., Paradigms and
Paradoxes: The Philosophical Challenge of the Quantum Domain. Univ. of Pittsburgh Press: 303-66.
John A. Wheeler and Wojciech Hubert Zurek (eds), Quantum Theory and Measurement, Princeton: Princeton
University Press, ISBN 0-691-08316-9, LoC QC174.125.Q38 1983.

References
[1] For a discussion of the provenance of the phrase "shut up and calculate", see (http:/ / scitation. aip. org/ journals/ doc/ PHTOAD-ft/ vol_57/
iss_5/ 10_1. shtml)
[2] Vaidman, L. (2002, March 24). Many-Worlds Interpretation of Quantum Mechanics. Retrieved March 19, 2010, from Stanford Encyclopedia
of Philosophy: http:/ / plato. stanford. edu/ entries/ qm-manyworlds/ #Teg98
[3] "Who believes in many-worlds?" (http:/ / www. hedweb. com/ everett/ everett. htm#believes). Hedweb.com. . Retrieved 2011-01-24.
[4] Quantum theory as a universal physical theory, by David Deutsch, International Journal of Theoretical Physics, Vol 24 #1 (1985)
[5] Three connections between Everett's interpretation and experiment Quantum Concepts of Space and Time, by David Deutsch, Oxford
University Press (1986)
[6] La nouvelle cuisine, by John S. Bell, last article of Speakable and Unspeakable in Quantum Mechanics, second edition.
[7] A. Einstein, B. Podolsky and N. Rosen, 1935, "Can quantum-mechanical description of physical reality be considered complete?" Phys. Rev.
47: 777.
[8] "An experiment illustrating the ensemble interpretation" (http:/ / www. hitachi. com/ rd/ research/ em/ doubleslit. html). Hitachi.com. .
Retrieved 2011-01-24.
[9] Why Bohm's Theory Solves the Measurement Problem by T. Maudlin, Philosophy of Science 62, pp. 479-483 (September, 1995).
[10] Bohmian Mechanics as the Foundation of Quantum Mechanics by D. Durr, N. Zanghi, and S. Goldstein in Bohmian Mechanics and
Quantum Theory: An Appraisal, edited by J.T. Cushing, A. Fine, and S. Goldstein, Boston Studies in the Philosophy of Science 184, 21-44
(Kluwer, 1996) 1997 arXiv:quant-ph/9511016

67

Interpretations of Quantum Mechanics


[11] "Relational Quantum Mechanics (Stanford Encyclopedia of Philosophy)" (http:/ / plato. stanford. edu/ entries/ qm-relational/ ).
Plato.stanford.edu. . Retrieved 2011-01-24.
[12] For more information, see Carlo Rovelli (1996). "Relational Quantum Mechanics". International Journal of Theoretical Physics 35 (8):
1637. arXiv:quant-ph/9609002. doi:10.1007/BF02302261.
[13] David Bohm, The Special Theory of Relativity, Benjamin, New York, 1965
[14] (http:/ / www. quantum-relativity. org/ Quantum-Relativity. pdf). For a full account (http:/ / www. quantum-relativity. org/
Quantum_Optics_as_a_Relativistic_Theory_of_Light. pdf), see Q. Zheng and T. Kobayashi, 1996, "Quantum Optics as a Relativistic Theory
of Light," Physics Essays 9: 447. Annual Report, Department of Physics, School of Science, University of Tokyo (1992) 240.
[15] "Quantum Nocality - Cramer" (http:/ / www. npl. washington. edu/ npl/ int_rep/ qm_nl. html). Npl.washington.edu. . Retrieved 2011-01-24.
[16] Nelson,E. (1966) Derivation of the Schrdinger Equation from Newtonian Mechanics, Phys. Rev. 150, 1079-1085
[17] M. Pavon, Stochastic mechanics and the Feynman integral, J. Math. Phys. 41, 6060-6078 (2000)
[18] Roumen Tsekov (2009). "Bohmian Mechanics versus Madelung Quantum Hydrodynamics". arXiv:0904.0723[quant-ph].
[19] "Frigg, R. GRW theory" (http:/ / www. romanfrigg. org/ writings/ GRW Theory. pdf) (PDF). . Retrieved 2011-01-24.
[20] "Review of Penrose's Shadows of the Mind" (http:/ / www. thymos. com/ mind/ penrose. html). Thymos.com. . Retrieved 2011-01-24.
[21] von Neumann, John. (1932/1955). Mathematical Foundations of Quantum Mechanics. Princeton: Princeton University Press. Translated
by Robert T. Beyer.
[22] Zvi Schreiber (1995). "The Nine Lives of Schroedinger's Cat". arXiv:quant-ph/9501014[quant-ph].
[23] Dick J. Bierman and Stephen Whitmarsh. (2006). Consciousness and Quantum Physics: Empirical Research on the Subjective Reduction of
the State Vector. in Jack A. Tuszynski (Ed). The Emerging Physics of Consciousness. p. 27-48.
[24] C. M. H. Nunn et. al. (1994). Collapse of a Quantum Field may Affect Brain Function. Journal of Consciousness Studies. 1(1):127-139.
[25] "- The anthropic universe" (http:/ / www. abc. net. au/ rn/ scienceshow/ stories/ 2006/ 1572643. htm). Abc.net.au. 2006-02-18. . Retrieved
2011-01-24.
[26] "Modal Interpretations of Quantum Mechanics (Stanford Encyclopedia of Philosophy)" (http:/ / www. science. uva. nl/ ~seop/ entries/
qm-modal/ ). Science.uva.nl. . Retrieved 2011-01-24.
[27] Aharonov, Y. and Vaidman, L. "On the Two-State Vector Reformulation of Quantum Mechanics." Physica Scripta, Volume T76, pp. 85-92
(1998).
[28] Wharton, K. B. "Time-Symmetric Quantum Mechanics." Foundations of Physics, 37(1), pp. 159-168 (2007).
[29] Wharton, K. B. "A Novel Interpretation of the Klein-Gordon Equation." Foundations of Physics, 40(3), pp. 313-332 (2010).
[30] de Muynck, Willem M (2002). Foundations of quantum mechanics: an empiricist approach (http:/ / books. google. com/
?id=k3rUe8XVjJUC& printsec=frontcover& dq=an+ empiricist+ approach#v=onepage& q=& f=false). Klower Academic Publishers.
ISBN1402009321. . Retrieved 2011-01-24.

Further reading
Almost all authors below are professional physicists.
David Z Albert, 1992. Quantum Mechanics and Experience. Harvard Univ. Press. ISBN 0674741129.
John S. Bell, 1987. Speakable and Unspeakable in Quantum Mechanics. Cambridge Univ. Press, ISBN
0-521-36869-3. The 2004 edition (ISBN 0-521-52338-9) includes two additional papers and an introduction by
Alain Aspect.
Dmitrii Ivanovich Blokhintsev, 1968. The Philosophy of Quantum Mechanics. D. Reidel Publishing Company.
ISBN 9027701059.
David Bohm, 1980. Wholeness and the Implicate Order. London: Routledge. ISBN 0-7100-0971-2.
Adan Cabello (15 November 2004). "Bibliographic guide to the foundations of quantum mechanics and quantum
information". arXiv:quant-ph/0012089[quant-ph].
David Deutsch, 1997. The Fabric of Reality. London: Allen Lane. ISBN 014027541X; ISBN 0713990619.
Argues forcefully against instrumentalism. For general readers.
Bernard d'Espagnat, 1976. Conceptual Foundation of Quantum Mechanics, 2nd ed. Addison Wesley. ISBN
081334087X.
--------, 1983. In Search of Reality. Springer. ISBN 0387113991.
--------, 2003. Veiled Reality: An Analysis of Quantum Mechanical Concepts. Westview Press.
--------, 2006. On Physics and Philosophy. Princeton Univ. Press.
Arthur Fine, 1986. The Shaky Game: Einstein Realism and the Quantum Theory. Science and its Conceptual
Foundations. Univ. of Chicago Press. ISBN 0226249484.
Ghirardi, Giancarlo, 2004. Sneaking a Look at Gods Cards. Princeton Univ. Press.

68

Interpretations of Quantum Mechanics


Gregg Jaeger (2009) Entanglement, Information, and the Interpretation of Quantum Mechanics. (http://www.
springer.com/physics/quantum+physics/book/978-3-540-92127-1) Springer. ISBN 9783540921271.
N. David Mermin (1990) Boojums all the way through. (http://www.cambridge.org/catalogue/catalogue.
asp?isbn=0521388805) Cambridge Univ. Press. ISBN 0521388805.
Roland Omnes, 1994. The Interpretation of Quantum Mechanics. Princeton Univ. Press. ISBN 0691036691.
--------, 1999. Understanding Quantum Mechanics. Princeton Univ. Press.
--------, 1999. Quantum Philosophy: Understanding and Interpreting Contemporary Science. Princeton Univ.
Press.
Roger Penrose, 1989. The Emperor's New Mind. Oxford Univ. Press. ISBN 0-198-51973-7. Especially chpt. 6.
--------, 1994. Shadows of the Mind. Oxford Univ. Press. ISBN 0-19-853978-9.
--------, 2004. The Road to Reality. New York: Alfred A. Knopf. Argues that quantum theory is incomplete.

External links
Stanford Encyclopedia of Philosophy:
" Bohmian mechanics (http://plato.stanford.edu/entries/qm-bohm/)" by Sheldon Goldstein.
" Collapse Theories. (http://plato.stanford.edu/entries/qm-collapse/)" by Giancarlo Ghirardi.

" Copenhagen Interpretation of Quantum Mechanics (http://plato.stanford.edu/entries/qm-copenhagen/)"


by Jan Faye.
" Everett's Relative State Formulation of Quantum Mechanics (http://plato.stanford.edu/entries/qm-everett/
)" by Jeffrey Barrett.
" Many-Worlds Interpretation of Quantum Mechanics (http://plato.stanford.edu/entries/qm-manyworlds/)"
by Lev Vaidman.
" Modal Interpretation of Quantum Mechanics (http://plato.stanford.edu/entries/qm-modal/)" by Michael
Dickson and Dennis Dieks.
" Quantum Entanglement and Information (http://plato.stanford.edu/entries/qt-entangle/)" by Jeffrey Bub.
" Quantum mechanics (http://plato.stanford.edu/entries/qm/)" by Jenann Ismael.
" Relational Quantum Mechanics (http://plato.stanford.edu/entries/qm-relational/)" by Federico Laudisa
and Carlo Rovelli.
" The Role of Decoherence in Quantum Mechanics (http://plato.stanford.edu/entries/qm-decoherence/)" by
Guido Bacciagaluppi.
Willem M. de Muynck, Broad overview (http://www.phys.tue.nl/ktn/Wim/muynck.htm#quantum) of the
realist vs. empiricist interpretations, against oversimplified view of the measurement process.
Schreiber, Z., " The Nine Lives of Schrodinger's Cat. (http://arxiv.org/abs/quant-ph/9501014)" Overview of
competing interpretations.
Interpretations of quantum mechanics on arxiv.org. (http://xstructure.inr.ac.ru/x-bin/subthemes3.
py?level=2&index1=362483&skip=0)
The many worlds of quantum mechanics. (http://www.johnsankey.ca/qm.html)
Erich Joos' Decoherence Website. (http://www.decoherence.de/)
Quantum Mechanics for Philosophers. (http://home.sprynet.com/~owl1/qm.htm) Argues for the superiority of
the Bohm interpretation.
Hidden Variables in Quantum Theory: The Hidden Cultural Variables of their Rejection. (http://www.
miguel-montenegro.com/Hidden_cultural_variables.htm)
Numerous Many Worlds-related Topics and Articles. (http://www.station1.net/DouglasJones/many.htm)
Relational Approach to Quantum Physics. (http://www.quantum-relativity.org/)

Theory of incomplete measurements. (http://cc3d.free.fr/tim.pdf) Deriving quantum mechanics axioms from


properties of acceptable measurements.
Alfred Neumaier's FAQ. (http://www.mat.univie.ac.at/~neum/physics-faq.txt)

69

Interpretations of Quantum Mechanics


Measurement in Quantum Mechanics FAQ. (http://www.mtnmath.com/faq/meas-qm.html)

The Copenhagen Interpretation


The Copenhagen interpretation is an interpretation of quantum mechanics. Amongst physicists, it is the most
widely accepted of the various competing interpretations of quantum mechanics.
Classical physics draws a distinction between particles and energy, holding that only the latter exhibit waveform
characteristics, whereas quantum mechanics is based on the fact that matter has both wave and particle aspects and
postulates that the state of every subatomic particle can be described by a wavefunctiona mathematical
representation used to calculate the probability that the particle, if measured, will be in a given location or state of
motion.
In the early work of Max Planck, Albert Einstein and Niels Bohr, the existence of energy in discrete quantities had
been postulated, in order to avoid certain paradoxes that arise when classical physics is pushed to extremes. Also,
while elementary particles showed predictable properties in many experiments, they became highly unpredictable in
certain contexts; for example, if one attempted to measure their individual trajectories through a simple physical
apparatus.
The Copenhagen interpretation is an attempt to explain the results of the experiments and their mathematical
formulations, in terms of quantum mechanics. It was devised by Bohr, Werner Heisenberg and others in the years
192427. They theorised a new world of energy quanta, entities which fit neither the classical idea of particles nor
the classical idea of waves. They thereby stepped beyond the world of empirical experiments and pragmatic
predictions of such phenomena as the frequencies of light emitted under various conditions. According to their
interpretation, the act of measurement causes the calculated set of probabilities to "collapse" to the value defined by
the measurement. This feature of the mathematics is known as wavefunction collapse.
The Copenhagen interpretation is, in form, a composite of those statements which can be legitimately made in
natural language to complement the statements and predictions made in the language of instrument readings and
mathematical operations. In substance, it attempts to answer the question, "What do these amazing experimental
results really mean?" The concept that quantum mechanics does not yield an objective description of microscopic
reality but deals only with probabilities, and that measurement plays an ineradicable role, is the most significant
characteristic of the Copenhagen interpretation. One consequence of this, derived by Heisenberg, is that knowledge
of the position of a particle limits how accurately its momentum can be knownand vice versa.

Background
Early twentieth-century experiments on the physics of very small-scale phenomena led to the discovery of
phenomena which could not be predicted on the basis of classical physics, and to new models (theories) that
described and predicted very accurately these micro-scale phenomena. These models of the real world, being
observed at this micro scale, could not easily be reconciled with the way objects are observed to behave on the macro
scale of everyday life. The predictions they offered often appeared counter-intuitive to observers. Indeed, they
touched off much consternationeven in the minds of their discoverers.

70

The Copenhagen Interpretation

71

Overview
Because it consists of the views developed by a number of scientists and philosophers during the second quarter of
the 20th Century, there is no definitive statement of the Copenhagen Interpretation.[1] Thus, various ideas have been
associated with it; Asher Peres remarked that very different, sometimes opposite, views are presented as "the
Copenhagen interpretation" by different authors.[2]

Principles
1. A system is completely described by a wave function

, representing an observer's subjective knowledge of the

system. (Heisenberg)
2. The description of nature is essentially probabilistic, with the probability of an event related to the square of the
amplitude of the wave function related to it. (The Born rule, after Max Born)
3. It is not possible to know the value of all the properties of the system at the same time; those properties that are
not known with precision must be described by probabilities. (Heisenberg's uncertainty principle)
4. Matter exhibits a waveparticle duality. An experiment can show the particle-like properties of matter, or the
wave-like properties; in some experiments both of these complementary viewpoints must be invoked to explain
the results, according to the complementarity principle of Niels Bohr.
5. Measuring devices are essentially classical devices, and measure only classical properties such as position and
momentum.
6. The quantum mechanical description of large systems will closely approximate the classical description. (The
correspondence principle of Bohr and Heisenberg.)

Meaning of the wave function


The Copenhagen Interpretation denies that the wave function is anything more than a theoretical concept, or is at
least non-committal about its being a discrete entity or a discernible component of some discrete entity.
The subjective view, that the wave function is merely a mathematical tool for calculating the probabilities in a
specific experiment, is a similar approach to the Ensemble interpretation.
There are some who say that there are objective variants of the Copenhagen Interpretation that allow for a "real"
wave function, but it is questionable whether that view is really consistent with logical positivism and/or with some
of Bohr's statements. Bohr emphasized that science is concerned with predictions of the outcomes of experiments,
and that any additional propositions offered are not scientific but meta-physical. Bohr was heavily influenced by
positivism. On the other hand, Bohr and Heisenberg were not in complete agreement, and they held different views
at different times. Heisenberg in particular was prompted to move towards realism.[3]
Even if the wave function is not regarded as real, there is still a divide between those who treat it as definitely and
entirely subjective, and those who are non-committal or agnostic about the subject. An example of the agnostic view
is given by Carl Friedrich von Weizscker, who, while participating in a colloquium at Cambridge, denied that the
Copenhagen interpretation asserted: "What cannot be observed does not exist". He suggested instead that the
Copenhagen interpretation follows the principle: "What is observed certainly exists; about what is not observed we
are still free to make suitable assumptions. We use that freedom to avoid paradoxes."[4]

Nature of collapse
All versions of the Copenhagen interpretation include at least a formal or methodological version of wave function
collapse,[5] in which unobserved eigenvalues are removed from further consideration. (In other words,
Copenhagenists have always made the assumption of collapse, even in the early days of quantum physics, in the way
that adherents of the Many-worlds interpretation have not.) In more prosaic terms, those who hold to the
Copenhagen understanding are willing to say that a wave function involves the various probabilities that a given

The Copenhagen Interpretation


event will proceed to certain different outcomes. But when one or another of those more- or less-likely outcomes
becomes manifest the other probabilities cease to have any function in the real world. So if an electron passes
through a double slit apparatus there are various probabilities for where on the detection screen that individual
electron will hit. But once it has hit, there is no longer any probability whatsoever that it will hit somewhere else.
Many-worlds interpretations say that an electron hits wherever there is a possibility that it might hit, and that each of
these hits occurs in a separate universe.
An adherent of the subjective view, that the wave function represents nothing but knowledge, would take an equally
subjective view of "collapse".
Some argue that the concept of the collapse of a "real" wave function was introduced by Heisenberg and later
developed by John Von Neumann in 1932.[6]

Acceptance among physicists


According to a poll at a Quantum Mechanics workshop in 1997,[7] the Copenhagen interpretation is the most
widely-accepted specific interpretation of quantum mechanics, followed by the many-worlds interpretation.[8]
Although current trends show substantial competition from alternative interpretations, throughout much of the
twentieth century the Copenhagen interpretation had strong acceptance among physicists. Astrophysicist and science
writer John Gribbin describes it as having fallen from primacy after the 1980s.[9]

Consequences
The nature of the Copenhagen Interpretation is exposed by considering a number of experiments and paradoxes.
1. Schrdinger's Cat
This thought experiment highlights the implications that accepting uncertainty at the microscopic level has on
macroscopic objects. A cat is put in a sealed box, with its life or death made dependent on the state of a
subatomic particle. Thus a description of the cat during the course of the experimenthaving been entangled
with the state of a subatomic particlebecomes a "blur" of "living and dead cat." But this can't be accurate
because it implies the cat is actually both dead and alive until the box is opened to check on it. But the cat, if
he survives, will only remember being alive. Schrdinger resists "so naively accepting as valid a 'blurred
model' for representing reality."[10] How can the cat be both alive and dead?
The Copenhagen Interpretation: The wave function reflects our knowledge of the system. The wave function
means that, once the cat is observed, there is a 50% chance it will be dead, and
50% chance it will be alive.
2. Wigner's Friend
Wigner puts his friend in with the cat. The external observer believes the system is in the state
. His friend however is convinced that cat is alive, i.e. for him, the cat is in the
state

. How can Wigner and his friend see different wave functions?

The Copenhagen Interpretation: Wigner's friend highlights the subjective nature of probability. Each observer
(Wigner and his friend) has different information and therefore different wave functions. The distinction
between the "objective" nature of reality and the subjective nature of probability has led to a great deal of
controversy. Cf. Bayesian versus Frequentist interpretations of probability.
3. Double-Slit Diffraction
Light passes through double slits and onto a screen resulting in a diffraction pattern. Is light a particle or a
wave?
The Copenhagen Interpretation: Light is neither. A particular experiment can demonstrate particle (photon) or
wave properties, but not both at the same time (Bohr's Complementarity Principle).

72

The Copenhagen Interpretation


The same experiment can in theory be performed with any physical system: electrons, protons, atoms,
molecules, viruses, bacteria, cats, humans, elephants, planets, etc. In practice it has been performed for light,
electrons, buckminsterfullerene,[11] [12] and some atoms. Due to the smallness of Planck's constant it is
practically impossible to realize experiments that directly reveal the wave nature of any system bigger than a
few atoms but, in general, quantum mechanics considers all matter as possessing both particle and wave
behaviors. The greater systems (like viruses, bacteria, cats, etc.) are considered as "classical" ones but only as
an approximation.
4. EPR (EinsteinPodolskyRosen) paradox
Entangled "particles" are emitted in a single event. Conservation laws ensure that the measured spin of one
particle must be the opposite of the measured spin of the other, so that if the spin of one particle is measured,
the spin of the other particle is now instantaneously known. The most discomforting aspect of this paradox is
that the effect is instantaneous so that something that happens in one galaxy could cause an instantaneous
change in another galaxy. But, according to Einstein's theory of special relativity, no information-bearing
signal or entity can travel at or faster than the speed of light, which is finite. Thus, it seems as if the
Copenhagen interpretation is inconsistent with special relativity.
The Copenhagen Interpretation: Assuming wave functions are not real, wave-function collapse is interpreted
subjectively. The moment one observer measures the spin of one particle, he knows the spin of the other.
However, another observer cannot benefit until the results of that measurement have been relayed to him, at
less than or equal to the speed of light.
Copenhagenists claim that interpretations of quantum mechanics where the wave function is regarded as real
have problems with EPR-type effects, since they imply that the laws of physics allow for influences to
propagate at speeds greater than the speed of light. However, proponents of Many worlds[13] and the
Transactional interpretation[14] [15] maintain that Copenhagen interpretation is fatally non-local.
The claim that EPR effects violate the principle that information cannot travel faster than the speed of light can
be avoided by noting that they cannot be used for signaling because neither observer can control, or
predetermine, what he observes, and therefore cannot manipulate what the other observer measures.
Relativistic difficulties about establishing which measurement occurred first also undermine the idea that one
observer is causing what the other is measuring.

Criticism
The completeness of quantum mechanics (thesis 1) was attacked by the Einstein-Podolsky-Rosen thought
experiment which was intended to show that quantum physics could not be a complete theory.
Experimental tests of Bell's inequality using particles have supported the quantum mechanical prediction of
entanglement.
The Copenhagen Interpretation gives special status to measurement processes without clearly defining them or
explaining their peculiar effects. In his article entitled "Criticism and Counterproposals to the Copenhagen
Interpretation of Quantum Theory," countering the view of Alexandrov that (in Heisenberg's paraphrase) "the wave
function in configuration space characterizes the objective state of the electron." Heisenberg says,
Of course the introduction of the observer must not be misunderstood to imply that some kind of subjective
features are to be brought into the description of nature. The observer has, rather, only the function of
registering decisions, i.e., processes in space and time, and it does not matter whether the observer is an
apparatus or a human being; but the registration, i.e., the transition from the "possible" to the "actual," is
absolutely necessary here and cannot be omitted from the interpretation of quantum theory.
Heisenberg, Physics and Philosophy, p. 137

73

The Copenhagen Interpretation


Many physicists and philosophers have objected to the Copenhagen interpretation, both on the grounds that it is
non-deterministic and that it includes an undefined measurement process that converts probability functions into
non-probabilistic measurements. Einstein's comments "I, at any rate, am convinced that He (God) does not throw
dice."[16] and "Do you really think the moon isn't there if you aren't looking at it?"[17] exemplify this. Bohr, in
response, said "Einstein, don't tell God what to do".
Steven Weinberg in "Einstein's Mistakes", Physics Today, November 2005, page 31, said:
All this familiar story is true, but it leaves out an irony. Bohr's version of quantum mechanics was deeply
flawed, but not for the reason Einstein thought. The Copenhagen interpretation describes what happens when
an observer makes a measurement, but the observer and the act of measurement are themselves treated
classically. This is surely wrong: Physicists and their apparatus must be governed by the same quantum
mechanical rules that govern everything else in the universe. But these rules are expressed in terms of a wave
function (or, more precisely, a state vector) that evolves in a perfectly deterministic way. So where do the
probabilistic rules of the Copenhagen interpretation come from?
Considerable progress has been made in recent years toward the resolution of the problem, which I cannot go
into here. It is enough to say that neither Bohr nor Einstein had focused on the real problem with quantum
mechanics. The Copenhagen rules clearly work, so they have to be accepted. But this leaves the task of
explaining them by applying the deterministic equation for the evolution of the wave function, the Schrdinger
equation, to observers and their apparatus.
The problem of thinking in terms of classical measurements of a quantum system becomes particularly acute in the
field of quantum cosmology, where the quantum system is the universe.[18]

Alternatives
The Ensemble Interpretation is similar; it offers an interpretation of the wave function, but not for single particles.
The consistent histories interpretation advertises itself as "Copenhagen done right". Consciousness causes collapse is
often confused with the Copenhagen interpretation.
If the wave function is regarded as ontologically real, and collapse is entirely rejected, a many worlds theory results.
If wave function collapse is regarded as ontologically real as well, an objective collapse theory is obtained. Dropping
the principle that the wave function is a complete description results in a hidden variable theory.
Many physicists have subscribed to the instrumentalist interpretation of quantum mechanics, a position often equated
with eschewing all interpretation. It is summarized by the sentence "Shut up and calculate!". While this slogan is
sometimes attributed to Paul Dirac[19] or Richard Feynman, it is in fact due to the lesser known David Mermin.[20]

Notes and references


[1] In fact Bohr and Heisenberg never totally agreed on how to understand the mathematical formalism of quantum mechanics, and none of them
ever used the term "the Copenhagen interpretation" as a joint name for their ideas. Bohr once distanced himself from what he considered to be
Heisenberg's more subjective interpretation Stanford Encyclopedia of Philosophy (http:/ / plato. stanford. edu/ entries/ qm-copenhagen/ )
[2] "There seems to be at least as many different Copenhagen interpretations as people who use that term, probably there are more. For example,
in two classic articles on the foundations of quantum mechanics, Ballentine (1970) and Stapp(1972) give diametrically opposite definitions of
'Copenhagen.'", Asher Peres (2002). "Popper's experiment and the Copenhagen interpretation". Stud. History Philos. Modern Physics 33 (23).
arXiv:quant-ph/9910078.
[3] "Historically, Heisenberg wanted to base quantum theory solely on observable quantities such as the intensity of spectral lines, getting rid of
all intuitive (anschauliche) concepts such as particle trajectories in space-time. This attitude changed drastically with his paper in which he
introduced the uncertainty relations there he put forward the point of view that it is the theory which decides what can be observed. His
move from positivism to operationalism can be clearly understood as a reaction on the advent of Schrdingers wave mechanics which, in
particular due to its intuitiveness, became soon very popular among physicists. In fact, the word anschaulich (intuitive) is contained in the title
of Heisenbergs paper.", from Claus Kiefer (2002). "On the interpretation of quantum theory - from Copenhagen to the present day".
arXiv:quant-ph/0210152[quant-ph].
[4] John Cramer on the Copenhagen Interpretation (http:/ / www. npl. washington. edu/ npl/ int_rep/ tiqm/ TI_20. html#2. 0)

74

The Copenhagen Interpretation


[5] "To summarize, one can identify the following ingredients as being characteristic for the Copenhagen interpretation(s)[...]Reduction of the
wave packet as a formal rule without dynamical significance", Claus Kiefer (2002). "On the interpretation of quantum theory - from
Copenhagen to the present day". arXiv:quant-ph/0210152[quant-ph].
[6] "the collapse or reduction of the wave function. This was introduced by Heisenberg in his uncertainty paper [3] and later postulated by
von Neumann as a dynamical process independent of the Schrodinger equation", Claus Kiefer (2002). "On the interpretation of quantum
theory - from Copenhagen to the present day". arXiv:quant-ph/0210152[quant-ph].
[7] Max Tegmark (1998). "The Interpretation of Quantum Mechanics: Many Worlds or Many Words?". Fortsch.Phys. 46 (68): 855862.
arXiv:quant-ph/9709032. doi:10.1002/(SICI)1521-3978(199811)46:6/8<855::AID-PROP855>3.0.CO;2-Q.
[8] The Many Worlds Interpretation of Quantum Mechanics (http:/ / www. hep. upenn. edu/ ~max/ everett. ps)
[9] Gribbin, J. Q for Quantum
[10] Erwin Schrdinger, in an article in the Proceedings of the American Philosophical Society, 124, 323-38.
[11] Nairz, Olaf; Brezger, Bjrn; Arndt, Markus; Zeilinger, Anton (2001). "Diffraction of Complex Molecules by Structures Made of Light".
Physical Review Letters 87 (16). doi:10.1103/PhysRevLett.87.160401.
[12] Brezger, Bjrn; Hackermller, Lucia; Uttenthaler, Stefan; Petschinka, Julia; Arndt, Markus; Zeilinger, Anton (2002). "Matter-Wave
Interferometer for Large Molecules". Physical Review Letters 88 (10). doi:10.1103/PhysRevLett.88.100404.
[13] Michael price on nonlocality in Many Worlds (http:/ / www. hedweb. com/ manworld. htm#local)
[14] Relativity and Causality in the Transactional Interpretation (http:/ / www. npl. washington. edu/ npl/ int_rep/ tiqm/ TI_38. html#3. 9)
[15] Collapse and Nonlocality in the Transactional Interpretation (http:/ / www. npl. washington. edu/ npl/ int_rep/ tiqm/ TI_33. html#3. 7)
[16] "God does not throw dice" quote
[17] A. Pais, Einstein and the quantum theory, Reviews of Modern Physics 51, 863-914 (1979), p. 907.
[18] 'Since the Universe naturally contains all of its observers, the problem arises to come up with an interpretation of quantum theory that
contains no classical realms on the fundamental level.', Claus Kiefer (2002). "On the interpretation of quantum theory - from Copenhagen to
the present day". arXiv:quant-ph/0210152[quant-ph].
[19] http:/ / home. fnal. gov/ ~skands/ slides/ A-Quantum-Journey. ppt
[20] N. David Mermin. "Could Feynman Have Said This?" (http:/ / scitation. aip. org/ journals/ doc/ PHTOAD-ft/ vol_57/ iss_5/ 10_1. shtml).
Physics Today 57 (5). .

Further reading

G. Weihs et al., Phys. Rev. Lett. 81 (1998) 5039


M. Rowe et al., Nature 409 (2001) 791.
J.A. Wheeler & W.H. Zurek (eds) , Quantum Theory and Measurement, Princeton University Press 1983
A. Petersen, Quantum Physics and the Philosophical Tradition, MIT Press 1968
H. Margeneau, The Nature of Physical Reality, McGraw-Hill 1950
M. Chown, Forever Quantum, New Scientist No. 2595 (2007) 37.
T. Schrmann, A Single Particle Uncertainty Relation, Acta Physica Polonica B39 (2008) 587. (http://th-www.
if.uj.edu.pl/acta/vol39/pdf/v39p0587.pdf)

External links
Copenhagen Interpretation (Stanford Encyclopedia of Philosophy) (http://plato.stanford.edu/entries/
qm-copenhagen)
Physics FAQ section about Bell's inequality (http://math.ucr.edu/home/baez/physics/Quantum/
bells_inequality.html)
The Copenhagen Interpretation of Quantum Mechanics (http://www.benbest.com/science/quantum.html)
Preprint of Afshar Experiment (http://www.irims.org/quant-ph/030503/)
The Quantum Illusion (http://knol.google.com/k/andy-biddulph/the-quantum-illusion/2na7zaaxgtohe/2/)

75

76

4. Einstein's Objections
Principle of Locality
In physics, the principle of locality states that an object is influenced directly only by its immediate surroundings.
Experiments have shown that quantum mechanically entangled particles must violate either the principle of locality
or the form of philosophical realism known as counterfactual definiteness.

Pre-quantum mechanics
In the 17th Century Newton's law of universal gravitation was formulated in terms of action at a distance, thereby
violating the principle of locality. Coulomb's law of electric forces was initially also formulated as instantaneous
action at a distance, but was later superseded by Maxwell's Equations of electromagnetism which obey locality.
In 1905 Albert Einstein's Special Theory of Relativity postulated that no material or energy can travel faster than the
speed of light, and Einstein thereby sought to reformulate physical laws in a way which obeyed the principle of
locality. He later succeeded in producing an alternative theory of gravitation, General Relativity, which obeys the
principle of locality.
However a different challenge to the principle of locality subsequently emerged from the theory of Quantum
Mechanics, which Einstein himself had helped to create.

Quantum mechanics
Einstein's view
EPR Paradox
Albert Einstein felt that there was something fundamentally incorrect with quantum mechanics since it predicted
violations of the principle of locality. Seeking to undermine quantum mechanics, in a famous paper he and his
co-authors articulated the Einstein-Podolsky-Rosen Paradox. Thirty years later John Stewart Bell responded with a
paper which posited (paraphrased) that no physical theory of local hidden variables can ever reproduce all of the
predictions of quantum mechanics (known as Bell's theorem).
Philosophical view
Einstein assumed that the principle of locality was necessary, and that there could be no violations of it. He said[1] :
The following idea characterises the relative independence of objects far apart in space, A and B: external influence on A has no direct
influence on B; this is known as the Principle of Local Action, which is used consistently only in field theory. If this axiom were to be
completely abolished, the idea of the existence of quasienclosed systems, and thereby the postulation of laws which can be checked
empirically in the accepted sense, would become impossible.

Principle of Locality

Local realism
Local realism is the combination of the principle of locality with the "realistic" assumption that all objects must
objectively have a pre-existing value for any possible measurement before the measurement is made. Einstein liked
to say that the Moon is "out there" even when no one is observing it.

Realism
Realism in the sense used by physicists does not equate to realism in metaphysics.[2] The latter is the claim that the
world is in some sense mind-independent: that even if the results of a possible measurement do not pre-exist the act
of measurement, that does not require that they are the creation of the observer (contrary to the "consciousness
causes collapse" interpretation of quantum mechanics). Furthermore, a mind-independent property does not have to
be the value of some physical variable such as position or momentum. A property can be dispositional (or potential),
i.e. it can be a tendency: in the way that glass objects tend to break, or are disposed to break, even if they do not
actually break. Likewise, the mind-independent properties of quantum systems could consist of a tendency to
respond to particular measurements with particular values with ascertainable probability.[3] Such an ontology would
be metaphysically realistic, without being realistic in the physicist's sense of "local realism" (which would require
that a single value be produced with certainty).
Local realism is a significant feature of classical mechanics, of general relativity, and of electrodynamics; but
quantum mechanics largely rejects this principle due to the theory of distant quantum entanglements, an
interpretation rejected by Einstein in the EPR paradox but subsequently apparently quantified by Bell's
inequalities.[4] Any theory, such as quantum mechanics, that violates Bell's inequalities must abandon either local
realism or counterfactual definiteness; but some physicists dispute that experiments have demonstrated Bell's
violations, on the grounds that the sub-class of inhomogeneous Bell inequalities has not been tested or due to
experimental limitations in the tests. Different interpretations of quantum mechanics violate different parts of local
realism and/or counterfactual definiteness.

Copenhagen interpretation
In most of the conventional interpretations, such as the Copenhagen interpretation and the interpretation based on
Consistent Histories, where the wavefunction is not assumed to be a direct physical interpretation of reality, it is
local realism that is rejected. These interpretations propose that actual definite properties of a physical system "do
not exist" prior to the measurement; and the wavefunction has a restricted interpretation, as nothing more than a
mathematical tool used to calculate the probabilities of experimental outcomes, hence in agreement with positivism
in philosophy as the only topic that science should discuss.
In the version of the Copenhagen interpretation where the wavefunction is assumed to be a physical interpretation of
reality (the nature of which is unspecified) the principle of locality is violated during the measurement process via
wavefunction collapse. This is a non-local process because Born's Rule, when applied to the system's wavefunction,
yields a probability density for all regions of space and time. Upon actual measurement of the physical system, the
probability density vanishes everywhere instantaneously, except where (and when) the measured entity is found to
exist. This "vanishing" is postulated to be a real physical process, and clearly non-local (i.e. faster than light) if the
wavefunction is considered physically real and the probability density has converged to zero at arbitrarily far
distances during the finite time required for the measurement process.

77

Principle of Locality

Bohm interpretation
The Bohm interpretation preserves realism, hence it needs to violate the principle of locality in order to achieve the
required correlations.

Many-worlds interpretation
In the many-worlds interpretation both realism and locality are retained, but counterfactual definiteness is rejected
by the extension of the notion of reality to allow the existence of parallel universes.
Because the differences between the different interpretations are mostly philosophical ones (except for the Bohm and
many-worlds interpretations), physicists usually employ language in which the important statements are neutral with
regard to all of the interpretations. In this framework, only the measurable action at a distance - a superluminal
propagation of real, physical information - would usually be considered in violation of the principle of locality by
physicists. Such phenomena have never been seen, and they are not predicted by the current theories.

Relativity
Locality is one of the axioms of relativistic quantum field theory, as required for causality. The formalization of
locality in this case is as follows: if we have two observables, each localized within two distinct space-time regions
which happen to be at a spacelike separation from each other, the observables must commute. Alternatively, a
solution to the field equations is local if the underlying equations are either Lorentz invariant or, more generally,
generally covariant or locally Lorentz invariant.

Notes
[1]
[2]
[3]
[4]

"Quantum Mechanics and Reality" ("Quanten-Mechanik und Wirklichkeit", Dialectica 2:320-324, 1948)
Norsen, T. - Against "Realism" (http:/ / arxiv. org/ abs/ quant-ph/ 0607057v2)
Ian Thomson's dispositional quantum mechanics (http:/ / www. generativescience. org/ )
Ben Dov, Y. Local Realism and the Crucial experiment. (http:/ / bendov. info/ eng/ crucial. htm)

78

EPR Paradox

EPR Paradox
The EPR paradox (or EinsteinPodolskyRosen paradox) is a topic in quantum physics and the philosophy of
science concerning the measurement and description of microscopic systems (such as individual photons, electrons
or atoms) by the methods of quantum physics. It refers to the dichotomy that either the measurement of a physical
quantity in one system must affect the measurement of a physical quantity in another, spatially separate, system or
the description of reality given by a wave function must be incomplete.
This challenge to the Copenhagen interpretation of quantum physics (that only the position or momentum of a
particle, but not both, can be known with certainty) originated from the consequences of a thought experiment in
1935 by Einstein, Podolsky and Rosen. The paper they authored indicated what seemed to be a flaw in the
interpretation. The experiment involved two systems that initially interact with each other and are then separated.
Then the position or momentum of one of the systems is measured, and due to the known relationship between the
(measured) value of the first particle and the value of the second particle, the observer is aware of that value in the
second particle. A measurement of the other value is then made on the second particle, and, once again, due to the
relationship between the two particles, that value is then known in the first particle. This outcome seems to violate
the uncertainty principle, as both the position and momentum of a single particle would be known with certainty.[1]
Einstein struggled to the end of his life for a theory that could better comply with causality, protesting against the
view that there exists no objective physical reality other than that which is revealed through measurement interpreted
in terms of quantum mechanical formalism. However, since Einstein's death, experiments analogous to that of the
EPR paradox have been carried out, starting in 1976 by French scientists at the Saclay Nuclear Research Centre.
These experiments appear to show that the local realism theory is false.[2]

Quantum mechanics and its interpretation


Since the early twentieth century, quantum theory has proved to be successful in describing accurately the physical
reality of the mesoscopic and microscopic world, in multiple reproducible physics experiments.
Quantum mechanics was developed with the aim of describing atoms and explaining the observed spectral lines in a
measurement apparatus. Although disputed, it has yet to be seriously challenged. Philosophical interpretations of
quantum phenomena, however, are another matter: the question of how to interpret the mathematical formulation of
quantum mechanics has given rise to a variety of different answers from people of different philosophical
backgrounds (see Interpretation of quantum mechanics).
Quantum theory and quantum mechanics do not provide single measurement outcomes in a deterministic way.
According to the theory of quantum mechanics known as the Copenhagen interpretation, measurement causes an
instantaneous collapse of the wave function describing the quantum system into an eigenstate of the observable state
that was measured.
The most prominent opponent of the Copenhagen interpretation is Albert Einstein. In his view, quantum mechanics
is incomplete. Commenting on this, other writers (such as John von Neumann[3] and David Bohm[4] ) have suggested
that consequently there would have to be 'hidden' variables responsible for random measurement results, something
which was not expressly claimed in the original paper.
That paper, "Can Quantum-Mechanical Description of Physical Reality Be Considered Complete?"[5], authored by
Einstein, Podolsky and Rosen in 1935, condensed the philosophical discussion into a physical argument. They claim
that given a specific experiment, in which the outcome of a measurement is known before the measurement takes
place, there must exist something in the real world, an "element of reality", which determines the measurement
outcome. They postulate that these elements of reality are local, in the sense that each belongs to a certain point in
spacetime. Each element may only be influenced by events which are located in the backward light cone of its point
in spacetime (i.e. the past). These claims are founded on assumptions about nature which constitute what is now

79

EPR Paradox
known as local realism.
Though the EPR paper has often been taken as an exact expression of Einstein's views, it was primarily authored by
Podolsky, based on discussions at the Institute for Advanced Study with Einstein and Rosen. Einstein later expressed
to Erwin Schrdinger that, "it did not come out as well as I had originally wanted; rather, the essential thing was, so
to speak, smothered by the formalism."[6] In 1948 Einstein presented a less formal account of his local realist ideas.

Description of the paradox


The original EPR paradox challenges the prediction of quantum mechanics that it is impossible to know both the
position and the momentum of a quantum particle. This can be extended to other pairs of physical properties.

EPR paper
The original paper describes what happens to "two systems I and II, which we permit to interact ...", and, after some
time, "we suppose that there is no longer any interaction between the two parts." In the words of Kumar (2009), it
has "Two particles, A and B, [which] interact briefly and then move off in opposite directions."[7] According to
Heisenberg's uncertainty principle, it is impossible to measure both the momentum and the position of particle B,
say, exactly. However, it is possible to measure the exact position of particle A and the exact momentum of particle
B. By calculation, therefore, with the exact position of particle A known, the exact position of particle B can be
known. Also, with the exact momentum of particle B known, the exact momentum of particle A can be worked out.
"EPR argued that they had proved that ... particle B can have simultaneously exact values of position and
momentum."
This is a paradox in Quantum Mechanics: The theory predicts that both values cannot be known for a particle, and
yet the EPR experiment shows that they can. "Therefore, the quantum mechanical description of physical reality,
EPR conclude, is incomplete."[8] The paper says: "We are thus forced to conclude that the quantum-mechanical
description of physical reality given by wave functions is not complete."
The EPR paper ends with:
While we have thus shown that the wave function does not provide a complete description of the
physical reality, we left open the question of whether or not such a description exists. We believe,
however, that such a theory is possible.

Greene version
The paradox was explained in a different way by Greene and others, using electron spin.
A positron and an electron are emitted from a source, by pion decay, so that their spins are opposite; one particles
spin about any axis is the negative of the other's. But making a measurement of a particles spin about one axis
disturbs the particle, so that its spin about any other axis cannot be measured.
However, the measurement of the electrons spin about the x-axis discloses the positrons spin about the x-axis. Since
this measurement has not disturbed the positron in any way, it cannot be that the positron only came to have that
state when it was measured, because it was not measured.
The positrons spin about the y-axis can also be measured, so that it is known the positron had a definite spin about
two axes much more information than the positron is capable of holding, and a "hidden variable" according to
some interpretations of EPR.

80

EPR Paradox

Measurements on an entangled state


We have a source that emits electron-positron pairs, with the electron sent to destination A, where there is an
observer named Alice, and the positron sent to destination B, where there is an observer named Bob. According to
quantum mechanics, we can arrange our source so that each emitted pair occupies a quantum state called a spin
singlet. The particles are thus said to be entangled. This can be viewed as a quantum superposition of two states,
which we call state I and state II. In state I, the electron has spin pointing upward along the z-axis (+z) and the
positron has spin pointing downward along the z-axis (-z). In state II, the electron has spin -z and the positron has
spin +z. Therefore, it is impossible (without measuring) to know the definite state of spin of either particle in the
spin singlet.

The EPR thought experiment, performed with electron-positron pairs. A source (center) sends particles toward two observers, electrons to
Alice (left) and positrons to Bob (right), who can perform spin measurements.

Alice now measures the spin along the z-axis. She can obtain one of two possible outcomes: +z or -z. Suppose she
gets +z. According to quantum mechanics, the quantum state of the system collapses into state I (different
interpretations of quantum mechanics have different ways of saying this, but the basic result is the same). The
quantum state determines the probable outcomes of any measurement performed on the system. In this case, if Bob
subsequently measures spin along the z-axis, there is 100% probability that he will obtain -z. Similarly, if Alice gets
-z, Bob will get +z.
There is, of course, nothing special about choosing the z-axis: according to quantum mechanics the spin singlet state
may equally well be expressed as a superposition of spin states pointing in the x direction. Suppose that Alice and
Bob had decided to measure spin along the x-axis. We'll call these states Ia and IIa. In state Ia, Alice's electron has
spin +x and Bob's positron has spin -x. In state IIa, Alice's electron has spin -x and Bob's positron has spin +x.
Therefore, if Alice measures +x, the system 'collapses' into state Ia, and Bob will get -x. If Alice measures -x, the
system collapses into state IIa, and Bob will get +x.
Whatever axis their spins are measured along, they are always found to be opposite. This can only be explained if the
particles are linked in some way. Either they were created with a definite (opposite) spin about every axisa
"hidden variable" argument or they are linked so that one electron "feels" which axis the other is having its spin
measured along, and becomes its opposite about that one axisan "entanglement" argument. Moreover, if the two
particles have their spins measured about different axes, once the electron's spin has been measured about the x-axis
(and the positron's spin about the x-axis deduced), the positron's spin about the y-axis will no longer be certain, as if
(a) it knows that the measurement has taken place, or (b) it has a definite spin already, about a second axisa hidden
variable.
In quantum mechanics, the x-spin and z-spin are "incompatible observables", meaning there is a Heisenberg
uncertainty principle operating between them: a quantum state cannot possess a definite value for both of these
variables. Suppose Alice measures the z-spin and obtains +z, so that the quantum state collapses into state I. Now,

81

EPR Paradox
instead of measuring the z-spin as well, Bob measures the x-spin. According to quantum mechanics, when the system
is in state I, Bob's x-spin measurement will have a 50% probability of producing +x and a 50% probability of -x. It is
impossible to predict which outcome will appear until Bob actually performs the measurement.
Here is the crux of the matter. You might imagine that, when Bob measures the x-spin of his positron, he would
get an answer with absolute certainty, since prior to this he hasn't disturbed his particle at all. But Bob's positron has
a 50% probability of producing +x and a 50% probability of -xso the outcome is not certain. Bob's positron
"knows" that Alice's electron has been measured, and its z-spin detected, and hence B's z-spin calculated, so its
x-spin is uncertain.
Put another way, how does Bob's positron know which way to point if Alice decides (based on information
unavailable to Bob) to measure x (i.e. to be the opposite of Alice's electron's spin about the x-axis) and also how to
point if Alice measures z, since it is only supposed to know one thing at a time? The Copenhagen interpretation rules
that say the wave function "collapses" at the time of measurement, so there must be action at a distance
(entanglement) or the positron must know more than it's supposed to (hidden variables).
Here is the paradox summed up:
It is one thing to say that physical measurement of the first particle's momentum affects uncertainty in its own
position, but to say that measuring the first particle's momentum affects the uncertainty in the position of the other is
another thing altogether. Einstein, Podolsky and Rosen asked how can the second particle "know" to have precisely
defined momentum but uncertain position? Since this implies that one particle is communicating with the other
instantaneously across space, i.e. faster than light, this is the "paradox".
Incidentally, Bell used spin as his example, but many types of physical quantitieswhat quantum mechanics refer to
as "observables"can be used. The EPR paper used momentum for the observable. Experimental realisations of the
EPR scenario often use photon polarization, because polarized photons are easy to prepare and measure.

Locality in the EPR experiment


The principle of locality states that physical processes occurring at one place should have no immediate effect on the
elements of reality at another location. At first sight, this appears to be a reasonable assumption to make, as it seems
to be a consequence of special relativity, which states that information can never be transmitted faster than the speed
of light without violating causality. It is generally believed that any theory which violates causality would also be
internally inconsistent, and thus deeply unsatisfactory.
It turns out that the usual rules for combining quantum mechanical and classical descriptions violate the principle of
locality without violating causality. Causality is preserved because there is no way for Alice to transmit messages
(i.e. information) to Bob by manipulating her measurement axis. Whichever axis she uses, she has a 50% probability
of obtaining "+" and 50% probability of obtaining "-", completely at random; according to quantum mechanics, it is
fundamentally impossible for her to influence what result she gets. Furthermore, Bob is only able to perform his
measurement once: there is a fundamental property of quantum mechanics, known as the "no cloning theorem",
which makes it impossible for him to make a million copies of the electron he receives, perform a spin measurement
on each, and look at the statistical distribution of the results. Therefore, in the one measurement he is allowed to
make, there is a 50% probability of getting "+" and 50% of getting "-", regardless of whether or not his axis is
aligned with Alice's.
However, the principle of locality appeals powerfully to physical intuition, and Einstein, Podolsky and Rosen were
unwilling to abandon it. Einstein derided the quantum mechanical predictions as "spooky action at a distance". The
conclusion they drew was that quantum mechanics is not a complete theory.
In recent years, however, doubt has been cast on EPR's conclusion due to developments in understanding locality
and especially quantum decoherence. The word locality has several different meanings in physics. For example, in
quantum field theory "locality" means that quantum fields at different points of space do not interact with one

82

EPR Paradox
another. However, quantum field theories that are "local" in this sense appear to violate the principle of locality as
defined by EPR, but they nevertheless do not violate locality in a more general sense. Wavefunction collapse can be
viewed as an epiphenomenon of quantum decoherence, which in turn is nothing more than an effect of the
underlying local time evolution of the wavefunction of a system and all of its environment. Since the underlying
behaviour doesn't violate local causality, it follows that neither does the additional effect of wavefunction collapse,
whether real or apparent. Therefore, as outlined in the example above, neither the EPR experiment nor any quantum
experiment demonstrates that faster-than-light signaling is possible.

Resolving the paradox


Hidden variables
There are several ways to resolve the EPR paradox. The one suggested by EPR is that quantum mechanics, despite
its success in a wide variety of experimental scenarios, is actually an incomplete theory. In other words, there is
some yet undiscovered theory of nature to which quantum mechanics acts as a kind of statistical approximation
(albeit an exceedingly successful one). Unlike quantum mechanics, the more complete theory contains variables
corresponding to all the "elements of reality". There must be some unknown mechanism acting on these variables to
give rise to the observed effects of "non-commuting quantum observables", i.e. the Heisenberg uncertainty principle.
Such a theory is called a hidden variable theory.
To illustrate this idea, we can formulate a very simple hidden variable theory for the above thought experiment. One
supposes that the quantum spin-singlet states emitted by the source are actually approximate descriptions for "true"
physical states possessing definite values for the z-spin and x-spin. In these "true" states, the electron going to Bob
always has spin values opposite to the electron going to Alice, but the values are otherwise completely random. For
example, the first pair emitted by the source might be "(+z, -x) to Alice and (-z, +x) to Bob", the next pair "(-z, -x) to
Alice and (+z, +x) to Bob", and so forth. Therefore, if Bob's measurement axis is aligned with Alice's, he will
necessarily get the opposite of whatever Alice gets; otherwise, he will get "+" and "-" with equal probability.
Assuming we restrict our measurements to the z and x axes, such a hidden variable theory is experimentally
indistinguishable from quantum mechanics. In reality, of course, there is an (uncountably) infinite number of axes
along which Alice and Bob can perform their measurements, so there has to be an infinite number of independent
hidden variables. However, this is not a serious problem; we have formulated a very simplistic hidden variable
theory, and a more sophisticated theory might be able to patch it up. It turns out that there is a much more serious
challenge to the idea of hidden variables.
Bell's inequality
In 1964, John Bell showed that the predictions of quantum mechanics in the EPR thought experiment are
significantly different from the predictions of a very broad class of hidden variable theories (the local hidden
variable theories). Roughly speaking, quantum mechanics has a much stronger statistical correlation with
measurement results performed on different axes than do the hidden variable theories. These differences, expressed
using inequality relations known as "Bell's inequalities", are in principle experimentally detectable. Later work by
Eberhard showed that the key properties of local hidden variable theories which lead to Bell's inequalities are locality
and counter-factual definiteness. Any theory in which these principles apply produces the inequalities. A Fine
subsequently showed that any theory satisfying the inequalities can be modeled by a local hidden variable theory.
After the publication of Bell's paper, a variety of experiments were devised to test Bell's inequalities (experiments
which generally rely on photon polarization measurement). All the experiments conducted to date have found
behaviour in line with the predictions of standard quantum mechanics theory.
However, Bell's theorem does not apply to all possible philosophically realist theories. It is a common misconception
that quantum mechanics is inconsistent with all notions of philosophical realism, but realist interpretations of

83

EPR Paradox
quantum mechanics are possible, although, as discussed above, such interpretations must reject either locality or
counter-factual definiteness. Mainstream physics prefers to keep locality, while striving also to maintain a notion of
realism that nevertheless rejects counter-factual definiteness. Examples of such mainstream realist interpretations are
the consistent histories interpretation and the transactional interpretation. Fine's work showed that, taking locality as
a given, there exist scenarios in which two statistical variables are correlated in a manner inconsistent with
counter-factual definiteness, and that such scenarios are no more mysterious than any other, despite the inconsistency
with counter-factual definiteness seeming 'counter-intuitive'.
Violation of locality is difficult to reconcile with special relativity, and is thought to be incompatible with the
principle of causality. On the other hand the Bohm interpretation of quantum mechanics keeps counter-factual
definiteness while introducing a conjectured non-local mechanism called the 'quantum potential'. Some workers in
the field have also attempted to formulate hidden variable theories that exploit loopholes in actual experiments, such
as the assumptions made in interpreting experimental data, although no theory has been proposed that can reproduce
all the results of quantum mechanics.
There are also individual EPR-like experiments that have no local hidden variables explanation. Examples have been
suggested by David Bohm and by Lucien Hardy.

Einstein's hope for a purely algebraic theory


The Bohm interpretation of quantum mechanics hypothesizes that the state of the universe evolves smoothly through
time with no collapsing of quantum wavefunctions. One problem for the Copenhagen interpretation is to precisely
define wavefunction collapse. Einstein maintained that quantum mechanics is physically incomplete and logically
unsatisfactory. In "The Meaning of Relativity," Einstein wrote, "One can give good reasons why reality cannot at all
be represented by a continuous field. From the quantum phenomena it appears to follow with certainty that a finite
system of finite energy can be completely described by a finite set of numbers (quantum numbers). This does not
seem to be in accordance with a continuum theory and must lead to an attempt to find a purely algebraic theory for
the representation of reality. But nobody knows how to find the basis for such a theory." If time, space, and energy
are secondary features derived from a substrate below the Planck scale, then Einstein's hypothetical algebraic system
might resolve the EPR paradox (although Bell's theorem would still be valid). Edward Fredkin in the Fredkin Finite
Nature Hypothesis has suggested an informational basis for Einstein's hypothetical algebraic system. If physical
reality is totally finite, then the Copenhagen interpretation might be an approximation to an information processing
system below the Planck scale.

"Acceptable theories" and the experiment


According to the present view of the situation, quantum mechanics flatly contradicts Einstein's philosophical
postulate that any acceptable physical theory must fulfill "local realism".
In the EPR paper (1935) the authors realised that quantum mechanics was inconsistent with their assumptions, but
Einstein nevertheless thought, erroneously, that quantum mechanics might simply be augmented by 'hidden
variables' (i.e. variables which were, at that point, still obscure to him), without any other change, to achieve an
acceptable theory. He pursued these ideas until the end of his life, in 1955, over twenty years.
In contrast, John Bell, in his 1964 paper, showed "once and for all" that quantum mechanics and Einstein's
assumptions lead to different results: different by a factor of 32 for certain correlations. So the issue of
"acceptability", up to that time mainly concerning theory (even philosophy), finally became experimentally
decidable.
There are many Bell test experiments, e.g. those of Alain Aspect and others. They all tend to show that pure quantum
mechanics, and not Einstein's "local realism", is acceptable.[9] Thus, according to Karl Popper these experiments
falsify Einstein's philosophical assumptions, especially his ideas of "hidden variables", whereas quantum mechanics
itself remains a good candidate for a theory that is acceptable in a wider context.

84

EPR Paradox

85

Implications for quantum mechanics


Most physicists today believe that quantum mechanics is correct, and that the EPR paradox is a "paradox" only
because classical intuitions do not correspond to physical reality. How EPR is interpreted regarding locality depends
on the interpretation of quantum mechanics one uses. In the Copenhagen interpretation, it is usually understood that
instantaneous wavefunction collapse does occur. However, the view that there is no causal instantaneous effect has
also been proposed within the Copenhagen interpretation: in this alternate view, measurement affects our ability to
define (and measure) quantities in the physical system, not the system itself. In the many-worlds interpretation, a
kind of locality is preserved, since the effects of irreversible operations such as measurement arise from the
relativization of a global state to a subsystem such as that of an observer.
The EPR paradox has deepened our understanding of quantum mechanics by exposing the fundamentally
non-classical characteristics of the measurement process. Prior to the publication of the EPR paper, a measurement
was often visualized as a physical disturbance inflicted directly upon the measured system. For instance, when
measuring the position of an electron, one imagines shining a light on it, thus disturbing the electron and producing
the quantum mechanical uncertainties in its position. Such explanations, which are still encountered in popular
expositions of quantum mechanics, are debunked by the EPR paradox, which shows that a "measurement" can be
performed on a particle without disturbing it directly, by performing a measurement on a distant entangled particle.
Technologies relying on quantum entanglement are now being developed. In quantum cryptography, entangled
particles are used to transmit signals that cannot be eavesdropped upon without leaving a trace. In quantum
computation, entangled quantum states are used to perform computations in parallel, which may allow certain
calculations to be performed much more quickly than they ever could be with classical computers.

Mathematical formulation
The above discussion can be expressed mathematically using the quantum mechanical formulation of spin. The spin
degree of freedom for an electron is associated with a two-dimensional complex Hilbert space H, with each quantum
state corresponding to a vector in that space. The operators corresponding to the spin along the x, y, and z direction,
denoted Sx, Sy, and Sz respectively, can be represented using the Pauli matrices:

where

stands for Planck's constant divided by 2.

The eigenstates of Sz are represented as

and the eigenstates of Sx are represented as

The Hilbert space of the electron pair is

, the tensor product of the two electrons' Hilbert spaces. The spin

singlet state is

where the two terms on the right hand side are what we have referred to as state I and state II above.
From the above equations, it can be shown that the spin singlet can also be written as

where the terms on the right hand side are what we have referred to as state Ia and state IIa.

EPR Paradox
To illustrate how this leads to the violation of local realism, we need to show that after Alice's measurement of Sz (or
Sx), Bob's value of Sz (or Sx) is uniquely determined, and therefore corresponds to an "element of physical reality".
This follows from the principles of measurement in quantum mechanics. When Sz is measured, the system state
collapses into an eigenvector of Sz. If the measurement result is +z, this means that immediately after measurement
the system state undergoes an orthogonal projection of onto the space of states of the form

For the spin singlet, the new state is

Similarly, if Alice's measurement result is -z, the system undergoes an orthogonal projection onto

which means that the new state is

This implies that the measurement for Sz for Bob's electron is now determined. It will be -z in the first case or +z in
the second case.
It remains only to show that Sx and Sz cannot simultaneously possess definite values in quantum mechanics. One may
show in a straightforward manner that no possible vector can be an eigenvector of both matrices. More generally,
one may use the fact that the operators do not commute,

along with the Heisenberg uncertainty relation

References
Selected papers
A. Aspect, Bell's inequality test: more ideal than ever, Nature 398 189 (1999). [10]
J.S. Bell, On the Einstein-Poldolsky-Rosen paradox [11], Physics 1
195bbcv://prola.aps.org/abstract/PR/v48/i8/p696_1]
P.H. Eberhard, Bell's theorem without hidden variables. Nuovo Cimento 38B1 75 (1977).
P.H. Eberhard, Bell's theorem and the different concepts of locality. Nuovo Cimento 46B 392 (1978).
A. Einstein, B. Podolsky, and N. Rosen, Can quantum-mechanical description of physical reality be considered
complete? [12] Phys. Rev. 47 777 (1935). [5]
A. Fine, Hidden Variables, Joint Probability, and the Bell Inequalities. Phys. Rev. Lett. 48, 291 (1982).[13]
A. Fine, Do Correlations need to be explained?, in Philosophical Consequences of Quantum Theory: Reflections
on Bell's Theorem, edited by Cushing & McMullin (University of Notre Dame Press, 1986).
L. Hardy, Nonlocality for two particles without inequalities for almost all entangled states. Phys. Rev. Lett. 71
1665 (1993).[14]
M. Mizuki, A classical interpretation of Bell's inequality. Annales de la Fondation Louis de Broglie 26 683
(2001).
P. Pluch, "Theory for Quantum Probability", PhD Thesis University of Klagenfurt (2006)
M. A. Rowe, D. Kielpinski, V. Meyer, C. A. Sackett, W. M. Itano, C. Monroe and D. J. Wineland, Experimental
violation of a Bell's inequality with efficient detection, Nature 409, 791-794 (15 February 2001). [15]
M. Smerlak, C. Rovelli, Relational EPR [16]

86

EPR Paradox
Notes
[1] Einstein, A; B Podolsky, N Rosen (1935-05-15). "Can Quantum-Mechanical Description of Physical Reality be Considered Complete?".
Physical Review 47 (10): 777780. doi:10.1103/PhysRev.47.777.
[2] Gribbin, J (1984). In Search of Schroedinger's cat. Black Swan. ISBN0704530716.
[3] von Neumann, J. (1932/1955). In Mathematische Grundlagen der Quantenmechanik, Springer, Berlin, translated into English by Beyer, R.T.,
Princeton University Press, Princeton, cited by Baggott, J. (2004) Beyond Measure: Modern physics, philosophy, and the meaning of quantum
theory, Oxford University Press, Oxford, ISBN 0-19-852927-9, pages 144-145.
[4] Bohm, D. (1951). Quantum Theory (http:/ / books. google. com. au/ books?id=9DWim3RhymsC& printsec=frontcover& dq=david+ bohm+
quantum+ theory& source=bl& ots=6G-2u1wtav& sig=Q1GcoVDLFRmKOmDYFAJte6LzrZU& hl=en& ei=Pv45TNSnLYffcfnS6foO&
sa=X& oi=book_result& ct=result& resnum=7& ved=0CEEQ6AEwBg#v=onepage& q& f=false), Prentice-Hall, Englewood Cliffs, page 29,
and Chapter 5 section 3, and Chapter 22 Section 19.
[5] http:/ / prola. aps. org/ abstract/ PR/ v47/ i10/ p777_1
[6] Quoted in Kaiser, David. "Bringing the human actors back on stage: the personal context of the Einstein-Bohr debate," British Journal for the
History of Science 27 (1994): 129-152, on page 147.
[7] Kumar, M., Quantum, Icon Books, 2009, p. 305.
[8] Kumar, M., Quantum, Icon Books, 2009, p. 306.
[9] Aspect A (1999-03-18). "Bells inequality test: more ideal than ever" (http:/ / www-ece. rice. edu/ ~kono/ ELEC565/ Aspect_Nature. pdf).
Nature 398: 18990. doi:10.1038/18296. . Retrieved 2010-09-08.
[10] http:/ / www-ece. rice. edu/ ~kono/ ELEC565/ Aspect_Nature. pdf
[11] http:/ / www. drchinese. com/ David/ Bell_Compact. pdf
[12] http:/ / www. drchinese. com/ David/ EPR. pdf
[13]
[14]
[15]
[16]

http:/ / prola. aps. org/ abstract/ PRL/ v48/ i5/ p291_1


http:/ / prola. aps. org/ abstract/ PRL/ v71/ i11/ p1665_1
http:/ / www. nature. com/ nature/ journal/ v409/ n6822/ full/ 409791a0. html
http:/ / arxiv. org/ abs/ quant-ph/ 0604064

Books
John S. Bell (1987) Speakable and Unspeakable in Quantum Mechanics. Cambridge University Press. ISBN
0-521-36869-3.
Arthur Fine (1996) The Shaky Game: Einstein, Realism and the Quantum Theory, 2nd ed. Univ. of Chicago Press.
J.J. Sakurai, J. J. (1994) Modern Quantum Mechanics. Addison-Wesley: 174187, 223-232. ISBN
0-201-53929-2.
Selleri, F. (1988) Quantum Mechanics Versus Local Realism: The Einstein-Podolsky-Rosen Paradox. New York:
Plenum Press. ISBN 0-306-42739-7
Leon Lederman, L., Teresi, D. (1993). The God Particle: If the Universe is the Answer, What is the Question?
Houghton Mifflin Company, pages 21, 187 to 189.
John Gribbin (1984) In Search of Schroedinger's Cat. Black Swan. ISBN 9780552125550

External links
The Einstein-Podolsky-Rosen Argument in Quantum Theory; 1.2 The argument in the text;
http://plato.stanford.edu/entries/qt-epr/#1.2
The original EPR paper. (http://prola.aps.org/abstract/PR/v47/i10/p777_1)
Stanford Encyclopedia of Philosophy: " The Einstein-Podolsky-Rosen Argument in Quantum Theory (http://
plato.stanford.edu/entries/qt-epr/)" by Arthur Fine.
Abner Shimony (2004) " Bells Theorem. (http://plato.stanford.edu/entries/bell-theorem/)"
EPR, Bell & Aspect: The Original References. (http://www.drchinese.com/David/EPR_Bell_Aspect.htm)
Does Bell's Inequality Principle rule out local theories of quantum mechanics? (http://math.ucr.edu/home/
baez/physics/Quantum/bells_inequality.html) From the Usenet Physics FAQ.
Theoretical use of EPR in teleportation. (http://www.research.ibm.com/journal/rd/481/brassard.html)
Effective use of EPR in cryptography. (http://www.dhushara.com/book/quantcos/aq/qcrypt.htm)
EPR experiment with single photons interactive. (http://www.QuantumLab.de)

87

Bell's Theorem

Bell's Theorem
In theoretical physics, Bell's theorem (a.k.a. Bell's inequality) is a no-go theorem, loosely stating that:
no physical theory of local hidden variables can reproduce all of the predictions of quantum mechanics.
It is the most famous legacy of the late physicist John Stewart Bell.
Bell's theorem has important implications for physics and the philosophy of science, as it implies that quantum
theory must necessarily violate either the Principle of locality or counterfactual definiteness. In conjunction with
experiments verifying the quantum mechanical predictions of Bell-type systems, Bell's theorem posits that certain
quantum effects travel faster than light, hence limits the class of tenable 'hidden variable' theories to the non-local
variety.
However, no test of the theorem has ever fulfilled all of the conditions implicit in it, due to Bell test loopholes.
Accordingly, the issue is not finally settled.

Overview
Bells theorem implies that the local hidden variable interpretation of quantum mechanics favoured by Einstein, also
known as local realism, necessitates certain conditions which are violated by the findings of experiments performed
on entangled quantum systems. A variety of experiments appear to have verified the predictions of the Copenhagen
interpretation of quantum mechanics, thereby ruling out the local hidden variables interpretation while also proving
the existence of superluminal effects.
The theorem has several variations, but the most common involves a system comprising two Spin- particles (or
equivalently, any pair of 2-state quantum systems or qubits), which were the particles used in Bell's original proof
and will be the primary focus of this article.
As in the EinsteinPodolskyRosen (EPR)
paradox, Bell considered an experiment in
which a source produces pairs of correlated
particles. For example, a pair of particles
produced in a Bell state, such that if the
spins are measured along the same axes they
are certain to produce identical results, are
sent to two distant observers: Alice and Bob.
Illustration of Bell test for massless spin-1 particles, such as photons (see
In each trial, both of the observers
polarization of light). Source produces spin singlet pair, one particle sent to Alice
independently chooses an axis along which
and the other to Bob. Each performs one of the two spin measurements.
to measure the spin of their respective
particle, and each measurement yields a
result of either spin-up (+1) or spin-down (1). Whether or not Alice and Bob obtain the same result depends on the
relationship between the orientations of the two spin measurements, and in general is subject to some uncertainty.
The classical incarnation of Bell's theorem is derived from the statistical properties observed over many runs of this
experiment.
Mathematically the correlation between results is represented by their product (thus taking on values of 1 for a
single run). While measuring the spin of these entangled particles along the same axis will always result in identical
(i.e. perfectly correlated) results, measurement at perpendicular directions will have only a 50% chance of matching
(i.e. will have a 50% probability of an uncorrelated result). These basic cases are illustrated in the table to the right.

88

Bell's Theorem

89

Same axis:

pair 1

pair 2

pair 3

pair 4

... pair n

Alice, 0:

...

Bob, 0:

...

Correlation: (

+1

+1

+1

+1

... ) / n = +1
(100% identical)

Orthogonal axes:

pair 1

pair 2

pair 3

pair 4

... pair n

Alice, 0:

...

Bob, 90:

...

Correlation: (

+1

+1

... ) / n = 0
(50% identical)

With the measurements oriented at arbitrary angles, the mathematics of quantum mechanics predicts that the
probability of the two measurements matching (i.e. the "quantum mechanical expectation of a correlation") is equal
to the cosine of the angle, an outcome which closely agrees with what is observed in practice.
Bell achieved his breakthrough by first assuming that a theory of local hidden variables could reproduce these
results. Without making any assumptions about the specific form of the theory beyond requirements of basic
consistency, he was able to derive a mathematical inequality that was clearly at odds with the results (described
above) predicted by quantum mechanics and observed experimentally. Thus, Bell's theorem rules out local realism as
a viable interpretation of quantum mechanics, though it still leaves the door open for non-local realism.
Over the years, Bell's theorem has undergone a wide variety of experimental tests. Two possible loopholes in the
original argument have been proposed, the detection loophole[1] and the communication loophole,[1] each prompting
a new round of experiments that re-verified the integrity of the result.[1] To date, Bell's theorem is supported by a
substantial body of evidence and is treated as a fundamental principle of physics in mainstream quantum mechanics
textbooks.[2] [3] However, no principle of physics can ever be absolutely beyond question, and some physicists do
not accept the theorem's validity.[4] Yet Kumar (2009) writes: "Although no experiment had been conducted in
which every loophole has been closed, most physicists accept that Bell's inequality has been violated." [5]

Importance of the theorem


Bell's theorem, derived in his seminal 1964 paper titled On the Einstein Podolsky Rosen paradox,[6] has been called
"the most profound in science."[7] The title of the article refers to the famous paper by Einstein, Podolsky and
Rosen[8] purporting to prove the incompleteness of quantum mechanics. In his paper, Bell started from the same two
assumptions as did EPR, namely (i) reality (that microscopic objects have real properties determining the outcomes
of quantum mechanical measurements), and (ii) locality (that reality is not influenced by measurements performed
simultaneously at a large distance). Bell was able to derive from those two assumptions an important result, namely
Bell's inequality, implying that at least one of the assumptions must be false.
In two respects Bell's 1964 paper was a step forward compared to the EPR paper: firstly, it considered more hidden
variables than merely the element of physical reality in the EPR paper; and, more importantly, Bell's inequality was
liable to be experimentally tested, thus yielding the opportunity to lift the discussion on the completeness of quantum
mechanics from metaphysics to physics. Whereas Bell's paper deals only with deterministic hidden variable theories,
Bell's theorem was later generalized to stochastic theories[9] as well, and it was also realised[10] that the theorem can
even be proven without introducing hidden variables.
After the EPR paper, quantum mechanics was in an unsatisfactory position: either it was incomplete, in the sense
that it failed to account for some elements of physical reality, or it violated the principle of a finite propagation speed
of physical effects. In a modified version of the EPR thought experiment, two observers, now commonly referred to

Bell's Theorem
as Alice and Bob, perform independent measurements of spin on a pair of electrons, prepared at a source in a special
state called a spin singlet state. It is the conclusion of EPR that once Alice measures spin in one direction (e.g. on the
x axis), Bob's measurement in that direction is determined with certainty, as being the opposite outcome to that of
Alice, whereas immediately before Alice's measurement Bob's outcome was only statistically determined (i.e. was
only a probability, not a certainty); thus, either the spin in each direction is an element of physical reality, or the
effects travel from Alice to Bob instantly.
In QM, predictions are formulated in terms of probabilities for example, the probability that an electron will be
detected in a particular place, or the probability that its spin is up or down. The idea persisted, however, that the
electron in fact has a definite position and spin, and that QM's weakness is its inability to predict those values
precisely. The possibility existed that some unknown theory, such as a hidden variables theory, might be able to
predict those quantities exactly, while at the same time also being in complete agreement with the probabilities
predicted by QM. If such a hidden variables theory exists, then because the hidden variables are not described by
QM the latter would be an incomplete theory.
The desire to find a local realist theory was driven by two assumptions:
1. That objects have a definite state which determines the values of all other measurable properties, such as position
and momentum.
2. That the effects of local actions, such as measurements, cannot travel faster than the speed of light (in
consequence of special relativity). Thus if the observers are sufficiently far apart, a measurement made by one can
have no effect on a measurement made by the other.
In the form of local realism used by Bell, the predictions of the theory result from the application of classical
probability theory to an underlying parameter space. By a simple argument based on classical probability, he showed
that correlations between measurements are bounded in a way that is violated by QM.
Bell's theorem seemed to put an end to local realism. Per Bell's theorem, either quantum mechanics or local realism
is wrong, as they are mutually exclusive. Experiments were needed to determine which of them is correct, but it took
many years and many improvements in technology to perform them.
The Bell test experiments have been interpreted as showing that the Bell inequalities are violated in favour of QM.
The no-communication theorem shows that the observers cannot use the effect to communicate (classical)
information to each other faster than the speed of light, but the fair sampling and no enhancement assumptions
require more careful consideration (below).
John Bell's paper examines both John von Neumann's 1932 proof of the incompatibility of hidden variables with QM
and the seminal 1935 EPR paper on the subject by Albert Einstein and his colleagues.

Bell inequalities
Bell inequalities concern measurements made by observers on pairs of particles that have interacted and then
separated. According to quantum mechanics they are entangled, while local realism would limit the correlation of
subsequent measurements of the particles.
Different authors subsequently derived inequalities similar to Bells original inequality, and these are here
collectively termed Bell inequalities. All Bell inequalities describe experiments in which the predicted result from
quantum entanglement differs from that flowing from local realism. The inequalities assume that each quantum-level
object has a well defined state that accounts for all its measurable properties and that distant objects do not exchange
information faster than the speed of light. These well defined states are typically called hidden variables, the
properties that Einstein posited when he stated his famous objection to quantum mechanics: "God does not play
dice."
Bell showed that under quantum mechanics, the mathematics of which contains no local hidden variables, the Bell
inequalities can nevertheless be violated: the properties of a particle are not clear, but may be correlated with those

90

Bell's Theorem

91

of another particle due to quantum entanglement, allowing their state to be well defined only after a measurement is
made on either particle. That restriction agrees with the Heisenberg uncertainty principle, a fundamental concept in
quantum mechanics.
In Bell's words:
Theoretical physicists live in a classical world, looking out into a quantum-mechanical world. The latter we
describe only subjectively, in terms of procedures and results in our classical domain. (...) Now nobody knows
just where the boundary between the classical and the quantum domain is situated. (...) More plausible to me is
that we will find that there is no boundary. The wave functions would prove to be a provisional or incomplete
description of the quantum-mechanical part. It is this possibility, of a homogeneous account of the world,
which is for me the chief motivation of the study of the so-called "hidden variable" possibility.
(...) A second motivation is connected with the statistical character of quantum-mechanical predictions. Once
the incompleteness of the wave function description is suspected, it can be conjectured that random statistical
fluctuations are determined by the extra "hidden" variables "hidden" because at this stage we can only
conjecture their existence and certainly cannot control them.
(...) A third motivation is in the peculiar character of some quantum-mechanical predictions, which seem
almost to cry out for a hidden variable interpretation. This is the famous argument of Einstein, Podolsky and
Rosen. (...) We will find, in fact, that no local deterministic hidden-variable theory can reproduce all the
experimental predictions of quantum mechanics. This opens the possibility of bringing the question into the
experimental domain, by trying to approximate as well as possible the idealized situations in which local
hidden variables and quantum mechanics cannot agree
In probability theory, repeated measurements of system properties can be regarded as repeated sampling of random
variables. In Bell's experiment, Alice can choose a detector setting to measure either
or
and Bob can
choose a detector setting to measure either

or

. Measurements of Alice and Bob may be somehow

correlated with each other, but the Bell inequalities say that if the correlation stems from local random variables,
there is a limit to the amount of correlation one might expect to see.

Original Bell's inequality


The original inequality that Bell derived was:[6]

where C is the "correlation" of the particle pairs and a, b and c settings of the apparatus. This inequality is not used
in practice. For one thing, it is true only for genuinely "two-outcome" systems, not for the "three-outcome" ones
(with possible outcomes of zero as well as +1 and 1) encountered in real experiments. For another, it applies only to
a very restricted set of hidden variable theories, namely those for which the outcomes on both sides of the
experiment are always exactly anticorrelated when the analysers are parallel, in agreement with the quantum
mechanical prediction.
There is a simple limit of Bell's inequality which has the virtue of being completely intuitive. If the result of three
different statistical coin-flips A, B, and C have the property that:
1. A and B are the same (both heads or both tails) 99% of the time
2. B and C are the same 99% of the time
then A and C are the same at least 98% of the time. The number of mismatches between A and B (1/100) plus the
number of mismatches between B and C (1/100) are together the maximum possible number of mismatches between
A and C.
In quantum mechanics, by letting A, B, and C be the values of the spin of two entangled particles measured relative
to some axis at 0 degrees, degrees, and 2 degrees respectively, the overlap of the wavefunction between the
different angles is proportional to
. The probability that A and B give the same answer is

Bell's Theorem

92

, where is proportional to . This is also the probability that B and C give the same answer. But A and C are the same
1(2)2 of the time. Choosing the angle so that
, A and B are 99% correlated, B and C are 99% correlated and A
and C are only 96% correlated.
Imagine that two entangled particles in a spin singlet are shot out to two distant locations, and the spins of both are
measured in the direction A. The spins are 100% correlated (actually, anti-correlated but for this argument that is
equivalent). The same is true if both spins are measured in directions B or C. It is safe to conclude that any hidden
variables which determine the A,B and C measurements in the two particles are 100% correlated and can be used
interchangeably.
If A is measured on one particle and B on the other, the correlation between them is 99%. If B is measured on one
and C on the other, the correlation is 99%. This allows us to conclude that the hidden variables determining A and B
are 99% correlated and B and C are 99% correlated. But if A is measured in one particle and C in the other, the
results are only 96% correlated, which is a contradiction. The intuitive formulation is due to David Mermin, while
the small-angle limit is emphasized in Bell's original article.

CHSH inequality
In addition to Bell's original inequality,[6] the form given by John Clauser, Michael Horne, Abner Shimony and R. A.
Holt,[11] (the CHSH form) is especially important,[11] as it gives classical limits to the expected correlation for the
above experiment conducted by Alice and Bob:

where C denotes correlation.


Correlation of observables X, Y is defined as

This is a non-normalized form of the correlation coefficient considered in statistics (see Quantum correlation).
In order to formulate Bell's theorem, we formalize local realism as follows:
1. There is a probability space

and the observed outcomes by both Alice and Bob result by random sampling of

the parameter
.
2. The values observed by Alice or Bob are functions of the local detector settings and the hidden parameter only.
Thus
Value observed by Alice with detector setting is
Value observed by Bob with detector setting is
Implicit in assumption 1) above, the hidden parameter space
random variable X on

with respect to

has a probability measure

and the expectation of a

is written

where for accessibility of notation we assume that the probability measure has a density.
Bell's inequality. The CHSH inequality (1) holds under the hidden variables assumptions above.
For simplicity, let us first assume the observed values are +1 or 1; we remove this assumption in Remark 1 below.
Let

is 0. Thus

. Then at least one of

Bell's Theorem

93

and therefore

Remark 1. The correlation inequality (1) still holds if the variables

are allowed to take on any

real values between 1 and +1. Indeed, the relevant idea is that each summand in the above average is bounded
above by 2. This is easily seen to be true in the more general case:

To justify the upper bound 2 asserted in the last inequality, without loss of generality, we can assume that

In that case

Remark 2. Though the important component of the hidden parameter

in Bell's original proof is associated with

the source and is shared by Alice and Bob, there may be others that are associated with the separate detectors, these
others being independent. This argument was used by Bell in 1971, and again by Clauser and Horne in 1974,[12] to
justify a generalisation of the theorem forced on them by the real experiments, in which detectors were never 100%
efficient. The derivations were given in terms of the averages of the outcomes over the local detector variables. The
formalisation of local realism was thus effectively changed, replacing A and B by averages and retaining the symbol
but with a slightly different meaning. It was henceforth restricted (in most theoretical work) to mean only those
components that were associated with the source.
However, with the extension proved in Remark 1, CHSH inequality still holds even if the instruments themselves
contain hidden variables. In that case, averaging over the instrument hidden variables gives new variables:

on

which still have values in the range [1,+1] to which we can apply the previous result.

Bell inequalities are violated by quantum mechanical predictions


In the usual quantum mechanical formalism, the observables X and Y are represented as self-adjoint operators on a
Hilbert space. To compute the correlation, assume that X and Y are represented by matrices in a finite dimensional
space and that X and Y commute; this special case suffices for our purposes below. The von Neumann measurement
postulate states: a series of measurements of an observable X on a series of identical systems in state produces a
distribution of real values. By the assumption that observables are finite matrices, this distribution is discrete. The
probability of observing is non-zero if and only if is an eigenvalue of the matrix X and moreover the probability

Bell's Theorem

94

is

where EX () is the projector corresponding to the eigenvalue . The system state immediately after the measurement
is

From this, we can show that the correlation of commuting observables X and Y in a pure state

is

We apply this fact in the context of the EPR paradox. The measurements performed by Alice and Bob are spin
measurements on electrons. Alice can choose between two detector settings labelled a and a; these settings
correspond to measurement of spin along the z or the x axis. Bob can choose between two detector settings labelled b
and b; these correspond to measurement of spin along the z or x axis, where the x z coordinate system is rotated
135 relative to the x z coordinate system. The spin observables are represented by the 2 2 self-adjoint matrices:

These are the Pauli spin matrices normalized so that the corresponding eigenvalues are +1, 1. As is customary, we
denote the eigenvectors of Sx by
Let

be the spin singlet state for a pair of electrons discussed in the EPR paradox. This is a specially constructed

state described by the following vector in the tensor product

Now let us apply the CHSH formalism to the measurements that can be performed by Alice and Bob.

Illustration of Bell test for spin 1/2 particles. Source produces spin singlet pairs, one
particle of each pair is sent to Alice and the other to Bob. Each performs one of the two
spin measurements.

The operators

correspond to Bob's spin measurements along x and z. Note that the A operators

commute with the B operators, so we can apply our calculation for the correlation. In this case, we can show that the
CHSH inequality fails. In fact, a straightforward calculation shows that

Bell's Theorem

95

and

so that

Bell's Theorem: If the quantum mechanical formalism is correct, then the system consisting of a pair of entangled
electrons cannot satisfy the principle of local realism. Note that
is indeed the upper bound for quantum
mechanics called Tsirelson's bound. The operators giving this maximal value are always isomorphic to the Pauli
matrices.

Practical experiments testing Bell's theorem


Experimental tests can determine
whether the Bell inequalities required
by local realism hold up to the
empirical evidence.

Scheme of a "two-channel" Bell test


The source S produces pairs of "photons", sent in opposite directions. Each photon
encounters a two-channel polariser whose orientation (a or b) can be set by the
experimenter. Emerging signals from each channel are detected and coincidences of four
types (++, , + and +) counted by the coincidence monitor.

Bell's inequalities are tested by


"coincidence counts" from a Bell test
experiment such as the optical one
shown in the diagram. Pairs of
particles are emitted as a result of a
quantum process, analysed with
respect to some key property such as
polarisation direction, then detected.
The setting (orientations) of the
analysers are selected by the

experimenter.
Bell test experiments to date overwhelmingly violate Bell's inequality. Indeed, a table of Bell test experiments
performed prior to 1986 is given in 4.5 of Redhead, 1987.[13] Of the thirteen experiments listed, only two reached
results contradictory to quantum mechanics; moreover, according to the same source, when the experiments were
repeated, "the discrepancies with QM could not be reproduced".
Nevertheless, the issue is not conclusively settled. According to Shimony's 2004 Stanford Encyclopedia overview
article:[1]
Most of the dozens of experiments performed so far have favored Quantum Mechanics, but not decisively
because of the 'detection loopholes' or the 'communication loophole.' The latter has been nearly decisively
blocked by a recent experiment and there is a good prospect for blocking the former.
To explore the 'detection loophole', one must distinguish the classes of homogeneous and inhomogeneous Bell
inequality.
The standard assumption in Quantum Optics is that "all photons of given frequency, direction and polarization are
identical" so that photodetectors treat all incident photons on an equal basis. Such a fair sampling assumption
generally goes unacknowledged, yet it effectively limits the range of local theories to those which conceive of the
light field as corpuscular. The assumption excludes a large family of local realist theories, in particular, Max Planck's
description. We must remember the cautionary words of Albert Einstein[14] shortly before he died: "Nowadays every
Tom, Dick and Harry ('jeder Kerl' in German original) thinks he knows what a photon is, but he is mistaken".

Bell's Theorem
Objective physical properties for Bells analysis (local realist theories) include the wave amplitude of a light signal.
Those who maintain the concept of duality, or simply of light being a wave, recognize the possibility or actuality that
the emitted atomic light signals have a range of amplitudes and, furthermore, that the amplitudes are modified when
the signal passes through analyzing devices such as polarizers and beam splitters. It follows that not all signals have
the same detection probability.[15]

Two classes of Bell inequalities


The fair sampling problem was faced openly in the 1970s. In early designs of their 1973 experiment, Freedman and
Clauser[16] used fair sampling in the form of the Clauser-Horne-Shimony-Holt (CHSH[11] ) hypothesis. However,
shortly afterwards Clauser and Horne[12] made the important distinction between inhomogeneous (IBI) and
homogeneous (HBI) Bell inequalities. Testing an IBI requires that we compare certain coincidence rates in two
separated detectors with the singles rates of the two detectors. Nobody needed to perform the experiment, because
singles rates with all detectors in the 1970s were at least ten times all the coincidence rates. So, taking into account
this low detector efficiency, the QM prediction actually satisfied the IBI. To arrive at an experimental design in
which the QM prediction violates IBI we require detectors whose efficiency exceeds 82% for singlet states, but have
very low dark rate and short dead and resolving times. This is well above the 30% achievable[17] so Shimonys
optimism in the Stanford Encyclopedia, quoted in the preceding section, appears over-stated.

Practical challenges
Because detectors don't detect a large fraction of all photons, Clauser and Horne[12] recognized that testing Bell's
inequality requires some extra assumptions. They introduced the No Enhancement Hypothesis (NEH):
a light signal, originating in an atomic cascade for example, has a certain probability of activating a
detector. Then, if a polarizer is interposed between the cascade and the detector, the detection
probability cannot increase.
Given this assumption, there is a Bell inequality between the coincidence rates with polarizers and coincidence rates
without polarizers.
The experiment was performed by Freedman and Clauser,[16] who found that the Bell's inequality was violated. So
the no-enhancement hypothesis cannot be true in a local hidden variables model. The Freedman-Clauser experiment
reveals that local hidden variables imply the new phenomenon of signal enhancement:
In the total set of signals from an atomic cascade there is a subset whose detection probability increases
as a result of passing through a linear polarizer.
This is perhaps not surprising, as it is known that adding noise to data can, in the presence of a threshold, help reveal
hidden signals (this property is known as stochastic resonance[18] ). One cannot conclude that this is the only
local-realist alternative to Quantum Optics, but it does show that the word loophole is biased. Moreover, the analysis
leads us to recognize that the Bell-inequality experiments, rather than showing a breakdown of realism or locality,
are capable of revealing important new phenomena.

Theoretical challenges
Most advocates of the hidden variables idea believe that experiments have ruled out local hidden variables. They are
ready to give up locality, explaining the violation of Bell's inequality by means of a "non-local" hidden variable
theory, in which the particles exchange information about their states. This is the basis of the Bohm interpretation of
quantum mechanics, which requires that all particles in the universe be able to instantaneously exchange information
with all others. A recent experiment ruled out a large class of non-Bohmian "non-local" hidden variable theories.[19]
If the hidden variables can communicate with each other faster than light, Bell's inequality can easily be violated.
Once one particle is measured, it can communicate the necessary correlations to the other particle. Since in relativity

96

Bell's Theorem
the notion of simultaneity is not absolute, this is unattractive. One idea is to replace instantaneous communication
with a process which travels backwards in time along the past Light cone. This is the idea behind a transactional
interpretation of quantum mechanics, which interprets the statistical emergence of a quantum history as a gradual
coming to agreement between histories that go both forward and backward in time.[20]
A few advocates of deterministic models have not given up on local hidden variables. For example, Gerard 't Hooft
has argued that the superdeterminism loophole cannot be dismissed.[21]
The quantum mechanical wavefunction can also provide a local realistic description, if the wavefunction values are
interpreted as the fundamental quantities that describe reality. Such an approach is called a many-worlds
interpretation of quantum mechanics. In this view, two distant observers both split into superpositions when
measuring a spin. The Bell inequality violations are no longer counterintuitive, because it is not clear which copy of
the observer B observer A will see when going to compare notes. If reality includes all the different outcomes,
locality in physical space (not outcome space) places no restrictions on how the split observers can meet up.
This implies that there is a subtle assumption in the argument that realism is incompatible with quantum mechanics
and locality. The assumption, in its weakest form, is called counterfactual definiteness. This states that if the results
of an experiment are always observed to be definite, there is a quantity which determines what the outcome would
have been even if you don't do the experiment.
Many worlds interpretations are not only counterfactually indefinite, they are factually indefinite. The results of all
experiments, even ones that have been performed, are not uniquely determined.
Another type of challenge is to reproduce the correlations of quantum entanglement while respecting the requirement
of locality. One possible way to accomplish this is by a simulation in which particle pairs are generated from a
computer, separated, and sent to two separate computers which play the role of detectors. At each detector computer,
any local algorithm may be executed to decide the outcome. Just such an approach has been reported as successful
by the use of Malus' law to generate non-deterministic outcomes at each separate detector computer.[22] This
approach is consistent with Bell's paper in which he stated that hidden variables may fall into two sets, one set
dependent on one detector and the other set dependent on the other detector, with the "dependencies thereon"
unrestricted.[23]

Final remarks
The phenomenon of quantum entanglement that is behind violation of Bell's inequality is just one element of
quantum physics which cannot be represented by any classical picture of physics; other non-classical elements are
complementarity and wavefunction collapse. The problem of interpretation of quantum mechanics is intended to
provide a satisfactory picture of these non-classical elements of quantum physics.
The EPR paper "pinpointed" the unusual properties of the entangled states, e.g. the above-mentioned singlet state,
which is the foundation for present-day applications of quantum physics, such as quantum cryptography; one
application involves the measurement of quantum entanglement as a physical source of bits for Rabin's oblivious
transfer protocol. This strange non-locality was originally supposed to be a Reductio ad absurdum, because the
standard interpretation could easily do away with action-at-a-distance by simply assigning to each particle definite
spin-states. Bell's theorem showed that the "entangledness" prediction of quantum mechanics have a degree of
non-locality that cannot be explained away by any local theory.
In well-defined Bell experiments (see the paragraph on "test experiments") one can now falsify either quantum
mechanics or Einstein's quasi-classical assumptions: currently many experiments of this kind have been performed,
and the experimental results support quantum mechanics, though some believe that detectors give a biased sample of
photons, so that until nearly every photon pair generated is observed there will be loopholes.
What is powerful about Bell's theorem is that it doesn't come from any particular physical theory. What makes Bell's
theorem unique and powerful is that it relies only on the general properties of quantum mechanics. Physical theories

97

Bell's Theorem
which assume a deterministic variable inside the particle can account for the experimental results only by assuming
that this variable cannot causally change other variables far away.

Notes
[1] Article on Bell's Theorem (http:/ / plato. stanford. edu/ entries/ bell-theorem) by Abner Shimony in the Stanford Encyclopedia of Philosophy,
(2004).
[2] Griffiths, David J. Introduction to Quantum Mechanics: Second Edition. Pearson Prentice Hall, 1998. p. 423.
[3] Merzbacher, Eugene Quantum Mechanics: Third Edition. John Wiley & Sons Inc., 2005. p. 18, 362.
[4] Caroline H. Thompson The Chaotic Ball: An Intuitive Analogy for EPR Experiments Found.Phys.Lett. 9 (1996) 357-382
arXiv:quant-ph/9611037
[5] Kumar, M., Quantum, Icon Books, 2009, p. 350.
[6] J. S. Bell, On the Einstein Podolsky Rosen Paradox (http:/ / www. drchinese. com/ David/ Bell_Compact. pdf), Physics 1, 195-200 (1964)
[7] Stapp, 1975
[8] A. Einstein, B. Podolsky and N. Rosen, Can quantum-mechanical description of physical reality be considered complete? Phys. Rev. 47,
777--780 (1935).
[9] J.F. Clauser and M.A. Horne, Experimental consequences of objective local theories, Phys. Rev. D 10, 526-535 (1974).
[10] P.H. Eberhard, Bell's theorem without hidden variables, Nuovo Cimento 38B, 75-80 (1977).
[11] J. F. Clauser, M. A. Horne, A. Shimony and R. A. Holt, Proposed experiment to test local hidden-variable theories, Physical Review Letters
23, 880884 (1969)
[12] J. F. Clauser and M. A. Horne, Experimental consequences of objective local theories, Physical Review D, 10, 52635 (1974)
[13] M. Redhead, Incompleteness, Nonlocality and Realism, Clarendon Press (1987)
[14] A. Einstein in Correspondance EinsteinBesso, p.265 (Herman, Paris, 1979)
[15] Marshall and Santos, Semiclassical optics as an alternative to nonlocality (http:/ / www. crisisinphysics. co. uk/ optrev. pdf) Recent Research
Developments in Optics 2:683-717 (2002)
[16] S. J. Freedman and J. F. Clauser, Experimental test of local hidden-variable theories, Phys. Rev. Lett. 28, 938 (1972)
[17] Brida et al. Experimental tests of hidden variable theories from dBB to Stochastic Electrodynamics ournal of Physics: Conference Series 67
(2007) 012047, arXiv:quant-ph/0612075
[18] Gammaitoni et al., Stochastic resonance (http:/ / prola. aps. org/ abstract/ RMP/ v70/ i1/ p223_1) Rev. Mod. Phys. 70, 223 - 287 (1998)
[19] S. Grblacher et al., An experimental test of non-local realism (http:/ / www. nature. com/ nature/ journal/ v446/ n7138/ abs/ nature05677.
html) Nature 446, 871875, 2007
[20] Cramer, John G. "The Transactional Interpretation of Quantum Mechanics", Reviews of Modern Physics 58, 647688, July 1986
[21] G 't Hooft, Entangled quantum states in a local deterministic theory (http:/ / arxiv. org/ abs/ 0908. 3408); The Free-Will Postulate in
Quantum Mechanics (http:/ / arxiv. org/ abs/ quant-ph/ 0701097)
[22] Kracklauer, A.F. Non-locality:what are the odds?, (http:/ / arxiv. org/ abs/ quant-ph/ 0302113), Mar. 2004
[23] Bell, John. On the Einstein Podolsky Rosen Paradox, Physics 1 3, 195-200, Nov. 1964

References
A. Aspect et al., Experimental Tests of Realistic Local Theories via Bell's Theorem, Phys. Rev. Lett. 47, 460
(1981)
A. Aspect et al., Experimental Realization of Einstein-Podolsky-Rosen-Bohm Gedankenexperiment: A New
Violation of Bell's Inequalities, Phys. Rev. Lett. 49, 91 (1982).
A. Aspect et al., Experimental Test of Bell's Inequalities Using Time-Varying Analyzers, Phys. Rev. Lett. 49, 1804
(1982).
A. Aspect and P. Grangier, About resonant scattering and other hypothetical effects in the Orsay atomic-cascade
experiment tests of Bell inequalities: a discussion and some new experimental data, Lettere al Nuovo Cimento 43,
345 (1985)
B. D'Espagnat, The Quantum Theory and Reality (http://www.sciam.com/media/pdf/197911_0158.pdf),
Scientific American, 241, 158 (1979)
J. S. Bell, On the problem of hidden variables in quantum mechanics, Rev. Mod. Phys. 38, 447 (1966)
J. S. Bell, On the Einstein Podolsky Rosen Paradox, Physics 1, 3, 195-200 (1964)
J. S. Bell, Introduction to the hidden variable question, Proceedings of the International School of Physics 'Enrico
Fermi', Course IL, Foundations of Quantum Mechanics (1971) 17181

98

Bell's Theorem
J. S. Bell, Bertlmanns socks and the nature of reality, Journal de Physique, Colloque C2, suppl. au numero 3,
Tome 42 (1981) pp C2 4161
J. S. Bell, Speakable and Unspeakable in Quantum Mechanics (Cambridge University Press 1987) [A collection
of Bell's papers, including all of the above.]
J. F. Clauser and A. Shimony, Bell's theorem: experimental tests and implications, Reports on Progress in Physics
41, 1881 (1978)
J. F. Clauser and M. A. Horne, Phys. Rev D 10, 526535 (1974)
E. S. Fry, T. Walther and S. Li, Proposal for a loophole-free test of the Bell inequalities, Phys. Rev. A 52, 4381
(1995)
E. S. Fry, and T. Walther, Atom based tests of the Bell Inequalities the legacy of John Bell continues, pp
103117 of Quantum [Un]speakables, R.A. Bertlmann and A. Zeilinger (eds.) (Springer, Berlin-Heidelberg-New
York, 2002)
R. B. Griffiths, Consistent Quantum Theory', Cambridge University Press (2002).
L. Hardy, Nonlocality for 2 particles without inequalities for almost all entangled states. Physical Review Letters
71 (11) 16651668 (1993)
M. A. Nielsen and I. L. Chuang, Quantum Computation and Quantum Information, Cambridge University Press
(2000)

P. Pearle, Hidden-Variable Example Based upon Data Rejection, Physical Review D 2, 141825 (1970)
A. Peres, Quantum Theory: Concepts and Methods, Kluwer, Dordrecht, 1993.
P. Pluch, Theory of Quantum Probability, PhD Thesis, University of Klagenfurt, 2006.
B. C. van Frassen, Quantum Mechanics, Clarendon Press, 1991.
M.A. Rowe, D. Kielpinski, V. Meyer, C.A. Sackett, W.M. Itano, C. Monroe, and D.J. Wineland, Experimental
violation of Bell's inequalities with efficient detection,(Nature, 409, 791794, 2001).
S. Sulcs, The Nature of Light and Twentieth Century Experimental Physics, Foundations of Science 8, 365391
(2003)
S. Grblacher et al., An experimental test of non-local realism,(Nature, 446, 871875, 2007).
D. N. Matsukevich, P. Maunz, D. L. Moehring, S. Olmschenk, and C. Monroe, Bell Inequality Violation with Two
Remote Atomic Qubits, Phys. Rev. Lett. 100, 150404 (2008).
The comic Dilbert, by Scott Adams, refers to Bell's Theorem in the 1992-09-21 (http://www.dilbert.com/strips/
comic/1992-09-21/) and 1992-09-22 (http://www.dilbert.com/strips/comic/1992-09-22/) strips.

Further reading
The following are intended for general audiences.

Amir D. Aczel, Entanglement: The greatest mystery in physics (Four Walls Eight Windows, New York, 2001).
A. Afriat and F. Selleri, The Einstein, Podolsky and Rosen Paradox (Plenum Press, New York and London, 1999)
J. Baggott, The Meaning of Quantum Theory (Oxford University Press, 1992)
N. David Mermin, "Is the moon there when nobody looks? Reality and the quantum theory", in Physics Today,
April 1985, pp.3847.
Louisa Gilder, The Age of Entanglement: When Quantum Physics Was Reborn (New York: Alfred A. Knopf,
2008)
Brian Greene, The Fabric of the Cosmos (Vintage, 2004, ISBN 0-375-72720-5)
Nick Herbert, Quantum Reality: Beyond the New Physics (Anchor, 1987, ISBN 0-385-23569-0)
D. Wick, The infamous boundary: seven decades of controversy in quantum physics (Birkhauser, Boston 1995)
R. Anton Wilson, Prometheus Rising (New Falcon Publications, 1997, ISBN 1-56184-056-4)

Gary Zukav "The Dancing Wu Li Masters" (Perennial Classics, 2001, ISBN 0-06-095968-1)

99

Bell's Theorem

External links
An explanation of Bell's Theorem (http://www.ncsu.edu/felder-public/kenny/papers/bell.html), based on N.
D. Mermin's article, " Bringing Home the Atomic World: Quantum Mysteries for Anybody (http://dx.doi.org/
10.1119/1.12594)," Am. J. of Phys. 49 (10), 940 (October 1981)
Quantum Entanglement (http://www.ipod.org.uk/reality/reality_entangled.asp) Includes a simple explanation
of Bell's Inequality.
Bell's theorem on arXiv.org (http://xstructure.inr.ac.ru/x-bin/theme3.py?level=2&index1=369244)
Disproofs of Bell, GHZ, and Hardy Type Theorems and the Illusion of Entanglement (http://arxiv.org/abs/
0904.4259) Disproof of Bell's Theorem
Interactive experiments with single photons: entanglement and Bells theorem (http://www.didaktik.physik.
uni-erlangen.de/quantumlab/english/index.html)

100

101

5. Schrdinger's Objections
Schrdinger's Cat
Schrdinger's cat is a thought
experiment, usually described as a
paradox, devised by Austrian physicist
Erwin Schrdinger in 1935. It illustrates
what he saw as the problem of the
Copenhagen interpretation of quantum
mechanics applied to everyday objects.
The thought experiment presents a cat
that might be alive or dead, depending
on an earlier random event. In the
course of developing this experiment,
he coined the term Verschrnkung
(entanglement).

Origin and motivation

Schrdinger's Cat: A cat, along with a flask containing a poison and a radioactive source,
is placed in a sealed box shielded against environmentally induced quantum
decoherence. If an internal Geiger counter detects radiation, the flask is shattered,
releasing the poison that kills the cat. The Copenhagen interpretation of quantum
mechanics implies that after a while, the cat is simultaneously alive and dead. Yet, when
we look in the box, we see the cat either alive or dead, not both alive and dead.

Schrdinger's thought experiment was


intended as a discussion of the EPR article, named after its authorsEinstein, Podolsky, and Rosenin 1935.[1] The
EPR article had highlighted the strange nature of quantum entanglement, which is a characteristic of a quantum state
that is a combination of the states of two systems (for example, two subatomic particles), that once interacted but
were then separated and are not each in a definite state. The Copenhagen interpretation implies that the state of the
two systems undergoes collapse into a definite state when one of the systems is measured.
Schrdinger and Einstein had exchanged letters about Einstein's EPR article, in the course of which Einstein had
pointed out that the state of an unstable keg of gunpowder will, after a while, contain a superposition of both
exploded and unexploded states.
To further illustrate the putative incompleteness of quantum mechanics, Schrdinger describes how one could, in
principle, transpose the superposition of an atom to large-scale systems of a live and dead cat by coupling cat and
atom with the help of a "diabolical mechanism". He proposed a scenario with a cat in a sealed box, wherein the cat's
life or death was dependent on the state of a subatomic particle. According to Schrdinger, the Copenhagen
interpretation implies that the cat remains both alive and dead (to the universe outside the box) until the box is
opened.
Schrdinger did not wish to promote the idea of dead-and-alive cats as a serious possibility; quite the reverse, the
paradox is a classic reductio ad absurdum.[2] The thought experiment serves to illustrate the bizarreness of quantum
mechanics and the mathematics necessary to describe quantum states. Intended as a critique of just the Copenhagen
interpretation (the prevailing orthodoxy in 1935), the Schrdinger cat thought experiment remains a topical
touchstone for all interpretations of quantum mechanics. How each interpretation deals with Schrdinger's cat is
often used as a way of illustrating and comparing each interpretation's particular features, strengths, and weaknesses.

Schrdinger's Cat

The thought experiment


Schrdinger wrote:
One can even set up quite ridiculous cases. A cat is penned up in a steel chamber, along with the following
device (which must be secured against direct interference by the cat): in a Geiger counter, there is a tiny bit of
radioactive substance, so small that perhaps in the course of the hour, one of the atoms decays, but also, with
equal probability, perhaps none; if it happens, the counter tube discharges, and through a relay releases a
hammer that shatters a small flask of hydrocyanic acid. If one has left this entire system to itself for an hour,
one would say that the cat still lives if meanwhile no atom has decayed. The psi-function of the entire system
would express this by having in it the living and dead cat (pardon the expression) mixed or smeared out in
equal parts.
It is typical of these cases that an indeterminacy originally restricted to the atomic domain becomes
transformed into macroscopic indeterminacy, which can then be resolved by direct observation. That prevents
us from so naively accepting as valid a "blurred model" for representing reality. In itself, it would not embody
anything unclear or contradictory. There is a difference between a shaky or out-of-focus photograph and a
snapshot of clouds and fog banks.[3]
The above text is a translation of two paragraphs from a much larger original article that appeared in the German
magazine Naturwissenschaften ("Natural Sciences") in 1935.[2]
Schrdinger's famous thought experiment poses the question, when does a quantum system stop existing as a
superposition of states and become one or the other? (More technically, when does the actual quantum state stop
being a linear combination of states, each of which resembles different classical states, and instead begins to have a
unique classical description?) If the cat survives, it remembers only being alive. But explanations of the EPR
experiments that are consistent with standard microscopic quantum mechanics require that macroscopic objects, such
as cats and notebooks, do not always have unique classical descriptions. The purpose of the thought experiment is to
illustrate this apparent paradox. Our intuition says that no observer can be in a mixture of states; yet the cat, it seems
from the thought experiment, can be such a mixture. Is the cat required to be an observer, or does its existence in a
single well-defined classical state require another external observer? Each alternative seemed absurd to Albert
Einstein, who was impressed by the ability of the thought experiment to highlight these issues. In a letter to
Schrdinger dated 1950, he wrote:
You are the only contemporary physicist, besides Laue, who sees that one cannot get around the assumption of
reality, if only one is honest. Most of them simply do not see what sort of risky game they are playing with
realityreality as something independent of what is experimentally established. Their interpretation is,
however, refuted most elegantly by your system of radioactive atom + amplifier + charge of gunpowder + cat
in a box, in which the psi-function of the system contains both the cat alive and blown to bits. Nobody really
doubts that the presence or absence of the cat is something independent of the act of observation.[4]
Note that no charge of gunpowder is mentioned in Schrdinger's setup, which uses a Geiger counter as an amplifier
and hydrocyanic poison instead of gunpowder. The gunpowder had been mentioned in Einstein's original suggestion
to Schrdinger 15 years before, and apparently Einstein had carried it forward to the present discussion.

102

Schrdinger's Cat

103

Interpretations of the experiment


Since Schrdinger's time, other interpretations of quantum mechanics have been proposed that give different answers
to the questions posed by Schrodinger's cat of how long superpositions last and when (or if) they collapse.

Copenhagen interpretation
In the Copenhagen interpretation of quantum mechanics, a system stops being a superposition of states and becomes
either one or the other when an observation takes place. This experiment makes apparent the fact that the nature of
measurement, or observation, is not well-defined in this interpretation. The experiment can be interpreted to mean
that while the box is closed, the system simultaneously exists in a superposition of the states "decayed nucleus/dead
cat" and "undecayed nucleus/living cat", and that only when the box is opened and an observation performed does
the wave function collapse into one of the two states.
However, one of the main scientists associated with the Copenhagen interpretation, Niels Bohr, never had in mind
the observer-induced collapse of the wave function, so that Schrdinger's Cat did not pose any riddle to him. The cat
would be either dead or alive long before the box is opened by a conscious observer.[5] Analysis of an actual
experiment found that measurement alone (for example by a Geiger counter) is sufficient to collapse a quantum
wave function before there is any conscious observation of the measurement.[6] The view that the "observation" is
taken when a particle from the nucleus hits the detector can be developed into objective collapse theories. In
contrast, the many worlds approach denies that collapse ever occurs.

Many-worlds interpretation and consistent histories


In 1957, Hugh Everett formulated the
many-worlds interpretation of quantum
mechanics, which does not single out
observation as a special process. In the
many-worlds interpretation, both alive
and dead states of the cat persist after
the box is opened, but are decoherent
from each other. In other words, when
the box is opened, the observer and the
already-split cat split into an observer
looking at a box with a dead cat, and
an observer looking at a box with a
live cat. But since the dead and alive
states are decoherent, there is no
effective communication or interaction
between them.

The quantum-mechanical "Schrdinger's cat" paradox according to the many-worlds


interpretation. In this interpretation, every event is a branch point; the cat is both alive and
dead, irrespective of whether the box is opened, but the "alive" and "dead" cats are in
different branches of the universe, both of which are equally real, but which cannot
interact with each other.

When opening the box, the observer becomes entangled with the cat, so "observer states" corresponding to the cat's
being alive and dead are formed; each observer state is entangled or linked with the cat so that the "observation of
the cat's state" and the "cat's state" correspond with each other. Quantum decoherence ensures that the different
outcomes have no interaction with each other. The same mechanism of quantum decoherence is also important for
the interpretation in terms of consistent histories. Only the "dead cat" or "alive cat" can be a part of a consistent
history in this interpretation.
Roger Penrose criticises this:
"I wish to make it clear that, as it stands, this is far from a resolution of the cat paradox. For there is nothing in
the formalism of quantum mechanics that demands that a state of consciousness cannot involve the

Schrdinger's Cat
simultaneous perception of a live and a dead cat",[7]
Although the mainstream view (without necessarily endorsing many-worlds) is that decoherence is the mechanism
that forbids such simultaneous perception.[8] [9]
A variant of the Schrdinger's Cat experiment, known as the quantum suicide machine, has been proposed by
cosmologist Max Tegmark. It examines the Schrdinger's Cat experiment from the point of view of the cat, and
argues that by using this approach, one may be able to distinguish between the Copenhagen interpretation and
many-worlds.

Ensemble interpretation
The ensemble interpretation states that superpositions are nothing but subensembles of a larger statistical ensemble.
The state vector would not apply to individual cat experiments, but only to the statistics of many similarly prepared
cat experiments. Proponents of this interpretation state that this makes the Schrdinger's Cat paradox a trivial
non-issue.
This interpretation serves to discard the idea that a single physical system in quantum mechanics has a mathematical
description that corresponds to it in any way.

Relational interpretation
The relational interpretation makes no fundamental distinction between the human experimenter, the cat, or the
apparatus, or between animate and inanimate systems; all are quantum systems governed by the same rules of
wavefunction evolution, and all may be considered "observers." But the relational interpretation allows that different
observers can give different accounts of the same series of events, depending on the information they have about the
system.[10] The cat can be considered an observer of the apparatus; meanwhile, the experimenter can be considered
another observer of the system in the box (the cat plus the apparatus). Before the box is opened, the cat, by nature of
it being alive or dead, has information about the state of the apparatus (the atom has either decayed or not decayed);
but the experimenter does not have information about the state of the box contents. In this way, the two observers
simultaneously have different accounts of the situation: To the cat, the wavefunction of the apparatus has appeared to
"collapse"; to the experimenter, the contents of the box appear to be in superposition. Not until the box is opened,
and both observers have the same information about what happened, do both system states appear to "collapse" into
the same definite result, a cat that is either alive or dead.

Objective collapse theories


According to objective collapse theories, superpositions are destroyed spontaneously (irrespective of external
observation) when some objective physical threshold (of time, mass, temperature, irreversibility, etc.) is reached.
Thus, the cat would be expected to have settled into a definite state long before the box is opened. This could loosely
be phrased as "the cat observes itself", or "the environment observes the cat".
Objective collapse theories require a modification of standard quantum mechanics to allow superpositions to be
destroyed by the process of time evolution.

104

Schrdinger's Cat

Practical applications
The experiment is a purely theoretical one, and the machine proposed is not known to have been constructed. In
quantum computing, however, the phrase "cat state" often refers to the special entanglement of qubits wherein the
qubits are in an equal superposition of all being 0 and all being 1; i.e.,
+
.

Extensions
A variant on the experiment is Wigner's friend, in which there are two external observers, the first of whom opens
and inspects the box and then communicates his observations to a second observer. The issue here is, does the wave
function "collapse" when the first observer opens the box, or only when the second observer is informed of the first
observer's observations?
In another extension, prominent physicists have gone so far as to suggest that astronomers observing dark energy in
the universe in 1998 may have "reduced its life expectancy" through a pseudo-Schrdinger's Cat scenario, although
this is a controversial viewpoint.[11] [12]

References
[1] EPR article: Can Quantum-Mechanical Description Reality Be Considered Complete? (http:/ / prola. aps. org/ abstract/ PR/ v47/ i10/ p777_1)
[2] Schrdinger, Erwin (November 1935). "Die gegenwrtige Situation in der Quantenmechanik (The present situation in quantum mechanics)".
Naturwissenschaften.
[3] Schroedinger: "The Present Situation in Quantum Mechanics" (http:/ / www. tu-harburg. de/ rzt/ rzt/ it/ QM/ cat. html#sect5)
[4] Pay link to Einstein letter (http:/ / www. jstor. org/ pss/ 687649)
[5] Faye, J (2008-01-24). "Copenhagen Interpretation of Quantum Mechanics" (http:/ / plato. stanford. edu/ entries/ qm-copenhagen/ ). Stanford
Encyclopedia of Philosophy. The Metaphysics Research Lab Center for the Study of Language and Information, Stanford University. .
Retrieved 2010-09-19.
[6] Carpenter RHS, Anderson AJ (2006). "The death of Schroedinger's Cat and of consiousness-based wave-function collapse" (http:/ / web.
archive. org/ web/ 20061130173850/ http:/ / www. ensmp. fr/ aflb/ AFLB-311/ aflb311m387. pdf). Annales de la Fondation Louis de Broglie
(http:/ / web. archive. org/ web/ 20080618174026/ http:/ / www. ensmp. fr/ aflb/ AFLB-Web/ en-annales-index. htm) 31 (1): 4552. Archived
from the original (http:/ / www. ensmp. fr/ aflb/ AFLB-311/ aflb311m387. pdf) on 2006-11-30. . Retrieved 2010-09-10.
[7] Penrose, R. The Road to Reality, p 807.
[8] Wojciech H. Zurek, Decoherence, einselection, and the quantum origins of the classical, Reviews of Modern Physics 2003, 75, 715 or (http:/ /
arxiv. org/ abs/ quant-ph/ 0105127)
[9] Wojciech H. Zurek, "Decoherence and the transition from quantum to classical", Physics Today, 44, pp 3644 (1991)
[10] Rovelli, Carlo (1996). "Relational Quantum Mechanics". International Journal of Theoretical Physics 35: 16371678.
arXiv:quant-ph/9609002. doi:10.1007/BF02302261.
[11] Chown, Marcus (2007-11-22). "Has observing the universe hastened its end?" (http:/ / www. newscientist. com/ channel/ fundamentals/
mg19626313. 800-has-observing-the-universe-hastened-its-end. html). New Scientist. . Retrieved 2007-11-25.
[12] Krauss, Lawrence M.; James Dent (April 30, 2008). "Late Time Behavior of False Vacuum Decay: Possible Implications for Cosmology
and Metastable Inflating States". Phys. Rev. Lett. (USA: APS) 100 (17). arXiv:0711.1821.

External links
Erwin Schrdinger, The Present Situation in Quantum Mechanics (Translation) (http://www.tu-harburg.de/rzt/
rzt/it/QM/cat.html)
The EPR paper (http://prola.aps.org/abstract/PR/v47/i10/p777_1)
Viennese Meow (the cat's perspective - short story) (http://primastoria.com/story/viennese-meow/)
The story of Schroedinger's cat (an epic poem) (http://www.straightdope.com/classics/a1_122.html); The
Straight Dope
Tom Leggett (Aug. 1, 2000) New life for Schrdinger's cat, Physics World, UK (http://physicsworld.com/cws/
article/print/525) Experiments at two universities claim to observe superposition in large scale systems
Information Philosopher on Schrdinger's cat (http://www.informationphilosopher.com/solutions/
experiments/schrodingerscat/) More diagrams and an information creation explanation.

105

Schrdinger's Cat
A YouTube video explaining Schrdingers cat (http://www.youtube.com/watch?v=CrxqTtiWxs4)

106

107

6. Measurement Problems
The Measurement Problem
The measurement problem in quantum mechanics is the unresolved problem of how (or if) wavefunction collapse
occurs. The inability to observe this process directly has given rise to different interpretations of quantum mechanics,
and poses a key set of questions that each interpretation must answer. The wavefunction in quantum mechanics
evolves according to the Schrdinger equation into a linear superposition of different states, but actual measurements
always find the physical system in a definite state. Any future evolution is based on the state the system was
discovered to be in when the measurement was made, meaning that the measurement "did something" to the process
under examination. Whatever that "something" may be does not appear to be explained by the basic theory.
To express matters differently (to paraphrase Steven Weinberg [1] [2] ), the wave function evolves deterministically
knowing the wave function at one moment, the Schrdinger equation determines the wave function at any later time.
If observers and their measuring apparatus are themselves described by a deterministic wave function, why can we
not predict precise results for measurements, but only probabilities? As a general question: How can one establish a
correspondence between quantum and classical reality?[3]

Example
The best known is the "paradox" of the Schrdinger's cat: a cat is apparently evolving into a linear superposition of
basis vectors that can be characterized as an "alive cat" and states that can be described as a "dead cat". Each of these
possibilities is associated with a specific nonzero probability amplitude; the cat seems to be in a "mixed" state.
However, a single, particular observation of the cat does not measure the probabilities: it always finds either a living
cat, or a dead cat. After the measurement the cat is definitively alive or dead. The question is: How are the
probabilities converted into an actual, sharply well-defined outcome?

Interpretations
Some interpretations claim that the latter approach was put on firm ground in the 1980s by the phenomenon of
quantum decoherence.[4] It is claimed that decoherence makes it possible to identify the fuzzy boundary between the
quantum microworld and the world where the classical intuition is applicable.[5] Quantum decoherence was proposed
in the context of the many-worlds interpretation, but it has also become an important part of some modern updates of
the Copenhagen interpretation based on consistent histories ("Copenhagen done right"). Quantum decoherence does
not describe the actual process of the wavefunction collapse, but it explains the conversion of the quantum
probabilities (that exhibit interference effects) to the ordinary classical probabilities. See, for example, Zurek,[3]
Zeh[5] and Schlosshauer.[6]
Hugh Everett's many-worlds interpretation attempts to avoid the problem by suggesting there is only one
wavefunction, the superposition of the entire universe, and it never collapsesso there is no measurement problem.
Instead the act of measurement is actually an interaction between two quantum entities, which entangle to form a
single larger entity, for instance living cat/happy scientist. Everett also attempted to demonstrate the way that in
measurements the probabilistic nature of quantum mechanics would appear; work later extended by Bryce DeWitt.
De BroglieBohm theory tries to solve the measurement problem very differently: this interpretation contains not
only the wavefunction, but also the information about the position of the particle(s). The role of the wavefunction is
to generate the velocity field for the particles. These velocities are such that the probability distribution for the

The Measurement Problem


particle remains consistent with the predictions of the orthodox quantum mechanics. According to de BroglieBohm
theory, interaction with the environment during a measurement procedure separates the wave packets in
configuration space which is where apparent wavefunction collapse comes from even though there is no actual
collapse. Decoherence analysis is one way to view this.
The present situation is slowly clarifying, as described in a recent paper by Schlosshauer as follows:[7]
Several decoherence-unrelated proposals have been put forward in the past to elucidate the meaning of
probabilities and arrive at the Born rule It is fair to say that no decisive conclusion appears to have been
reached as to the success of these derivations.
As it is well known, [many papers by Bohr insist upon] the fundamental role of classical concepts. The
experimental evidence for superpositions of macroscopically distinct states on increasingly large length scales
counters such a dictum. Only the physical interactions between systems then determine a particular
decomposition into classical states from the view of each particular system. Thus classical concepts are to be
understood as locally emergent in a relative-state sense and should no longer claim a fundamental role in the
physical theory.

References and notes


[1] Steven Weinberg (1998). The Oxford History of the Twentieth Century (http:/ / books. google. com/ ?id=uYTW5ZWrwWAC& pg=PA22&
dq=observer+ measurement+ "S+ Weinberg") (Michael Howard & William Roger Louis, editors ed.). Oxford University Press. p.26.
ISBN0198204280. .
[2] Steven Weinberg: Einstein's Mistakes (http:/ / scitation. aip. org/ journals/ doc/ PHTOAD-ft/ vol_58/ iss_11/ 31_1. shtml) in Physics Today
(2005); see subsection "Contra quantum mechanics"
[3] Wojciech Hubert Zurek Decoherence, einselection, and the quantum origins of the classical Reviews of Modern Physics, Vol. 75, July 2003
(http:/ / hubcap. clemson. edu/ ~daw/ D_PHYS455/ RevModPhys. v75p715y03. pdf)
[4] Joos, E., and H. D. Zeh, "The emergence of classical properties through interaction with the environment" (1985), Z. Phys. B 59, 223.
[5] H D Zeh (http:/ / arxiv. org/ abs/ quant-ph/ 9506020v3) in E. Joos .... (2003). Decoherence and the Appearance of a Classical World in
Quantum Theory (http:/ / books. google. com/ ?id=6eTHcxeNxdUC& printsec=frontcover& dq=isbn=3540613943#PPT21,M1) (2nd Edition;
Erich Joos, H. D. Zeh, C. Kiefer, Domenico Giulini, J. Kupsch, I. O. Stamatescu (editors) ed.). Springer-Verlag. Chapter 2.
ISBN3540003908. .
[6] Maximilian Schlosshauer (2005). "Decoherence, the measurement problem, and interpretations of quantum mechanics". Rev. Mod. Phys. 76:
12671305. doi:10.1103/RevModPhys.76.1267. arXiv:quant-ph/0312059v4.
[7] M Schlosshauer: Experimental motivation and empirical consistency in minimal no-collapse quantum mechanics, Annals of Physics, Volume
321, Issue 1, January 2006, Pages 112-149 (http:/ / www. citebase. org/ fulltext?format=application/ pdf& identifier=oai:arXiv. org:quant-ph/
0506199)

Further reading
R. Buniy, S. Hsu and A. Zee On the origin of probability in quantum mechanics (2006) (http://duende.uoregon.
edu/~hsu/talks/probability_qm.pdf)

External links
The Quantum Measurement Problem (http://www.shantena.com/en/physicslectures/quantummeasurement)
Two presentations: a non-technical and a more technical presentation.

108

Measurement in Quantum Mechanics

Measurement in Quantum Mechanics


The framework of quantum mechanics requires a careful definition of measurement. The issue of measurement lies
at the heart of the problem of the interpretation of quantum mechanics, for which there is currently no consensus.

Measurement from a practical point of view


Measurement is viewed in different ways in the many interpretations of quantum mechanics; however, despite the
considerable philosophical differences, they almost universally agree on the practical question of what results from a
routine quantum-physics laboratory measurement. To describe this, a simple framework to use is the Copenhagen
interpretation, and it will be implicitly used in this section; the utility of this approach has been verified countless
times, and all other interpretations are necessarily constructed so as to give the same quantitative predictions as this
in almost every case.

Qualitative overview
The quantum state of a system is a mathematical object that fully describes the quantum system. One typically
imagines some experimental apparatus and procedure which "prepares" this quantum state; the mathematical object
then reflects the setup of the apparatus. Once the quantum state has been prepared, some aspect of it is measured (for
example, its position or energy). If the experiment is repeated, so as to measure the same aspect of the same quantum
state prepared in the same way, the result of the measurement will often be different.
The expected result of the measurement is in general described by a probability distribution that specifies the
likelihoods that the various possible results will be obtained. (This distribution can be either discrete or continuous,
depending on what is being measured.)
The measurement process is often said to be random and indeterministic. (However, there is considerable dispute
over this issue; in some interpretations of quantum mechanics, the result merely appears random and indeterministic,
in other interpretations the indeterminism is core and irreducible.) This is because an important aspect of
measurement is wavefunction collapse, the nature of which varies according to the interpretation adopted.
What is universally agreed, however, is that if the measurement is repeated, without re-preparing the state, one finds
the same result as the first measurement. As a result, after measuring some aspect of the quantum state, we normally
update the quantum state to reflect the result of the measurement; it is this updating that ensures that if an immediate
re-measurement is repeated without re-preparing the state, one finds the same result as the first measurement. The
updating of the quantum state model is called wavefunction collapse.

Quantitative details
The mathematical relationship between the quantum state and the probability distribution is, again, widely accepted
among physicists, and has been experimentally confirmed countless times. This section summarizes this relationship,
which is stated in terms of the mathematical formulation of quantum mechanics.
Measurable quantities ("observables") as operators
It is a postulate of quantum mechanics that all measurements have an associated operator (called an observable
operator, or just an observable), with the following properties:
1. The observable is a Hermitian (self-adjoint) operator mapping a Hilbert space (namely, the state space, which
consists of all possible quantum states) into itself.
2. The observable's eigenvalues are real. The possible outcomes of the measurement are precisely the eigenvalues of
the given observable.

109

Measurement in Quantum Mechanics

110

3. For each eigenvalue there are one or more corresponding eigenvectors (which in this context are called
eigenstates), which will make up the state of the system after the measurement.
4. The observable has a set of eigenvectors which span the state space. It follows that each observable generates an
orthonormal basis of eigenvectors (called an eigenbasis). Physically, this is the statement that any quantum state
can always be represented as a superposition of the eigenstates of an observable.
Important examples of observables are:
The Hamiltonian operator, representing the total energy of the system; with the special case of the nonrelativistic
Hamiltonian operator:

The momentum operator:


The position operator:

(in the position basis).

, where

(in the momentum basis).

Operators can be noncommuting. Two Hermitian operators commute if (and only if) there is at least one basis of
vectors, each of which is an eigenvector of both operators (this is sometimes called a simultaneous eigenbasis).
Noncommuting observables are said to be incompatible and cannot in general be measured simultaneously. In fact,
they are related by an uncertainty principle, as a consequence of the Robertson-Schrdinger relation.
Measurement probabilities and wavefunction collapse
There are a few possible ways to mathematically describe the measurement process (both the probability distribution
and the collapsed wavefunction). The most convenient description depends on the spectrum (i.e., set of eigenvalues)
of the observable.
Discrete, nondegenerate spectrum
Let

be an observable, and suppose that it has discrete eigenstates

and corresponding eigenvalues

(in bra-ket notation) for

, no two of which are equal.

Assume the system is prepared in state

. Since the eigenstates of an observable form a basis (the eigenbasis), it

follows that

can be written in terms of the eigenstates as

(where

are complex numbers). Then measuring

can yield any of the results

, with

corresponding probabilities given by

Usually

is assumed to be normalized, in which case this expression reduces to

If the result of the measurement is

, then the system's quantum state after the measurement is

so any repeated measurement of

will yield the same result

collapse.)

. (This phenomenon is called wavefunction

Measurement in Quantum Mechanics

111

Continuous, nondegenerate spectrum


Let

be an observable, and suppose that it has a continuous spectrum of eigenvalues filling the interval (a,b).

Assume further that each eigenvalue x in this range is associated with a unique eigenstate
Assume the system is prepared in state

(where

, which can be written in terms of the eigenbasis as

is a complex-valued function). Then measuring

with probability density function

Again,

can yield a result anywhere in the interval (a,b),

; i.e., a result between y and z will occur with probability

is often assumed to be normalized, in which case this expression reduces to

If the result of the measurement is x, then the new wave function will be

Alternatively, it is often possible and convenient to analyze a continuous-spectrum measurement by taking it to be


the limit of a different measurement with a discrete spectrum. For example, an analysis of scattering involves a
continuous spectrum of energies, but by adding a "box" potential (which bounds the volume in which the particle can
be found), the spectrum becomes discrete. By considering larger and larger boxes, this approach need not involve
any approximation, but rather can be regarded as an equally valid formalism in which this problem can be analyzed.
Degenerate spectra
If there are multiple eigenstates with the same eigenvalue (called degeneracies), the analysis is a bit less simple to
state, but not essentially different. In the discrete case, for example, instead of finding a complete eigenbasis, it is a
bit more convenient to write the Hilbert space as a direct sum of eigenspaces. The probability of measuring a
particular eigenvalue is the squared component of the state vector in the corresponding eigenspace, and the new state
after measurement is the projection of the original state vector into the appropriate eigenspace.
Density matrix formulation
Instead of performing quantum-mechanics computations in terms of wavefunctions (kets), it is sometimes necessary
to describe a quantum-mechanical system in terms of a density matrix. The analysis in this case is formally slightly
different, but the physical content is the same, and indeed this case can be derived from the wavefunction
formulation above. The result for the discrete, degenerate case, for example, is as follows:
Let

be an observable, and suppose that it has discrete eigenvalues


respectively. Let

be the projection operator into the space

, associated with eigenspaces


.

Assume the system is prepared in the state described by the density matrix . Then measuring
the results

, with corresponding probabilities given by

where Tr denotes trace. If the result of the measurement is n, then the new density matrix will be

Alternatively, one can say that the measurement process results in the new density matrix

can yield any of

Measurement in Quantum Mechanics

112

where the difference is that ' ' is the density matrix describing the entire ensemble, whereas ' is the density matrix
describing the sub-ensemble whose measurement result was n.
Statistics of measurement
As detailed above, the result of measuring a quantum-mechanical system is described by a probability distribution.
Some properties of this distribution are as follows:
Suppose we take a measurement corresponding to observable

, on a state whose quantum state is

The mean (average) value of the measurement is (see Expectation value (quantum mechanics))
.
The variance of the measurement is

The standard deviation of the measurement is

These are direct consequences of the above formulas for measurement probabilities.
Example
Suppose that we have a particle in a 1-dimensional box, set up initially in the ground state
computed from the time-independent Schrdinger equation, the energy of this state is
particle's mass and L is the box length), and the spatial wavefunction is

. As can be
(where m is the
. If the energy

is now measured, the result will always certainly be


, and this measurement will not affect the wavefunction.
Next suppose that the particle's position is measured. The position x will be measured with probability density

If the measurement result was x=S, then the wavefunction after measurement will be the position eigenstate
. If the particle's position is immediately measured again, the same position will be obtained.
The new wavefunction

can, like any wavefunction, be written as a superposition of eigenstates of any

observable. In particular, using energy eigenstates,

, we have

If we now leave this state alone, it will smoothly evolve in time according to the Schrdinger equation. But suppose
instead that an energy measurement is immediately taken. Then the possible energy values
will be measured
with relative probabilities:

and moreover if the measurement result is

, then the new state will be the energy eigenstate

So in this example, due to the process of wavefunction collapse, a particle initially in the ground state can end up in
any energy level, after just two subsequent non-commuting measurements are made.

Measurement in Quantum Mechanics

113

Wavefunction collapse
The process in which a quantum state becomes one of the eigenstates of the operator corresponding to the measured
observable is called "collapse", or "wavefunction collapse". The final eigenstate appears randomly with a probability
equal to the square of its overlap with the original state. The process of collapse has been studied in many
experiments, most famously in the double-slit experiment. The wavefunction collapse raises serious questions
regarding "the measurement problem",[1] as well as, questions of determinism and locality, as demonstrated in the
EPR paradox and later in GHZ entanglement. (See below.)
In the last few decades, major advances have been made toward a theoretical understanding of the collapse process.
This new theoretical framework, called quantum decoherence, supersedes previous notions of instantaneous collapse
and provides an explanation for the absence of quantum coherence after measurement. While this theory correctly
predicts the form and probability distribution of the final eigenstates, it does not explain the randomness inherent in
the choice of final state.

von Neumann measurement scheme


The von Neumann measurement scheme, an ancestor of quantum decoherence theory, describes measurements by
taking into account the measuring apparatus which is also treated as a quantum object. Let the quantum state be in
the superposition
, where
are eigenstates of the operator that needs to be measured. In order to
make the measurement, the measured system described by
described by the quantum state

needs to interact with the measuring apparatus

, so that the total wave function before the interaction is

. During the

interaction of object and measuring instrument the unitary evolution is supposed to realize the following transition
from the initial to the final total wave function:

where

are orthonormal states of the measuring apparatus. The unitary evolution above is referred to as

premeasurement. The relation with wave function collapse is established by calculating from the final total wave
function the final density operator of the object as
This density operator is interpreted by von
Neumann as describing an ensemble of objects being after the measurement with probability

in the state

The transition

is often referred to as weak von Neumann projection, the wave function collapse or strong von Neumann projection

being thought to correspond to an additional selection of a subensemble by means of observation.


In case the measured observable has a degenerate spectrum, weak von Neumann projection is generalized to Lders
projection

in which the vectors

for fixed n are the degenerate eigenvectors of the measured observable. For an arbitrary

state described by a density operator Lders projection is given by

Measurement in Quantum Mechanics

114

Measurements of the second kind


In a measurement of the second kind the unitary evolution during the interaction of object and measuring instrument
is supposed to be given by

in which the states

of the object are determined by specific properties of the interaction between object and

measuring instrument. They are normalized but not necessarily mutually orthogonal. The relation with wave function
collapse is analogous to that obtained for measurements of the first kind, the final state of the object now being
with probability

Note that many present-day measurement procedures are measurements of the second kind,

some even functioning correctly only as a consequence of being of the second kind (for instance, a photon counter,
detecting a photon by absorbing and hence annihilating it, thus ideally leaving the electromagnetic field in the
vacuum state rather than in the state corresponding to the number of detected photons; also the Stern-Gerlach
[2]
experiment would not function at all if it really were a measurement of the first kind ). [pdf]

Decoherence in quantum measurement


One can also introduce the interaction with the environment

, so that, in a measurement of the first kind, after the

interaction the total wave function takes a form

which is related to the phenomenon of decoherence.


The above is completely described by the Schrdinger equation and there are not any interpretational problems with
this. Now the problematic wavefunction collapse does not need to be understood as a process
on the level
of the measured system, but can also be understood as a process
or as a process

on the level of the measuring apparatus,

on the level of the environment. Studying these processes provides considerable insight into

the measurement problem by avoiding the arbitrary boundary between the quantum and classical worlds, though it
does not explain the presence of randomness in the choice of final eigenstate. If the set of states
,

, or

represents a set of states that do not overlap in space, the appearance of collapse can be generated by either the Bohm
interpretation or the Everett interpretation which both deny the reality of wavefunction collapse. Both of these are
stated to predict the same probabilities for collapses to various states as the conventional interpretation by their
supporters. The Bohm interpretation is held to be correct only by a small minority of physicists, since there are
difficulties with the generalization for use with relativistic quantum field theory. However, there is no proof that the
Bohm interpretation is inconsistent with quantum field theory, and work to reconcile the two is ongoing. The Everett
interpretation easily accommodates relativistic quantum field theory.

Philosophical problems of quantum measurements


What physical interaction constitutes a measurement?
Until the advent of quantum decoherence theory in the late 20th century, a major conceptual problem of quantum
mechanics and especially the Copenhagen interpretation was the lack of a distinctive criterion for a given physical
interaction to qualify as "a measurement" and cause a wavefunction to collapse. This is best illustrated by the
Schrdinger's cat paradox. Certain aspects of this question are now well understood in the framework of quantum
decoherence theory, such as an understanding of weak measurements, and quantifying what measurements or
interactions are sufficient to destroy quantum coherence. Nevertheless, there remains less than universal agreement
among physicists on some aspects of the question of what constitutes a measurement.

Measurement in Quantum Mechanics


(One particularly well-known aspect of this question is whether a conscious observer is necessary for a
measurement; see the article Consciousness causes collapse.)

Does measurement actually determine the state?


The question of whether (and in what sense) a measurement actually determines the state is one which differs among
the different interpretations of quantum mechanics. (It is also closely related to the understanding of wavefunction
collapse.) For example, in most versions of the Copenhagen interpretation, the measurement determines the state,
and after measurement the state is definitely what was measured. But according to the Many-worlds interpretation,
measurement determines the state in a more restricted sense: In other "worlds", other measurement results were
obtained, and the other possible states still exist.

Is the measurement process random or deterministic?


As described above, there is universal agreement that quantum mechanics appears random, in the sense that all
experimental results yet uncovered can be predicted and understood in the framework of quantum mechanics
measurements being fundamentally random. Nevertheless, it is not settled[3] whether this is true, fundamental
randomness, or merely "emergent" randomness resulting from underlying hidden variables which deterministically
cause measurement results to happen a certain way each time. This continues to be an area of active research.[4]
(If there are hidden variables, they would have to be "nonlocal", see below.)

Does the measurement process violate locality?


In physics, the Principle of locality is the concept that information cannot travel faster than the speed of light (also
see special relativity). It is known experimentally (see Bell's theorem, which is related to the EPR paradox) that if
quantum mechanics is deterministic (due to hidden variables, as described above), then it is nonlocal (i.e. violates
the principle of locality). Nevertheless, there is not universal agreement among physicists on whether quantum
mechanics is nondeterministic, nonlocal, or both.[3]

External links

"The Double Slit Experiment [5]". (physicsweb.org)


"Measurement in Quantum Mechanics [6]" Henry Krips in the Stanford Encyclopedia of Philosophy
Decoherence, the measurement problem, and interpretations of quantum mechanics [7]
Measurements and Decoherence [8]
The conditions for discrimination between quantum states with minimum error [9]
Quantum behavior of measurement apparatus [10]
Yonina C. Eldar, Alexandre Megretski, and George C. Verghese. Designing optimal quantum detectors via
semidefinite programming. IEEE Transactions on Information Theory, Vol. 49, No. 4, 10071012, 2003.

115

Measurement in Quantum Mechanics

Further reading
John A. Wheeler and Wojciech Hubert Zurek (eds), Quantum Theory and Measurement, Princeton University
Press, (1983), ISBN 0-691-08316-9
Vladimir B. Braginsky and Farid Ya. Khalili, Quantum Measurement, Camebridge University Press, (1992),
ISBN 0-521-41928-X
Greenstein, G. and Zajonc, A.G., The Quantum Challenge, Jones and Bartlett Publishers, (2006), ISBN
0-7367-2470-X

References
[1] http:/ / books. google. com/ books?id=5t0tm0FB1CsC& pg=PA215& lpg=PA215& dq=wave+ function+ collapse& source=bl&
ots=a7iUGurRDC& sig=o1ddjY7lQrj4EQdvS49xcceWq2M& hl=en& ei=RfgtSsDNL4WgM8u-rf4J& sa=X& oi=book_result& ct=result&
resnum=7#PPA215,M1
[2] M.O. Scully, W.E. Lamb, A. Barut, On the theory of the Stern-Gerlach apparatus, Foundations of Physics 17, 575-583 (1987).
[3] Quantum mechanics: Myths and facts (http:/ / arxiv. org/ pdf/ quant-ph/ 0609163)
[4] S. Grblacher et al., An experimental test of non-local realism, Nature 446, 871 (2007). Direct web link (http:/ / dx. doi. org/ 10. 1038/
nature05677)
[5] http:/ / physicsweb. org/ article/ world/ 15/ 9/ 1
[6] http:/ / plato. stanford. edu/ entries/ qt-measurement/
[7] http:/ / arxiv. org/ abs/ quant-ph/ 0312059
[8] http:/ / arxiv. org/ abs/ quant-ph/ 0505070
[9] http:/ / arxiv. org/ pdf/ 0810. 1919
[10] http:/ / arxiv1. library. cornell. edu/ abs/ 1001. 3032v1

116

117

7. Advanced Concepts
Quantum Number
Quantum numbers describe values of conserved quantities in the dynamics of the quantum system. Perhaps the
most peculiar aspect of quantum mechanics is the quantization of observable quantities. This is distinguished from
classical mechanics where the values can range continuously. They often describe specifically the energies of
electrons in atoms, but other possibilities include angular momentum, spin etc. Any quantum system can have one or
more quantum numbers, it is thus rigorous to list all possible quantum numbers.

How many quantum numbers?


The question of how many quantum numbers are needed to describe any given system has no universal answer,
although for each system one must find the answer for a full analysis of the system. Obviously, a quantized system
requires at least one quantum number. The dynamics of any quantum system are described by a quantum
Hamiltonian, H. There is one quantum number of the system corresponding to the energy, i.e., the eigenvalue of the
Hamiltonian. There is also one quantum number for each operator O that commutes with the Hamiltonian (i.e.
satisfies the relation HO=OH). These are all the quantum numbers that the system can have. Note that the
operators O defining the quantum numbers should be independent of each other. Often there is more than one way to
choose a set of independent operators. Consequently, in different situations different sets of quantum numbers may
be used for the description of the same system.
To completely describe an electron in an atom, four quantum numbers are needed.

Traditional nomenclature
Many different models have been proposed throughout the history of quantum mechanics, but the most prominent
system of nomenclature spawned from the Hund-Mulliken molecular orbital theory of Friedrich Hund, Robert S.
Mulliken, and contributions from Schrdinger, Slater and John Lennard-Jones. This system of nomenclature
incorporated Bohr energy levels, Hund-Mulliken orbital theory, and observations on electron spin based on
spectroscopy and Hund's rules.
This model describes electrons using four quantum numbers,

, and

. It is also the common

nomenclature in the classical description of nuclear particle states (e.g., proton and neutrons.)
The first,

, describes the electron shell, or energy level.

The value of ranges from 1 to "n", where "n" is the shell containing the outermost electron of that atom. For
example, in cesium (Cs), the outermost valence electron is in the shell with energy level 6, so an electron in
cesium can have an value from 1 to 6. This is known as the principal quantum number.
The second, , describes the subshell (0 = s orbital, 1 = p orbital, 2 = d orbital, 3 = f orbital, etc.).
The value of

ranges from

to

. This is because the first p orbital (l=1) appears in the second

electron shell (n=2), the first d orbital (l=2) appears in the third shell (n=3), and so on. A quantum number
beginning in 3,0,... describes an electron in the s orbital of the third electron shell of an atom.
The third,
, describes the specific orbital (or "cloud") within that subshell.*
The values of

range from

to

. The s subshell (l=0) contains only one orbital, and therefore the ml

of an electron in an s subshell will always be 0. The p subshell (l=1) contains three orbitals (in some systems,
depicted as three "dumbbell-shaped" clouds), so the ml of an electron in a p subshell will be -1, 0, or 1. The d

Quantum Number

118

subshell (l=2) contains five orbitals, with ml values of -2,-1,0,1, and 2.


The fourth,
, describes the spin of the electron within that orbital.*
An electron can have a spin of either or

will be either , corresponding with "spin" and "opposite

spin." Each electron in any individual orbital must have different spins, therefore, an orbital never contains
more than two electrons.
* Note that, since atoms and electrons are in a state of constant motion, there is no universal fixed value for ml and
ms values. Therefore, the ml and ms values are defined somewhat arbitrarily. The only requirement is that the
naming schematic used within a particular set of calculations or descriptions must be consistent (e.g. the orbital
occupied by the first electron in a p subshell could be described as ml=-1 or ml=0, or ml=1, but the ml value of the
other electron in that orbital must be the same, and the ml assigned to electrons in other orbitals must be different).
These rules are summarized as follows:
name

symbol

orbital meaning

range of values

value example

principal quantum number

shell

azimuthal quantum number (angular


momentum)

subshell (s orbital is listed as 0, p orbital as


1 etc.)

for

magnetic quantum number, (projection


of angular momentum)

energy shift (orientation of the subshell's


shape)

for

spin projection quantum number

spin of the electron (-1/2 =


counter-clockwise, 1/2 = clockwise)

for an electron, either:

Example: The quantum numbers used to refer to the outermost valence electrons of the Carbon (C) atom, which are
located in the 2p atomic orbital, are; n = 2 (2nd electron shell), l = 1 (p orbital subshell), ml = 1, 0 or 1, ms = 1/2
(parallel spins).
As applied to the Hamiltonian and Schrdinger equation
The principal quantum number (n = 1, 2, 3, 4 ...) denotes the eigenvalue of H with the J2 part removed. This
number therefore has a dependence only on the distance between the electron and the nucleus (i.e., the radial
coordinate, r). The average distance increases with n, and hence quantum states with different principal quantum
numbers are said to belong to different shells.
The azimuthal quantum number (l = 0, 1 ... n1) (also known as the angular quantum number or orbital
quantum number) gives the orbital angular momentum through the relation
. In chemistry,
this quantum number is very important, since it specifies the shape of an atomic orbital and strongly influences
chemical bonds and bond angles. In some contexts, l=0 is called an s orbital, l=1 a p orbital, l=2 a d orbital, and
l=3 an f orbital.
The magnetic quantum number (ml = l, l+1 ... 0 ... l1, l) yields the projection of the orbital angular momentum
along a specified axis.
.
The spin projection quantum number (ms = 1/2 or +1/2), is the intrinsic angular momentum of the electron or
nucleon. This is the projection of the spin s=1/2 along the specified axis.
Results from spectroscopy indicated that up to two electrons can occupy a single orbital. However two
electrons can never have the same exact quantum state nor the same set of quantum numbers according to
Hund's Rules, which addresses the Pauli exclusion principle. A fourth quantum number with two possible
values was added as an ad hoc assumption to resolve the conflict; this supposition could later be explained in
detail by relativistic quantum mechanics and from the results of the renowned Stern-Gerlach experiment.
Molecular orbitals require different quantum numbers, because the Hamiltonian and its symmetries are quite
different.

Quantum Number

119

Quantum numbers with spin-orbit interaction


When one takes the spin-orbit interaction into consideration, the l-, m- and s-operators no longer commute with the
Hamiltonian, and their eigenvalues therefore change over time. Thus another set of quantum numbers should be
used. This set includes
The total angular momentum quantum number

gives the total angular momentum through the relation

.
The projection of the total angular momentum along a specified axis (mj = -j,-j+1... j), which is analogous to m,
and satisfies
where
Parity. This is the eigenvalue under reflection, and is positive (i.e. +1) for states which came from even l and
negative (i.e. -1) for states which came from odd l. The former is also known as even parity and the latter as odd
parity
For example, consider the following eight states, defined by their quantum numbers:
n

l ml

ms

l + s l - s ml + ms

#1.

2 1

1 +1/2

3/2

1/2

3/2

#2.

2 1

1 -1/2

3/2

1/2

1/2

#3.

2 1

0 +1/2

3/2

1/2

1/2

#4.

2 1

0 -1/2

3/2

1/2

-1/2

#5.

2 1 -1 +1/2

3/2

1/2

-1/2

#6.

2 1 -1 -1/2

3/2

1/2

-3/2

#7.

2 0

0 +1/2

1/2 -1/2

1/2

#8.

2 0

0 -1/2

1/2 -1/2

-1/2

The quantum states in the system can be described as linear combination of these eight states. However, in the
presence of spin-orbit interaction, if one wants to describe the same system by eight states which are eigenvectors of
the Hamiltonian (i.e. each represents a state which does not mix with others over time), we should consider the
following eight states:
j = 3/2, mj =

3/2,

odd parity (coming from state (1) above)

j = 3/2, mj =

1/2,

odd parity (coming from states (2) and (3) above)

j = 3/2, mj = -1/2,

odd parity (coming from states (4) and (5) above)

j = 3/2, mj = -3/2,

odd parity (coming from state (6))

j = 1/2, mj =

1/2,

odd parity (coming from states (2) and (3) above)

j = 1/2, mj = -1/2,

odd parity (coming from states (4) and (5) above)

j = 1/2, mj =

1/2, even parity (coming from state (7) above)

j = 1/2, mj = -1/2, even parity (coming from state (8) above)

Elementary particles
Elementary particles contain many quantum numbers which are usually said to be intrinsic to them. However, it
should be understood that the elementary particles are quantum states of the standard model of particle physics, and
hence the quantum numbers of these particles bear the same relation to the Hamiltonian of this model as the quantum
numbers of the Bohr atom does to its Hamiltonian. In other words, each quantum number denotes a symmetry of the
problem. It is more useful in field theory to distinguish between spacetime and internal symmetries.

Quantum Number
Typical quantum numbers related to spacetime symmetries are spin (related to rotational symmetry), the parity,
C-parity and T-parity (related to the Poincare symmetry of spacetime). Typical internal symmetries are lepton
number and baryon number or the electric charge. (For a full list of quantum numbers of this kind see the article on
flavour.)
It is worth mentioning here a minor but often confusing point. Most conserved quantum numbers are additive. Thus,
in an elementary particle reaction, the sum of the quantum numbers should be the same before and after the reaction.
However, some, usually called a parity, are multiplicative; i.e., their product is conserved. All multiplicative
quantum numbers belong to a symmetry (like parity) in which applying the symmetry transformation twice is
equivalent to doing nothing. These are all examples of an abstract group called Z2.

References and external links


General principles
Dirac, Paul A.M. (1982). Principles of quantum mechanics. Oxford University Press. ISBN0-19-852011-5.

Atomic physics
Quantum numbers for the hydrogen atom (http://hyperphysics.phy-astr.gsu.edu/hbase/qunoh.html)

Particle physics
Griffiths, David J. (2004). Introduction to Quantum Mechanics (2nd ed.). Prentice Hall. ISBN0-13-805326-X.
Halzen, Francis and Martin, Alan D. (1984). QUARKS AND LEPTONS: An Introductory Course in Modern
Particle Physics. John Wiley & Sons. ISBN0-471-88741-2.
The particle data group (http://pdg.lbl.gov/)
Lecture notes on quantum numbers (http://www.physics.byu.edu/faculty/durfee/courses/Summer2009/
physics222/AtomicQuantumNumbers.pdf)

120

Quantum Information

121

Quantum Information
In quantum mechanics, quantum information is physical information that is held in the "state" of a quantum
system. The most popular unit of quantum information is the qubit, a two-level quantum system. However, unlike
classical digital states (which are discrete), a two-state quantum system can actually be in a superposition of the two
states at any given time.
Quantum information differs from classical information in several respects, among which we note the following:
It cannot be read without the state becoming the measured value,
An arbitrary state cannot be cloned,
The state may be in a superposition of basis values.
However, despite this, the amount of information that can be retrieved in a single qubit is equal to one bit. It is in the
processing of information (quantum computation) that the differentiation occurs.
The ability to manipulate quantum information enables us to perform tasks that would be unachievable in a classical
context, such as unconditionally secure transmission of information. Quantum information processing is the most
general field that is concerned with quantum information. There are certain tasks which classical computers cannot
perform "efficiently" (that is, in polynomial time) according to any known algorithm. However, a quantum computer
can compute the answer to some of these problems in polynomial time; one well-known example of this is Shor's
factoring algorithm. Other algorithms can speed up a task less dramatically - for example, Grover's search algorithm
which gives a quadratic speed-up over the best possible classical algorithm.
Quantum information, and changes in quantum information, can be quantitatively measured by using an analogue of
Shannon entropy, called the von Neumann entropy. Given a statistical ensemble of quantum mechanical systems
with the density matrix , it is given by

Many of the same entropy measures in classical information theory can also be generalized to the quantum case,
such as Holevo entropy [1] and the conditional quantum entropy.

Quantum information theory


The theory of quantum information is a result of the effort to generalise classical information theory to the quantum
world. Quantum information theory aims to answer the following question:
What happens if information is stored in a state of a quantum system?
One of the strengths of classical information theory is that physical representation of information can be disregarded:
There is no need for an 'ink-on-paper' information theory or a 'DVD information' theory. This is because it is always
possible to efficiently transform information from one representation to another. However, this is not the case for
quantum information: it is not possible, for example, to write down on paper the previously unknown information
contained in the polarisation of a photon.
In general, quantum mechanics does not allow us to read out the state of a quantum system with arbitrary precision.
The existence of Bell correlations between quantum systems cannot be converted into classical information. It is
only possible to transform quantum information between quantum systems of sufficient information capacity. The
information content of a message
can, for this reason, be measured in terms of the minimum number n of
two-level systems which are needed to store the message:

consists of n qubits. In its original theoretical sense,

the term qubit is thus a measure for the amount of information. A two-level quantum system can carry at most one
qubit, in the same sense a classical binary digit can carry at most one classical bit.
As a consequence of the noisy-channel coding theorem, noise limits the information content of an analog
information carrier to be finite. It is very difficult to protect the remaining finite information content of analog

Quantum Information
information carriers against noise. The example of classical analog information shows that quantum information
processing schemes must necessarily be tolerant against noise, otherwise there would not be a chance for them to be
useful. It was a big breakthrough for the theory of quantum information, when quantum error correction codes and
fault-tolerant quantum computation schemes were discovered.

Journals
Among the journals in this field are
International Journal of Quantum Information
Journal of Quantum Chemistry
Applied Mathematics & Information Sciences

External links and references


Lectures at the Institut Henri Poincar (slides and videos) [2]
Quantum Information Theory at ETH Zurich [3]
Quantum Information [4] Perimeter Institute for Theoretical Physics
Center for Quantum Computation [5] - The CQC, part of Cambridge University, is a group of researchers studying
quantum information, and is a useful portal for those interested in this field.
Quantum Information Group [6] The quantum information research group at the University of Nottingham.
Qwiki [7] - A quantum physics wiki devoted to providing technical resources for practicing quantum information
scientists.
Quantiki [8] - A wiki portal for quantum information with introductory tutorials.
Charles H. Bennett and Peter W. Shor, "Quantum Information Theory," IEEE Transactions on Information
Theory, Vol 44, pp 27242742, Oct 1998
Institute for Quantum Computing [9] - The Institute for Quantum Computing, based in Waterloo, ON Canada, is a
research institute working in conjunction with the University of Waterloo [10] and Perimeter Institute [11] on the
subject of Quantum Information.
Quantum information can be negative [12]
Gregg Jaeger's book on Quantum Information [13](published by Springer, New York, 2007, ISBN 0-387-35725-4)
The International Conference on Quantum Information (ICQI) [14]
New Trends in Quantum Computation, Stony Brook, 2010 [15]
Institute of Quantum Information [16] Caltech
Quantum Information Theory [17] Imperial College
Quantum Information Technology [18] Toshiba Research
International Journal of Quantum Information [19] World Scientific
Quantum Information Processing [20] Springer
USC Center for Quantum Information Science & Technology [21]
Center for Quantum Information and Control [22] Theoretical and experimental groups from University of New
Mexico and University of Arizona.

122

Quantum Information

References
[1] http:/ / www. mi. ras. ru/ ~holevo/ eindex. html
[2] http:/ / www. quantware. ups-tlse. fr/ IHP2006/
[3] http:/ / www. qit. ethz. ch/
[4] http:/ / www. perimeterinstitute. ca/ Outreach/ What_We_Research/ Quantum_Information/
[5] http:/ / cam. qubit. org/
[6] http:/ / www. maths. nottingham. ac. uk/ research/ appliedmathematics/ quantuminformation/
[7] http:/ / qwiki. caltech. edu/
[8] http:/ / www. quantiki. org
[9] http:/ / www. iqc. ca/
[10] http:/ / www. uwaterloo. ca
[11] http:/ / www. perimeterinstitute. ca/
[12] http:/ / www. damtp. cam. ac. uk/ user/ jono/ negative-information. html
[13] http:/ / www. springer. com/ east/ home?SGWID=5-102-22-173664707-0& changeHeader=true
[14] http:/ / osa. org/ meetings/ topicalmeetings/ icqi/ default. aspx
[15] http:/ / insti. physics. sunysb. edu/ itp/ conf/ simons-qcomputation2/ program. html
[16] http:/ / www. iqi. caltech. edu/
[17] http:/ / www3. imperial. ac. uk/ quantuminformation
[18] http:/ / www. toshiba-europe. com/ research/ crl/ qig/ index. html
[19] http:/ / www. worldscinet. com/ ijqi/ ijqi. shtml
[20] http:/ / www. springer. com/ new+ %26+ forthcoming+ titles+ %28default%29/ journal/ 11128
[21] http:/ / cqist. usc. edu/
[22] http:/ / www. cquic. org/

Quantum Statistical Mechanics


Quantum statistical mechanics is the study of statistical ensembles of quantum mechanical systems. A statistical
ensemble is described by a density operator S, which is a non-negative, self-adjoint, trace-class operator of trace 1 on
the Hilbert space H describing the quantum system. This can be shown under various mathematical formalisms for
quantum mechanics. One such formalism is provided by quantum logic.

Expectation
From classical probability theory, we know that the expectation of a random variable X is completely determined by
its distribution DX by

assuming, of course, that the random variable is integrable or that the random variable is non-negative. Similarly, let
A be an observable of a quantum mechanical system. A is given by a densely defined self-adjoint operator on H. The
spectral measure of A defined by

uniquely determines A and conversely, is uniquely determined by A. EA is a boolean homomorphism from the Borel
subsets of R into the lattice Q of self-adjoint projections of H. In analogy with probability theory, given a state S, we
introduce the distribution of A under S which is the probability measure defined on the Borel subsets of R by

Similarly, the expected value of A is defined in terms of the probability distribution DA by

123

Quantum Statistical Mechanics

124

Note that this expectation is relative to the mixed state S which is used in the definition of DA.
Remark. For technical reasons, one needs to consider separately the positive and negative parts of A defined by the
Borel functional calculus for unbounded operators.
One can easily show:

Note that if S is a pure state corresponding to the vector ,

Von Neumann entropy


Of particular significance for describing randomness of a state is the von Neumann entropy of S formally defined by
.
Actually, the operator S log2 S is not necessarily trace-class. However, if S is a non-negative self-adjoint operator not
of trace class we define Tr(S) = +. Also note that any density operator S can be diagonalized, that it can be
represented in some orthonormal basis by a (possibly infinite) matrix of the form

and we define

The convention is that

, since an event with probability zero should not contribute to the entropy. This

value is an extended real number (that is in [0, ]) and this is clearly a unitary invariant of S.
Remark. It is indeed possible that H(S) = + for some density operator S. In fact T be the diagonal matrix

T is non-negative trace class and one can show T log2 T is not trace-class.
Theorem. Entropy is a unitary invariant.
In analogy with classical entropy (notice the similarity in the definitions), H(S) measures the amount of randomness
in the state S. The more dispersed the eigenvalues are, the larger the system entropy. For a system in which the space
H is finite-dimensional, entropy is maximized for the states S which in diagonal form have the representation

For such an S, H(S) = log2 n. The state S is called the maximally mixed state.
Recall that a pure state is one of the form

Quantum Statistical Mechanics


for a vector of norm 1.
Theorem. H(S) = 0 if and only if S is a pure state.
For S is a pure state if and only if its diagonal form has exactly one non-zero entry which is a 1.
Entropy can be used as a measure of quantum entanglement.

Gibbs canonical ensemble


Consider an ensemble of systems described by a Hamiltonian H with average energy E. If H has pure-point spectrum
and the eigenvalues
of H go to + sufficiently fast, e-r H will be a non-negative trace-class operator for every
positive r.
The Gibbs canonical ensemble is described by the state

where is such that the ensemble average of energy satisfies

,and

is the quantum mechanical version of the canonical partition function. The probability that a system chosen at
random from the ensemble will be in a state corresponding to energy eigenvalue
is

Under certain conditions, the Gibbs canonical ensemble maximizes the von Neumann entropy of the state subject to
the energy conservation requirement.

References
J. von Neumann, Mathematical Foundations of Quantum Mechanics, Princeton University Press, 1955.
F. Reif, Statistical and Thermal Physics, McGraw-Hill, 1965.

125

126

8. Advanced Topics
Quantum Field Theory
Quantum field theory (QFT)[1] provides a theoretical framework for constructing quantum mechanical models of
systems classically parametrized (represented) by an infinite number of dynamical degrees of freedom, that is, fields
and (in a condensed matter context) many-body systems. It is the natural and quantitative language of particle
physics and condensed matter physics. Most theories in modern particle physics, including the Standard Model of
elementary particles and their interactions, are formulated as relativistic quantum field theories. Quantum field
theories are used in many contexts, elementary particle physics being the most vital example, where the particle
count/number going into a reaction fluctuates and changes, differing from the count/number going out, for example,
and for the description of critical phenomena and quantum phase transitions, such as in the BCS theory of
superconductivity, also see phase transition, quantum phase transition, critical phenomena. Quantum field theory is
thought by many to be the unique and correct outcome of combining the rules of quantum mechanics with special
relativity.
In perturbative quantum field theory, the forces between particles are mediated by other particles. The
electromagnetic force between two electrons is caused by an exchange of photons. Intermediate vector bosons
mediate the weak force and gluons mediate the strong force. There is currently no complete quantum theory of the
remaining fundamental force, gravity, but many of the proposed theories postulate the existence of a graviton
particle that mediates it. These force-carrying particles are virtual particles and, by definition, cannot be detected
while carrying the force, because such detection will imply that the force is not being carried. In addition, the notion
of "force mediating particle" comes from perturbation theory, and thus does not make sense in a context of bound
states.
In QFT photons are not thought of as 'little billiard balls', they are considered to be field quanta necessarily
chunked ripples in a field, or "excitations", that 'look like' particles. Fermions, like the electron, can also be described
as ripples/excitations in a field, where each kind of fermion has its own field. In summary, the classical visualisation
of "everything is particles and field", in quantum field theory, resolves into "everything is particles", which then
resolves into "everything is fields". In the end, particles are regarded as excited states of a field (field quanta). The
gravitational field and the electromagnetic field are the only two fundamental fields in Nature that have infinite range
and a corresponding classical low-energy limit, which greatly diminishes and hides their "particle-like" excitations.
Albert Einstein, in 1905, attributed "particle-like" and discrete exchanges of momenta and energy, characteristic of
"field quanta", to the electromagnetic field. Originally, his principal motivation was to explain the thermodynamics
of radiation. Although it is often claimed that the photoelectric and Compton effects require a quantum description of
the EM field, this is now understood to be untrue, and proper proof of the quantum nature of radiation is now taken
up into modern quantum optics as in the antibunching effect.[2] The word "photon" was coined in 1926 by the great
physical chemist Gilbert Newton Lewis (see also the articles photon antibunching and laser).
The "low-energy limit" of the correct quantum field-theoretic description of the electromagnetic field, quantum
electrodynamics, is believed to become James Clerk Maxwell's 1864 theory, although the "classical limit" of
quantum electrodynamics has not been as widely explored as that of quantum mechanics. Presumably, the as yet
unknown correct quantum field-theoretic treatment of the gravitational field will become and "look exactly like"
Einstein's general theory of relativity in the "low-energy limit". Indeed, quantum field theory itself is possibly the
low-energy-effective-field-theory limit of a more fundamental theory such as superstring theory. Compare in this
context the article effective field theory.

Quantum Field Theory

History
Quantum field theory originated in the 1920s from the problem of creating a quantum mechanical theory of the
electromagnetic field. In 1925, Werner Heisenberg, Max Born, and Pascual Jordan constructed such a theory by
expressing the field's internal degrees of freedom as an infinite set of harmonic oscillators and by employing the
canonical quantization procedure to those oscillators. This theory assumed that no electric charges or currents were
present and today would be called a free field theory. The first reasonably complete theory of quantum
electrodynamics, which included both the electromagnetic field and electrically charged matter (specifically,
electrons) as quantum mechanical objects, was created by Paul Dirac in 1927.[3] This quantum field theory could be
used to model important processes such as the emission of a photon by an electron dropping into a quantum state of
lower energy, a process in which the number of particles changesone atom in the initial state becomes an atom
plus a photon in the final state. It is now understood that the ability to describe such processes is one of the most
important features of quantum field theory.
It was evident from the beginning that a proper quantum treatment of the electromagnetic field had to somehow
incorporate Einstein's relativity theory, which had grown out of the study of classical electromagnetism. This need to
put together relativity and quantum mechanics was the second major motivation in the development of quantum field
theory. Pascual Jordan and Wolfgang Pauli showed in 1928 that quantum fields could be made to behave in the way
predicted by special relativity during coordinate transformations (specifically, they showed that the field
commutators were Lorentz invariant). A further boost for quantum field theory came with the discovery of the Dirac
equation, which was originally formulated and interpreted as a single-particle equation analogous to the Schrdinger
equation, but unlike the Schrdinger equation, the Dirac equation satisfies both the Lorentz invariance, that is, the
requirements of special relativity, and the rules of quantum mechanics. The Dirac equation accommodated the
spin-1/2 value of the electron and accounted for its magnetic moment as well as giving accurate predictions for the
spectra of hydrogen. The attempted interpretation of the Dirac equation as a single-particle equation could not be
maintained long, however, and finally it was shown that several of its undesirable properties (such as
negative-energy states) could be made sense of by reformulating and reinterpreting the Dirac equation as a true field
equation, in this case for the quantized "Dirac field" or the "electron field", with the "negative-energy solutions"
pointing to the existence of anti-particles. This work was performed first by Dirac himself with the invention of hole
theory in 1930 and by Wendell Furry, Robert Oppenheimer, Vladimir Fock, and others. Schrdinger, during the
same period that he discovered his famous equation in 1926, also independently found the relativistic generalization
of it known as the Klein-Gordon equation but dismissed it since, without spin, it predicted impossible properties for
the hydrogen spectrum. (See Oskar Klein and Walter Gordon.) All relativistic wave equations that describe spin-zero
particles are said to be of the Klein-Gordon type.
Of great importance are the studies of soviet physicists, Viktor Ambartsumian and Dmitri Ivanenko, in particular the
Ambarzumian-Ivanenko hypothesis of creation of massive particles (published in 1930) which is the corner stone of
the contemporary quantum field theory.[4] The idea that not only the quanta of the electromagnetic field, photons, but
also other particles (including particles having nonzero rest mass) may be born and disappear as a result of their
interaction with other particles. This idea of Ambartsumian and Ivanenko formed the basis of modern quantum field
theory and theory of elementary particles.[5] [6]
A subtle and careful analysis in 1933 and later in 1950 by Niels Bohr and Leon Rosenfeld showed that there is a
fundamental limitation on the ability to simultaneously measure the electric and magnetic field strengths that enter
into the description of charges in interaction with radiation, imposed by the uncertainty principle, which must apply
to all canonically conjugate quantities. This limitation is crucial for the successful formulation and interpretation of a
quantum field theory of photons and electrons(quantum electrodynamics),and indeed,any perturbative quantum field
theory. The analysis of Bohr and Rosenfeld explains fluctuations in the values of the electromagnetic field that differ
from the classically "allowed" values distant from the sources of the field. Their analysis was crucial to showing that
the limitations and physical implications of the uncertainty principle apply to all dynamical systems, whether fields

127

Quantum Field Theory


or material particles. Their analysis also convinced most people that any notion of returning to a fundamental
description of nature based on classical field theory, such as what Einstein aimed at with his numerous and failed
attempts at a classical unified field theory, was simply out of the question.
The third thread in the development of quantum field theory was the need to handle the statistics of many-particle
systems consistently and with ease. In 1927, Jordan tried to extend the canonical quantization of fields to the
many-body wave functions of identical particles, a procedure that is sometimes called second quantization. In 1928,
Jordan and Eugene Wigner found that the quantum field describing electrons, or other fermions, had to be expanded
using anti-commuting creation and annihilation operators due to the Pauli exclusion principle. This thread of
development was incorporated into many-body theory and strongly influenced condensed matter physics and nuclear
physics.
Despite its early successes quantum field theory was plagued by several serious theoretical difficulties. Basic
physical quantities, such as the self-energy of the electron, the energy shift of electron states due to the presence of
the electromagnetic field, gave infinite, divergent contributionsa nonsensical resultwhen computed using the
perturbative techniques available in the 1930s and most of the 1940s. The electron self-energy problem was already
a serious issue in the classical electromagnetic field theory, where the attempt to attribute to the electron a finite size
or extent (the classical electron-radius) led immediately to the question of what non-electromagnetic stresses would
need to be invoked, which would presumably hold the electron together against the Coulomb repulsion of its
finite-sized "parts". The situation was dire, and had certain features that reminded many of the "Rayleigh-Jeans
difficulty". What made the situation in the 1940s so desperate and gloomy, however, was the fact that the correct
ingredients (the second-quantized Maxwell-Dirac field equations) for the theoretical description of interacting
photons and electrons were well in place, and no major conceptual change was needed analogous to that which was
necessitated by a finite and physically sensible account of the radiative behavior of hot objects, as provided by the
Planck radiation law.
This "divergence problem" was solved in the case of quantum electrodynamics during the late 1940s and early 1950s
by Hans Bethe, Tomonaga, Schwinger, Feynman, and Dyson, through the procedure known as renormalization.
Great progress was made after realizing that ALL infinities in quantum electrodynamics are related to two effects:
the self-energy of the electron/positron and vacuum polarization. Renormalization concerns the business of paying
very careful attention to just what is meant by, for example, the very concepts "charge" and "mass" as they occur in
the pure, non-interacting field-equations. The "vacuum" is itself polarizable and, hence, populated by virtual particle
(on shell and off shell) pairs, and, hence, is a seething and busy dynamical system in its own right. This was a critical
step in identifying the source of "infinities" and "divergences". The "bare mass" and the "bare charge" of a particle,
the values that appear in the free-field equations (non-interacting case), are abstractions that are simply not realized
in experiment (in interaction). What we measure, and hence, what we must take account of with our equations, and
what the solutions must account for, are the "renormalized mass" and the "renormalized charge" of a particle. That is
to say, the "shifted" or "dressed" values these quantities must have when due care is taken to include all deviations
from their "bare values" is dictated by the very nature of quantum fields themselves.
The first approach that bore fruit is known as the "interaction representation", (see the article Interaction picture) a
Lorentz covariant and gauge-invariant generalization of time-dependent perturbation theory used in ordinary
quantum mechanics, and developed by Tomonaga and Schwinger, generalizing earlier efforts of Dirac, Fock and
Podolsky. Tomonaga and Schwinger invented a relativistically covariant scheme for representing field commutators
and field operators intermediate between the two main representations of a quantum system, the Schrdinger and the
Heisenberg representations (see the article on quantum mechanics). Within this scheme, field commutators at
separated points can be evaluated in terms of "bare" field creation and annihilation operators. This allows for keeping
track of the time-evolution of both the "bare" and "renormalized", or perturbed, values of the Hamiltonian and
expresses everything in terms of the coupled, gauge invariant "bare" field-equations. Schwinger gave the most
elegant formulation of this approach. The next and most famous development is due to Feynman, who, with his

128

Quantum Field Theory


brilliant rules for assigning a "graph"/"diagram" to the terms in the scattering matrix (See S-Matrix Feynman
diagrams). These directly corresponded (through the Schwinger-Dyson equation) to the measurable physical
processes (cross sections, probability amplitudes, decay widths and lifetimes of excited states) one needs to be able
to calculate. This revolutionized how quantum field theory calculations are carried-out in practice.
Two classic text-books from the 1960s, J.D. Bjorken and S.D. Drell, Relativistic Quantum Mechanics (1964) and J.J.
Sakurai, Advanced Quantum Mechanics (1967), thoroughly developed the Feynman graph expansion techniques
using physically intuitive and practical methods following from the correspondence principle, without worrying
about the technicalities involved in deriving the Feynman rules from the superstructure of quantum field theory
itself. Although both Feynman's heuristic and pictorial style of dealing with the infinities, as well as the formal
methods of Tomonaga and Schwinger, worked extremely well, and gave spectacularly accurate answers, the true
analytical nature of the question of "renormalizability", that is, whether ANY theory formulated as a "quantum field
theory" would give finite answers, was not worked-out till much later, when the urgency of trying to formulate finite
theories for the strong and electro-weak (and gravitational interactions) demanded its solution.
Renormalization in the case of QED was largely fortuitous due to the smallness of the coupling constant, the fact that
the coupling has no dimensions involving mass, the so-called fine structure constant, and also the zero-mass of the
gauge boson involved, the photon, rendered the small-distance/high-energy behavior of QED manageable. Also,
electromagnetic processess are very "clean" in the sense that they are not badly suppressed/damped and/or hidden by
the other gauge interactions. By 1958 Sidney Drell observed: "Quantum electrodynamics (QED) has achieved a
status of peaceful coexistence with its divergences...".
The unification of the electromagnetic force with the weak force encountered initial difficulties due to the lack of
accelerator energies high enough to reveal processes beyond the Fermi interaction range. Additionally, a satisfactory
theoretical understanding of hadron substructure had to be developed, culminating in the quark model.
In the case of the strong interactions, progress concerning their short-distance/high-energy behavior was much
slower and more frustrating. For strong interactions with the electro-weak fields, there were difficult issues regarding
the strength of coupling, the mass generation of the force carriers as well as their non-linear, self interactions.
Although there has been theoretical progress toward a grand unified quantum field theory incorporating the
electro-magnetic force, the weak force and the strong force, empirical verification is still pending. Superunification,
incorporating the gravitational force, is still very speculative, and is under intensive investigation by many of the best
minds in contemporary theoretical physics. Gravitation is a tensor field description of a spin-2 gauge-boson, the
"graviton", and is further discussed in the articles on general relativity and quantum gravity.
From the point of view of the techniques of (four-dimensional) quantum field theory, and as the numerous and heroic
efforts to formulate a consistent quantum gravity theory by some very able minds attests, gravitational quantization
was, and is still, the reigning champion for bad behavior. There are problems and frustrations stemming from the fact
that the gravitational coupling constant has dimensions involving inverse powers of mass, and as a simple
consequence, it is plagued by badly behaved (in the sense of perturbation theory) non-linear and violent
self-interactions. Gravity, basically, gravitates, which in turn...gravitates...and so on, (i.e., gravity is itself a source of
gravity,...,) thus creating a nightmare at all orders of perturbation theory. Also, gravity couples to all energy equally
strongly, as per the equivalence principle, so this makes the notion of ever really "switching-off", "cutting-off" or
separating, the gravitational interaction from other interactions ambiguous and impossible since, with gravitation, we
are dealing with the very structure of space-time itself. (See general covariance and, for a modest, yet highly
non-trivial and significant interplay between (QFT) and gravitation (spacetime), see the article Hawking radiation
and references cited therein. Also quantum field theory in curved spacetime).
Thanks to the somewhat brute-force, clanky and heuristic methods of Feynman, and the elegant and abstract methods
of Tomonaga/Schwinger, from the period of early renormalization, we do have the modern theory of quantum
electrodynamics (QED). It is still the most accurate physical theory known, the prototype of a successful quantum
field theory. Beginning in the 1950s with the work of Yang and Mills, as well as Ryoyu Utiyama, following the

129

Quantum Field Theory


previous lead of Weyl and Pauli, deep explorations illuminated the types of symmetries and invariances any field
theory must satisfy. QED, and indeed, all field theories, were generalized to a class of quantum field theories known
as gauge theories. Quantum electrodynamics is the most famous example of what is known as an Abelian gauge
theory. It relies on the symmetry group U(1) and has one massless gauge field, the U(1) gauge symmetry, dictating
the form of the interactions involving the electromagnetic field, with the photon being the gauge boson. That
symmetries dictate, limit and necessitate the form of interaction between particles is the essence of the "gauge theory
revolution". Yang and Mills formulated the first explicit example of a non-Abelian gauge theory, Yang-Mills theory,
with an attempted explanation of the strong interactions in mind. The strong interactions were then (incorrectly)
understood in the mid-1950s, to be mediated by the pi-mesons, the particles predicted by Hideki Yukawa in 1935,
based on his profound reflections concerning the reciprocal connection between the mass of any force-mediating
particle and the range of the force it mediates. This was allowed by the uncertainty principle. The 1960s and 1970s
saw the formulation of a gauge theory now known as the Standard Model of particle physics, which systematically
describes the elementary particles and the interactions between them.
The electroweak interaction part of the standard model was formulated by Sheldon Glashow in the years 1958-60
with his discovery of the SU(2)xU(1) group structure of the theory. Steven Weinberg and Abdus Salam brilliantly
invoked the Anderson-Higgs mechanism for the generation of the W's and Z masses (the intermediate vector
boson(s) responsible for the weak interactions and neutral-currents) and keeping the mass of the photon zero. The
Goldstone/Higgs idea for generating mass in gauge theories was sparked in the late 1950s and early 1960s when a
number of theoreticians (including Yoichiro Nambu, Steven Weinberg, Jeffrey Goldstone, Franois Englert, Robert
Brout, G. S. Guralnik, C. R. Hagen, Tom Kibble and Philip Warren Anderson) noticed a possibly useful analogy to
the (spontaneous) breaking of the U(1) symmetry of electromagnetism in the formation of the BCS ground-state of a
superconductor. The gauge boson involved in this situation, the photon, behaves as though it has acquired a finite
mass. There is a further possibility that the physical vacuum (ground-state) does not respect the symmetries implied
by the "unbroken" electroweak Lagrangian (see the article Electroweak interaction for more details) from which one
arrives at the field equations. The electroweak theory of Weinberg and Salam was shown to be renormalizable
(finite) and hence consistent by Gerardus 't Hooft and Martinus Veltman. The Glashow-Weinberg-Salam theory
(GWS-Theory) is a triumph and, in certain applications, gives an accuracy on a par with quantum electrodynamics.
Also during the 1970s, parallel developments in the study of phase transitions in condensed matter physics led Leo
Kadanoff, Michael Fisher and Kenneth Wilson (extending work of Ernst Stueckelberg, Andre Peterman, Murray
Gell-Mann, and Francis Low) to a set of ideas and methods known as the renormalization group. By providing a
better physical understanding of the renormalization procedure invented in the 1940s, the renormalization group
sparked what has been called the "grand synthesis" of theoretical physics, uniting the quantum field theoretical
techniques used in particle physics and condensed matter physics into a single theoretical framework.

Principles of quantum field theory


Classical fields and quantum fields
Quantum mechanics, in its most general formulation, is a theory of abstract operators (observables) acting on an
abstract state space (Hilbert space), where the observables represent physically observable quantities and the state
space represents the possible states of the system under study. Furthermore, each observable corresponds, in a
technical sense, to the classical idea of a degree of freedom. For instance, the fundamental observables associated
with the motion of a single quantum mechanical particle are the position and momentum operators and .
Ordinary quantum mechanics deals with systems such as this, which possess a small set of degrees of freedom.
(It is important to note, at this point, that this article does not use the word "particle" in the context of waveparticle
duality. In quantum field theory, "particle" is a generic term for any discrete quantum mechanical entity, such as an
electron or photon, which can behave like classical particles or classical waves under different experimental

130

Quantum Field Theory

131

conditions, such that one could say 'this "particle" can behave like a wave or a particle'.)
A quantum field is a quantum mechanical system containing a large, and possibly infinite, number of degrees of
freedom. A classical field contains a set of degrees of freedom at each point of space; for instance, the classical
electromagnetic field defines two vectors the electric field and the magnetic field that can in principle take on
distinct values for each position . When the field as a whole is considered as a quantum mechanical system, its
observables form an infinite (in fact uncountable) set, because is continuous.
Furthermore, the degrees of freedom in a quantum field are arranged in "repeated" sets. For example, the degrees of
freedom in an electromagnetic field can be grouped according to the position , with exactly two vectors for each
. Note that is an ordinary number that "indexes" the observables; it is not to be confused with the position
operator
encountered in ordinary quantum mechanics, which is an observable. (Thus, ordinary quantum
mechanics is sometimes referred to as "zero-dimensional quantum field theory", because it contains only a single set
of observables.)
It is also important to note that there is nothing special about
one way of indexing the degrees of freedom in the field.

because, as it turns out, there is generally more than

In the following sections, we will show how these ideas can be used to construct a quantum mechanical theory with
the desired properties. We will begin by discussing single-particle quantum mechanics and the associated theory of
many-particle quantum mechanics. Then, by finding a way to index the degrees of freedom in the many-particle
problem, we will construct a quantum field and study its implications.

Single-particle and many-particle quantum mechanics


In quantum mechanics, the time-dependent Schrdinger equation for a single particle is

where

is the particle's mass,

is the applied potential, and

We wish to consider how this problem generalizes to

denotes the wavefunction.

particles. There are two motivations for studying the

many-particle problem. The first is a straightforward need in condensed matter physics, where typically the number
of particles is on the order of Avogadro's number (6.0221415 x 1023). The second motivation for the many-particle
problem arises from particle physics and the desire to incorporate the effects of special relativity. If one attempts to
include the relativistic rest energy into the above equation (in quantum mechanics where position is an observable),
the result is either the Klein-Gordon equation or the Dirac equation. However, these equations have many
unsatisfactory qualities; for instance, they possess energy eigenvalues that extend to , so that there seems to be no
easy definition of a ground state. It turns out that such inconsistencies arise from relativistic wavefunctions having a
probabilistic interpretation in position space, as probability conservation is not a relativistically covariant concept. In
quantum field theory, unlike in quantum mechanics, position is not an observable, and thus, one does not need the
concept of a position-space probability density. For quantum fields whose interaction can be treated perturbatively,
this is equivalent to neglecting the possibility of dynamically creating or destroying particles, which is a crucial
aspect of relativistic quantum theory. Einstein's famous mass-energy relation allows for the possibility that
sufficiently massive particles can decay into several lighter particles, and sufficiently energetic particles can combine
to form massive particles. For example, an electron and a positron can annihilate each other to create photons. This
suggests that a consistent relativistic quantum theory should be able to describe many-particle dynamics.
Furthermore, we will assume that the

particles are indistinguishable. As described in the article on identical

particles, this implies that the state of the entire system must be either symmetric (bosons) or antisymmetric
(fermions) when the coordinates of its constituent particles are exchanged. These multi-particle states are rather
complicated to write. For example, the general quantum state of a system of
bosons is written as

Quantum Field Theory

where

132

are the single-particle states,

over all possible permutations

acting on

is the number of particles occupying state


elements. In general, this is a sum of

terms, which quickly becomes unmanageable as

, and the sum is taken


(

factorial) distinct

increases. The way to simplify this problem is to turn it into a

quantum field theory.

Second quantization
In this section, we will describe a method for constructing a quantum field theory called second quantization. This
basically involves choosing a way to index the quantum mechanical degrees of freedom in the space of multiple
identical-particle states. It is based on the Hamiltonian formulation of quantum mechanics; several other approaches
exist, such as the Feynman path integral,[7] which uses a Lagrangian formulation. For an overview, see the article on
quantization.
Second quantization of bosons
For simplicity, we will first discuss second quantization for bosons, which form perfectly symmetric quantum states.
Let us denote the mutually orthogonal single-particle states by
and so on. For example, the
3-particle state with one particle in state

and two in state

is

The first step in second quantization is to express such quantum states in terms of occupation numbers, by listing
the number of particles occupying each of the single-particle states
etc. This is simply another way of
labelling the states. For instance, the above 3-particle state is denoted as
The next step is to expand the

-particle state space to include the state spaces for all possible values of

. This

extended state space, known as a Fock space, is composed of the state space of a system with no particles (the
so-called vacuum state), plus the state space of a 1-particle system, plus the state space of a 2-particle system, and so
forth. It is easy to see that there is a one-to-one correspondence between the occupation number representation and
valid boson states in the Fock space.
At this point, the quantum mechanical system has become a quantum field in the sense we described above. The
field's elementary degrees of freedom are the occupation numbers, and each occupation number is indexed by a
number
, indicating which of the single-particle states
it refers to.
The properties of this quantum field can be explored by defining creation and annihilation operators, which add and
subtract particles. They are analogous to "ladder operators" in the quantum harmonic oscillator problem, which
added and subtracted energy quanta. However, these operators literally create and annihilate particles of a given
quantum state. The bosonic annihilation operator

and creation operator

have the following effects:

It can be shown that these are operators in the usual quantum mechanical sense, i.e. linear operators acting on the
Fock space. Furthermore, they are indeed Hermitian conjugates, which justifies the way we have written them. They
can be shown to obey the commutation relation

Quantum Field Theory


where

133

stands for the Kronecker delta. These are precisely the relations obeyed by the ladder operators for an

infinite set of independent quantum harmonic oscillators, one for each single-particle state. Adding or removing
bosons from each state is therefore analogous to exciting or de-exciting a quantum of energy in a harmonic
oscillator.
The Hamiltonian of the quantum field (which, through the Schrdinger equation, determines its dynamics) can be
written in terms of creation and annihilation operators. For instance, the Hamiltonian of a field of free
(non-interacting) bosons is

where

is the energy of the

-th single-particle energy eigenstate. Note that

Second quantization of fermions


It turns out that a different definition of creation and annihilation must be used for describing fermions. According to
the Pauli exclusion principle, fermions cannot share quantum states, so their occupation numbers
can only take
on the value 0 or 1. The fermionic annihilation operators

and creation operators

are defined by their actions on

a Fock state thus

These obey an anticommutation relation:

One may notice from this that applying a fermionic creation operator twice gives zero, so it is impossible for the
particles to share single-particle states, in accordance with the exclusion principle.
Field operators
We have previously mentioned that there can be more than one way of indexing the degrees of freedom in a quantum
field. Second quantization indexes the field by enumerating the single-particle quantum states. However, as we have
discussed, it is more natural to think about a "field", such as the electromagnetic field, as a set of degrees of freedom
indexed by position.
To this end, we can define field operators that create or destroy a particle at a particular point in space. In particle
physics, these operators turn out to be more convenient to work with, because they make it easier to formulate
theories that satisfy the demands of relativity.
Single-particle states are usually enumerated in terms of their momenta (as in the particle in a box problem.) We can
construct field operators by applying the Fourier transform to the creation and annihilation operators for these states.
For example, the bosonic field annihilation operator
is

The bosonic field operators obey the commutation relation

where

stands for the Dirac delta function. As before, the fermionic relations are the same, with the

commutators replaced by anticommutators.

Quantum Field Theory

134

The field operator is not the same thing as a single-particle wavefunction. The former is an operator acting on the
Fock space, and the latter is a quantum-mechanical amplitude for finding a particle in some position. However, they
are closely related, and are indeed commonly denoted with the same symbol. If we have a Hamiltonian with a space
representation, say

where the indices

and

run over all particles, then the field theory Hamiltonian (in the non-relativistic limit and

for negligible self-interactions) is

This looks remarkably like an expression for the expectation value of the energy, with

playing the role of the

wavefunction. This relationship between the field operators and wavefunctions makes it very easy to formulate field
theories starting from space-projected Hamiltonians.

Implications of quantum field theory


Unification of fields and particles
The "second quantization" procedure that we have outlined in the previous section takes a set of single-particle
quantum states as a starting point. Sometimes, it is impossible to define such single-particle states, and one must
proceed directly to quantum field theory. For example, a quantum theory of the electromagnetic field must be a
quantum field theory, because it is impossible (for various reasons) to define a wavefunction for a single photon. In
such situations, the quantum field theory can be constructed by examining the mechanical properties of the classical
field and guessing the corresponding quantum theory. For free (non-interacting) quantum fields, the quantum field
theories obtained in this way have the same properties as those obtained using second quantization, such as
well-defined creation and annihilation operators obeying commutation or anticommutation relations.
Quantum field theory thus provides a unified framework for describing "field-like" objects (such as the
electromagnetic field, whose excitations are photons) and "particle-like" objects (such as electrons, which are treated
as excitations of an underlying electron field), so long as one can treat interactions as "perturbations" of free fields.
There are still unsolved problems relating to the more general case of interacting fields that may or may not be
adequately described by perturbation theory. For more on this topic, see Haag's theorem.
Physical meaning of particle indistinguishability
The second quantization procedure relies crucially on the particles being identical. We would not have been able to
construct a quantum field theory from a distinguishable many-particle system, because there would have been no
way of separating and indexing the degrees of freedom.
Many physicists prefer to take the converse interpretation, which is that quantum field theory explains what identical
particles are. In ordinary quantum mechanics, there is not much theoretical motivation for using symmetric
(bosonic) or antisymmetric (fermionic) states, and the need for such states is simply regarded as an empirical fact.
From the point of view of quantum field theory, particles are identical if and only if they are excitations of the same
underlying quantum field. Thus, the question "why are all electrons identical?" arises from mistakenly regarding
individual electrons as fundamental objects, when in fact it is only the electron field that is fundamental.

Quantum Field Theory

135

Particle conservation and non-conservation


During second quantization, we started with a Hamiltonian and state space describing a fixed number of particles (
), and ended with a Hamiltonian and state space for an arbitrary number of particles. Of course, in many common
situations

is an important and perfectly well-defined quantity, e.g. if we are describing a gas of atoms sealed in a

box. From the point of view of quantum field theory, such situations are described by quantum states that are
eigenstates of the number operator
, which measures the total number of particles present. As with any quantum
mechanical observable,
trapped in the
ordinary

is conserved if it commutes with the Hamiltonian. In that case, the quantum state is

-particle subspace of the total Fock space, and the situation could equally well be described by

-particle quantum mechanics. (Strictly speaking, this is only true in the noninteracting case or in the low

energy density limit of renormalized quantum field theories)


For example, we can see that the free-boson Hamiltonian described above conserves particle number. Whenever the
Hamiltonian operates on a state, each particle destroyed by an annihilation operator
is immediately put back by
the creation operator

On the other hand, it is possible, and indeed common, to encounter quantum states that are not eigenstates of

which do not have well-defined particle numbers. Such states are difficult or impossible to handle using ordinary
quantum mechanics, but they can be easily described in quantum field theory as quantum superpositions of states
having different values of
. For example, suppose we have a bosonic field whose particles can be created or
destroyed by interactions with a fermionic field. The Hamiltonian of the combined system would be given by the
Hamiltonians of the free boson and free fermion fields, plus a "potential energy" term such as

where

and

denotes the bosonic creation and annihilation operators,

creation and annihilation operators, and

denotes the fermionic

is a parameter that describes the strength of the interaction. This

"interaction term" describes processes in which a fermion in state


kicked into a different eigenstate

and

either absorbs or emits a boson, thereby being

. (In fact, this type of Hamiltonian is used to describe interaction between

conduction electrons and phonons in metals. The interaction between electrons and photons is treated in a similar
way, but is a little more complicated because the role of spin must be taken into account.) One thing to notice here is
that even if we start out with a fixed number of bosons, we will typically end up with a superposition of states with
different numbers of bosons at later times. The number of fermions, however, is conserved in this case.
In condensed matter physics, states with ill-defined particle numbers are particularly important for describing the
various superfluids. Many of the defining characteristics of a superfluid arise from the notion that its quantum state is
a superposition of states with different particle numbers. In addition, the concept of a coherent state (used to model
the laser and the BCS ground state) refers to a state with an ill-defined particle number but a well-defined phase.

Axiomatic approaches
The preceding description of quantum field theory follows the spirit in which most physicists approach the subject.
However, it is not mathematically rigorous. Over the past several decades, there have been many attempts to put
quantum field theory on a firm mathematical footing by formulating a set of axioms for it. These attempts fall into
two broad classes.
The first class of axioms, first proposed during the 1950s, include the Wightman, Osterwalder-Schrader, and
Haag-Kastler systems. They attempted to formalize the physicists' notion of an "operator-valued field" within the
context of functional analysis, and enjoyed limited success. It was possible to prove that any quantum field theory
satisfying these axioms satisfied certain general theorems, such as the spin-statistics theorem and the CPT theorem.
Unfortunately, it proved extraordinarily difficult to show that any realistic field theory, including the Standard
Model, satisfied these axioms. Most of the theories that could be treated with these analytic axioms were physically
trivial, being restricted to low-dimensions and lacking interesting dynamics. The construction of theories satisfying

Quantum Field Theory


one of these sets of axioms falls in the field of constructive quantum field theory. Important work was done in this
area in the 1970s by Segal, Glimm, Jaffe and others.
During the 1980s, a second set of axioms based on geometric ideas was proposed. This line of investigation, which
restricts its attention to a particular class of quantum field theories known as topological quantum field theories, is
associated most closely with Michael Atiyah and Graeme Segal, and was notably expanded upon by Edward Witten,
Richard Borcherds, and Maxim Kontsevich. However, most of the physically relevant quantum field theories, such
as the Standard Model, are not topological quantum field theories; the quantum field theory of the fractional
quantum Hall effect is a notable exception. The main impact of axiomatic topological quantum field theory has been
on mathematics, with important applications in representation theory, algebraic topology, and differential geometry.
Finding the proper axioms for quantum field theory is still an open and difficult problem in mathematics. One of the
Millennium Prize Problemsproving the existence of a mass gap in Yang-Mills theoryis linked to this issue.

Phenomena associated with quantum field theory


In the previous part of the article, we described the most general properties of quantum field theories. Some of the
quantum field theories studied in various fields of theoretical physics possess additional special properties, such as
renormalizability, gauge symmetry, and supersymmetry. These are described in the following sections.

Renormalization
Early in the history of quantum field theory, it was found that many seemingly innocuous calculations, such as the
perturbative shift in the energy of an electron due to the presence of the electromagnetic field, give infinite results.
The reason is that the perturbation theory for the shift in an energy involves a sum over all other energy levels, and
there are infinitely many levels at short distances that each give a finite contribution.
Many of these problems are related to failures in classical electrodynamics that were identified but unsolved in the
19th century, and they basically stem from the fact that many of the supposedly "intrinsic" properties of an electron
are tied to the electromagnetic field that it carries around with it. The energy carried by a single electron its self
energy is not simply the bare value, but also includes the energy contained in its electromagnetic field, its
attendant cloud of photons. The energy in a field of a spherical source diverges in both classical and quantum
mechanics, but as discovered by Weisskopf with help from Wendell Furry, in quantum mechanics the divergence is
much milder, going only as the logarithm of the radius of the sphere.
The solution to the problem, presciently suggested by Stueckelberg, independently by Bethe after the crucial
experiment by Lamb, implemented at one loop by Schwinger, and systematically extended to all loops by Feynman
and Dyson, with converging work by Tomonaga in isolated postwar Japan, comes from recognizing that all the
infinities in the interactions of photons and electrons can be isolated into redefining a finite number of quantities in
the equations by replacing them with the observed values: specifically the electron 's mass and charge: this is called
renormalization. The technique of renormalization recognizes that the problem is essentially purely mathematical,
that extremely short distances are at fault. In order to define a theory on a continuum, first place a cutoff on the
fields, by postulating that quanta cannot have energies above some extremely high value. This has the effect of
replacing continuous space by a structure where very short wavelengths do not exist, as on a lattice. Lattices break
rotational symmetry, and one of the crucial contributions made by Feynman, Pauli and Villars, and modernized by 't
Hooft and Veltman, is a symmetry-preserving cutoff for perturbation theory (this process is called regularization).
There is no known symmetrical cutoff outside of perturbation theory, so for rigorous or numerical work people often
use an actual lattice.
On a lattice, every quantity is finite but depends on the spacing. When taking the limit of zero spacing, we make sure
that the physically observable quantities like the observed electron mass stay fixed, which means that the constants
in the Lagrangian defining the theory depend on the spacing. Hopefully, by allowing the constants to vary with the

136

Quantum Field Theory


lattice spacing, all the results at long distances become insensitive to the lattice, defining a continuum limit.
The renormalization procedure only works for a certain class of quantum field theories, called renormalizable
quantum field theories. A theory is perturbatively renormalizable when the constants in the Lagrangian only
diverge at worst as logarithms of the lattice spacing for very short spacings. The continuum limit is then well defined
in perturbation theory, and even if it is not fully well defined non-perturbatively, the problems only show up at
distance scales that are exponentially small in the inverse coupling for weak couplings. The Standard Model of
particle physics is perturbatively renormalizable, and so are its component theories (quantum
electrodynamics/electroweak theory and quantum chromodynamics). Of the three components, quantum
electrodynamics is believed to not have a continuum limit, while the asymptotically free SU(2) and SU(3) weak
hypercharge and strong color interactions are nonperturbatively well defined.
The renormalization group describes how renormalizable theories emerge as the long distance low-energy effective
field theory for any given high-energy theory. Because of this, renormalizable theories are insensitive to the precise
nature of the underlying high-energy short-distance phenomena. This is a blessing because it allows physicists to
formulate low energy theories without knowing the details of high energy phenomenon. It is also a curse, because
once a renormalizable theory like the standard model is found to work, it gives very few clues to higher energy
processes. The only way high energy processes can be seen in the standard model is when they allow otherwise
forbidden events, or if they predict quantitative relations between the coupling constants.

Gauge freedom
A gauge theory is a theory that admits a symmetry with a local parameter. For example, in every quantum theory the
global phase of the wave function is arbitrary and does not represent something physical. Consequently, the theory is
invariant under a global change of phases (adding a constant to the phase of all wave functions, everywhere); this is a
global symmetry. In quantum electrodynamics, the theory is also invariant under a local change of phase, that is
one may shift the phase of all wave functions so that the shift may be different at every point in space-time. This is a
local symmetry. However, in order for a well-defined derivative operator to exist, one must introduce a new field,
the gauge field, which also transforms in order for the local change of variables (the phase in our example) not to
affect the derivative. In quantum electrodynamics this gauge field is the electromagnetic field. The change of local
gauge of variables is termed gauge transformation.
In quantum field theory the excitations of fields represent particles. The particle associated with excitations of the
gauge field is the gauge boson, which is the photon in the case of quantum electrodynamics.
The degrees of freedom in quantum field theory are local fluctuations of the fields. The existence of a gauge
symmetry reduces the number of degrees of freedom, simply because some fluctuations of the fields can be
transformed to zero by gauge transformations, so they are equivalent to having no fluctuations at all, and they
therefore have no physical meaning. Such fluctuations are usually called "non-physical degrees of freedom" or gauge
artifacts; usually some of them have a negative norm, making them inadequate for a consistent theory. Therefore, if
a classical field theory has a gauge symmetry, then its quantized version (i.e. the corresponding quantum field
theory) will have this symmetry as well. In other words, a gauge symmetry cannot have a quantum anomaly. If a
gauge symmetry is anomalous (i.e. not kept in the quantum theory) then the theory is non-consistent: for example, in
quantum electrodynamics, had there been a gauge anomaly, this would require the appearance of photons with
longitudinal polarization and polarization in the time direction, the latter having a negative norm, rendering the
theory inconsistent; another possibility would be for these photons to appear only in intermediate processes but not
in the final products of any interaction, making the theory non unitary and again inconsistent (see optical theorem).
In general, the gauge transformations of a theory consist of several different transformations, which may not be
commutative. These transformations are together described by a mathematical object known as a gauge group.
Infinitesimal gauge transformations are the gauge group generators. Therefore the number of gauge bosons is the
group dimension (i.e. number of generators forming a basis).

137

Quantum Field Theory

138

All the fundamental interactions in nature are described by gauge theories. These are:
Quantum electrodynamics, whose gauge transformation is a local change of phase, so that the gauge group is
U(1). The gauge boson is the photon.
Quantum chromodynamics, whose gauge group is SU(3). The gauge bosons are eight gluons.
The electroweak Theory, whose gauge group is
,(a direct product of U(1) and SU(2)).
Gravity, whose classical theory is general relativity, admits the equivalence principle, which is a form of gauge
symmetry. However, it is explicitly non-renormalizable.

Multivalued Gauge Transformations


The gauge transformations which leave the theory invariant involve by definition only single-valued gauge functions
which satisfy the Schwarz integrability criterion

An interesting extension of gauge transformations arises if the gauge functions

are allowed to be multivalued

functions which violate the integrability criterion. These are capable of changing the physical field strengths and are
therefore no proper symmetry transformations. Nevertheless, the transformed field equations describe correctly the
physical laws in the presence of the newly generated field strengths. See the textbook by H. Kleinert cited below for
the applications to phenomena in physics.

Supersymmetry
Supersymmetry assumes that every fundamental fermion has a superpartner that is a boson and vice versa. It was
introduced in order to solve the so-called Hierarchy Problem, that is, to explain why particles not protected by any
symmetry (like the Higgs boson) do not receive radiative corrections to its mass driving it to the larger scales (GUT,
Planck...). It was soon realized that supersymmetry has other interesting properties: its gauged version is an
extension of general relativity (Supergravity), and it is a key ingredient for the consistency of string theory.
The way supersymmetry protects the hierarchies is the following: since for every particle there is a superpartner with
the same mass, any loop in a radiative correction is cancelled by the loop corresponding to its superpartner,
rendering the theory UV finite.
Since no superpartners have yet been observed, if supersymmetry exists it must be broken (through a so-called soft
term, which breaks supersymmetry without ruining its helpful features). The simplest models of this breaking require
that the energy of the superpartners not be too high; in these cases, supersymmetry is expected to be observed by
experiments at the Large Hadron Collider.

Notes
[1] Weinberg, S. Quantum Field Theory, Vols. I to III, 2000, Cambridge University Press: Cambridge, UK.
[2] People.whitman.edu (http:/ / people. whitman. edu/ ~beckmk/ QM/ grangier/ Thorn_ajp. pdf)
[3] Dirac, P.A.M. (1927). The Quantum Theory of the Emission and Absorption of Radiation, Proceedings of the Royal Society of London, Series
A, Vol. 114, p. 243.
[4] G-sardanashvily.ru (http:/ / www. g-sardanashvily. ru/ ivanenko1. html)
[5] Vaprize.sci.am (http:/ / vaprize. sci. am/ results. html)
[6] Sciteclibrary.ru (http:/ / www. sciteclibrary. ru/ texsts/ rus/ stat/ st2718. pdf)
[7] Abraham Pais, Inward Bound: Of Matter and Forces in the Physical World ISBN 0-19-851997-4. Pais recounts how his astonishment at the
rapidity with which Feynman could calculate using his method. Feynman's method is now part of the standard methods for physicists.

Quantum Field Theory

Further reading
General readers:
Feynman, R.P. (2001) [1964]. The Character of Physical Law. MIT Press. ISBN0262560038.
Feynman, R.P. (2006) [1985]. QED: The Strange Theory of Light and Matter. Princeton University Press.
ISBN0691125759.
Gribbin, J. (1998). Q is for Quantum: Particle Physics from A to Z. Weidenfeld & Nicolson. ISBN0297817523.
Schumm, Bruce A. (2004) Deep Down Things. Johns Hopkins Univ. Press. Chpt. 4.
Introductory texts:

Bogoliubov, N.; Shirkov, D. (1982). Quantum Fields. Benjamin-Cummings. ISBN0805309837.


Frampton, P.H. (2000). Gauge Field Theories. Frontiers in Physics (2nd ed.). Wiley.
Greiner, W; Mller, B. (2000). Gauge Theory of Weak Interactions. Springer. ISBN3-540-67672-4.
Itzykson, C.; Zuber, J.-B. (1980). Quantum Field Theory. McGraw-Hill. ISBN0-07-032071-3.
Kane, G.L. (1987). Modern Elementary Particle Physics. Perseus Books. ISBN0-201-11749-5.
Kleinert, H.; Schulte-Frohlinde, Verena (2001). Critical Properties of 4-Theories (http://users.physik.
fu-berlin.de/~kleinert/re.html#B6). World Scientific. ISBN981-02-4658-7.

Kleinert, H. (2008). Multivalued Fields in Condensed Matter, Electrodynamics, and Gravitation (http://users.
physik.fu-berlin.de/~kleinert/public_html/kleiner_reb11/psfiles/mvf.pdf). World Scientific.
ISBN978-981-279-170-2.
Loudon, R (1983). The Quantum Theory of Light. Oxford University Press. ISBN0-19-851155-8.
Mandl, F.; Shaw, G. (1993). Quantum Field Theory. John Wiley & Sons. ISBN0-0471-94186-7.
Peskin, M.; Schroeder, D. (1995). An Introduction to Quantum Field Theory. Westview Press.
ISBN0-201-50397-2.
Ryder, L.H. (1985). Quantum Field Theory. Cambridge University Press. ISBN0-521-33859-X.
Srednicki, Mark (2007) Quantum Field Theory. (http://www.cambridge.org/us/catalogue/catalogue.
asp?isbn=0521864496) Cambridge Univ. Press.
Yndurain, F.J. (1996). Relativistic Quantum Mechanics and Introduction to Field Theory (1st ed.). Springer.
ISBN978-3540604532.
Zee, A. (2003). Quantum Field Theory in a Nutshell. Princeton University Press. ISBN0-691-01019-6.
Advanced texts:
Bogoliubov, N.; Logunov, A.A.; Oksak, A.I.; Todorov, I.T. (1990). General Principles of Quantum Field Theory.
Kluwer Academic Publishers. ISBN978-0792305408.
Weinberg, S. (1995). The Quantum Theory of Fields. 13. Cambridge University Press.
Articles:
Gerard 't Hooft (2007) " The Conceptual Basis of Quantum Field Theory (http://www.phys.uu.nl/~thooft/
lectures/basisqft.pdf)" in Butterfield, J., and John Earman, eds., Philosophy of Physics, Part A. Elsevier:
661-730.
Frank Wilczek (1999) " Quantum field theory (http://arxiv.org/abs/hep-th/9803075)", Reviews of Modern
Physics 71: S83-S95. Also doi=10.1103/Rev. Mod. Phys. 71.

139

Quantum Field Theory

External links
Stanford Encyclopedia of Philosophy: " Quantum Field Theory (http://plato.stanford.edu/entries/
quantum-field-theory/)", by Meinard Kuhlmann.
Siegel, Warren, 2005. Fields. (http://insti.physics.sunysb.edu/~siegel/errata.html) A free text, also available
from arXiv:hep-th/9912205.
Pedagogic Aids to Quantum Field Theory (http://quantumfieldtheory.info). Click on "Introduction" for a
simplified introduction suitable for someone familiar with quantum mechanics.
Free condensed matter books and notes (http://www.freebookcentre.net/Physics/Condensed-Matter-Books.
html).
Quantum field theory texts (http://motls.blogspot.com/2006/01/qft-didactics.html), a list with links to
amazon.com.
Quantum Field Theory (http://www.nat.vu.nl/~mulders/QFT-0.pdf) by P. J. Mulders
Quantum Field Theory (http://damtp.cam.ac.uk/user/tong/qft/qft.pdf) by David Tong
Quantum Field Theory Video Lectures (http://pirsa.org/index.php?p=speaker&name=David_Tong) by David
Tong
Quantum Field Theory Lecture Notes (http://www.physics.utoronto.ca/~luke/PHY2403F_10/
References_files/lecturenotes.pdf) by Michael Luke
Quantum Field Theory Video Lectures (http://www.physics.harvard.edu/about/Phys253.html) by Sidney R.
Coleman
Quantum Field Theory Lecture Notes (http://www.physics.gla.ac.uk/~dmiller/lectures/RQF_1-6_2010.pdf)
by D.J. Miller
Quantum Field Theory Lecture Notes II (http://www.physics.gla.ac.uk/~dmiller/lectures/RQF_7-9_2010.
pdf) by D.J. Miller

String Theory
String theory is a developing theory in particle physics that attempts to reconcile quantum mechanics and general
relativity.[1] It is a contender for the theory of everything (TOE), a manner of describing the known fundamental
forces and matter in a mathematically complete system. The theory has yet to make testable experimental
predictions, which a theory must do in order to be considered a part of science.
String theory mainly posits that the electrons and quarks within an atom are not 0-dimensional objects, but rather
1-dimensional oscillating lines ("strings"). The earliest string model, the bosonic string, incorporated only bosons,
although this view developed to the superstring theory, which posits that a connection (a "supersymmetry") exists
between bosons and fermions. String theories also require the existence of several extra, unobservable dimensions to
the universe, in addition to the usual four spacetime dimensions.
The theory has its origins in the dual resonance model (1969). Since that time, the term string theory has developed
to incorporate any of a group of related superstring theories. Five major string theories were formulated. The main
differences among them were the number of dimensions in which the strings developed and their characteristics. All
of them appeared to be correct, however. In the mid 1990s a unification of all previous superstring theories, called
M-theory, was proposed, which asserted that strings are really 1-dimensional slices of a 2-dimensional membrane
vibrating in 11-dimensional space.
As a result of the many properties and principles shared by these approaches (such as the holographic principle),
their mutual logical consistency, and the fact that some easily include the standard model of particle physics, some
mathematical physicists (i.e. Witten, Maldacena and Susskind) believe that string theory is a step towards the correct
fundamental description of nature.[2] [3] [4] [5] Nevertheless, other prominent physicists (e.g. Feynman and Glashow)

140

String Theory

141

have criticized string theory for not providing any quantitative experimental predictions.[6] [7]

Overview
String theory posits that the electrons and quarks within an atom are not 0-dimensional objects, but 1-dimensional
strings. These strings can move and vibrate, giving the observed particles their flavor, charge, mass and spin. String
theories also include objects more general than strings, called branes. The word brane, derived from "membrane",
refers to a variety of interrelated objects, such as D-branes, black p-branes and NeveuSchwarz 5-branes. These are
extended objects that are charged sources for differential form generalizations of the vector potential electromagnetic
field. These objects are related to one another by a variety of dualities. Black hole-like black p-branes are identified
with D-branes, which are endpoints for strings, and this identification is called Gauge-gravity duality. Research on
this equivalence has led to new insights on quantum chromodynamics, the fundamental theory of the strong nuclear
force.[8] [9] [10] [11] The strings make closed loops unless they encounter D-branes, where they can open up into
1-dimensional lines. The endpoints of the string cannot break off the D-brane, but they can slide around on it.
Since the string theory is widely believed to be a consistent theory of
quantum gravity, many hope that it correctly describes our universe,
making it a theory of everything. There are known configurations which
describe all the observed fundamental forces and matter but with a zero
cosmological constant and some new fields.[12] There are other
configurations with different values of the cosmological constant, which
are metastable but long-lived. This leads many to believe that there is at
least one metastable solution which is quantitatively identical with the
standard model, with a small cosmological constant, which contains
dark matter and a plausible mechanism for cosmic inflation. It is not yet
known whether string theory has such a solution, nor how much
freedom the theory allows to choose the details.
The full theory does not yet have a satisfactory definition in all
circumstances, since the scattering of strings is most straightforwardly
defined by a perturbation theory. The complete quantum mechanics of
high dimensional branes is not easily defined, and the behavior of string
theory in cosmological settings (time-dependent backgrounds) is not
fully worked out. It is also not clear if there is any principle by which
string theory selects its vacuum state, the spacetime configuration which
determines the properties of our universe (see string theory landscape).
As is the case for any other quantum theory of gravity, it is widely
believed that testing the theory directly would require prohibitively
expensive feats of engineering. Although direct experimental testing of
string theory involves grand explorations and development in
engineering, there are several indirect experiments that may prove
partial truth to string theory.

Basic properties

Levels of magnification:
1. Macroscopic level Matter
2. Molecular level
3. Atomic level Protons, neutrons, and
electrons
4. Subatomic level Electron
5. Subatomic level Quarks
6. String level

String Theory
String theory can be formulated in terms of an action principle, either the Nambu-Goto action or the Polyakov
action, which describes how strings move through space and time. In the absence of external interactions, string
dynamics are governed by tension and kinetic energy, which combine to produce oscillations. The quantum
mechanics of strings implies these oscillations take on discrete vibrational modes, the spectrum of the theory.
On distance scales larger than the string radius, each oscillation mode behaves as a different species of particle, with
its mass, spin and charge determined by the string's dynamics. Splitting and recombination of strings correspond to
particle emission and absorption, giving rise to the interactions between particles.
An analogy for strings' modes of vibration is a guitar string's production of multiple but distinct musical notes. In the
analogy, different notes correspond to different particles. The only difference is the guitar is only 2-dimensional; you
can strum it up, and down. In actuality the guitar strings would be every dimension, and the strings could vibrate in
any direction, meaning that the particles could move through not only our dimension, but other dimensions as well.
String theory includes both open strings, which have two distinct endpoints, and closed strings making a complete
loop. The two types of string behave in slightly different ways, yielding two different spectra. For example, in most
string theories, one of the closed string modes is the graviton, and one of the open string modes is the photon.
Because the two ends of an open string can always meet and connect, forming a closed string, there are no string
theories without closed strings.
The earliest string model, the bosonic string, incorporated only bosons. This model describes, in low enough
energies, a quantum gravity theory, which also includes (if open strings are incorporated as well) gauge fields such
as the photon (or, more generally, any gauge theory). However, this model has problems. Most importantly, the
theory has a fundamental instability, believed to result in the decay (at least partially) of spacetime itself.
Additionally, as the name implies, the spectrum of particles contains only bosons, particles which, like the photon,
obey particular rules of behavior. Roughly speaking, bosons are the constituents of radiation, but not of matter,
which is made of fermions. Investigating how a string theory may include fermions in its spectrum led to the
invention of supersymmetry, a mathematical relation between bosons and fermions. String theories which include
fermionic vibrations are now known as superstring theories; several different kinds have been described, but all are
now thought to be different limits of M-theory.
Some qualitative properties of quantum strings can be understood in a fairly simple fashion. For example, quantum
strings have tension, much like regular strings made of twine; this tension is considered a fundamental parameter of
the theory. The tension of a quantum string is closely related to its size. Consider a closed loop of string, left to move
through space without external forces. Its tension will tend to contract it into a smaller and smaller loop. Classical
intuition suggests that it might shrink to a single point, but this would violate Heisenberg's uncertainty principle. The
characteristic size of the string loop will be a balance between the tension force, acting to make it small, and the
uncertainty effect, which keeps it "stretched". Consequently, the minimum size of a string is related to the string
tension.

World-sheet
A point-like particle's motion may be described by drawing a graph of its position (in one or two dimensions of
space) against time. The resulting picture depicts the worldline of the particle (its 'history') in spacetime. By analogy,
a similar graph depicting the progress of a string as time passes by can be obtained; the string (a one-dimensional
object a small line by itself) will trace out a surface (a two-dimensional manifold), known as the worldsheet.
The different string modes (representing different particles, such as photon or graviton) are surface waves on this
manifold.
A closed string looks like a small loop, so its worldsheet will look like a pipe or, more generally, a Riemann surface
(a two-dimensional oriented manifold) with no boundaries (i.e. no edge). An open string looks like a short line, so its
worldsheet will look like a strip or, more generally, a Riemann surface with a boundary.

142

String Theory

Strings can split and connect. This is


reflected by the form of their worldsheet
(more accurately, by its topology). For
example, if a closed string splits, its
worldsheet will look like a single pipe
splitting (or connected) to two pipes (often
referred to as a pair of pants see drawing
at right). If a closed string splits and its two
parts later reconnect, its worldsheet will
look like a single pipe splitting to two and
then reconnecting, which also looks like a
torus connected to two pipes (one
Interaction in the subatomic world: world lines of point-like particles in the
representing the ingoing string, and the
Standard Model or a world sheet swept up by closed strings in string theory
other the outgoing one). An open string
doing the same thing will have its
worldsheet looking like a ring connected to two strips.
Note that the process of a string splitting (or strings connecting) is a global process of the worldsheet, not a local
one: locally, the worldsheet looks the same everywhere and it is not possible to determine a single point on the
worldsheet where the splitting occurs. Therefore these processes are an integral part of the theory, and are described
by the same dynamics that controls the string modes.
In some string theories (namely, closed strings in Type I and some versions of the bosonic string), strings can split
and reconnect in an opposite orientation (as in a Mbius strip or a Klein bottle). These theories are called unoriented.
Formally, the worldsheet in these theories is a non-orientable surface.

Dualities
Before the 1990s, string theorists believed there were five distinct superstring theories: open type I, closed type I,
closed type IIA, closed type IIB, and the two flavors of heterotic string theory (SO(32) and E8E8).[13] The thinking
was that out of these five candidate theories, only one was the actual correct theory of everything, and that theory
was the one whose low energy limit, with ten spacetime dimensions compactified down to four, matched the physics
observed in our world today. It is now believed that this picture was incorrect and that the five superstring theories
are connected to one another as if they are each a special case of some more fundamental theory (thought to be
M-theory). These theories are related by transformations that are called dualities. If two theories are related by a
duality transformation, it means that the first theory can be transformed in some way so that it ends up looking just
like the second theory. The two theories are then said to be dual to one another under that kind of transformation. Put
differently, the two theories are mathematically different descriptions of the same phenomena.
These dualities link quantities that were also thought to be separate. Large and small distance scales, as well as
strong and weak coupling strengths, are quantities that have always marked very distinct limits of behavior of a
physical system in both classical field theory and quantum particle physics. But strings can obscure the difference
between large and small, strong and weak, and this is how these five very different theories end up being related.
T-duality relates the large and small distance scales between string theories, whereas S-duality relates strong and
weak coupling strengths between string theories. U-duality links T-duality and S-duality.

143

String Theory

144

String theories
Type

Spacetime
dimensions

Details

Bosonic

26

Only bosons, no fermions, meaning only forces, no matter, with both open and closed strings; major flaw: a particle
with imaginary mass, called the tachyon, representing an instability in the theory.

10

Supersymmetry between forces and matter, with both open and closed strings; no tachyon; group symmetry is
SO(32)

IIA

10

Supersymmetry between forces and matter, with only closed strings bound to D-branes; no tachyon; massless
fermions are non-chiral

IIB

10

Supersymmetry between forces and matter, with only closed strings bound to D-branes; no tachyon; massless
fermions are chiral

HO

10

Supersymmetry between forces and matter, with closed strings only; no tachyon; heterotic, meaning right moving
and left moving strings differ; group symmetry is SO(32)

HE

10

Supersymmetry between forces and matter, with closed strings only; no tachyon; heterotic, meaning right moving
and left moving strings differ; group symmetry is E8E8

Note that in the type IIA and type IIB string theories closed strings are allowed to move everywhere throughout the
ten-dimensional spacetime (called the bulk), while open strings have their ends attached to D-branes, which are
membranes of lower dimensionality (their dimension is odd 1, 3, 5, 7 or 9 in type IIA and even 0, 2, 4, 6 or
8 in type IIB, including the time direction).

Extra dimensions
Number of dimensions
An intriguing feature of string theory is that it involves the prediction of extra dimensions. The number of
dimensions is not fixed by any consistency criterion, but flat spacetime solutions do exist in the so-called "critical
dimension". Cosmological solutions exist in a wider variety of dimensionalities, and these different
dimensionsmore precisely different values of the "effective central charge", a count of degrees of freedom which
reduces to dimensionality in weakly curved regimesare related by dynamical transitions.[14]
One such theory is the 11-dimensional M-theory, which requires spacetime to have eleven dimensions,[15] as
opposed to the usual three spatial dimensions and the fourth dimension of time. The original string theories from the
1980s describe special cases of M-theory where the eleventh dimension is a very small circle or a line, and if these
formulations are considered as fundamental, then string theory requires ten dimensions. But the theory also describes
universes like ours, with four observable spacetime dimensions, as well as universes with up to 10 flat space
dimensions, and also cases where the position in some of the dimensions is not described by a real number, but by a
completely different type of mathematical quantity. So the notion of spacetime dimension is not fixed in string
theory: it is best thought of as different in different circumstances.[16]
Nothing in Maxwell's theory of electromagnetism or Einstein's theory of relativity makes this kind of prediction;
these theories require physicists to insert the number of dimensions "by both hands", and this number is fixed and
independent of potential energy. String theory allows one to relate the number of dimensions to scalar potential
energy. Technically, this happens because a gauge anomaly exists for every separate number of predicted
dimensions, and the gauge anomaly can be counteracted by including nontrivial potential energy into equations to
solve motion. Furthermore, the absence of potential energy in the "critical dimension" explains why flat spacetime
solutions are possible.
This can be better understood by noting that a photon included in a consistent theory (technically, a particle carrying
a force related to an unbroken gauge symmetry) must be massless. The mass of the photon which is predicted by

String Theory
string theory depends on the energy of the string mode which represents the photon. This energy includes a
contribution from the Casimir effect, namely from quantum fluctuations in the string. The size of this contribution
depends on the number of dimensions since for a larger number of dimensions; there are more possible fluctuations
in the string position. Therefore, the photon in flat spacetime will be masslessand the theory consistentonly for a
particular number of dimensions.[17] When the calculation is done, the critical dimensionality is not four as one may
expect (three axes of space and one of time). The subset of X is equal to the relation of photon fluctuations in a linear
dimension. Flat space string theories are 26-dimensional in the bosonic case, while superstring and M-theories turn
out to involve 10 or 11 dimensions for flat solutions. In bosonic string theories, the 26 dimensions come from the
Polyakov equation.[18] Starting from any dimension greater than four, it is necessary to consider how these are
reduced to four dimensional spacetime.
Compact dimensions
Two different ways have been proposed to resolve this apparent
contradiction. The first is to compactify the extra dimensions; i.e.,
the 6 or 7 extra dimensions are so small as to be undetectable by
present day experiments.
To retain a high degree of supersymmetry, these compactification
spaces must be very special, as reflected in their holonomy. A
6-dimensional manifold must have SU(3) structure, a particular
case (torsionless) of this being SU(3) holonomy, making it a
CalabiYau space, and a 7-dimensional manifold must have G2
structure, with G2 holonomy again being a specific, simple, case.
Such spaces have been studied in attempts to relate string theory to
the 4-dimensional Standard Model, in part due to the
computational simplicity afforded by the assumption of
CalabiYau manifold (3D projection)
supersymmetry. More recently, progress has been made
constructing more realistic compactifications without the degree of symmetry of CalabiYau or G2 manifolds.
A standard analogy for this is to consider multidimensional space as a garden hose. If the hose is viewed from a
sufficient distance, it appears to have only one dimension, its length. Indeed, think of a ball just small enough to
enter the hose. Throwing such a ball inside the hose, the ball would move more or less in one dimension; in any
experiment we make by throwing such balls in the hose, the only important movement will be one-dimensional, that
is, along the hose. However, as one approaches the hose, one discovers that it contains a second dimension, its
circumference. Thus, an ant crawling inside it would move in two dimensions (and a fly flying in it would move in
three dimensions). This "extra dimension" is only visible within a relatively close range to the hose, or if one "throws
in" small enough objects. Similarly, the extra compact dimensions are only "visible" at extremely small distances, or
by experimenting with particles with extremely small wavelengths (of the order of the compact dimension's radius),
which in quantum mechanics means very high energies (see wave-particle duality).
Brane-world scenario
Another possibility is that we are "stuck" in a 3+1 dimensional (i.e. three spatial dimensions plus the time
dimension) subspace of the full universe. If such sub-spacetimes are exceptional sets within a larger-dimensional
one, there typically exist properly localized matter and YangMills gauge fields.[19] These "exceptional sets" are
ubiquitous in CalabiYau n-folds and may be described as subspaces without local deformations, akin to a crease in
a sheet of paper or a crack in a crystal, the neighborhood of which is markedly different from the exceptional
subspace itself. However, until the work of Randall and Sundrum,[20] it was not known that gravity too can be
properly localized to a sub-spacetime; their proof that it can made such cosmological scenarios realistic. In addition,

145

String Theory
spacetime may well be stratified, containing strata of various dimensions so that we may be inhabiting a 3+1
dimensional stratum; such geometries occur naturally in CalabiYau compactifications.[21] Such sub-spacetimes are
supposed to be D-branes, hence such models are known as a brane-world scenarios.
Effect of the hidden dimensions
In either case, gravity acting in the hidden dimensions affects other non-gravitational forces such as
electromagnetism. In fact, Kaluza's early work demonstrated that general relativity in five dimensions actually
predicts the existence of electromagnetism. However, because of the nature of CalabiYau manifolds, no new forces
appear from the small dimensions, but their shape has a profound effect on how the forces between the strings appear
in our four-dimensional universe. In principle, therefore, it is possible to deduce the nature of those extra dimensions
by requiring consistency with the standard model, but this is not yet a practical possibility. It is also possible to
extract information regarding the hidden dimensions by precision tests of gravity, but so far these have only put
upper limitations on the size of such hidden dimensions.

D-branes
Another key feature of string theory is the existence of D-branes. These are membranes of different dimensionality
(anywhere from a zero dimensional membranewhich is in fact a point and up, including 2-dimensional
membranes, 3-dimensional volumes and so on).
D-branes are defined by the fact that worldsheet boundaries are attached to them. D-branes have mass, since they
emit and absorb closed strings which describe gravitons, and in superstring theories charge as well, since they
couple to open strings which describe gauge interactions.
From the point of view of open strings, D-branes are objects to which the ends of open strings are attached. The open
strings attached to a D-brane are said to "live" on it, and they give rise to gauge theories "living" on it (since one of
the open string modes is a gauge boson such as the photon). In the case of one D-brane there will be one type of a
gauge boson and we will have an Abelian gauge theory (with the gauge boson being the photon). If there are
multiple parallel D-branes there will be multiple types of gauge bosons, giving rise to a non-Abelian gauge theory.
D-branes are thus gravitational sources, on which a gauge theory "lives". This gauge theory is coupled to gravity
(which is said to exist in the bulk), so that normally each of these two different viewpoints is incomplete.

Gauge-gravity duality
Gauge-gravity duality is a conjectured duality between a quantum theory of gravity in certain cases and gauge theory
in a lower number of dimensions. This means that each predicted phenomenon and quantity in one theory has an
analogue in the other theory, with a "dictionary" translating from one theory to the other.

Description of the duality


In certain cases the gauge theory on the D-branes is decoupled from the gravity living in the bulk; thus open strings
attached to the D-branes are not interacting with closed strings. Such a situation is termed a decoupling limit.
In those cases, the D-branes have two independent alternative descriptions. As discussed above, from the point of
view of closed strings, the D-branes are gravitational sources, and thus we have a gravitational theory on spacetime
with some background fields. From the point of view of open strings, the physics of the D-branes is described by the
appropriate gauge theory. Therefore in such cases it is often conjectured that the gravitational theory on spacetime
with the appropriate background fields is dual (i.e. physically equivalent) to the gauge theory on the boundary of this
spacetime (since the subspace filled by the D-branes is the boundary of this spacetime). So far, this duality has not
been proven in any cases, so there is also disagreement among string theorists regarding how strong the duality
applies to various models.

146

String Theory

147

Examples and intuition


The best known example and the first one to be studied is the duality between Type IIB superstring on AdS5

S5 (a

product space of a five-dimensional Anti de Sitter space and a five-sphere) on one hand, and N = 4 supersymmetric
YangMills theory on the four-dimensional boundary of the Anti de Sitter space (either a flat four-dimensional
spacetime R3,1 or a three-sphere with time S3
R). This is known as the AdS/CFT correspondence,[22] [23] [24] [25]
a name often used for Gauge / gravity duality in general.
This duality can be thought of as follows: suppose there is a spacetime with a gravitational source, for example an
extremal black hole.[26] When particles are far away from this source, they are described by closed strings (i.e. a
gravitational theory, or usually supergravity). As the particles approach the gravitational source, they can still be
described by closed strings; alternatively, they can be described by objects similar to QCD strings,[27] [28] [29] which
are made of gauge bosons (gluons) and other gauge theory degrees of freedom.[30] So if one is able (in a decoupling
limit) to describe the gravitational system as two separate regions one (the bulk) far away from the source, and the
other close to the source then the latter region can also be described by a gauge theory on D-branes. This latter
region (close to the source) is termed the near-horizon limit, since usually there is an event horizon around (or at) the
gravitational source.
In the gravitational theory, one of the directions in spacetime is the radial direction, going from the gravitational
source and away (towards the bulk). The gauge theory lives only on the D-brane itself, so it does not include the
radial direction: it lives in a spacetime with one less dimension compared to the gravitational theory (in fact, it lives
on a spacetime identical to the boundary of the near-horizon gravitational theory). Let us understand how the two
theories are still equivalent:
The physics of the near-horizon gravitational theory involves only on-shell states (as usual in string theory), while
the field theory includes also off-shell correlation function. The on-shell states in the near-horizon gravitational
theory can be thought of as describing only particles arriving from the bulk to the near-horizon region and interacting
there between themselves. In the gauge theory these are "projected" onto the boundary, so that particles which arrive
at the source from different directions will be seen in the gauge theory as (off-shell) quantum fluctuations far apart
from each other, while particles arriving at the source from almost the same direction in space will be seen in the
gauge theory as (off-shell) quantum fluctuations close to each other. Thus the angle between the arriving particles in
the gravitational theory translates to the distance scale between quantum fluctuations in the gauge theory. The angle
between arriving particles in the gravitational theory is related to the radial distance from the gravitational source at
which the particles interact: the larger the angle, the closer the particles have to get to the source in order to interact
with each other. On the other hand, the scale of the distance between quantum fluctuations in a quantum field theory
is related (inversely) to the energy scale in this theory, so small radius in the gravitational theory translates to low
energy scale in the gauge theory (i.e. the IR regime of the field theory) while large radius in the gravitational theory
translates to high energy scale in the gauge theory (i.e. the UV regime of the field theory).
A simple example to this principle is that if in the gravitational theory there is a setup in which the dilaton field
(which determines the strength of the coupling) is decreasing with the radius, then its dual field theory will be
asymptotically free, i.e. its coupling will grow weaker in high energies.

String Theory

Contact with experiment


The mathematics of string theory may lead to new insights on quantum chromodynamics, a gauge theory which is
the fundamental theory of the strong nuclear force. To this end, it is hoped that a gravitational theory dual to
quantum chromodynamics will be found.[31]
A mathematical technique from string theory (the AdS/CFT correspondence) has been used to describe qualitative
features of quarkgluon plasma behavior in relativistic heavy-ion collisions;[8] [9] [10] [11] the physics, however, is
strictly that of standard quantum chromodynamics, which has been quantitatively modeled by lattice QCD methods
with good results.[32]
Other potential forms of evidence for string theory have been proposed. One would be the discovery of large cosmic
strings in space, formed when the Big Bang "stretched" some strings to astronomical proportions. Other theories,
however, predict cosmic strings with a different physical basis (topological defects in various fields). Novel
phenomena consistent with versions of string theory may be observed with the newly built Large Hadron Collider.
One would be indications of an anomalously large strength of gravity on a microscopic scale, which would be
consistent with branes interacting across extra dimensions through leakage of gravitons from one to the other. The
discovery of supersymmetry could also be considered evidence since string theory was the first theory to require it,
though other theories incorporate supersymmetry as well. The absence of supersymmetric particles at energies
accessible to the LHC would not necessarily disprove string theory, since supersymmetry could exist but still be
outside the accelerator's range.

Problems and controversy


Although string theory comes from physics, some say that string theory's currently untested status means that it
should be classified as more of a mathematical framework for building models as opposed to a physical theory.[33]
Some go further, and say that string theory as a theory of everything is a failure.[34] [35] This led to a public debate in
2007,[36] [37] with one journalist expressing this opinion:
"For more than a generation, physicists have been chasing a will-o-the-wisp called string theory. The
beginning of this chase marked the end of what had been three-quarters of a century of progress. Dozens of
string-theory conferences have been held, hundreds of new Ph.D.s have been minted, and thousands of papers
have been written. Yet, for all this activity, not a single new testable prediction has been made, not a single
theoretical puzzle has been solved. In fact, there is no theory so farjust a set of hunches and calculations
suggesting that a theory might exist. And, even if it does, this theory will come in such a bewildering number
of versions that it will be of no practical use: a Theory of Nothing."
Jim Holt[38]

Is string theory predictive?


String theory as a theory of everything has been criticized as unscientific because it is so difficult to test by
experiments. The controversy concerns two properties:
1. It is widely believed that any theory of quantum gravity would require extremely high energies to probe directly,
higher by orders of magnitude than those that current experiments such as the Large Hadron Collider[39] can
attain.
2. String theory as it is currently understood has a huge number of solutions, called string vacua,[40] and these vacua
might be sufficiently diverse to accommodate almost any phenomena we might observe at lower energies.
If these properties are true, string theory as a theory of everything would have little or no predictive power for low
energy particle physics experiments.[41] [42] Because the theory is so difficult to test, some theoretical physicists have
asked if it can even be called a scientific theory. Notable critics include Peter Woit, Lee Smolin, Philip Warren
Anderson,[43] Sheldon Glashow,[44] Lawrence Krauss,[45] and Carlo Rovelli.[46]

148

String Theory
All string theory models are quantum mechanical, Lorentz invariant, unitary and contain Einstein's General
Relativity as a low energy limit.[47] Therefore to falsify string theory, it would suffice to falsify quantum mechanics,
Lorentz invariance, or general relativity.[48] Hence string theory is falsifiable and meets the definition of scientific
theory according to the Popperian criterion. However to constitute a convincing potential verification of string
theory, a prediction should be specific to it, not shared by any quantum field theory model or by General Relativity.
One such unique prediction is string harmonics: at sufficiently high energiesprobably near the quantum gravity
scalethe string-like nature of particles would become obvious. There should be heavier copies of all particles
corresponding to higher vibrational states of the string. But it is not clear how high these energies are. In the most
likely case, they would be 1014 times higher than those accessible in the newest particle accelerator, the LHC,
making this prediction impossible to test with any particle accelerator in the foreseeable future.

Background independence
A separate and older criticism of string theory is that it is background-dependent string theory describes
perturbative expansions about fixed spacetime backgrounds. Although the theory has some
background-independence topology change is an established process in string theory, and the exchange of
gravitons is equivalent to a change in the background mathematical calculations in the theory rely on preselecting
a background as a starting point. This is because, like many quantum field theories, much of string theory is still only
formulated perturbatively, as a divergent series of approximations. Although nonperturbative techniques have
progressed considerably including conjectured complete definitions in spacetimes satisfying certain asymptotics
a full non-perturbative definition of the theory is still lacking. Some see background independence as a
fundamental requirement of a theory of quantum gravity, particularly since general relativity is already background
independent. Some hope that M-theory, or a non-perturbative treatment of string theory (string field theory was
thought to be non-perturbative in the 1980s) have a background-independent formulation.

Supersymmetry breaking
A central problem for applications is that the best understood backgrounds of string theory preserve much of the
supersymmetry of the underlying theory, which results in time-invariant spacetimes: currently string theory cannot
deal well with time-dependent, cosmological backgrounds. However, several models have been proposed to explain
supersymmetry breaking, most notably the KKLT model,[40] which incorporates branes and fluxes to make a
metastable compactification.

String theory landscape


The vacuum structure of the theory, called the string theory landscape (or the anthropic portion of string theory
vacua), is not well understood. String theory contains an infinite number of distinct meta-stable vacua, and perhaps
10500 of these or more correspond to a universe roughly similar to ours with four dimensions, a high planck scale,
gauge groups, and chiral fermions. Each of these corresponds to a different possible universe, with a different
collection of particles and forces.[40] What principle, if any, can be used to select among these vacua is an open
issue. While there are no continuous parameters in the theory, there is a very large set of possible universes, which
may be radically different from each other. It is also suggested that the landscape is surrounded by an even more vast
swampland of consistent-looking semiclassical effective field theories, which are actually inconsistent.
Some physicists believe this is a good thing, because it may allow a natural anthropic explanation of the observed
values of physical constants, in particular the small value of the cosmological constant.[49] [50] The argument is that
most universes contain values for physical constants which do not lead to habitable universes (at least for humans),
and so we happen to live in the most "friendly" universe. This principle is already employed to explain the existence
of life on earth as the result of a life-friendly orbit around the medium-sized sun among an infinite number of
possible orbits (as well as a relatively stable location in the galaxy). However, the cosmological version of the

149

String Theory
anthropic principle remains highly controversial because it would be difficult if not impossible to Popper falsify; so
many do not accept it as scientific.

Other testability criteria


Many physicists strongly oppose the idea that string theory is not falsifiable, among them Sylvester James Gates:
"So, the next time someone tells you that string theory is not testable, remind them of the AdS/CFT connection
..."[51] AdS/CFT relates string theory to gauge theory, and allows contact with low energy experiments in quantum
chromodynamics. This type of string theory, which only describes the strong interactions, is much less controversial
today than string theories of everything (although two decades ago, it was the other way around).
In addition, Gates points out that the grand unification natural in string theories of everything requires that the
coupling constants of the four forces meet at one point under renormalization group rescaling. This is also a
falsifiable statement, but it is not restricted to string theory, but is shared by grand unified theories.[52] The LHC will
be used both for testing AdS/CFT, and to check if the electroweakstrong unification does happen as predicted.[53]

History
Some of the structures reintroduced by string theory arose for the first time much earlier as part of the program of
classical unification started by Albert Einstein. The first person to add a fifth dimension to general relativity was
German mathematician Theodor Kaluza in 1919, who noted that gravity in five dimensions describes both gravity
and electromagnetism in four. In 1926, the Swedish physicist Oskar Klein gave a physical interpretation of the
unobservable extra dimension--- it is wrapped into a small circle. Einstein introduced a non-symmetric metric tensor,
while much later Brans and Dicke added a scalar component to gravity. These ideas would be revived within string
theory, where they are demanded by consistency conditions.
String theory was originally developed during the late 1960s and early 1970s as a never completely successful theory
of hadrons, the subatomic particles like the proton and neutron which feel the strong interaction. In the 1960s,
Geoffrey Chew and Steven Frautschi discovered that the mesons make families called Regge trajectories with
masses related to spins in a way that was later understood by Yoichiro Nambu, Holger Bech Nielsen and Leonard
Susskind to be the relationship expected from rotating strings. Chew advocated making a theory for the interactions
of these trajectories which did not presume that they were composed of any fundamental particles, but would
construct their interactions from self-consistency conditions on the S-matrix. The S-matrix approach was started by
Werner Heisenberg in the 1940s as a way of constructing a theory which did not rely on the local notions of space
and time, which Heisenberg believed break down at the nuclear scale. While the scale was off by many orders of
magnitude, the approach he advocated was ideally suited for a theory of quantum gravity.
Working with experimental data, R. Dolen, D. Horn and C. Schmid[54] developed some sum rules for hadron
exchange. When a particle and antiparticle scatter, virtual particles can be exchanged in two qualitatively different
ways. In the s-channel, the two particles annihilate to make temporary intermediate states which fall apart into the
final state particles. In the t-channel, the particles exchange intermediate states by emission and absorption. In field
theory, the two contributions add together, one giving a continuous background contribution, the other giving peaks
at certain energies. In the data, it was clear that the peaks were stealing from the background--- the authors
interpreted this as saying that the t-channel contribution was dual to the s-channel one, meaning both described the
whole amplitude and included the other.
The result was widely advertised by Murray Gell-Mann, leading Gabriele Veneziano to construct a scattering
amplitude which had the property of Dolen-Horn-Schmid duality, later renamed world-sheet duality. The amplitude
needed poles where the particles appear, on straight line trajectories, and there is a special mathematical function
whose poles are evenly spaced on half the real line--- the Gamma function--- which was widely used in Regge
theory. By manipulating combinations of Gamma functions, Veneziano was able to find a consistent scattering
amplitude with poles on straight lines, with mostly positive residues, which obeyed duality and had the appropriate

150

String Theory
Regge scaling at high energy. The amplitude could fit near-beam scattering data as well as other Regge type fits, and
had a suggestive integral representation which could be used for generalization.
Over the next years, hundreds of physicists worked to complete the bootstrap program for this model, with many
surprises. Veneziano himself discovered that for the scattering amplitude to describe the scattering of a particle
which appears in the theory, an obvious self-consistency condition, the lightest particle must be a tachyon. Miguel
Virasoro and Joel Shapiro found a different amplitude now understood to be that of closed strings, while Ziro Koba
and Holger Nielsen generalized Veneziano's integral representation to multiparticle scattering. Veneziano and Sergio
Fubini introduced an operator formalism for computing the scattering amplitudes which was a forerunner of
world-sheet conformal theory, while Virasoro understood how to remove the poles with wrong-sign residues using a
constraint on the states. Claud Lovelace calculated a loop amplitude, and noted that there is an inconsistency unless
the dimension of the theory is 26. Charles Thorn, Peter Goddard and Richard Brower went on to prove that there are
no wrong-sign propagating states in dimensions less than or equal to 26.
In 1969 Yoichiro Nambu, Holger Bech Nielsen and Leonard Susskind recognized that the theory could be given a
description in space and time in terms of strings. The scattering amplitudes were derived systematically from the
action principle by Peter Goddard, Jeffrey Goldstone, Claudio Rebbi and Charles Thorn, giving a space-time picture
to the vertex operators introduced by Veneziano and Fubini and a geometrical interpretation to the Virasoro
conditions.
In 1970, Pierre Ramond added fermions to the model, which led him to formulate a two-dimensional supersymmetry
to cancel the wrong-sign states. John Schwarz and Andr Neveu added another sector to the fermi theory a short time
later. In the fermion theories, the critical dimension was 10. Stanley Mandelstam formulated a world sheet conformal
theory for both the bose and fermi case, giving a two-dimensional field theoretic path-integral to generate the
operator formalism. Michio Kaku and Keiji Kikkawa gave a different formulation of the bosonic string, as a string
field theory, with infinitely many particle types and with fields taking values not on points, but on loops and curves.
In 1974, Tamiaki Yoneya discovered that all the known string theories included a massless spin-two particle which
obeyed the correct Ward identities to be a graviton. John Schwarz and Joel Scherk came to the same conclusion and
made the bold leap to suggest that string theory was a theory of gravity, not a theory of hadrons. They reintroduced
KaluzaKlein theory as a way of making sense of the extra dimensions. At the same time, quantum chromodynamics
was recognized as the correct theory of hadrons, shifting the attention of physicists and apparently leaving the
bootstrap program in the dustbin of history.
String theory eventually made it out of the dustbin, but for the following decade all work on the theory was
completely ignored. Still, the theory continued to develop at a steady pace thanks the work of a handful of devotees.
Ferdinando Gliozzi, Joel Scherk, and David Olive realized in 1976 that the original Ramond and Neveu
Schwarz-strings were separately inconsistent and needed to be combined. The resulting theory did not have a
tachyon, and was proven to have space-time supersymmetry by John Schwarz and Michael Green in 1981. The same
year, Alexander Polyakov gave the theory a modern path integral formulation, and went on to develop conformal
field theory extensively. In 1979, Daniel Friedan showed that the equations of motions of string theory, which are
generalizations of the Einstein equations of General Relativity, emerge from the Renormalization group equations
for the two-dimensional field theory. Schwarz and Green discovered T-duality, and constructed two different
superstring theories--- IIA and IIB related by T-duality, and type I theories with open strings. The consistency
conditions had been so strong, that the entire theory was nearly uniquely determined, with only a few discrete
choices.
In the early 1980s, Edward Witten discovered that most theories of quantum gravity could not accommodate chiral
fermions like the neutrino. This led him, in collaboration with Luis Alvarez-Gaum to study violations of the
conservation laws in gravity theories with anomalies, concluding that type I string theories were inconsistent. Green
and Schwarz discovered a contribution to the anomaly that Witten and Alvarez-Gaum had missed, which restricted
the gauge group of the type I string theory to be SO(32). In coming to understand this calculation, Edward Witten

151

String Theory
became convinced that string theory was truly a consistent theory of gravity, and he became a high-profile advocate.
Following Witten's lead, between 1984 and 1986, hundreds of physicists started to work in this field, and this is
sometimes called the first superstring revolution.
During this period, David Gross, Jeffrey Harvey, Emil Martinec, and Ryan Rohm discovered heterotic strings. The
gauge group of these closed strings was two copies of E8, and either copy could easily and naturally include the
standard model. Philip Candelas, Gary Horowitz, Andrew Strominger and Edward Witten found that the Calabi-Yau
manifolds are the compactifications which preserve a realistic amount of supersymmetry, while Lance Dixon and
others worked out the physical properties of orbifolds, distinctive geometrical singularities allowed in string theory.
Cumrun Vafa generalized T-duality from circles to arbitrary manifolds, creating the mathematical field of mirror
symmetry. David Gross and Vipul Periwal discovered that string perturbation theory was divergent in a way that
suggested that new non-perturbative objects were missing.
In the 1990s, Joseph Polchinski discovered that the theory requires higher-dimensional objects, called D-branes and
identified these with the black-hole solutions of supergravity. These were understood to be the new objects suggested
by the perturbative divergences, and they opened up a new field with rich mathematical structure. It quickly became
clear that D-branes and other p-branes, not just strings, formed the matter content of the string theories, and the
physical interpretation of the strings and branes was revealed--- they are a type of black hole. Leonard Susskind had
incorporated the holographic principle of Gerardus 't Hooft into string theory, identifying the long highly-excited
string states with ordinary thermal black hole states. As suggested by 't Hooft, the fluctuations of the black hole
horizon, the world-sheet or world-volume theory, describes not only the degrees of freedom of the black hole, but all
nearby objects too.
In 1995, at the annual conference of string theorists at the University of Southern California (USC), Edward Witten
gave a speech on string theory that essentially united the five string theories that existed at the time, and giving birth
to a new 11-dimensional theory called M-theory. M-theory was also foreshadowed in the work of Paul Townsend at
approximately the same time. The flurry of activity which began at this time is sometimes called the second
superstring revolution.
During this period, Tom Banks, Willy Fischler Stephen Shenker and Leonard Susskind formulated a full holographic
description of M-theory on IIA D0 branes,[55] the first definition of string theory that was fully non-perturbative and
a concrete mathematical realization of the holographic principle. Andrew Strominger and Cumrun Vafa calculated
the entropy of certain configurations of D-branes and found agreement with the semi-classical answer for extreme
charged black holes. Petr Horava and Edward Witten found the eleven-dimensional formulation of the heterotic
string theories, showing that orbifolds solve the chirality problem. Witten noted that the effective description of the
physics of D-branes at low energies is by a supersymmetric gauge theory, and found geometrical interpretations of
mathematical structures in gauge theory that he and Nathan Seiberg had earlier discovered in terms of the location of
the branes.
In 1997 Juan Maldacena noted that the low energy excitations of a theory near a black hole consist of objects close to
the horizon, which for extreme charged black holes looks like an anti de Sitter space. He noted that in this limit the
gauge theory describes the string excitations near the branes. So he hypothesized that string theory on a near-horizon
extreme-charged black-hole geometry, an anti-deSitter space times a sphere with flux, is equally well described by
the low-energy limiting gauge theory, the N=4 supersymmetric Yang-Mills theory. This hypothesis, which is called
the AdS/CFT correspondence, was further developed by Steven Gubser, Igor Klebanov and Alexander Polyakov,
and by Edward Witten, and it is now well-accepted. It is a concrete realization of the holographic principle, which
has far-reaching implications for black holes, locality and information in physics, as well as the nature of the
gravitational interaction. Through this relationship, string theory has been shown to be related to gauge theories like
quantum chromodynamics and this has led to more quantitative understanding of the behavior of hadrons, bringing
string theory back to its roots.

152

String Theory

References
[1] Sunil Mukhi(1999)" The Theory of Strings: A Detailed Introduction (http:/ / theory. tifr. res. in/ ~mukhi/ Physics/ string2. html)"
[2] Joseph Polchinski, "All Strung Out?", American Scientist, January-February 2007 Volume 95, Number 1 (http:/ / www. americanscientist.
org/ bookshelf/ pub/ all-strung-out)
[3] "On the right track. Interview with Professor Edward Witten. ", Frontline, Volume 18 - Issue 03, Feb. 03 - 16, 2001 (http:/ / www.
hinduonnet. com/ fline/ fl1803/ 18030830. htm)
[4] Leonard Susskind, "Hold fire! This epic vessel has only just set sail...", Times Higher Education, 25 August 2006 (http:/ / www.
timeshighereducation. co. uk/ story. asp?storyCode=204991& sectioncode=26)
[5] Geoff Brumfiel, "Our Universe: Outrageous fortune", Nature, Nature 439, 10-12 (5 January 2006) | doi:10.1038/439010a; Published online 4
January 2006 (http:/ / www. nature. com/ nature/ journal/ v439/ n7072/ full/ 439010a. html)
[6] "NOVA - The elegant Universe" (http:/ / www. pbs. org/ wgbh/ nova/ elegant/ view-glashow. html)
[7] Jim Holt, "Unstrung", The New Yorker, October 2, 2006 (http:/ / www. newyorker. com/ archive/ 2006/ 10/ 02/ 061002crat_atlarge)
[8] H. Nastase, More on the RHIC fireball and dual black holes, BROWN-HET-1466, arXiv:hep-th/0603176, March 2006,
[9] H. Liu, K. Rajagopal, U. A. Wiedemann, An AdS/CFT Calculation of Screening in a Hot Wind, MIT-CTP-3757, arXiv:hep-ph/0607062 July
2006,
[10] H. Liu, K. Rajagopal, U. A. Wiedemann, Calculating the Jet Quenching Parameter from AdS/CFT, Phys.Rev.Lett.97:182301,2006
arXiv:hep-ph/0605178
[11] H. Nastase, The RHIC fireball as a dual black hole, BROWN-HET-1439, arXiv:hep-th/0501068, January 2005,
[12] Burt A. Ovrut (2006). "A Heterotic Standard Model". Fortschritte der Physik 54-(2-3): 160164.
[13] S. James Gates, Jr., Ph.D., Superstring Theory: The DNA of Reality (http:/ / www. teach12. com/ ttcx/ coursedesclong2. aspx?cid=1284)
"Lecture 23 Can I Have That Extra Dimension in the Window?", 0:04:54, 0:21:00.
[14] Simeon Hellerman and Ian Swanson(2006): " Dimension-changing exact solutions of string theory (http:/ / arxiv. org/ abs/ hep-th/
0612051v3)".; Ofer Aharony and Eva Silverstein(2006):" Supercritical stability, transitions and (pseudo)tachyons (http:/ / arxiv. org/ abs/
hep-th/ 0612031v2)".
[15] M. J. Duff, James T. Liu and R. Minasian Eleven Dimensional Origin of String/String Duality: A One Loop Test (http:/ / arxiv. org/ abs/
hep-th/ 9506126v2) Center for Theoretical Physics, Department of Physics, Texas A&M University
[16] Polchinski, Joseph (1998). String Theory, Cambridge University Press.
[17] The calculation of the number of dimensions can be circumvented by adding a degree of freedom which compensates for the "missing"
quantum fluctuations. However, this degree of freedom behaves similar to spacetime dimensions only in some aspects, and the produced
theory is not Lorentz invariant, and has other characteristics which don't appear in nature. This is known as the linear dilaton or non-critical
string.
[18] "Quantum Geometry of Bosonic Strings Revisited" (ftp:/ / ftp2. biblioteca. cbpf. br/ pub/ apub/ 1999/ nf/ nf_zip/ nf04299. pdf)
[19] See, for example, T. Hbsch, " A Hitchhikers Guide to Superstring Jump Gates and Other Worlds (http:/ / homepage. mac. com/ thubsch/
HSProc. pdf)", in Proc. SUSY 96 Conference, R. Mohapatra and A. Rasin (eds.), Nucl. Phys. (Proc. Supl.) 52A (1997) 347351
[20] L. Randall and R. Sundrum, " An Alternative to compactification (http:/ / arXiv. org/ pdf/ hep-th/ 9906064)" Phys. Rev. Lett. 83 (1999)
46904693
[21] P. Aspinwall, D. Morrison and B. Greene, " CalabiYau moduli space, mirror manifolds and space-time topology change in string theory
(http:/ / arXiv. org/ pdf/ hep-th/ 9309097)", Nucl. Phys. B416 (1994) 414480
[22] J. Maldacena, The Large N Limit of Superconformal Field Theories and Supergravity, arXiv:hep-th/9711200
[23] S. S. Gubser, I. R. Klebanov and A. M. Polyakov (1998). "Gauge theory correlators from non-critical string theory". Physics Letters B428:
105114. arXiv:hep-th/9802109.
[24] Edward Witten (1998). "Anti-de Sitter space and holography". Advances in Theoretical and Mathematical Physics 2: 253291.
arXiv:hep-th/9802150.
[25] Aharony, O.; S.S. Gubser, J. Maldacena, H. Ooguri, Y. Oz (2000). "Large N Field Theories, String Theory and Gravity". Phys. Rept. 323
(3-4): 183386. arXiv:hep-th/9905111. doi:10.1016/S0370-1573(99)00083-6..
[26] Robbert Dijkgraaf,Erik Verlinde and Herman Verlinde(1997)" 5D Black Holes and Matrix Strings (http:/ / arxiv. org/ abs/ hep-th/
9704018v2)"
[27] Minoru Eto,Koji Hashimoto and Seiji Terashima(2007)" QCD String as Vortex String in Seiberg-Dual Theory (http:/ / arxiv. org/ abs/ 0706.
2005v1)"
[28] Harvey B.Meyer(2005)" Vortices on the worldsheet of the QCD string (http:/ / arxiv. org/ abs/ hep-th/ 0506034v1)"
[29] Koji Hashimoto(2007)" Cosmic Strings, Strings and D-branes (http:/ / www2. yukawa. kyoto-u. ac. jp/ ~kiasyk2/ slides/ hashimoto.
pdf)"
[30] Piljin Yi(2007)" Story of baryons in a gravity dual of QCD (http:/ / www2. yukawa. kyoto-u. ac. jp/ ~gc2007/ pdf/ yi. pdf)"
[31] For example: T. Sakai and S. Sugimoto, Low energy hadron physics in holographic QCD, Prog.Theor.Phys.113:843882,2005,
arXiv:hep-th/0412141, December 2004
[32] See for example Recent Results (http:/ / www. physics. indiana. edu/ ~sg/ milc/ results. pdf) of the MILC research program, taken from the
MILC Collaboration homepage (http:/ / www. physics. indiana. edu/ ~sg/ milc. html)

153

String Theory
[33] David Gross, Perspectives (http:/ / video. tau. ac. il/ Lectures/ Exact_Sciences/ Physics/ stringfest/ OnDemand. html?11), String Theory:
Achievements and Perspectives - A conference
[34] Peter Woit's Not Even Wrong weblog (http:/ / www. math. columbia. edu/ ~woit/ wordpress/ ?cat=2)
[35] Lee Smolin's The Trouble With Physics webpage (http:/ / www. thetroublewithphysics. com)
[36] e.g. John Baez and responses on the group weblog The n-Category Cafe (http:/ / golem. ph. utexas. edu/ category/ 2007/ 02/
this_weeks_finds_in_mathematic_7. html)
[37] John Baez weblog (http:/ / math. ucr. edu/ home/ baez/ week246. html)
[38] Unstrung: The New Yorker (http:/ / www. newyorker. com/ archive/ 2006/ 10/ 02/ 061002crat_atlarge)
[39] Elias Kiritsis(2007)" String Theory in a Nutshell (http:/ / press. princeton. edu/ chapters/ s8456. pdf)"
[40] S. Kachru, R. Kallosh, A. Linde and S. P. Trivedi, de Sitter Vacua in String Theory, Phys.Rev.D68:046005,2003, arXiv:hep-th/0301240
[41] P. Woit (Columbia University), String theory: An Evaluation,February 2001, arXiv:physics/0102051
[42] P. Woit, Is String Theory Testable? (http:/ / www. math. columbia. edu/ ~woit/ testable. pdf) INFN Rome March 2007
[43] "string theory is the first science in hundreds of years to be pursued in pre-Baconian fashion, without any adequate experimental guidance",
New York Times, 4 January 2005
[44] "there ain't no experiment that could be done nor is there any observation that could be made that would say, `You guys are wrong.' The
theory is safe, permanently safe" NOVA interview (http:/ / pbs. org/ wgbh/ nova/ elegant/ view-glashow. html))
[45] "String theory [is] yet to have any real successes in explaining or predicting anything measurable" New York Times, 8 November 2005)
[46] see his Dialog on Quantum Gravity, arXiv:hep-th/0310077)
[47] J. Polchinski, String Theory, Cambridge University Press, Cambridge, UK (1998)
[48] N. Comins, W. Kaufmann, Discovering the Universe: From the Stars to the Planets, W.H. Freeman & Co., p. 357 (2008)
[49] N. Arkani-Hamed, S. Dimopoulos and S. Kachru, Predictive Landscapes and New Physics at a TeV, arXiv:hep-th/0501082,
SLAC-PUB-10928, HUTP-05-A0001, SU-ITP-04-44, January 2005
[50] L. Susskind The Anthropic Landscape of String Theory, arXiv:hep-th/0302219, February 2003
[51] S. James Gates, Jr., Ph.D., Superstring Theory: The DNA of Reality "Lecture 21 - Can 4D Forces (without Gravity) Love Strings?",
0:26:06-0:26:21, cf. 0:24:05-0:26-24.
[52] Idem, "Lecture 19 - Do-See-Do and Swing your Superpartner Part II" 0:16:05-0:24:29.
[53] Idem, Lecture 21, 0:20:10-0:21:20.
[54] Dolen, Horn, Schmid "Finite-Energy Sum Rules and Their Application to N Charge Exchange" http:/ / prola. aps. org/ abstract/ PR/ v166/
i5/ p1768_1
[55] Banks, Fischler, Shenker and Susskind "M Theory As A Matrix Model: A Conjecture" http:/ / arxiv. org/ abs/ hep-th/ 9610043v3

Further reading
Popular books and articles
Davies, Paul; Julian R. Brown (Eds.) (July 31 1992). Superstrings: A Theory of Everything? (Reprint ed.).
Cambridge: Cambridge University Press. p.244. ISBN0-521-43775-X.
Gefter, Amanda (December 2005). "Is string theory in trouble?" (http://www.newscientist.com/article/
mg18825305.800-is-string-theory-in-trouble.html?full=true). New Scientist. Retrieved December 19, 2005.
An interview with Leonard Susskind, the theoretical physicist who discovered that string theory is based on
one-dimensional objects and now is promoting the idea of multiple universes.
Green, Michael (September 1986). "Superstrings" (http://www.damtp.cam.ac.uk/user/mbg15/superstrings/
superstrings.html). Scientific American. Retrieved December 19, 2005.
Greene, Brian (October 20 2003). The Elegant Universe: Superstrings, Hidden Dimensions, and the Quest for the
Ultimate Theory (Reissue ed.). New York: W.W. Norton & Company. p.464. ISBN0-393-05858-1.
Greene, Brian (2004). The Fabric of the Cosmos: Space, Time, and the Texture of Reality. New York: Alfred A.
Knopf. p.569. ISBN0-375-41288-3.
Gribbin, John (1998). The Search for Superstrings, Symmetry, and the Theory of Everything. London: Little
Brown and Company. p.224. ISBN0-316-32975-4.
Halpern, Paul (2004). The Great Beyond: Higher Dimensions, Parallel Universes, and the Extraordinary Search
for a Theory of Everything. Hoboken, New Jersey: John Wiley & Sons, Inc.. p.326. ISBN0-471-46595-X.
Hooper, Dan (2006). Dark Cosmos: In Search of Our Universe's Missing Mass and Energy. New York:
HarperCollins. p.240. ISBN978-0-06-113032-8.

154

String Theory
Kaku, Michio (April 1994). Hyperspace: A Scientific Odyssey Through Parallel Universes, Time Warps, and the
Tenth Dimension. Oxford: Oxford University Press. p.384. ISBN0-19-508514-0.
Klebanov, Igor and Maldacena, Juan (January 2009). Solving Quantum Field Theories via Curved Spacetimes
(http://ptonline.aip.org/journals/doc/PHTOAD-ft/vol_62/iss_1/28_1.shtml). Physics Today.
Musser, George (2008). The Complete Idiot's Guide to String Theory. Indianapolis: Alpha. p.368.
ISBN978-1-59-257702-6.
Randall, Lisa (September 1 2005). Warped Passages: Unraveling the Mysteries of the Universe's Hidden
Dimensions. New York: Ecco Press. p.512. ISBN0-06-053108-8.
Susskind, Leonard (December 2006). The Cosmic Landscape: String Theory and the Illusion of Intelligent
Design. New York: Hachette Book Group/Back Bay Books. p.403. ISBN0-316-01333-1.
Taubes, Gary (November 1986). "Everything's Now Tied to Strings" Discover Magazine vol 7, #11. (Popular
article, probably the first ever written, on the first superstring revolution.)
Vilenkin, Alex (2006). Many Worlds in One: The Search for Other Universes. New York: Hill and Wang. p.235.
ISBN0-8090-9523-8.
Witten, Edward (June 2002). "The Universe on a String" (http://www.sns.ias.edu/~witten/papers/string.pdf)
(PDF). Astronomy Magazine. Retrieved December 19, 2005. An easy nontechnical article on the very basics of
the theory.
Two nontechnical books that are critical of string theory:
Smolin, Lee (2006). The Trouble with Physics: The Rise of String Theory, the Fall of a Science, and What Comes
Next. New York: Houghton Mifflin Co.. p.392. ISBN0-618-55105-0.
Woit, Peter (2006). Not Even Wrong - The Failure of String Theory And the Search for Unity in Physical Law.
London: Jonathan Cape &: New York: Basic Books. p.290. ISBN0-224-07605-1 & ISBN 978-0-465-09275-8.

Textbooks
Becker, Katrin, Becker, Melanie, and John H. Schwarz (2007) String Theory and M-Theory: A Modern
Introduction . Cambridge University Press. ISBN 0-521-86069-5
Bintruy, Pierre (2007) Supersymmetry: Theory, Experiment, and Cosmology. Oxford University Press. ISBN
978-0-19-850954-7.
Dine, Michael (2007) Supersymmetry and String Theory: Beyond the Standard Model. Cambridge University
Press. ISBN 0-521-85841-0.
Paul H. Frampton (1974). Dual Resonance Models. Frontiers in Physics. ISBN0-805-32581-6.
Gasperini, Maurizio (2007) Elements of String Cosmology. Cambridge University Press. ISBN
978-0-521-86875-4.
Michael Green, John H. Schwarz and Edward Witten (1987) Superstring theory. Cambridge University Press. The
original textbook.
Vol. 1: Introduction. ISBN 0-521-35752-7.
Vol. 2: Loop amplitudes, anomalies and phenomenology. ISBN 0-521-35753-5.
Kiritsis, Elias (2007) String Theory in a Nutshell. Princeton University Press. ISBN 978-0-691-12230-4.
Johnson, Clifford (2003). D-branes. Cambridge: Cambridge University Press. ISBN0-521-80912-6.
Polchinski, Joseph (1998) String Theory. Cambridge University Press.
Vol. 1: An introduction to the bosonic string. ISBN 0-521-63303-6.
Vol. 2: Superstring theory and beyond. ISBN 0-521-63304-4.
Szabo, Richard J. (Reprinted 2007) An Introduction to String Theory and D-brane Dynamics. Imperial College
Press. ISBN 978-1-86094-427-7.
Zwiebach, Barton (2004) A First Course in String Theory. Cambridge University Press. ISBN 0-521-83143-1.
Contact author for errata.

155

String Theory
Technical and critical:
Penrose, Roger (February 22 2005). The Road to Reality: A Complete Guide to the Laws of the Universe. Knopf.
p.1136. ISBN0-679-45443-8.

Online material
Schwarz, John H.. "Introduction to Superstring Theory". arXiv:hep-ex/0008017. Four lectures, presented at the
NATO Advanced Study Institute on Techniques and Concepts of High Energy Physics, St. Croix, Virgin Islands,
in June 2000, and addressed to an audience of graduate students in experimental high energy physics, survey
basic concepts in string theory.
Witten, Edward (1998). "Duality, Spacetime and Quantum Mechanics" (http://online.itp.ucsb.edu/online/
plecture/witten/). Kavli Institute for Theoretical Physics. Retrieved December 16, 2005. Slides and audio from
an Ed Witten lecture where he introduces string theory and discusses its challenges.
Kibble, Tom. "Cosmic strings reborn?". arXiv:astro-ph/0410073. Invited Lecture at COSLAB 2004, held at
Ambleside, Cumbria, United Kingdom, from 10 to 17 September 2004.
Marolf, Don. "Resource Letter NSST-1: The Nature and Status of String Theory". arXiv:hep-th/0311044. A
guide to the string theory literature.
Ajay, Shakeeb, Wieland et al. (2004). "The nth dimension" (http://thenthdimension.com/). Retrieved December
16, 2005. A comprehensive compilation of materials concerning string theory. Created by an international team
of students.
Woit, Peter (2002). "Is string theory even wrong?" (http://www.americanscientist.org/issues/pub/
is-string-theory-even-wrong). American Scientist. Retrieved December 16, 2005. A criticism of string theory.
Veneziano, Gabriele (May 2004). "The Myth of the Beginning of Time" (http://www.sciam.com/article.
cfm?chanID=sa006&articleID=00042F0D-1A0E-1085-94F483414B7F0000). Scientific American
McKie, Robin (2006-10-09). "Setback as string theory of the universe is de-bunked" (http://www.hindu.com/
thehindu/holnus/008200610091240.htm) ( Scholar search (http://scholar.google.co.uk/scholar?hl=en&lr=&
q=author:McKie+intitle:Setback+as+string+theory+of+the+universe+is+de-bunked&as_publication=&
as_ylo=&as_yhi=&btnG=Search)). The Hindu
Harris, Richard (2006-11-07). "Short of 'All,' String Theorists Accused of Nothing" (http://www.npr.org/
templates/story/story.php?storyId=6377252). National Public Radio. Retrieved 2007-03-05.
A website dedicated to creative writing inspired by string theory. (http://banyancollege.org/scriblerus/)
An Italian Website with various papers in English language concerning the mathematical connections between
String Theory and Number Theory. (http://nardelli.xoom.it//stringtheory/)
George Gardner (2007-01-24). " Theory of everything put to the test (news:ID109828243)".
[news:tech.blorge.com tech.blorge.com]. (Web link) (http://tech.blorge.com/Structure: /2007/01/24/
theory-of-everything-put-to-the-test/). Retrieved on 2007-03-03.
Minkel, J. R. (2006-03-02). "A Prediction from String Theory, with Strings Attached" (http://www.sciam.com/
article.cfm?chanId=sa003&articleId=1475A684-E7F2-99DF-355B95296BE6031C). Scientific American
Chalmers, Matthew (2007-09-03). "Stringscape" (http://physicsworld.com/cws/article/indepth/30940).
Physics World (http://physicsworld.com). Retrieved September 6, 2007. An up-to-date and thorough review
of string theory in a popular way.
Woit, Peter. Not Even Wrong: The Failure of String Theory & the Continuing Challenge to Unify the Laws of
Physics, 2006. ISBN 0-224-07605-1 (Jonathan Cape), ISBN 0-465-09275-6 (Basic Books)
Schwarz, John (2001). "Early History of String Theory: A Personal Perspective" (http://online.itp.ucsb.edu/
online/colloq/schwarz1/). Retrieved July 17, 2009.

156

String Theory

External links
Dialogue on the Foundations of String Theory (http://www.mathpages.com/home/kmath632/kmath632.htm)
at MathPages
Superstrings! String Theory Home Page (http://www.sukidog.com/jpierre/strings/) Online tutorial
CI.physics. STRINGS newsgroup (http://schwinger.harvard.edu/~sps/) A moderated newsgroup for
discussion of string theory (a theory of quantum gravity and unification of forces) and related fields of
high-energy physics.
Not Even Wrong (http://www.math.columbia.edu/~woit/blog/) A blog critical of string theory.
Superstring Theory (http://www.perimeterinstitute.ca/en/Outreach/What_We_Research/Superstring_Theory/
) Perimeter Institute for Theoretical Physics
The Official String Theory Web Site (http://superstringtheory.com/)
The Elegant Universe (http://www.pbs.org/wgbh/nova/elegant/) A Three-Hour Miniseries with Brian
Greene by NOVA (original PBS Broadcast Dates: October 28, 8-10 p.m. and November 4, 8-9 p.m., 2003).
Various images, texts, videos and animations explaining string theory.
Beyond String Theory (http://www.phys.ens.fr/~troost/beyondstringtheory/) A project by a string physicist
explaining aspects of string theory to a broad audience.
Spinning the Superweb: Essays on the History of Superstring Theory (http://www.spinningthesuperweb.
blogspot.com) A Science Studies' approach to the history of string theory (an elementary knowledge of string
theory is required).

Quantum Gravity
Quantum gravity (QG) is the field of theoretical physics attempting to unify quantum mechanics with general
relativity in a self-consistent manner, or more precisely, to formulate a self-consistent theory which reduces to
ordinary quantum mechanics in the limit of weak gravity (potentials much less than c2) and which reduces to
Einsteinian general relativity in the limit of large actions (action much larger than reduced Planck's constant). The
theory must be able to predict the outcome of situations where both quantum effects and strong-field gravity are
important (at the Planck scale, unless large extra dimension conjectures are correct). Motivation for quantizing
gravity comes from the remarkable success of the quantum theories of the other three fundamental interactions.
Although some quantum gravity theories such as string theory and other so-called theories of everything attempt to
unify gravity with the other fundamental forces, others such as loop quantum gravity make no such attempt; they
simply quantize the gravitational field while keeping it separate from the other forces.
Observed physical phenomena in the early 21st century can be described well by quantum mechanics or general
relativity, without needing both. This can be thought of as due to an extreme separation of mass scales at which they
are important. Quantum effects are usually important only for the "very small", that is, for objects no larger than
typical molecules. General relativistic effects, on the other hand, show up only for the "very large" bodies such as
collapsed stars. (Planets' gravitational fields, as of 2009, are well-described by linearized gravity; so strong-field
effectsany effects of gravity beyond lowest nonvanishing order in /c2have not been observed even in the
gravitational fields of planets and main sequence stars). There is a lack of experimental evidence relating to quantum
gravity, and classical physics adequately describes the observed effects of gravity over a range of 50 orders of
magnitude of mass, i.e. for masses of objects from about 1023 to 1030 kg.

157

Quantum Gravity

Overview
Much of the difficulty in meshing these theories at all energy scales
comes from the different assumptions that these theories make on how
the universe works. Quantum field theory depends on particle fields
embedded in the flat space-time of special relativity. General relativity
models gravity as a curvature within space-time that changes as a
gravitational mass moves. Historically, the most obvious way of
combining the two (such as treating gravity as simply another particle
field) ran quickly into what is known as the renormalization problem.
Diagram showing where quantum gravity sits in
In the old-fashioned understanding of renormalization, gravity particles
the hierarchy of physics theories
would attract each other and adding together all of the interactions
results in many infinite values which cannot easily be cancelled out
mathematically to yield sensible, finite results. This is in contrast with quantum electrodynamics where, while the
series still do not converge, the interactions sometimes evaluate to infinite results, but those are few enough in
number to be removable via renormalization.

Effective field theories


Quantum gravity can be treated as an effective field theory. Effective quantum field theories come with some
high-energy cutoff, beyond which we do not expect that the theory provides a good description of nature. The
"infinities" then become large but finite quantities proportional to this finite cutoff scale, and correspond to processes
that involve very high energies near the fundamental cutoff. These quantities can then be absorbed into an infinite
collection of coupling constants, and at energies well below the fundamental cutoff of the theory, to any desired
precision; only a finite number of these coupling constants need to be measured in order to make legitimate
quantum-mechanical predictions. This same logic works just as well for the highly successful theory of low-energy
pions as for quantum gravity. Indeed, the first quantum-mechanical corrections to graviton-scattering and Newton's
law of gravitation have been explicitly computed[1] (although they are so astronomically small that we may never be
able to measure them). In fact, gravity is in many ways a much better quantum field theory than the Standard Model,
since it appears to be valid all the way up to its cutoff at the Planck scale. (By comparison, the Standard Model is
expected to start to break down above its cutoff at the much smaller scale of around 1000 GeV.)
While confirming that quantum mechanics and gravity are indeed consistent at reasonable energies, it is clear that
near or above the fundamental cutoff of our effective quantum theory of gravity (the cutoff is generally assumed to
be of order the Planck scale), a new model of nature will be needed. Specifically, the problem of combining quantum
mechanics and gravity becomes an issue only at very high energies, and may well require a totally new kind of
model.

Quantum gravity theory for the highest energy scales


The general approach to deriving a quantum gravity theory that is valid at even the highest energy scales is to
assume that such a theory will be simple and elegant and, accordingly, to study symmetries and other clues offered
by current theories that might suggest ways to combine them into a comprehensive, unified theory. One problem
with this approach is that it is unknown whether quantum gravity will actually conform to a simple and elegant
theory, as it should resolve the dual conundrums of special relativity with regard to the uniformity of acceleration
and gravity, and general relativity with regard to spacetime curvature.
Such a theory is required in order to understand problems involving the combination of very high energy and very
small dimensions of space, such as the behavior of black holes, and the origin of the universe.

158

Quantum Gravity

159

Quantum mechanics and general relativity


The graviton
At present, one of the deepest problems in theoretical physics is harmonizing
the theory of general relativity, which describes gravitation, and applies to
large-scale structures (stars, planets, galaxies), with quantum mechanics,
which describes the other three fundamental forces acting on the atomic scale.
This problem must be put in the proper context, however. In particular,
contrary to the popular claim that quantum mechanics and general relativity
are fundamentally incompatible, one can demonstrate that the structure of
general relativity essentially follows inevitably from the quantum mechanics
of interacting theoretical spin-2 massless particles [2] [3] [4] [5] [6] (called
gravitons).
While there is no concrete proof of the existence of gravitons, quantized
theories of matter may necessitate their existence. Supporting this theory is
the observation that all other fundamental forces have one or more messenger
particles, except gravity, leading researchers to believe that at least one most
likely does exist; they have dubbed these hypothetical particles gravitons.
Many of the accepted notions of a unified theory of physics since the 1970s,
including string theory, superstring theory, M-theory, loop quantum gravity,
all assume, and to some degree depend upon, the existence of the graviton.
Many researchers view the detection of the graviton as vital to validating their
work.

Gravity Probe B (GP-B) has measured


spacetime curvature near Earth to test
related models in application of
Einstein's general theory of relativity.

The dilaton
The dilaton made its first appearance in Kaluza-Klein theory, a five-dimensional theory that combined gravitation
and electromagnetism. Generally, it appears in string theory. More recently, it has appeared in the lower-dimensional
many-bodied gravity problem[7] based on the field theoretic approach of Roman Jackiw. The impetus arose from the
fact that complete analytical solutions for the metric of a covariant N-body system have proven elusive in General
Relativity. To simplify the problem, the number of dimensions was lowered to (1+1) namely one spatial dimension
and one temporal dimension. This model problem, known as R=T theory[8] (as opposed to the general G=T theory)
was amenable to exact solutions in terms of a generalization of the Lambert W function. It was also found that the
field equation governing the dilaton (derived from differential geometry) was none other than the Schrdinger
equation and consequently amenable to quantization.[9] Thus, one had a theory which combined gravity, quantization
and even the electromagnetic interaction, promising ingredients of a fundamental physical theory. It is worth noting
that the outcome revealed a previously unknown and already existing natural link between general relativity and
quantum mechanics. However, this theory needs to be generalized in (2+1) or (3+1) dimensions although, in
principle, the field equations are amenable to such generalization. It is not yet clear what field equation will govern
the dilaton in higher dimensions. This is further complicated by the fact that gravitons can propagate in (3+1)
dimensions and consequently that would imply gravitons and dilatons exist in the real world. Moreover, detection of
the dilaton is expected to be even more elusive than the graviton. However, since this approach allows for the
combination of gravitational, electromagnetic and quantum effects, their coupling could potentially lead to a means
of vindicating the theory, through cosmology and perhaps even experimentally.

Quantum Gravity

Nonrenormalizability of gravity
General relativity, like electromagnetism, is a classical field theory. One might expect that, as with
electromagnetism, there should be a corresponding quantum field theory.
However, gravity is nonrenormalizable.[10] Also in one loop approximation ultraviolet divergencies cancel on mass
shell. For a quantum field theory to be well-defined according to this understanding of the subject, it must be
asymptotically free or asymptotically safe. The theory must be characterized by a choice of finitely many parameters,
which could, in principle, be set by experiment. For example, in quantum electrodynamics, these parameters are the
charge and mass of the electron, as measured at a particular energy scale.
On the other hand, in quantizing gravity, there are infinitely many independent parameters (counterterm coefficients)
needed to define the theory. For a given choice of those parameters, one could make sense of the theory, but since
we can never do infinitely many experiments to fix the values of every parameter, we do not have a meaningful
physical theory:
At low energies, the logic of the renormalization group tells us that, despite the unknown choices of these
infinitely many parameters, quantum gravity will reduce to the usual Einstein theory of general relativity.
On the other hand, if we could probe very high energies where quantum effects take over, then every one of the
infinitely many unknown parameters would begin to matter, and we could make no predictions at all.
As explained below, there is a way around this problem by treating QG as an effective field theory.
Any meaningful theory of quantum gravity that makes sense and is predictive at all energy scales must have some
deep principle that reduces the infinitely many unknown parameters to a finite number that can then be measured.
One possibility is that normal perturbation theory is not a reliable guide to the renormalizability of the theory, and
that there really is a UV fixed point for gravity. Since this is a question of non-perturbative quantum field theory,
it is difficult to find a reliable answer, but some people still pursue this option.
Another possibility is that there are new symmetry principles that constrain the parameters and reduce them to a
finite set. This is the route taken by string theory, where all of the excitations of the string essentially manifest
themselves as new symmetries.

QG as an effective field theory


In an effective field theory, all but the first few of the infinite set of parameters in a non-renormalizable theory are
suppressed by huge energy scales and hence can be neglected when computing low-energy effects. Thus, at least in
the low-energy regime, the model is indeed a predictive quantum field theory.[1] (A very similar situation occurs for
the very similar effective field theory of low-energy pions.) Furthermore, many theorists agree that even the
Standard Model should really be regarded as an effective field theory as well, with "nonrenormalizable" interactions
suppressed by large energy scales and whose effects have consequently not been observed experimentally.
Recent work[1] has shown that by treating general relativity as an effective field theory, one can actually make
legitimate predictions for quantum gravity, at least for low-energy phenomena. An example is the well-known
calculation of the tiny first-order quantum-mechanical correction to the classical Newtonian gravitational potential
between two masses.

Spacetime background dependence


A fundamental lesson of general relativity is that there is no fixed spacetime background, as found in Newtonian
mechanics and special relativity; the spacetime geometry is dynamic. While easy to grasp in principle, this is the
hardest idea to understand about general relativity, and its consequences are profound and not fully explored, even at
the classical level. To a certain extent, general relativity can be seen to be a relational theory,[11] in which the only
physically relevant information is the relationship between different events in space-time.

160

Quantum Gravity

161

On the other hand, quantum mechanics has depended since its inception on a fixed background (non-dynamic)
structure. In the case of quantum mechanics, it is time that is given and not dynamic, just as in Newtonian classical
mechanics. In relativistic quantum field theory, just as in classical field theory, Minkowski spacetime is the fixed
background of the theory.
String theory
String theory started out as a generalization of quantum field
theory where instead of point particles, string-like objects
propagate in a fixed spacetime background. Although string theory
had its origins in the study of quark confinement and not of
quantum gravity, it was soon discovered that the string spectrum
contains the graviton, and that "condensation" of certain vibration
modes of strings is equivalent to a modification of the original
background. In this sense, string perturbation theory exhibits
exactly the features one would expect of a perturbation theory that
may exhibit a strong dependence on asymptotics (as seen, for
example, in the AdS/CFT correspondence) which is a weak form
of background dependence.

Interaction in the subatomic world: world lines of


point-like particles in the Standard Model or a world
sheet swept up by closed strings in string theory

Background independent theories


Loop quantum gravity is the fruit of an effort to formulate a background-independent quantum theory.
Topological quantum field theory provided an example of background-independent quantum theory, but with no
local degrees of freedom, and only finitely many degrees of freedom globally. This is inadequate to describe gravity
in 3+1 dimensions which has local degrees of freedom according to general relativity. In 2+1 dimensions, however,
gravity is a topological field theory, and it has been successfully quantized in several different ways, including spin
networks.

Semi-classical quantum gravity


Quantum field theory on curved (non-Minkowskian) backgrounds, while not a full quantum theory of gravity, has
shown many promising early results. In an analogous way to the development of quantum electrodynamics in the
early part of the 20th century (when physicists considered quantum mechanics in classical electromagnetic fields),
the consideration of quantum field theory on a curved background has led to predictions such as black hole radiation.
Phenomena such as the Unruh effect, in which particles exist in certain accelerating frames but not in stationary
ones, do not pose any difficulty when considered on a curved background (the Unruh effect occurs even in flat
Minkowskian backgrounds). The vacuum state is the state with least energy (and may or may not contain particles).
See Quantum field theory in curved spacetime for a more complete discussion.

Quantum Gravity

162

Points of tension
There are other points of tension between quantum mechanics and general relativity.
First, classical general relativity breaks down at singularities, and quantum mechanics becomes inconsistent with
general relativity in the neighborhood of singularities (however, no one is certain that classical general relativity
applies near singularities in the first place).
Second, it is not clear how to determine the gravitational field of a particle, since under the Heisenberg
uncertainty principle of quantum mechanics its location and velocity cannot be known with certainty. The
resolution of these points may come from a better understanding of general relativity.[12]
Third, there is the Problem of Time in quantum gravity. Time has a different meaning in quantum mechanics and
general relativity and hence there are subtle issues to resolve when trying to formulate a theory which combines
the two.[13]

Candidate theories
There are a number of proposed quantum gravity theories.[14] Currently, there is still no complete and consistent
quantum theory of gravity, and the candidate models still need to overcome major formal and conceptual problems.
They also face the common problem that, as yet, there is no way to put quantum gravity predictions to experimental
tests, although there is hope for this to change as future data from cosmological observations and particle physics
experiments becomes available.[15] [16]

String theory
One suggested starting point is ordinary quantum field
theories which, after all, are successful in describing
the other three basic fundamental forces in the context
of the standard model of elementary particle physics.
However, while this leads to an acceptable effective
(quantum) field theory of gravity at low energies,[17]
gravity turns out to be much more problematic at
higher energies. Where, for ordinary field theories such
as quantum electrodynamics, a technique known as
renormalization is an integral part of deriving
predictions which take into account higher-energy
contributions,[18] gravity
turns
out
to
be
nonrenormalizable: at high energies, applying the
recipes of ordinary quantum field theory yields models
that are devoid of all predictive power.[19]
One attempt to overcome these limitations is to replace
ordinary quantum field theory, which is based on the
classical concept of a point particle, with a quantum
[20]
theory of one-dimensional extended objects: string theory.
At the energies reached in current experiments, these
strings are indistinguishable from point-like particles, but, crucially, different modes of oscillation of one and the
same type of fundamental string appear as particles with different (electric and other) charges. In this way, string
theory promises to be a unified description of all particles and interactions.[21] The theory is successful in that one
mode will always correspond to a graviton, the messenger particle of gravity; however, the price to pay are unusual
features such as six extra dimensions of space in addition to the usual three for space and one for time.[22]
Projection of a Calabi-Yau manifold, one of the ways of
compactifying the extra dimensions posited by string theory

Quantum Gravity

163

In what is called the second superstring revolution, it was conjectured that both string theory and a unification of
general relativity and supersymmetry known as supergravity[23] form part of a hypothesized eleven-dimensional
model known as M-theory, which would constitute a uniquely defined and consistent theory of quantum gravity.[24]
[25]
As presently understood, however, string theory is consistent with an estimated 10500 equally possible solutions
(the so-called "string landscape").

Loop quantum gravity


Another approach to quantum gravity starts with the
canonical quantization procedures of quantum theory.
Starting with the initial-value-formulation of general
relativity (cf. the section on evolution equations,
above), the result is an analogue of the Schrdinger
equation: the WheelerDeWitt equation, which some
argue is ill-defined.[26] A major break-through came
with the introduction of what are now known as
Ashtekar variables, which represent geometric gravity
using mathematical analogues of electric and magnetic
fields.[27] [28] The resulting candidate for a theory of
quantum gravity is Loop quantum gravity, in which
space is represented by a network structure called a
spin network, evolving over time in discrete steps.[29]
[30] [31] [32]

Simple spin network of the type used in loop quantum gravity

Other approaches
There are a number of other approaches to quantum gravity. The approaches differ depending on which features of
general relativity and quantum theory are accepted unchanged, and which features are modified.[33] [34] Examples
include:

Supergravity
Path-integral based models of quantum cosmology[35]
Causal Dynamical Triangulation[36]
Regge calculus
Causal sets[37]
Asymptotic safety
Twistor models[38]
Noncommutative geometry.
String-nets giving rise to gapless helicity 2 excitations with no other gapless excitations[39]
Acoustic metric and other analog models of gravity
MacDowellMansouri action
Group field theory[40]

Quantum Gravity

Weinberg-Witten theorem
In quantum field theory, the Weinberg-Witten theorem places some constraints on theories of composite
gravity/emergent gravity. However, recent developments attempt to show that if locality is only approximate and the
holographic principle is correct, the Weiberg-Witten theorem would not be valid.

References
[1] Donoghue (1995). "Introduction to the Effective Field Theory Description of Gravity". arXiv:gr-qc/9512024[gr-qc].
[2] Kraichnan, R. H. (1955). "Special-Relativistic Derivation of Generally Covariant Gravitation Theory". Physical Review 98 (4): 11181122.
doi:10.1103/PhysRev.98.1118.
[3] Gupta, S. N. (1954). "Gravitation and Electromagnetism". Physical Review 96 (6): 16831685. doi:10.1103/PhysRev.96.1683.
[4] Gupta, S. N. (1957). "Einstein's and Other Theories of Gravitation". Reviews of Modern Physics 29 (3): 334336.
doi:10.1103/RevModPhys.29.334.
[5] Gupta, S. N. (1962). "Quantum Theory of Gravitation". Recent Developments in General Relativity. Pergamon Press. pp.251258.
[6] Deser, S. (1970). "Self-Interaction and Gauge Invariance". General Relativity and Gravitation 1: 918. doi:10.1007/BF00759198.
[7] Ohta, Tadayuki; Mann, Robert (1996). "Canonical reduction of two-dimensional gravity for particle dynamics". Classical and Quantum
Gravity 13 (9): 2585. arXiv:gr-qc/9605004. doi:10.1088/0264-9381/13/9/022.
[8] Sikkema, A E; Mann, R B (1991). "Gravitation and cosmology in (1+1) dimensions". Classical and Quantum Gravity 8: 219.
doi:10.1088/0264-9381/8/1/022.
[9] Farrugia; Mann; Scott (2007). "N-body Gravity and the Schroedinger Equation". Classical and Quantum Gravity 24 (18): 46474659.
arXiv:gr-qc/0611144. doi:10.1088/0264-9381/24/18/006.
[10] Feynman, R. P.; Morinigo, F. B., Wagner, W. G., & Hatfield, B. (1995). Feynman lectures on gravitation. Addison-Wesley.
ISBN0201627345.
[11] Smolin, Lee (2001). Three Roads to Quantum Gravity. Basic Books. pp.2025. ISBN0-465-08735-4. Pages 220-226 are annotated
references and guide for further reading.
[12] Hunter Monroe (2005). "Singularity-Free Collapse through Local Inflation". arXiv:astro-ph/0506506[astro-ph].
[13] Edward Anderson (2010). "The Problem of Time in Quantum Gravity". arXiv:1009.2157[gr-qc].
[14] A timeline and overview can be found in Rovelli, Carlo (2000). "Notes for a brief history of quantum gravity".
arXiv:gr-qc/0006061[gr-qc]..
[15] Ashtekar, Abhay (2007). "Loop Quantum Gravity: Four Recent Advances and a Dozen Frequently Asked Questions". 11th Marcel
Grossmann Meeting on Recent Developments in Theoretical and Experimental General Relativity. pp.126. arXiv:0705.2222.
Bibcode2008mgm..conf..126A. doi:10.1142/9789812834300_0008.
[16] Schwarz, John H. (2007). "String Theory: Progress and Problems". Progress of Theoretical Physics Supplement 170: 214.
arXiv:hep-th/0702219. Bibcode2007PThPS.170..214S. doi:10.1143/PTPS.170.214.
[17] Donoghue, John F.(editor), (1995). "Introduction to the Effective Field Theory Description of Gravity". In Cornet, Fernando. Effective
Theories: Proceedings of the Advanced School, Almunecar, Spain, 26 June1 July 1995. Singapore: World Scientific. arXiv:gr-qc/9512024.
ISBN9810229089.
[18] Weinberg, Steven (1996). "17-18". The Quantum Theory of Fields II: Modern Applications. Cambridge University Press.
ISBN0-521-55002-5.
[19] Goroff, Marc H.; Sagnotti, Augusto (1985). "Quantum gravity at two loops". Physics Letters B 160: 8186.
doi:10.1016/0370-2693(85)91470-4.
[20] An accessible introduction at the undergraduate level can be found in Zwiebach, Barton (2004). A First Course in String Theory. Cambridge
University Press. ISBN0-521-83143-1., and more complete overviews in Polchinski, Joseph (1998). String Theory Vol. I: An Introduction to
the Bosonic String. Cambridge University Press. ISBN0-521-63303-6. and Polchinski, Joseph (1998b). String Theory Vol. II: Superstring
Theory and Beyond. Cambridge University Press. ISBN0-521-63304-4.
[21] Ibanez, L. E. (2000). "The second string (phenomenology) revolution". Classical & Quantum Gravity 17 (5): 11171128.
arXiv:hep-ph/9911499. doi:10.1088/0264-9381/17/5/321.
[22] For the graviton as part of the string spectrum, e.g. Green, Schwarz & Witten 1987, sec. 2.3 and 5.3; for the extra dimensions, ibid sec. 4.2.
[23] Weinberg, Steven (2000). "31" (http:/ / books. google. com/ books?id=aYDDRKqODpUC& printsec=frontcover). The Quantum Theory of
Fields II: Modern Applications. Cambridge University Press. ISBN0-521-55002-5. .
[24] Townsend, Paul K. (1996). Four Lectures on M-Theory. ICTP Series in Theoretical Physics. pp.385. arXiv:hep-th/9612121.
Bibcode1997hepcbconf..385T.
[25] Duff, Michael (1996). "M-Theory (the Theory Formerly Known as Strings)". International Journal of Modern Physics A 11 (32):
56235642. arXiv:hep-th/9608117. Bibcode1996IJMPA..11.5623D. doi:10.1142/S0217751X96002583.
[26] Kucha, Karel (1973). "Canonical Quantization of Gravity". In Israel, Werner. Relativity, Astrophysics and Cosmology. D. Reidel.
pp.237288 (section 3). ISBN90-277-0369-8.

164

Quantum Gravity
[27] Ashtekar, Abhay (1986). "New variables for classical and quantum gravity". Physical Review Letters 57 (18): 22442247.
doi:10.1103/PhysRevLett.57.2244. PMID10033673.
[28] Ashtekar, Abhay (1987). "New Hamiltonian formulation of general relativity". Physical Review D 36 (6): 15871602.
doi:10.1103/PhysRevD.36.1587.
[29] Thiemann, Thomas (2006). "Loop Quantum Gravity: An Inside View". Approaches to Fundamental Physics 721: 185.
arXiv:hep-th/0608210. Bibcode2007LNP...721..185T.
[30] Rovelli, Carlo (1998). "Loop Quantum Gravity" (http:/ / www. livingreviews. org/ lrr-1998-1). Living Reviews in Relativity 1. . Retrieved
2008-03-13.
[31] Ashtekar, Abhay; Lewandowski, Jerzy (2004). "Background Independent Quantum Gravity: A Status Report". Classical & Quantum
Gravity 21 (15): R53R152. arXiv:gr-qc/0404018. doi:10.1088/0264-9381/21/15/R01.
[32] Thiemann, Thomas (2003). "Lectures on Loop Quantum Gravity". Lecture Notes in Physics 631: 41135. arXiv:gr-qc/0210094.
Bibcode2003LNP...631...41T.
[33] Isham, Christopher J. (1994). "Prima facie questions in quantum gravity". In Ehlers, Jrgen; Friedrich, Helmut. Canonical Gravity: From
Classical to Quantum. Springer. arXiv:gr-qc/9310031. ISBN3-540-58339-4.
[34] Sorkin, Rafael D. (1997). "Forks in the Road, on the Way to Quantum Gravity". International Journal of Theoretical Physics 36 (12):
27592781. arXiv:gr-qc/9706002. doi:10.1007/BF02435709.
[35] Hawking, Stephen W. (1987). "Quantum cosmology". In Hawking, Stephen W.; Israel, Werner. 300 Years of Gravitation. Cambridge
University Press. pp.631651. ISBN0-521-37976-8..
[36] Loll, Renate (1998). "Discrete Approaches to Quantum Gravity in Four Dimensions" (http:/ / www. livingreviews. org/ lrr-1998-13). Living
Reviews in Relativity 1: 13. Bibcode1998LRR.....1...13L. . Retrieved 2008-03-09.
[37] Sorkin, Rafael D. (2005). "Causal Sets: Discrete Gravity". In Gomberoff, Andres; Marolf, Donald. Lectures on Quantum Gravity. Springer.
arXiv:gr-qc/0309009. ISBN0-387-23995-2.
[38] See ch. 33 in Penrose 2004 and references therein.
[39] Wen 2006
[40] See Daniele Oriti and references therein.

Further reading
Ahluwalia, D. V. (2002). "Interface of Gravitational and Quantum Realms". Modern Physics Letters A 17
(15-17): 1135. arXiv:gr-qc/0205121. Bibcode2002MPLA...17.1135A. doi:10.1142/S021773230200765X
Ashtekar, Abhay (2005). "The winding road to quantum gravity" (http://www.ias.ac.in/currsci/dec252005/
2064.pdf). Current Science 89: 20642074
Carlip, Steven (2001). "Quantum Gravity: a Progress Report". Reports on Progress in Physics 64 (8): 885.
arXiv:gr-qc/0108040. doi:10.1088/0034-4885/64/8/301
Kiefer, Claus (2007). Quantum Gravity. Oxford University Press. ISBN019921252X
Kiefer, Claus (2005). "Quantum Gravity: General Introduction and Recent Developments". Annalen der Physik
15: 129148. arXiv:gr-qc/0508120. doi:10.1002/andp.200510175
Lmmerzahl, Claus, ed (2003). Quantum Gravity: From Theory to Experimental Search. Lecture Notes in
Physics. Springer. ISBN354040810X
Trifonov, Vladimir (2008). "GR-friendly description of quantum systems". International Journal of Theoretical
Physics 47 (2): 492510. arXiv:math-ph/0702095. doi:10.1007/s10773-007-9474-3

165

166

Appendix
Quantum
In physics, a quantum (plural: quanta) is the minimum amount of any physical entity involved in an interaction.
Behind this, one finds the fundamental notion that a physical property may be "quantized," referred to as "the
hypothesis of quantization".[1] This means that the magnitude can take on only certain discrete values. There is a
related term of quantum number. An example of an entity that is quantized is the energy transfer of elementary
particles of matter (called fermions) and of photons and other bosons.
A photon is a single quantum of light, and is referred to as a "light quantum". The energy of an electron bound to an
atom (at rest) is said to be quantized, which results in the stability of atoms, and of matter in general.
As incorporated into the theory of quantum mechanics, this is regarded by physicists as part of the fundamental
framework for understanding and describing nature at the infinitesimal level.
Normally quanta are considered to be discrete packets with energy stored in them. Max Planck considered these
quanta to be particles that can change their form (meaning that they can be absorbed and released). This phenomenon
can be observed in the case of black body radiation, when it is being heated and cooled.

Etymology and discovery


The word "quantum" comes from the Latin "quantus," for "how much." "Quanta" meaning short for "quanta of
electricity" (or electron) was used in a 1902 article on the photoelectric effect by Philipp Lenard, who credited
Hermann von Helmholtz for using the word in the area of electricity. However, the word quantum in general was
well known before 1900.[2] It was often used by physicians, such as the term quantum satis. Both Helmholtz and
Julius von Mayer were physicians as well as physicists. Helmholtz used quantum with reference to heat in his article
[3]
on Mayer's work, and indeed, the word quantum can be found in the formulation of the first law of
thermodynamics by Mayer in his letter [4] dated July 24, 1841. Max Planck used "quanta" to mean "quanta of matter
and electricity",[5] gas, and heat.[6] In 1905, in response to Planck's work and the experimental work of Lenard, who
explained his results by using the term "quanta of electricity", Albert Einstein suggested that radiation existed in
spatially localized packets which he called "quanta of light" ("Lichtquanta").[7]
The concept of quantization of radiation was discovered in 1900 by Max Planck, who had been trying to understand
the emission of radiation from heated objects, known as black body radiation. By assuming that energy can only be
absorbed or released in tiny, differential, discrete packets he called "bundles" or "energy elements,",[8] Planck
accounted for the fact that certain objects change colour when heated.[9] On December 14, 1900, Planck reported his
revolutionary findings to the German Physical Society and introduced the idea of quantization for the first time as a
part of his research on black body radiation.[10] As a result of his experiments, Planck deduced the numerical value
of h, known as the Planck constant, and could also report a more precise value for the Avogadro-Loschmidt number,
the number of real molecules in a mole and the unit of electrical charge, to the German Physical Society. After his
theory was validated, Planck was awarded the Nobel Prize in Physics in 1918 for his discovery.

Quantum

Beyond electromagnetic radiation


While quantization was first discovered in electromagnetic radiation, it describes a fundamental aspect of energy not
just restricted to photons.[11]

References
[1] Wiener, N. (1966). Differential Space, Quantum Systems, and Prediction. Cambridge: The Massachusetts Institute of Technology Press
[2] E. Cobham Brewer 18101897. Dictionary of Phrase and Fable. 1898. (http:/ / www. bartleby. com/ 81/ 13830. html)
[3] E. Helmholtz, Robert Mayer's Prioritt (http:/ / www. ub. uni-heidelberg. de/ helios/ fachinfo/ www/ math/ edd/ helmholtz/ R-Mayer. pdf)

(German)
[4] Herrmann,A. Weltreich der Physik, GNT-Verlag (1991) (http:/ / fs. math. uni-frankfurt. de/ fsmath/ misc/ RobertMayer. html) (German)
[5] Planck, M. (1901). "Ueber die Elementarquanta der Materie und der Elektricitt". Annalen der Physik 309: 564566.
doi:10.1002/andp.19013090311. (German)
[6] Planck, Max (1883). "Ueber das thermodynamische Gleichgewicht von Gasgemengen". Annalen der Physik 255: 358.
doi:10.1002/andp.18832550612. (German)
[7] Einstein, A. (1905). "ber einen die Erzeugung und Verwandlung des Lichtes betreffenden heuristischen Gesichtspunkt" (http:/ / www.
physik. uni-augsburg. de/ annalen/ history/ einstein-papers/ 1905_17_132-148. pdf). Annalen der Physik 17: 132148.
doi:10.1002/andp.19053220607. . (German). A partial English translation is available from Wikisource.
[8] Max Planck (1901). "Ueber das Gesetz der Energieverteilung im Normalspectrum (On the Law of Distribution of Energy in the Normal
Spectrum)" (http:/ / web. archive. org/ web/ 20080418002757/ http:/ / dbhs. wvusd. k12. ca. us/ webdocs/ Chem-History/ Planck-1901/
Planck-1901. html). Annalen der Physik 309: 553. doi:10.1002/andp.19013090310. Archived from the original (http:/ / dbhs. wvusd. k12. ca.
us/ webdocs/ Chem-History/ Planck-1901/ Planck-1901. html) on 2008-04-18. .
[9] Brown, T., LeMay, H., Bursten, B. (2008). Chemistry: The Central Science Upper Saddle River, NJ: Pearson Education ISBN 0-13-600617-5
[10] Klein, Martin J. (1961). "Max Planck and the beginnings of the quantum theory". Archive for History of Exact Sciences 1: 459.
doi:10.1007/BF00327765.
[11] Melville, K. (2005, February 11). Real-World Quantum Effects Demonstrated (http:/ / www. scienceagogo. com/ news/
20050110221715data_trunc_sys. shtml)

Further reading
B. Hoffmann, The Strange Story of the Quantum, Pelican 1963.
Lucretius, "On the Nature of the Universe", transl. from the Latin by R.E. Latham, Penguin Books Ltd.,
Harmondsworth 1951. There are, of course, many translations, and the translation's title varies. Some put
emphasis on how things work, others on what things are found in nature.
J. Mehra and H. Rechenberg, The Historical Development of Quantum Theory, Vol.1, Part 1, Springer-Verlag
New York Inc., New York 1982.
M. Planck, A Survey of Physical Theory, transl. by R. Jones and D.H. Williams, Methuen & Co., Ltd., London
1925 (Dover editions 1960 and 1993) including the Nobel lecture.

167

Quantum state

168

Quantum state
In quantum physics, a quantum state is a set of mathematical variables that fully describes a quantum system. For
example, the set of 4 numbers { , ,
,
} define the state of an electron within a hydrogen atom and are
known as the electron's quantum numbers. More generally, the state of the system is represented by a single
mathematical object known as a ket.
Typically, one postulates some experimental apparatus and procedure which "prepares" this quantum state; the
mathematical object reflects the setup of the apparatus. Quantum states can be either pure or mixed. Pure states
cannot be described as a mixture of others. Mixed states correspond to an experiment involving a random process
that blends pure states together.[1]
When performing a particular measurement on a quantum state, the result is usually described by a probability
distribution, and the form that this distribution takes is completely determined by the quantum state and the
observable describing the measurement. However, unlike in classical mechanics, the result of a measurement on
even a pure quantum state is a probability, rather than a certainty. This reflects a core difference between classical
and quantum physics.
Mathematically, a pure quantum state is typically represented by a vector in a Hilbert space. In physics, bra-ket
notation is often used to denote such vectors. Linear combinations (superpositions) of vectors can describe
interference phenomena. Mixed quantum states are described by density matrices.
In a more general mathematical context, quantum states can be understood as positive normalized linear functionals
on a C* algebra; see GNS construction.

Conceptual description
The state of a physical system
The state of a physical system is a complete description of the parameters of the experiment. To understand this
rather abstract notion, it is useful to first explore it in an example from classical mechanics.
Consider an experiment with a (non-quantum) particle of mass

that moves freely, and without friction, in

one spatial direction.


We put the particle at initial position

and start the experiment at time

by pushing the particle with some

speed and in some direction. Doing this, we determine the initial momentum
conditions are what characterizes the state

of the system, formally denoted as

of the particle. These initial


. We say that we

prepare the state of the system by fixing its initial conditions.


At a later time

, we conduct measurements on the particle. The measurements we can perform on this simple

system are essentially its position


and

at time

, its momentum

, and combinations of these. Here

refer to the measurable quantities (observables) of the system as such, not the specific results they

produce in a certain run of the experiment.


However, knowing the state of the system, we can compute the value of the observables in the specific state, i.e.
the results that our measurements will produce, depending on and . We denote these values as
and
. In our simple example, it is well known that the particle moves with constant velocity; therefore,[2]
Now suppose that we start the particle with a random initial position and momentum. (For argument's sake, we may
suppose that the particle is pushed away at
by some apparatus which is controlled by a random number
generator.) The state

of the system is now not described by two numbers

distributions. The observables

and

and

, but rather by two probability

will produce random results now; they become random variables,

Quantum state

169

and their values in a single measurement cannot be predicted. However, if we repeat the experiment sufficiently
often, always preparing the same state , we can predict the expectation value of the observables (their statistical
mean) in the state . The expectation value of
is again denoted by
, etc.
These "statistical" states of the system are called mixed states, as opposed to the pure states

discussed

further above. Abstractly, mixed states arise as a statistical mixture of pure states.

Quantum states
In quantum systems, the conceptual
distinction between observables and
states persists just as described above.
The state of the system is fixed by
the way the physicist prepares the
experiment (e.g., how the physicist
adjusts the particle source). As above,
there is a distinction between pure
states and mixed states, the latter being
statistical mixtures of the former.
However, some important differences
arise in comparison with classical
mechanics.

Probability densities for the electron of a hydrogen atom in different quantum states.

In quantum theory, even pure states show statistical behaviour. Regardless of how carefully we prepare the state
of the system, measurement results are not repeatable in general, and we must understand the expectation value
of an observable
as a statistical mean. It is this mean that is predicted by physical theories.
For any fixed observable

, it is generally possible to prepare a pure state

this state: If we repeat the experiment several times, each time measuring
measurement result, without any random behaviour. Such pure states

such that

has a fixed value in

, we will always obtain the same

are called eigenstates of

However, it is impossible to prepare a simultaneous eigenstate for all observables. For example, we cannot prepare a
state such that both the position measurement
and the momentum measurement
(at the same time )
produce "sharp" results; at least one of them will exhibit random behaviour.[3] This is the content of the Heisenberg
uncertainty relation.
Moreover, in contrast to classical mechanics, it is unavoidable that performing a measurement on the system
generally changes its state. More precisely: After measuring an observable
, the system will be in an eigenstate
of

; thus the state has changed, unless the system was already in that eigenstate. This expresses a kind of logical

consistency: If we measure

twice in the same run of the experiment, the measurements being directly consecutive

in time, then they will produce the same results. This has some strange consequences however:
Consider two observables,

and

, where

that the system is in an eigenstate of


measure first

and then

corresponds to a measurement earlier in time than

. If we measure only

.[4] Suppose

, we will not notice statistical behaviour. If we

in the same run of the experiment, the system will transfer to an eigenstate of

after

Quantum state

170

the first measurement, and we will generally notice that the results of

are statistical. Thus, quantum mechanical

measurements influence one another, and it is important in which order they are performed.
Another feature of quantum states becomes relevant if we consider a physical system that consists of multiple
subsystems; for example, an experiment with two particles rather than one. Quantum physics allows for certain
states, called entangled states, that show certain statistical correlations between measurements on the two particles
which cannot be explained by classical theory. For details, see entanglement. These entangled states lead to
experimentally testable properties (Bell's theorem) that allow us to distinguish between quantum theory and
alternative classical (non-quantum) models.

Schrdinger picture vs. Heisenberg picture


In the discussion above, we have taken the observables

to be dependent on time, while the state

was fixed once at the beginning of the experiment. This approach is called the Heisenberg picture. One can,
equivalently, treat the observables as fixed, while the state of the system depends on time; that is known as the
Schrdinger picture. Conceptually (and mathematically), both approaches are equivalent; choosing one of them is a
matter of convention.
Both viewpoints are used in quantum theory. While non-relativistic quantum mechanics is usually formulated in
terms of the Schrdinger picture, the Heisenberg picture is often preferred in a relativistic context, that is, for
quantum field theory.

Formalism in quantum physics


Pure states as rays in a Hilbert space
Quantum physics is most commonly formulated in terms of linear algebra, as follows. Any given system is identified
with some Hilbert space, such that each vector in the Hilbert space (apart from the origin) corresponds to a pure
quantum state. In addition, two vectors that differ only by a nonzero complex scalar correspond to the same state (in
other words, each pure state is a ray in the Hilbert space; equivalently, a point in the projective Hilbert space.).
Alternatively, many authors choose to only consider normalized vectors (vectors of norm 1) as corresponding to
quantum states. In this case, the set of all pure states corresponds to the unit sphere of a Hilbert space, with the
proviso that two normalized vectors correspond to the same state if they differ only by a complex scalar of absolute
value 1, which is called the phase factor.

Bra-ket notation
Calculations in quantum mechanics make frequent use of linear operators, inner products, dual spaces, and
Hermitian conjugation. In order to make such calculations more straightforward, and to obviate the need (in some
contexts) to fully understand the underlying linear algebra, Paul Dirac invented a notation to describe quantum
states, known as bra-ket notation. Although the details of this are beyond the scope of this article (see the article
Bra-ket notation), some consequences of this are:
The variable name used to denote a vector (which corresponds to a pure quantum state) is chosen to be of the
form
(where the " " can be replaced by any other symbols, letters, numbers, or even words). This can be
contrasted with the usual mathematical notation, where vectors are usually bold, lower-case letters, or letters with
arrows on top.
Instead of vector, the term ket is used synonymously.
Each ket

is uniquely associated with a so-called bra, denoted

, which is also said to correspond to the

same physical quantum state. Technically, the bra is an element of the dual space, and related to the ket by the
Riesz representation theorem.

Quantum state

171

Inner products (also called brackets) are written so as to look like a bra and ket next to each other:

(Note that the phrase "bra-ket" is supposed to resemble "bracket".)

Spin, many-body states


It is important to note that in quantum mechanics besides, e.g., the usual position variable , a discrete variable m
exists, corresponding to the value of the z-component of the spin vector. This can be thought of as a kind of intrinsic
angular momentum. However, it does not appear at all in classical mechanics and arises from Dirac's relativistic
generalization of the theory. As a consequence, the quantum state of a system of N particles is described by a
function with four variables per particle, e.g.
. Here, the variables m assume values
from the set {

}, where

(in units of Planck's reduced constant

), is

either a non-negative integer (0,1,2...; bosons), or semi-integer (1/2,3/2,5/2,...; fermions). Moreover, in the case of
identical particles, the above N-particle function must either be symmetrized (in the bosonic case) or
anti-symmetrized (in the fermionic case) w.r.t. the particle numbers.
Electrons are fermions with S=1/2, photons (quanta of light) are bosons with S=1.
Apart from the symmetrization or anti-symmetrization, N-particle states can thus simply be obtained by tensor
products of one-particle states, to which we return herewith.

Basis states of one-particle systems


As with any vector space, if a basis is chosen for the Hilbert space of a system, then any ket can be expanded as a
linear combination of those basis elements. Symbolically, given basis kets
, any ket
can be written

where ci are complex numbers. In physical terms, this is described by saying that
quantum superposition of the states

has been expressed as a

. If the basis kets are chosen to be orthonormal (as is often the case), then

.
One property worth noting is that the normalized states

are characterized by

Expansions of this sort play an important role in measurement in quantum mechanics. In particular, if the
eigenstates (with eigenvalues

) of an observable, and that observable is measured on the normalized state


2

are
,

then the probability that the result of the measurement is ki is |ci| . (The normalization condition above mandates that
the total sum of probabilities is equal to one.)
A particularly important example is the position basis, which is the basis consisting of eigenstates of the observable
which corresponds to measuring position. If these eigenstates are nondegenerate (for example, if the system is a
single, spinless particle), then any ket
is associated with a complex-valued function of three-dimensional space:
.
This function is called the wavefunction corresponding to

Quantum state

172

Superposition of pure states


One aspect of quantum states, mentioned above, is that superpositions of them can be formed. If

and

are

two kets corresponding to quantum states, the ket


is a different quantum state (possibly not normalized). Note that which quantum state it is depends on both the
amplitudes and phases (arguments) of
and
. In other words, for example, even though
and
(for
real ) correspond to the same physical quantum state, they are not interchangeable, since for example
and

do not (in general) correspond to the same physical state. However,

and

do correspond to the same physical state. This is sometimes described by saying that "global"
phase factors are unphysical, but "relative" phase factors are physical and important.
One example of a quantum interference phenomenon that arises from superposition is the double-slit experiment.
The photon state is a superposition of two different states, one of which corresponds to the photon having passed
through the left slit, and the other corresponding to passage through the right slit. The relative phase of those two
states has a value which depends on the distance from each of the two slits. Depending on what that phase is, the
interference is constructive at some locations and destructive in others, creating the interference pattern.
Another example of the importance of relative phase in quantum superposition is Rabi oscillations, where the
relative phase of two states varies in time due to the Schrdinger equation. The resulting superposition ends up
oscillating back and forth between two different states.

Mixed states
A pure quantum state is a state which can be described by a single ket vector, as described above. A mixed quantum
state is a statistical ensemble of pure states (see quantum statistical mechanics). Equivalently, a mixed-quantum state
on a given quantum system described by a Hilbert space H naturally arises as a pure quantum state (called a
purification) on a larger bipartite system H tensor K, the other half of which is inaccessible to the observer.
A mixed state cannot be described as a ket vector. Instead, it is described by its associated density matrix (or density
operator), usually denoted . Note that density matrices can describe both mixed and pure states, treating them on
the same footing.
The density matrix is defined as

where

is the fraction of the ensemble in each pure state

Here, one typically uses a one-particle formalism

to describe the average behaviour of a N-particle system.


A simple criterion for checking whether a density matrix is describing a pure or mixed state is that the trace of 2 is
equal to 1 if the state is pure, and less than 1 if the state is mixed.[5] Another, equivalent, criterion is that the von
Neumann entropy is 0 for a pure state, and strictly positive for a mixed state.
The rules for measurement in quantum mechanics are particularly simple to state in terms of density matrices. For
example, the ensemble average (expectation value) of a measurement corresponding to an observable
is given by

where

are eigenkets and eigenvalues, respectively, for the operator

, and tr denotes trace. It is

important to note that two types of averaging are occurring, one being a quantum average over the basis kets
of the pure states, and the other being a statistical average with the probabilities
of those states.
W.r.t. these different types of averaging, i.e. to distinguish pure and/or mixed states, one often uses the expressions
'coherent' and/or 'incoherent superposition' of quantum states.

Quantum state

173

Mathematical formulation
For a mathematical discussion on states as functionals, see Gelfand-Naimark-Segal construction. There, the same
objects are described in a C*-algebraic context.

Notes
[1] http:/ / www. quantiki. org/ wiki/ Bloch_sphere
[2] In this example, momentum p = velocity, since p = m velocity, and m = 1. The system of units for this example has been defined such that
mass is unitless.
[3] To avoid misunderstandings: Here we mean that

and

are measured in the same state, but not in the same run of the

experiment.
[4] For concreteness' sake, you may suppose that
and
in the above example, with
.
[5] Blum, Density matrix theory and applications, page 39 (http:/ / books. google. com/ books?id=kl-pMd9Qx04C& pg=PA39). Note that this
criterion works when the density matrix is normalized so that the trace of is 1, as it is for the standard definition given in this section.
Occasionally a density matrix will be normalized differently, in which case the criterion is

Further reading
The concept of quantum states, in particular the content of the section Formalism in quantum physics above, is
covered in most standard textbooks on quantum mechanics.
For a discussion of conceptual aspects and a comparison with classical states, see:
Isham, Chris J (1995). Lectures on Quantum Theory: Mathematical and Structural Foundations. Imperial College
Press. ISBN978-1860940019.
For a more detailed coverage of mathematical aspects, see:
Bratteli, Ola; Robinson, Derek W (1987). Operator Algebras and Quantum Statistical Mechanics 1. Springer.
ISBN978-3540170938. 2nd edition. In particular, see Sec. 2.3.
For a discussion of purifications of mixed quantum states, see Chapter 2 of John Preskill's lecture notes for Physics
219 (http://www.theory.caltech.edu/~preskill/ph229/) at Caltech.

Article Sources and Contributors

Article Sources and Contributors


History of Quantum Mechanics Source: http://en.wikipedia.org/w/index.php?oldid=421259791 Contributors: Agge1000, BD2412, Bduke, BenRG, Betacommand, Boardhead, Bookalign,
Coppertwig, D.H, Dharmaraj.S, Dratman, DuncanHill, GeorgeLouis, Geremia, Graeme Bartlett, Grafen, Grimlock, Headbomb, Ignoranteconomist, Itub, JaGa, Jagged 85, Jbergquist, John
Vandenberg, Kbdank71, Kinema, Kusunose, Laurascudder, Lottamiata, LuchoX, Michael C Price, Michal Nebyla, Mild Bill Hiccup, Mpkannan, Myasuda, Nagualdesign, Okedem, Ospalh,
PSimeon, Rhetth, Riversider2008, Sadi Carnot, Shadowjams, Stephen Poppitt, StradivariusTV, Strait, Topherwhelan, Wikiklrsc, Yill577, Yvwv, Yym1997, 36 anonymous edits
Basic Concepts of Quantum Mechanics Source: http://en.wikipedia.org/w/index.php?oldid=418077184 Contributors: Awayforawhile, BenRG, Butwhatdoiknow, C. Trifle, Cardinality,
CristianCantoro, Disneyfreak96, Djr32, Dolphin51, GeorgeLouis, John Vandenberg, JohnJSal, Kbrose, Maurice Carbonaro, MuZemike, Nickj, Notolder, Patrick0Moran, R'n'B, RobinK, Skier
Dude, Stephen Poppitt, TimothyRias, UncleDouggie, 8 anonymous edits
Introduction to Quantum Mechanics Source: http://en.wikipedia.org/w/index.php?oldid=423526536 Contributors: 16@r, 1howardsr1, Accurrent, Adashiel, Alain10, Alansohn, Aliencam,
Altzinn, Amihaly, Ancheta Wis, Android79, AndyZ, Anoko moonlight, Aostrander, Arion 3x3, Arjen Dijksman, Army1987, Arpingstone, AussieBand1, Avb, Avramov, Awolf002, BabyNuke,
Baccyak4H, Barbara Shack, Barraki, Batmanand, Bearian, BenRG, Benhocking, Bevo, Bloodshedder, BohemianWikipedian, Brews ohare, Brina700, Butwhatdoiknow, C. Trifle, Calmer Waters,
Caltas, ChangChienFu, Changcho, CharlesHBennett, ChristopherWillis, Chzz, Clay Juicer, Closedmouth, Cmeyer1969, Colonies Chris, Comrade009, ConfuciusOrnis, Count Iblis, Craig Stuntz,
Csmiller, Curtmcd, Cybercobra, D-Notice, DESiegel, DJIndica, DVdm, Danaudick, Daniel Vollmer, Danny Rathjens, Dauto, David R. Ingham, David edwards, Deagle AP, Demoni, Dionyziz,
Discospinster, Djr32, DrFarhadn, Droll, ESkog, Easchiff, Editwiki, Edreher, Edward Z. Yang, El C, Elungtrings, EncycloPetey, Erkenbrack, Estrellador*, Everyking, Farrant, Fg2, Fones4jenke,
Gareth Jones, Gary King, Geffb, Gene Nygaard, George100, GeorgeLouis, Georgelulu, Gifani, Giftlite, Gill110951, Gmalivuk, Goethean, Gogo Dodo, Gommert, Gonzonoir, Grika, Gunderburg,
Hbar12, Headbomb, Hqb, IPSOS, Ian Pitchford, Ignoranteconomist, Ilmari Karonen, Inwind, J04n, JWSchmidt, Jaganath, Jak86, Janechii, JocK, John Vandenberg, JoseREMY, Jpkole, KTC,
Kasimant, Klonimus, KnowledgeOfSelf, Lambex, Legare, Likebox, LilHelpa, Linas, Lisatwo, Lonappan, Loom91, Lorne ipsum, Lue3378, Magog the Ogre, MarSch, Mario777Zelda,
Materialscientist, Maurice Carbonaro, Mbell, Mejor Los Indios, Michael C Price, Michael Courtney, Michael Hardy, Midnight Madness, Mike sa, Mobius Clock, Mogudo, Morten Isaksen, Mr
buick, Mteorman, Mukake, Mushin, OpenScience, Orton 4 President, Owl, P.wormer, Paeosl, Papa November, Patrick0Moran, Paulc1001, PedroFonini, Petersburg, Petri Krohn, PhySusie,
Pjacobi, Promethean, Provider uk, Proxmi, Psyche825, Ptpare, Qbnmichael, QuasiAbstract, Qwertyus, Qwyrxian, R'n'B, RG2, RJSampson, Ramanpotential, Ranjithsutari, Richard001, Richwil,
RobertG, Robth, Rocknrollanoah, Rose Garden, Rsocol, SCZenz, Sarbruis, SarekOfVulcan, Sciamanna, Scolobb, Sean Whitton, Sergio.ballestrero, Sheliak, Shlomi Hillel, Skizzik, Sneakums,
Sokari, Spoon!, Srleffler, Stephen Poppitt, Steve Quinn, Stevertigo, StradivariusTV, Svick, Swwright, T.M.M. Dowd, T2X, Tankred, Tcnuk, Tealish, Techdawg667, TheProject, Tide rolls,
Todayishere, Tone, Trackerseal, Twas Now, Tyler R, Ujm, UncleDouggie, VanishedUser314159, Voyajer, Vyznev Xnebara, WAS 4.250, Wafulz, Watchayakan, Welsh, Weregerbil,
Wernhervonbraun, WikHead, Wikaholic, Windowlicker, Woohookitty, Xt318, Yensin, Ytrewqt, Yumpis, Zarniwoot, Zundark, , 406 anonymous edits
Old Quantum Theory Source: http://en.wikipedia.org/w/index.php?oldid=422426834 Contributors: AManWithNoPlan, AndreHahn, BD2412, Bento00, Boardhead, Charles Matthews,
Corrector of Spelling, Crowsnest, Fradeve11, GoingBatty, J.delanoy, JoeBruno, Kient123, Likebox, Machine Elf 1735, Maedin, Myasuda, Otto.fox, Owlbuster, Ragesoss, Rjwilmsi, Thurth,
Vsmith, WMdeMuynck, Wikipe-tan, 30 anonymous edits
Quantum Mechanics after 1925 Source: http://en.wikipedia.org/w/index.php?oldid=422591103 Contributors: 06twalke, 100110100, 11341134a, 11341134b, 128.100.60.xxx, 172.133.159.xxx,
1howardsr1, 21655, 3peasants, 4RM0, 5Q5, 64.180.242.xxx, 65.24.178.xxx, A. di M., APH, Aaron Einstein, Abce2, Academic Challenger, Acalamari, Ackbeet, Acroterion, AdamJacobMuller,
Addionne, Addshore, AdevarTruth, Adhalanay, AdjustShift, Adolfman, Adventurer, AgadaUrbanit, Agge1000, Ahoerstemeier, Aitias, Ajcheema, Akriasas, Akubra, Alai, Alamadte, Alansohn,
Ale jrb, Alejo2083, AlexBG72, AlexiusHoratius, Alfio, Alison, Alvindclopez, Amareto2, Amarvc, Amatulic, AmiDaniel, Amit Moscovich, Ancheta Wis, Andonic, Andrej.westermann, Andris,
AndriyK, Andy chase, Angelsages, AnnaFrance, AnonGuy, Anonwhymus, Antandrus, Antixt, Antonio Lopez, Anville, Ap, ApolloCreed, Apparition11, Arespectablecitizen, Arjen Dijksman,
Arjun01, Army1987, Arnero, Arthur chos, Asdf1990, Ashujo, Asmackey, AstroNomer, Astronautics, Asyndeton, AtticusX, AugustinMa, Austin Maxwell, Awolf002, AxelBoldt, Axlrosen, BM,
Bad2101, Badgettrg, Baker APS, Barak Sh, Barbara Shack, Bart133, Batmanand, Bcasterline, Bci2, Bcorr, Bdesham, Beano, Beatnik8983, Beek man, Belsazar, Bender235, Bensaccount,
Bento00, Benvirg89, Benvogel, Beta Orionis, Betacommand, Betterusername, Bevo, Bfiene, Billcosbyislonelypart2, Billybobjow, Bkalafut, BlastOButter42, BlooddRose, Bmju, Bobo192,
Bongwarrior, Bookalign, Borki0, Bowlofknowledge, Brazmyth, Brettwats, Brews ohare, Brina700, Brougham96, Bsharvy, BuffaloBill90, Bustamonkey2003, Byrgenwulf, C.c. hopper, CALR,
CBMIBM, CIreland, CSTAR, CStar, CYD, Caiaffa, Cain47, Can't sleep, clown will eat me, CanadianLinuxUser, Canderson7, Capricorn42, Captain Quirk, Captain-tucker, Carborn1,
CardinalDan, Caroline Thompson, Carowinds, Casomerville, Catgut, Centic, Chairman S., Chao129, Chaos, CharlotteWebb, Charvest, Chaujie328, Chenyu, Cheryledbernard, Chetvorno,
ChooseAnother, Chris 73, Chris5858, Christian List, Christian75, ChristopherWillis, Chun-hian, Chzz, CiA10386, CieloEstrellado, Cigarettizer, Cipapuc, Ciro.santilli, Civil Engineer III,
Clarince63, Clarkbhm, Cleric12121, Closedmouth, ClovisPt, Cmichael, Cometstyles, Commander zander, Complexica, Connormah, Conversion script, CosineKitty, Couki, Cp111, Cpcheung,
Cpl Syx, Creativethought20, Crowsnest, CryptoDerk, Csdavis1, Curlymeatball38, Cybercobra, Cyclonenim, Cyp, DJIndica, DL5MDA, DV8 2XL, DVdm, Da monster under your bed, Dale
Ritter, Dalma112211221122, Dan Gluck, Dan100, DangApricot, DanielGlazer, Dannideak, Darguz Parsilvan, DarkFalls, Darrellpenta, Darth Panda, Das Nerd, Dataphile, Dauto, David Gerard,
David R. Ingham, David Woolley, Dbroesch, Dcooper, DeFoaBuSe, DeadEyeArrow, Deaddogwalking, Decumanus, Delldot, DenisDiderot, DerHexer, Derek Ross, Derval Sloan, DetlevSchm,
Diafanakrina, Dickwet89, Dingoatscritch, Discospinster, Djr32, Dmblub, Dmr2, Doc Quintana, Dod1, Dominic, Donbert, Donvinzk, Dougofborg, Dpotter, Dragon's Blood, Dreadstar, Dreftymac,
Drmies, Dspradau, Dtgm, Dylan Lake, E Wing, EMC125, Ed Poor, Edsanville, Edward Z. Yang, El C, El Chupacabra, ElectronicsEnthusiast, Elenseel, Elitedarklord dragonslyer 3.14159, Ellmist,
Elochai26, Eloquence, Enochlau, Enormousdude, Enviroboy, Epbr123, Eric-Wester, EricWesBrown, Eurobas, Evan Robidoux, Eveross, Ex nihil, Excirial, FF2010, FaTony, Factosis, FagChops,
Falcon8765, Falcorian, Farosdaughter, FastLizard4, Fastfission, Favonian, Fearingfearitself, Feezo, Feinoha, Femto, Figma, Final00123, Finalnight, Firas@user, Firewheel, Firien, Fizzackerly,
Floorsheim, FloridaSpaceCowboy, Foobar, Foober, Four Dog Night, FrancoGG, Frankenpuppy, Freakofnurture, Fredkinfollower, Fredrik, Freemarket, FreplySpang, Frosted14, Fvw, F,
GLaDOS, GTubio, GaaraMsg, Gabbe, GabrielF, Gaius Cornelius, Gandalf61, Gary King, Gcad92, Geek84, Geometry guy, GeorgeLouis, GeorgeOrr, Gfoley4, Giants27, Gierens22, Giftlite,
Givegains, Gjd001, Glennwells, Gnixon, Gogo Dodo, GoingBatty, Gpgarrettboast, Gr33k b0i, Grafnita, Graham87, GrahamHardy, GreenReflections, Greg HWWOK Shaw, GregorB, Grey
Shadow, Grim23, Grunt, Guitarspecs, Guy82914, Guybrush, Gwernol, Gyan, H.W. Clihor, H0riz0n, H1nomaru senshi, Hadal, HaeB, HamburgerRadio, Han Solar de Harmonics, Handsome Pete,
HappyCamper, Harddk, Harp, Harriemkali, Harryboyles, Harrytipper, Hdeasy, Headbomb, Hekoshi, Henry Delforn, Heptarchy of teh Anglo-Saxons, baby, Herakles01, Hermione1980,
Hickorybark, Hidaspal, Hikui87, Hobartimus, Holy Ganga, HounsGut, Hughgr, Huon, Husond, Hut 8.5, Hvitlys, II MusLiM HyBRiD II, IJMacD, IRP, IW.HG, Ilyushka88, Imalad, Imperial
Monarch, Info D, Iquseruniv, Iridescent, Islander, Itub, Ivan tambuk, Ixfd64, Izodman2012, J.delanoy, JForget, JIP, JMK, JSpudeman, JSpung, Jack O'Lantern, Jacksccsi, Jackthegrape,
Jaganath, Jagbag2, Jagged 85, Jake Wartenberg, JakeVortex, Jaleho, JamesBWatson, Jan eissfeldt, Jaquesthehunter, Jason Quinn, Javafreakin, JayJasper, Jc-S0CO, Jebba, Jecar, JeffreyVest,
JinJian, Jitse Niesen, Jj1236, Jmartinsson, Jni, JoJan, JocK, Joe hill, John Vandenberg, JohnCD, Johnleemk, Johnstone, Jon Cates, Jordan M, JorisvS, Joseph Solis in Australia, JoshC306,
Joshuafilmer, Jossi, Jostylr, Jouster, Juliancolton, Jusdafax, JzG, K-UNIT, KCinDC, KHamsun, KTC, Kadski, Kaesle, Kan8eDie, Karol Langner, Katalaveno, Katzmik, Kbrose, Kding, Ke4roh,
Keenan Pepper, Kenneth Dawson, Kesaloma, Ketsuekigata, Kevin aylward, Kevmitch, KeyStroke, Khalid Mahmood, Khatharr, Kingdon, KnowledgeOfSelf, Knutux, Koeplinger, Krackenback,
Krazywrath, Kri, Kroflin, Kross, Kyokpae, Kzollman, L Kensington, L-H, LWF, La hapalo, Lacrimosus, Lagalag, Latka, Laurascudder, Lavateraguy, Le fantome de l'opera, Lear's Fool, Leavage,
Leesy1106, Lemony123, LeoNomis, Leoadec, Lethe, Liarliar2009, Light current, Lightmouse, Ligulem, LilHelpa, Linas, Lindsaiv, Linuxbeak, Lithium cyanide, Lithium412, Lmp883, LokiClock,
Lontax, Looie496, Loom91, Loopism, Looxix, Lotje, Lowellian, LucianLachance, Luk, Luna Santin, Lupo, MBlue2020, MC ShAdYzOnE, MER-C, MK8, MKoltnow, MPS, MR87, Machdeep,
Machine Elf 1735, Mackafi92, Macy, Mafiaparty303, Magister Mathematicae, Magnus, Magog the Ogre, Mahommed alpac, Manuel Trujillo Berges, Marek69, Mark Swiggle, Marks87,
MarphyBlack, Martarius, Martial75, Martin Hedegaard, Materialscientist, Matt McIrvin, Mattb112885, Matthewsasse1, Matty j, Maurice Carbonaro, Mav, Maximillion Pegasus, Mbell, McSly,
Mcdutchie, Mckaysalisbury, Mebden, Megankerr, Mejor Los Indios, Melicans, Menchi, Meno25, Metaldev, Meznaric, Mgiganteus1, Michael C Price, Michael Courtney, Michael Hardy,
Michael9422, MichaelMaggs, MichaelVernonDavis, Michele123, Midnight Madness, Midom, Mifter, Mikaey, Mike Peel, Mike6271, Mikeo, Minimac's Clone, Mitchellsims08, MithrandirAgain,
Mjamja, Mjb, Mk2rhino, MoForce, Modusoperandi, Moeron, Moink, Mojska, Mollymop, Moncrief, Monedula, Moogoo, Moskvax, Mpatel, MrJones, Mrsastrochicken, Mrvanner, MtB, Muijz,
Mustbe, Mxn, Myasuda, Mygerardromance, N1RK4UDSK714, N5iln, Nabla, Nagy, Nakon, Nanobug, Nathan Hamblen, Naturelles, NawlinWiki, NellieBly, Netalarm, Neverquick,
NewEnglandYankee, Nibbleboob, Nick Mks, Nick Number, Nick125, NickBush24, NickJRocks95, Nielad, NightFalcon90909, NijaMunki, Nikai, Nin0rz4u 2nv, Niout, Niz, NoEdward, Noisy,
Norm, North Shoreman, NotAnonymous0, Notolder, Nsaa, Nturton, NuclearWarfare, Numbo3, Nunquam Dormio, Nv8200p, Nxtid, Octavianvs, Odenluna, Oleg Alexandrov, Olivier, Ondon,
Onesmoothlefty, Onorem, Oo64eva, Optim, OrgasGirl, Otolemur crassicaudatus, Ott, OwenX, Owl, Oxymoron83, P99am, PAR, PGWG, Palica, PandaSaver, Papadopc, Pascal.Tesson, Pasky,
Pathoschild, Patrick0Moran, Paul August, Paul Drye, Paul E T, Paul bunion, Paulc1001, Paulraine, Pcarbonn, Pchov, Peace love and feminism, Peter, Peterdjones, Petgraveyard, Petri Krohn,
Pfalstad, Pgb23, Pgk, Phil179, Philip Trueman, Philippe, Philmurray, Phoenix2, PhySusie, Phyguy03, Phys, PhysicsExplorer, Physika, Physprof, Piano non troppo, Pilotguy, Pippu d'Angelo,
Pj.de.bruin, Pjacobi, Pleasantville, Plotfeat, Pokemon274, Polgo, PorkHeart, Positivetruthintent, Possum, Prashanthns, Pratik.mallya, Professormeowington, Pxfbird, Qmtead, Quackbumper,
Quantummotion, Quercus basaseachicensis, Quuxplusone, Qwertyuio 132, Qxz, Qzd800, RG2, RJaguar3, Raaziq, Radzewicz, Rakniz, RandomXYZb, Randy Kryn, Razataza, RazorICE,
Rdsmith4, Reaper Eternal, Reconsider the static, Rednblu, Reinis, Rettetast, Rex the first, Rgamble, Rich Farmbrough, Richard L. Peterson, Richard001, Ridge Runner, Rightly, Rjwilmsi, Rnb,
Rnpg1014, Roadrunner, Robb37, Robbjedi, RobertG, Robertd514, Roberts Ken, RobinK, Rodier, Rogper, Romaioi, Ronhjones, Rory096, Roxbreak, RoyBoy, Royalbroil, Royalguard11, Rursus,
Russel Mcpigmin, RxS, Ryan ley, RyanB88, SCZenz, SFC9394, SJP, SWAdair, Sadi Carnot, Saibod, Sajendra, Samlyn.josfyn, SanfordEsq, Sango123, Sankalpdravid, Saperaud, Savant1984,
Sbandrews, Sbyrnes321, Schumi555, Scott3, ScudLee, Sean D Martin, Search4Lancer, Seawolf1111, Seba5618, SeriousGrinz, SeventyThree, Shadowjams, Shanemac, Shanes, Sheliak, Shipunits,
Shoyer, Signalhead, Silence, Silly rabbit, Sionus, Sir Vicious, SirFozzie, Sjoerd visscher, Skewwhiffy, Skizzik, SkyWalker, Slaniel, Smallpond, Smear, Smyth, SnappingTurtle, Snowolf, Snoyes,
Sokane, Sokratesinabasket, Soliloquial, Solipsist, Solo Zone, Someguy1221, Soporaeternus, SpaceMonkey, Spakwee, Spearhead, Specs112, Spellcast, Spiffy sperry, Spinningspark, Sriram sh,
Srleffler, St33lbird, Stanford96, StaticGull, StefanPernar, Stentor7, Stephan Leclercq, Stephen Poppitt, Stephen e nelson, Steve Quinn, Stevenganzburg, Stevenj, Stevenwagner, Stevertigo,
Stiangk, Stone, StonedChipmunk, Storm Rider, Stormie, StradivariusTV, Sukael, Sukaj, SunDragon34, Superior IQ Genius, Superlions123, Superstring, Supppersmart, Swegei, Swwright,
Syrthiss, TTGL, Ta bu shi da yu, TakuyaMurata, Tangotango, Tantalate, Taral, Tarotcards, Taylo9487, Tayste, Tcncv, Teardrop onthefire, Teles, Template namespace initialisation script, Terra

174

Article Sources and Contributors


Xin, Texture, The Anome, The Captain, The Firewall, The High Fin Sperm Whale, The Mad Genius, The Red, The Spam-a-nata, The Thing That Should Not Be, The beings, The ed17, The wub,
TheKoG, Theresa knott, Thingg, Thunderhead, Tide rolls, Tim Shuba, Tim Starling, TimVickers, Timwi, Titoxd, Tkirkman, Tohd8BohaithuGh1, Tommy2010, Toofy mcjack34, Tpbradbury,
TripLikeIDo, Tripodics, Tromer, Trusilver, Truthnlove, Ttasterul, Tycho, TylerM37, Tylerni7, UberCryxic, Ujm, Ultratomio, Uncle Dick, UncleDouggie, Updatehelper, Urbansky, User10 5,
Utcursch, V81, Van helsing, VandalCruncher, VanishedUser314159, Vb, Velvetron, Venu62, Vera.tetrix, Versus22, Viduoke, Vitalikk, Vlatkovedral, Voyajer, Vsmith, Vudujava, WAS 4.250,
WMdeMuynck, Wafulz, Waggers, WarEagleTH, Warbirdadmiral, Wareagles18, Wavelength, Waxigloo, Wayward, Weiguxp, Welsh-girl-Lowri, Werdna, WereSpielChequers, Weregerbil, White
Trillium, Wik, WikHead, Wikiklrsc, Wikipelli, Willieru18, WilliunWeales, Willking1979, Willtron, Wimt, Winston365, Wk muriithi, Wknight94, Woohookitty, Word2need, Wquester, Wwoods,
X41, XJamRastafire, XTRENCHARD29x, Xana's Servant, Xanzzibar, Xenfreak, XeniaKon, Xezlec, Xihr, Yamaguchi, Yath, Yeahyeahkickball, Yellowdesk, Yokabozeez, Zakuragi,
Zarniwoot, Zenbullets, Zero over zero, Zhinz, Zigger, ZiggyZig, Zolot, Zondor, Zooloo, Zslevi, , 1972 anonymous edits
Interpretations of Quantum Mechanics Source: http://en.wikipedia.org/w/index.php?oldid=421335311 Contributors: A. di M., Aarghdvaark, Afshar, Agondie, Airplaneman, Alessandro70,
Alienus, Alison, Alkivar, Amourfou, Arichnad, Arjen Dijksman, Ark, Army1987, Bardon Dornal, Bduke, Beetstra, Belsazar, Berteun, Bfiene, Billinghurst, Blainster, Bm gub, Bmju, Brazmyth,
Brews ohare, Burkhard.Plache, Byrgenwulf, C S, C h fleming, CSTAR, Camcolit, Carbon-16, Caroline Thompson, CesarB, Charles Matthews, Cholling, Cic, Citynoise, Crust, Cryptonaut,
CubsMan91, Dan Gluck, Dan Pelleg, Daniel Weissman, Danthewhale, David R. Ingham, Decumanus, Deniskrasnov, Djr32, Dragon's Blood, Dreamer08, Duff man2007, Dvtausk, ElBenevolente,
Elimisteve, Enormousdude, ErkDemon, Frau Holle, Freakofnurture, Gaius Cornelius, GangofOne, Gauge, Geremia, Giftlite, GoingBatty, Golbez, Graham87, Gramschmidt, Gregbard, Guevara's
Revenge, GuidoGer, Gurch, H0riz0n, H2g2bob, Hairy Dude, Headbomb, INic, Ilmari Karonen, Immer in Bewegung, InXistant, Iridescent, JEN9841, JIP, Jackol, Jefffire, John Broughton, John of
Reading, JohnBlackburne, JohnMcFerlane, JohnSankey, Jordgette, Jostylr, Kane5187, Karol Langner, KasugaHuang, Kbh3rd, Keithbowden, Kevin aylward, Kmarinas86, KnightMove,
KnightRider, Kripkenstein, L33tminion, LC, Lacatosias, Laurascudder, Lethe, Light current, Likebox, Linas, Luckstev, Lumidek, Machine Elf 1735, Materialscientist, Maurice Carbonaro, Mbell,
Michael C Price, Michael Hardy, Mike4ty4, Modocc, Mogudo, Naasking, Nordisk varg, Normxxx, Oleg Alexandrov, Omeganumber, Owl, Pablo X, PaddyLeahy, Pak21, Palindromatic, Paolo.dL,
Paul August, Paul D. Anderson, Pekka.virta, Peterdjones, Pfalstad, Physics108, Pjacobi, Pkeastman, Pleasantville, Plumbago, Pmcmahon, Portillo, Protohiro, RK, RQG, Rafaelgr, Rami radwan,
Randallbsmith, Randommuser, Rasmus Faber, Reddi, Redsolarearth, Reedy, Resuna, Rjwilmsi, Roadrunner, Ryan.vilbig, SQL, Samboy, Samlyn.josfyn, Sbyrnes321, Sdornan, Sebastianlutz,
Sheliak, Simetrical, Sokane, Spireguy, Stephen Poppitt, Stevertigo, Tabletop, Tanath, Tarotcards, Tesseract2, Texture, Tgr, Tito-, TonyW, Tparameter, Truthnlove, Ujm, Vashti, Vroman,
WMdeMuynck, Weregerbil, William M. Connolley, Wooster, WuTheFWasThat, Yensin, Yuttadhammo, Zarniwoot, 252 anonymous edits
The Copenhagen Interpretation Source: http://en.wikipedia.org/w/index.php?oldid=423572273 Contributors: 24.93.53.xxx, Afshar, Agge1000, Agger, Akriasas, Alienus, Alma Pater,
Amakuha, Andybiddulph, Ark, Atomota, AxelBoldt, Barbara Shack, Barraki, Batmanand, Bazzargh, Belsazar, Ben Thuronyi, Berkay0652, Bigmantonyd, Blaine Steinert, Bm gub, Bmatthewshea,
Boatteeth, Bonaovox, Bookalign, Boozerker, Bryan Derksen, CLW, CYD, Carey Evans, Cfp, Charles Matthews, Chas zzz brown, Chjoaygame, Complexica, ConradPino, Conversion script,
Cortonin, Cpiral, Crochet, Cybercobra, DAGwyn, DJIndica, Dan Gluck, David R. Ingham, David R. Ingharn, Dicklyon, Dirac66, Djr32, Dominic, Dragon's Blood, Drunken Pirate, Duendeverde,
Edward, Eequor, El C, Ewlyahoocom, Fazalmajid, Fernbom2, Fph, Freakofnurture, Gcbirzan, Geremia, Giftlite, Glenn, Graham87, Gregbard, Gretyl, Guoguo12, Haham hanuka, Hairouna, Harry
Potter, Hbayat, Headbomb, Heron, Igni, Jamshearer, Jeltz, Joel7687, JohnArmagh, Joriki, K-UNIT, KHamsun, Kahn, Kate, Kevin aylward, KillerChihuahua, Kriegman, Kwertii, LC, Light
current, Likebox, Linas, Looie496, Looxix, Lotu, Lumidek, Mac Dreamstate, Manicsleeper, Marcus Qwertyus, Maurice Carbonaro, Mav, Mbell, McGinnis, Mdwh, Michael Hardy, Murf42,
Nabla, Nbarth, Nercromancy?, Nick Number, OlEnglish, ParticleMan, Patrick0Moran, Pcbene, Peterdjones, Phys, Pjacobi, Prickus, Quadell, RCSZach, Rbedford, Realneyc, Rich Farmbrough,
Richard001, Rickard Vogelberg, Roadrunner, Rorro, Ross Fraser, Runningonbrains, Rursus, SalineBrain, Samboy, SaveTheWhales, Sbiki, Scapermoya, Schaefer, Schneelocke, Schuermann,
Sheliak, Sokane, Sperling, Stanford96, Stephen Poppitt, Stevenwagner, Stuidge, Superpaul3000, TYelliot, Tabish q, Tanuki Z, The Anome, TheMathinator, Thecheesykid, TheloniousBonk,
TimidGuy, Timo Honkasalo, Timwi, Tparameter, Ujm, Unyoyega, Uppland, User2004, Venny85, Victor Gijsbers, Vssun, Vyznev Xnebara, WMdeMuynck, Wjmallard, Xezbeth, Yonatan, 200
anonymous edits
Principle of Locality Source: http://en.wikipedia.org/w/index.php?oldid=421496878 Contributors: Agge1000, Avirab, Boson, Caroline Thompson, Charles Matthews, Clicketyclack,
ClockworkSoul, Cybercobra, Dave Kielpinski, Deathphoenix, Dugosz, Ehrenkater, Ettrig, Evercat, Grutter, Hairy Dude, Jmnbatista, Joriki, JorisvS, LarsPensjo, LedgendGamer, Len Raymond,
Leobold1, Linas, Lumidek, Maurice Carbonaro, Maxw41, Michael C Price, Mpatel, NOrbeck, Naasking, Nwbeeson, Onebyone, Peterdjones, Petri Krohn, Phys, Populus, Ruud Koot, Ryan Roos,
Stephen Poppitt, Stirling Newberry, Traitor, Tweisbach, Vyznev Xnebara, Wiggles007, Zvis, 58 anonymous edits
EPR Paradox Source: http://en.wikipedia.org/w/index.php?oldid=421489020 Contributors: -- April, .:Ajvol:., 207.171.93.xxx, 2over0, ASCWiki, Agge1000, Agger, Alan McBeth, Alkivar,
Amakuha, AmarChandra, Andejons, Apoc2400, Ark, Arpingstone, AxelBoldt, B4hand, Ballhausflip, Bevo, Bigbluefish, Binksternet, Blauhart, Bob K31416, Brews ohare, Brianga, Brrk.3001,
Bryan Derksen, CSTAR, CYD, Camrn86, Canadian-Bacon, Carbuncle, Cardmagic, Caroldermoid, Caroline Thompson, Chas zzz brown, Chjoaygame, Chris 73, Clarityfiend, CobbSalad,
Complexica, Cortonin, Cpiral, Craig Pemberton, Crowsnest, DBooth, Darktaco, David R. Ingham, Declare, Dogcow, Dpbsmith, Dr Smith, DrBob, DrChinese, DragonflySixtyseven, Dryke,
Dzhim, Dugosz, EdH, Edward, Eequor, Egmontaz, Ehn, Eiffel, Ejrh, Ettrig, Falcorian, Finemann, Finlay McWalter, Fredkinfollower, Frish, Fulldecent, Fwappler, GeorgeMoney, Goldfritter,
Graham87, GregorB, Gretyl, Hackwrench, Harald88, Headbomb, Hephaestos, Hirak 99, Houftermann, Hugo Dufort, Hydnjo, IAdem, Iamthedeus, J-Star, JaGa, JamesMLane, Jan-ke Larsson,
Jcajacob, Jinxman1, Jnc, JocK, JohnBlackburne, Jpittelo, Jwrosenzweig, KHamsun, Kapalama, Karada, Karol Langner, KasugaHuang, Keenan Pepper, Kuratowski's Ghost, Larsobrien, Lethe,
Lf89, LiDaobing, Linas, Linus M., Looxix, Lumidek, Maher27777, Marek.zukowski, Marie Poise, Mark J, Masudr, Maurice Carbonaro, Mav, Metron4, Michael C Price, Michael Hardy, Moink,
MrJones, Msridhar, Myrvin, Naddy, Natevw, NathanHurst, Nickyus, Noca2plus, ObsidianOrder, Orange Suede Sofa, Owen, Pace212, Paranoid, Pateblen, Peashy, Pekka.virta, Peter Erwin,
PeterBFZ, Peterdjones, Phancy Physicist, PierreAbbat, Prezbo, Publicly Visible, Pwjb, Prez, RTreacy, Radiofriendlyunitshifta, Rama, Razimantv, Rich Farmbrough, Roadrunner, Robert K S,
Robertd, Robertefields, Ronjoseph, Rracecarr, Ryft, SGBailey, Sanders muc, Schneelocke, Seb, Shadanan, Shakyshake, Shalom Yechiel, Sidasta, Skierpage, Snoyes, Srleffler, Stain, Stephen
Poppitt, Steve Quinn, Suisui, Sundar, Syko, Tarotcards, Tempshill, Tercer, Texture, ThomasK, Thrain2, Timwi, Tlabshier, Tsop, Ty8inf, VanishedUser314159, Vasi, Victor Gijsbers, Vodex,
Voyajer, WMdeMuynck, Weekwhom, Wik, Wile E. Heresiarch, William M. Connolley, WillowW, XJamRastafire, Xgrrr, Xnquist, YUL89YYZ, Yill577, Zeycus, Zootm, , 279
anonymous edits
Bell's Theorem Source: http://en.wikipedia.org/w/index.php?oldid=423646843 Contributors: A. di M., AC+79 3888, AdamSiska, Agge1000, Aldux, AmarChandra, Anakin101,
Andrewthomas10, Android Mouse, Antonielly, Aranel, ArnoldReinhold, Ashmoo, Avb, B9 hummingbird hovering, Ballhausflip, Barticus88, Baxxterr, BeatePaland, Bender235, BeteNoir, Bo
Jacoby, BobKawanaka, Bobo192, Bongwarrior, Brazmyth, BrianWren, Bryan Derksen, Byrgenwulf, CSTAR, CYD, Cacycle, Calvero JP, Caroline Thompson, Catchanil, Cgingold, Chalst,
Charles Matthews, Christopher Cooper, Complexica, Count Iblis, Cremepuff222, Crocodealer, Curious1i, Dauto, Dave Runger, Deathphoenix, Deniz195, Dfrg.msc, Dirac66, Dmr2,
DomenicDenicola, DrChinese, Drilnoth, Drmies, Dtgriscom, EcoMan, EdC, Eequor, Egg, Endlessmike 888, Fastily, Franl, Frish, Fulldecent, Fwappler, GangofOne, Gene Nygaard, Geremia,
Giftlite, Gmusser, Gregbard, GregorB, Grok42, Gsjaeger, Guy Harris, Gzornenplatz, HaeB, Hairy Dude, Harald88, Headbomb, Hmonroe, Hqb, IWillNeverLearn, Incnis Mrsi, Informatorium,
Interintel, Isaacdealey, Isocliff, J-Wiki, Jack who built the house, Jcobb, Jim E. Black, Jim.belk, Jncraton, Jpittelo, Jwpitts, KHamsun, Keenan Pepper, Keithbowden, Kingmundi, Kmmertes,
KnowledgeOfSelf, Konstable, Kyrisch, LC, Laudaka, Leobold1, Lethe, Likebox, Linas, Lumidek, Lycurgus, Malcohol, Maniatis, Marek.zukowski, MathKnight, Maxw41, Michael C Price,
Michael Hardy, Mike2vil, Mister fu-ck you, MisterSheik, Mmyotis, Monslucis, Myrvin, Nihil, Nilock, Nowakpl, Oakwood, Oleg Alexandrov, Olek Bijok, Orangedolphin, Ott2, Paddu,
Patrick0Moran, Paul August, Paulginz, Petri Krohn, Physicskev, Physis, Pjacobi, Plrk, Pokipsy76, Quibik, RJFJR, RK, Rbodgers, Rich Farmbrough, Richard75, Richwil, Rickard Vogelberg, Rl,
Roadrunner, Robert1947, Rodasmith, Roke, Rror, Ryanmcdaniel, SJLPP, Sabri Al-Safi, Salsb, Seth Bresnett, Simetrical, Skeppy, Stdjmax, Steady unit, Stephen Poppitt, StevenJohnston,
Stevertigo, Stirling Newberry, Suffusion of Yellow, Susvolans, Syncategoremata, Tethys sea, Tothebarricades.tk, Tox, TriTertButoxy, Ueit, Ulrich Utiger, Vessels42, WMdeMuynck,
Waleswatcher, Wfku, Wikikam, WookieInHeat, Wwoods, Ybband, Yecril, Yuttadhammo, Zeycus, Zvis, 238 anonymous edits
Schrdinger's Cat Source: http://en.wikipedia.org/w/index.php?oldid=422955766 Contributors: -jmac-, .:Ajvol:., 100110100, 198, 24.93.53.xxx, 64.12.104.xxx, ABCD, AEdwards, AVJP619,
Access Denied, Adashiel, Adpete, Af648, Agamemnon2, Agge1000, Albanharrison, Aleenf1, Alfredr, Amakuru, Amcbride, Amorymeltzer, AnOddName, Andrewjlockley, Andrewpmk, Angr,
Angry Lawyer, Anonymous Dissident, Antandrus, Anupamsr, Anville, Aperisic, Asbestos, Ashley Pomeroy, Ashmodai, AstroNomer, Avb, Bakkster Man, Beeblebrox, Benhocking, Bensmith53,
Betterusername, Bhoch14, BigFatBuddha, Bill Spencer5, Bjdehut, Bmicomp, Bob K31416, BobKawanaka, Bobby D. Bryant, Bobbydowns, Bobo192, BookgirlST, Borisblue, Bovineone, Brews
ohare, Brion VIBBER, Brookston, Broski, Brusegadi, Bryan Derksen, Bth, Bubba hotep, Bucephalus, CShippee, CYD, Calvinkrishy, Camembert, Canadian Paul, Canterbury Tail, Capricorn42,
Carbuncle, Catbag, Catgut, Ccscott, Centrx, Cgay88, Chaos, Charlesrkiss, Chase2032, Chemguy2, Chetvorno, Chitomcgee, Chris Longley, ClockwrokSoul, Closetsingle, Cmsreview, Cohesion,
Commonbrick, Complexica, Comrade009, Conversion script, Corporal Tunnel, Cpl Syx, Craw's Revenge, Crazynas, Crosbie Fitch, Crystallina, DAGwyn, DAID, DJ Clayworth, DJIndica, DV8
2XL, Danlibbo, Darkwhistle, Darth Panda, David R. Ingham, David R. Ingharn, David.Monniaux, Davidizer13, Davidpdx, Dc987, Den fjttrade ankan, Derwyk, Dfhussey, Dhatfield,
Discospinster, Djpadz, Dmoulton, Doorag1, Dr.Gonzo, Dr.K., Drmies, Dspradau, Eaglizard, Eblingdp, Ed Cormany, Eeekster, El C, Elation, Ennerk, Epbr123, Eric Shalov, Escape Orbit, Evil
Monkey, Falcon8765, Faseidman, Faseidrnan, Feldspaar, Fenway4201912, Fetchcomms, Fibonacci, Fieldday-sunday, Floria L, FlorianMarquardt, Foobar, Forai, Francis Ocoma,
FriendlyRiverOtter, Gabbe, Gagueci, Galaxiaad, Gawaxay, Giftlite, Giggy, Gilly3, Gore Physicist, Graham87, Greenleaf, Gregbard, Greudin, Grunt, Gruntree, Gurch, HDarke, HPRappaport,
Haakon, Hadal, HaeB, Hamiltondaniel, Harkleroad, Hasek is the best, Headbomb, Henning Makholm, Heracles31, Heron, Hillbrand, Hires an editor, Hmhur, Holylampposts, HorsePunchKid,
Hovden, Howellpm, I am a Wikipedian, IAmAndrewSoul, IRP, Intelligenttheorist, Ixfd64, J.delanoy, JE, JIP, JPX7, Jack Mickkelson, Jaganath, Jayden54, Jblatt, Jcfolsom, Jeltz, Jim Birkenshaw,
Jim McKeeth, Jjamison, Joao, John, JohnArmagh, JohnSankey, Jordgette, Josh Grosse, Jtmichcock, JustAddPeter, K, KagakuKyouju, Kanthoney, Karl, Kevin Forsyth, Kevin aylward,
KittySilvermoon, Klonimus, Koyaanis Qatsi, Kukec, Lambiam, Lancereq, LeadSongDog, Lear's Fool, LearningIsLiving, Lectonar, Lefty, Light current, Lignomontanus, Likebox, Linas,
Lkstrand, LookingGlass, Lord Anubis, Luckyherb, Lumidek, MER-C, Maelin, MafiaCapo, MagneticFlux, Mako098765, Marc711, MartinHarper, MathKnight, Max Terry, Maziotis, Me.play,
MercuryBlue, Michael C Price, Mieszko the first, Mike Rosoft, MikeBartlett, Mikoangelo, Mild Bill Hiccup, Mindmatrix, Minimac, Mion, Mipadi, Mjodonnell, Mono mano, Monsieur Prolong,
Mrawesomemanforeverwoo, Mrdude, Ms2ger, Mungukwachupi, Muu-karhu, NTK, Nabla, Nabokov, Nakon, Nard the Bard, Nessus, Nick Mks, Nile, No Guru, NormanEinstein, Northside777,
Nullx42, O18, Old Moonraker, Opelio, Optichan, OranL, OrangeAipom, Oski97, Otis182, Owen, PHDrillSergeant, Paste, Paul August, Personman, Peter Meyrathing, Peter Tomblin, Peterdjones,
Pfalstad, Pgan002, Pharaoh of the Wizards, Pharos, Philg88, Philip Trueman, Philosophus, Phoenixrod, Picaroon, PierreAbbat, Pinethicket, Piquan, Pitr C, Pladask, Polyparadigm, Prezbo,
Profugum, Protohiro, PseudoSudo, Psternenberg, Pureaddiction, Quaestor23, Quandon, RG2, RK, RMFan1, Raisesdead, RandMC, RandomCritic, Reach Out to the Truth, Rearete, RedWolf,

175

Article Sources and Contributors


Redirect fixer, Reportica, Rich Farmbrough, Richard Weil, RickDeNatale, Riconaps, Riki, Roadrunner, RobertG, Ronaldomundo, Ronhjones, Rwjensen78, Ryan4314, RyanCross, RyoGTO,
SMP, SabineLaGrande, Saebjorn, Saigon punkid, Sam Spade, Samaritan, Sampi, Sanders muc, Sango123, Saperaud, SchnitzelMannGreek, Scimitar, Scoutersig, Sean D Martin, Sean.hoyland,
Seaphoto, Seirscius, Serendipodous, Serpens, Sewblon, Shanes, Sheliak, Shervinafshar, Shizhao, Sidesteal, Silly rabbit, Sillybilly, SimonD, Sjorford, Skinnyweed, Skychrono, Smartneddy,
Smartyhall, Smitz, Smyth, Snillet, Snoyes, Sophitus, Spearhead, Splash, Spring1, Sprowl, Srnitz, Sten, Stepa, Steven J. Anderson, Stevertigo, Sumsum2010, Sweetser, TUF-KAT, Taejo,
TaintedMustard, Tannkrem, Teflon Don, Teutonic Tamer, The Cunctator, The Monster, The Rambling Man, The Thing That Should Not Be, The wub, TheSun, Theendofgravity, Thorncrag, Tide
rolls, Timwi, Tobias Bergemann, Tom Peters, Tommy2010, Toytoy, Tregoweth, Trevor MacInnis, Tufflaw, Twas Now, Typobox, Ukcreation, Ukexpat, Uogl, Useight, Vadim Makarov, Val42,
VanishedUser314159, Varuna, Veghead, Versehalf, VincentCalci, Violetriga, Virogtheconq, VolatileChemical, WCFrancis, WMdeMuynck, Wandall Schwartz, Wapcaplet, Waterfall117, Wenli,
Weregerbil, Wereon, Wetllama, Whitehorse1, Whosyourjudas, Wik, Wikiliki, Wikimancer, Wikipelli, Will2k, Wind-up Spirit, Wisq, Worthawholebean, Wwheaton, XJamRastafire, Yamamoto
Ichiro, Yar Kramer, Yeezil, Yngvarr, Ytrewq05, Yvwv, Zakhebeone, Zeerak88, Zenosparadox, ZeroOne, Zginder, ^demon, , 701 anonymous edits
The Measurement Problem Source: http://en.wikipedia.org/w/index.php?oldid=422392580 Contributors: Apoc2400, ArglebargleIV, Bm gub, Brews ohare, Charles Matthews, Chetvorno,
David R. Ingham, Dchristle, Doradus, Dragon's Blood, EMPE, Edward, El C, Erik Carson, Freakofnurture, Jdforrester, Jean-Christophe BENOIST, Jeffq, JocK, JohnBlackburne, Jostylr, Karol
Langner, Lendtuffz, Likebox, Linas, Lumidek, Machine Elf 1735, Maury Markowitz, Michael C Price, Michael Hardy, Mjszzz, Nihilo 01, Patrick0Moran, Pekka.virta, Peterdjones, Pfa alkemade,
Pfalstad, Phil Boswell, Rjwilmsi, Roberts Ken, RoyBoy, Sam Hocevar, Stevertigo, Tesseract2, Ujm, Victordelpanno, Vyznev Xnebara, Xihr, Zoegernitz, Zwikki, 41 anonymous edits
Measurement in Quantum Mechanics Source: http://en.wikipedia.org/w/index.php?oldid=420115282 Contributors: Alba, Algebraist, Ancheta Wis, Apoc2400, Aqerto, Arpingstone,
Auspex1729, Baltorn, Bdesham, Bm gub, Bobo192, Brenont, Brews ohare, Bronix, Bth, Cohesion, Commander Keane, CosmiCarl, DJIndica, Danlaycock, Dave Kielpinski, Fredrik,
Fuhghettaboutit, Fwappler, GangofOne, Grayshi, H2g2bob, Headbomb, Henry Delforn, Hudon, Incnis Mrsi, J S Lundeen, JFlav, Jake Wartenberg, Jantov, Jeepday, Jetherrie, JocK, Joriki,
Judgesurreal777, Karol Langner, KasugaHuang, Korepin, Kzollman, Likebox, Linas, Lofor, Macmelvino, Makiteo, MathKnight, MathMartin, Matthew Yeager, Maurice Carbonaro, Mct mht,
Michael C Price, Michael Hardy, MountainSplash, NJWard, Peterdjones, Philip Trueman, Phys, RCSB, Rjwilmsi, RobHar, Robert L, STHayden, SWAdair, Sam Hocevar, Sbyrnes321,
SeventyThree, Sheliak, Smack, Spoon!, StaticGull, Stevertigo, Stongli, TakuyaMurata, Tedickey, Ujm, WMdeMuynck, Who, William M. Connolley, Yaoyun.shi, Zarniwoot, 74 anonymous edits
Quantum Number Source: http://en.wikipedia.org/w/index.php?oldid=423530392 Contributors: Acalamari, Albrozdude, Aliekens, Allmightyduck, Arthena, Arthur Rubin, Azure8472,
Bambaiah, Barticus88, Brownsteve, BryanD, Bubbachuck, Chaos, Charles Baynham, Choihei, Christopherlin, Cobaro, Complexica, Craig Pemberton, Cubbi, DARTH SIDIOUS 2, DJIndica, Dan
Gluck, DaveTheRed, Davemcarlson, Decumanus, Demeza13, Dmmaus, Donarreiskoffer, Drova, EddEdmondson, Edsanville, Fred Bradstadt, Fresheneesz, Gfoley4, Giftlite, Go2slash, Hadal,
Hairy Dude, HappyCamper, Hidaspal, Im.a.lumberjack, Imnotoneofyou, Itub, Ivy martin08, J.delanoy, JabberWok, John, Joyonicity, Just granpa, Kanags, Kareeser, Katoa, Kushal one,
Larryisgood, Law, Maatghandi, ManoaChild, Mets501, Mpatel, Mrfair, Mushin, Musicality213, Mxn, NewEnglandYankee, Nsaa, PV=nRT, Phantombantam, Qmonkey, Ravilovefriends,
Redrose64, RetiredUser2, RobinK, Sealbock, SeventyThree, Somdebg, SpK, SpeedyGonsales, Spoon!, Stevertigo, Suspekt, Sverdrup, Syntax, Teply, That Guy, From That Show!, The Anome,
Tim Starling, Tommy2010, Trapezoidal, Trevor MacInnis, Tsemii, Venny85, Versus22, Voyajer, Wake chaser, Waxigloo, Waza, WikiUserPedia, Xavic69, Yevgeny Kats, Yoshigev, Zahd, 203
anonymous edits
Quantum Information Source: http://en.wikipedia.org/w/index.php?oldid=422148084 Contributors: 3malchio, Abdelaty, Alinihatekenwiki, Ammarsakaji, Antandrus, Ballhausflip, Bearcat,
CSTAR, Charles Matthews, Chemako0606, Cipapuc, Ckwongb, Cnilep, Conversion script, Creidieki, Crusio, DFRussia, DGG, Dave Kielpinski, Dcoetzee, Dgrant, Dojarca, Gene.arboit, Isnow,
Itchy, J S Lundeen, Jamelan, KathrynLybarger, Kenneth M Burke, Kilor, Kipton, Knotwork, Korepin, Linas, Looie496, Mandarax, Materialscientist, Maurice Carbonaro, Mct mht, Michael
Hardy, Naddy, OrgasGirl, Pengo, Phys, Pieterkonings, Robertvan1, Sanders muc, Seb, Shepelyansky, Shishir0610, Skippydo, Stevertigo, Template namespace initialisation script, The JPS, Tracy
Hall, Uukgoblin, Voorlandt, 46 anonymous edits
Quantum Statistical Mechanics Source: http://en.wikipedia.org/w/index.php?oldid=344774261 Contributors: B7582, CSTAR, Dani setiawan, Dyunvort, JamesTeterenko, Jheald, Linas, Mct
mht, Mets501, PAR, Passw0rd, Patrick, Phys, Sam Hocevar, Sietse Snel, 7 anonymous edits
Quantum Field Theory Source: http://en.wikipedia.org/w/index.php?oldid=420004469 Contributors: 130.64.83.xxx, 171.64.58.xxx, 9.86, APH, Acalamari, Ahoerstemeier, Albertod4, Alfred
Centauri, Alison, AmarChandra, Amareto2, Amarvc, Ancheta Wis, Andrea Allais, Archelon, Arnero, AstroPig7, Bambaiah, Bananan, Banus, Bci2, BenRG, Bevo, Bkalafut, Bktennis2006,
Blckavnger, Bob108, Brews ohare, Bt414, CALR, CYD, CambridgeBayWeather, Caracolillo, Charles Matthews, Ciphers, Complexica, ConCompS, ConradPino, Conversion script, Custos0,
Cybercobra, Cyp, D6, DJIndica, Dan Gluck, Dauto, Davidaedwards, DefLog, Dj thegreat, Dmhowarth26, Docboat, Drschawrz, Dylan1946, Egg, El C, EoGuy, Erik J, Etale, Ettrig, Fortune432,
Four Dog Night, Gandalf61, Garion96, Gcbirzan, GeorgeLouis, Giftlite, Gjsreejith, Glenn, GoingBatty, Graham87, GrahamHardy, HEL, Hairy Dude, Hanish.polavarapu, Headbomb, Helix84,
Henry Delforn, HexaChord, Heyheyheyhohoho, Hickorybark, HowardFrampton, Ht686rg90, Hyqeom, IRP, IZAK, Igodard, Igorivanov, Iridescent, Itinerant1, J.delanoy, JEH, Jamie Lokier, Jecht
(Final Fantasy X), Jeepday, Jim.belk, Jjspinorfield1, Jmnbatista, Joseph Solis in Australia, Juliancolton, KarlHallowell, Karol Langner, Kasimann, Kenneth Dawson, Knowandgive, Kromatol,
LLHolm, Lambiam, Leapold, Lerman, Lethe, Likebox, LilHelpa, Looxix, Lotje, Maliz, MarSch, Marie Poise, Markisgreen, Marksr, Martin Kostner, Masudr, Mav, Mbell, Mboverload, Mcld,
Melchoir, Mendicus, Mentifisto, Mgiganteus1, Michael C Price, Michael Hardy, Moltrix, Moose-32, Mpatel, Msebast, MuDavid, N shaji, Neparis, Newt Winkler, Nick Number, Nikolas Karalis,
Niout, Northgrove, Northryde, Nvj, Odddmonster, Opie, PSimeon, Palpher, Paolo.dL, Pcarbonn, Phys, Physicistjedi, Pinkgoanna, Pjacobi, Pt, Puksik, QFT, R.ductor, R.e.b., RE, RG2, Raphtee,
Reaper Eternal, Rebooted, Rjwilmsi, Roadrunner, Robert L, RogueNinja, Rotem Dan, Rudminjd, Rursus, RuudVisser, SCZenz, SebastianHelm, Semperf, Shanes, Sheliak, Shvav, Sietse Snel,
SirFozzie, Skou, Spellage, Srleffler, St3vo, Stephan Schneider, Stevertigo, StewartMH, Stupidmoron, Tamtamar, Template namespace initialisation script, Tetracube, The 80s chick, The Anome,
The Original Wildbear, The Wild West guy, Thialfi, Thingg, Threepounds, Tim Starling, TimNelson, Timwi, Tlabshier, Tobias Bergemann, Tom Lougheed, Truthnlove, Tugjob, UkPaolo,
Urvabara, Van helsing, Vanderdecken, Varuna, Victor Eremita, Wavelength, Wik, Witten Is God, Wwheaton, XJamRastafire, Yndurain, Zarniwoot, 298 anonymous edits
String Theory Source: http://en.wikipedia.org/w/index.php?oldid=423339760 Contributors: --=The Doctor=--, 100110100, 137.111.13.xxx, 195.5.70.xxx, 1exec1, 4lex, 4v4l0n42, 4wajzkd02,
62.104.216.xxx, 69gangsta420, A. di M., A23649, ABigGreenHippo, Aaron35510, Ab1, Academic Challenger, Adamstevenson, Addshore, Aero66, Aero77, Afteread, Ahoerstemeier, Aidan
Croft, Aigisthos, Aiyda, Aktsu, Alaithiran, Alansohn, Alex.muller, AlexGWU, Alexcalamaro, Alexdamaino9, Alexdeburca18, Algebraist, Algri, AlphaEta, AlphaNumeric, Amorim Parga,
Anaraug, Andrej.westermann, AndrewWTaylor, Andrewlp1991, Andrius.v, Angela, Angela26, AngryBacon, Angus Lepper, Anna Frodesiak, Anonymous Dissident, Anpecota, Antandrus,
Antimatter15, Anville, Arabani, Archanamiya, ArchonMagnus, Areldyb, ArielGold, Arjun01, Artaxiad, Arthurcprado, Aruben537, Asafavi, Astiburg, AstroGod, Athelwulf, Atomota, Audree,
Auspex1729, Avihu, Awolf002, AxelBoldt, Axl, Ayleuss, Az29, Azeraphale, Aziz1005, Aznhero3793, B9 hummingbird hovering, BVBede, Babawhitemoose, Babemachine, Babemonkey,
BahTab, Bandit5005, Baronnet, Bart133, Bassbonerocks, Batong, Bbatsell, Bdiah, Beland, Bemoeial, Ben Tibbetts, Benjamin1414141414141414, Bethovenn, Bevo, Bewildebeast, Bigdaddy4x4,
BillC, Bkazaz, Blackhall616, Bobo192, Bobrayner, Bongoo, Bongwarrior, Borgx, Borisblue, Born Gay, Bovineone, Brahmajnani, Branman515, Brews ohare, Brian0918, Brianlucas, Brina700,
Bryan Derksen, Btphelps, CBM, CHIAGEHYANG, COGDEN, CRGreathouse, Caidh, Cal 1234, Calwiki, Can't sleep, clown will eat me, CanadianLinuxUser, Canthusus, Capricorn24,
Capricorn42, Captain-tucker, Capudo, Carandol, CardinalDan, CaseyPenk, CatFiggy, Catgut, Causa sui, Caz34, Cenarium, CesarB, Ceyockey, Cg-realms, Charibdis, Charles Matthews,
Charlesvi, ChazBeckett, Cheesus, Chowbok, Chris 73, Chris Brennan, Chrispaps2413, Christopher Parham, Christopher Thomas, Ckatz, Cleared as filed, Cleeseheb, Closedmouth, CloudNine,
Coltonhs, Complexica, Computerjoe, Conorific, Conversion script, Cool3, Couldbenoway66, Courcelles, Cpl Syx, Craig Stuntz, Crazycomputers, Crazynas, Crazytales, Crohnie, Csigabi, Cst17,
Cureden, Curps, Cybercobra, Cyferx, Cyrus Andiron, D6, DAGwyn, DHN, DJBullfish, DKqwerty, DMacks, DV8 2XL, DVD R W, Dan Gluck, Danga1988, Daniel Olsen, Dante Alighieri, Dark
dude, David136a, David424, Davidmedlar, Daynightrader, Db63376, Dbelange, Dc987, DeletionUK, Den fjttrade ankan, Denali134, Dewayne76, Dhatfield, Dirac66, Directorstratton,
Discospinster, DoItAgain, Doc Tropics, Dotdotdotdash, DoubleBlue, Dougweller, Dpotter, Dr. Aakash Patel, Dreamafter, Drseudo, Drweetmola, Duchesserin, Duduong, Duncan McB,
Dunkaroo207, Dwielark, Dylan Lake, Dysepsion, Earthlyreason, Ed Fitzgerald, Ed270791, Editmyhandman, Edward, Eean, Eeekster, Egemont, Egmontaz, Ekwos, El C, ElBenevolente, ElTyrant,
Elvey, Emarv, Enauspeaker, Encyclopadia, Enirac Sum, Ensign beedrill, Enviroboy, Epastore, Epbr123, Ergative rlt, Eric Drexler, Erik4, ErikStewart, Erth64net, Escher26, Etov, Evil saltine,
Exert, Eyelidlessness, FF2010, Fadereu, Faigl.ladislav, Falcorian, Fanghong, Felix Wan, Fences and windows, Ferkelparade, FireMouseHQ, Fizzycyst, FlieGerFaUstMe262, Fortdj33, Fotoni,
Frankenpuppy, FredrickS, Fredrik, Fredsie, Froid, Fropuff, Fuhghettaboutit, Fuzzy Bob Saget, GSlicer, Gabbe, Gadfium, GaeusOctavius, Gaius Cornelius, Galwhaa, Gary King, Gautam10,
Gebrah, Geekdom04, Gene Nygaard, Gensanders, Geoffrey.landis, GeorgeOrr, Georgepowell2008, Geschichte, Ggb667, Giftlite, Giko, Gil Gamesh, Gil987, Gitman4, GlassCobra, Gmusser,
Gogo Dodo, Goldenglove, Green Lane, Greenrd, GregorB, Gtstricky, Gurch, Guy392, HEL, HFarmer, HJ Mitchell, Hadal, HaeB, Haemo, Hairy Dude, HamatoKameko, Harold f, Harp,
Harryboyles, Headbomb, Heidisql, Helicoptor, Hempfel, Hephaestos, Heqs, HerpesVirus, Hillbrand, Hillman, Hindol, Hitchhiker89, HowardFrampton, Hqb, Humanino, Huntscorpio, Huperphuff,
Hvitlys, Hydrogen Iodide, I dream of horses, I hate to register, IRP, Ikiroid, Ilmari Karonen, Immunize, ImperialismGo, Impunv, Insineratehymn, Introductory adverb clause, InvisibleK,
Invisifan, Ionfield, Iridescent, Irish Souffle, IronGargoyle, Isaac, It Is Me Here, J.A.Ireland, BA (IHPST), J.delanoy, JForget, JIP, JJ Harrison, JYolkowski, Jacobolus, Jagbag2, Jake Wartenberg,
James McNally, JamesBWatson, Jamesofur, Janus Shadowsong, JavierMC, Jay Litman, Jdforrester, Jdlambert, Jespinos, Jibal, Jimbo Mahoney, Jimp, Jitse Niesen, Jkl, Jmcc150, JoJan, John
Vandenberg, JohnBlackburne, Joke137, Jomoal99, JonezyKiDx, Jonnyk aus, Joriki, Joshua Davis, Jrcla2, Juan Marquez, Julesd, Jusdafax, Justanother, JustinHagstrom, K, Kaenneth, Kahananite,
Kartano, Keegan, Keegscee, KerathFreeman, Kevinmon, Khaosworks, Khlo, Kim Bruning, Kirbytime, Kjoonlee, Kjramesh, KnightLago, Knotwork, KnowledgeOfSelf, Kocio, Konstable, Kriak,
KrishSundaresan, Kronix35, Kurochka, Kusma, Kvdveer, Kwiki, Kylemew, Kzzl, L Kensington, Lacrimosus, Laurascudder, Lbr123, LeaveSleaves, LedgendGamer, Leflyman, Lemonflash,
Leujohn, Lightdarkness, Likebox, Linas, Lindmere, LindsayH, Linnell, Linuxlad, Lionelbrits, Little Mountain 5, Livajo, Looxix, Lostart, Lozeldafan, Luke Walkerson, Lulzprotuns, Lumidek,
Luna Santin, Lunch, M C Y 1008, M00npirate, M1k3 101, MER-C, MONGO, MSTCrow, MacsBug, Madagaskar07, Maedin, Magister Mathematicae, MagnaMopus, Malcolm Farmer,
Manheat84, Mani1, Manzeet, MaooaM, MarSch, Marcdean123, Martin451, Martti Muukkonen, Mashford, MathStuf, Mathgenius3141592, Matthew Sanders, Maurice Carbonaro, Mav,
Maximilian77, Mbell, Mboverload, Mcarling, Mdl53711, Mel Etitis, Melchoir, Mentifisto, Mephistophelian, Merenta, Merope, Mgiganteus1, Mhking, Michael C Price, Michael Hardy,
MichaelBillington, MichaelMaggs, Mick le pick, Mike Doughney, Mike18xx, Mikebernstein, Mikez, Milkocookie, Minimac, Mintrick, Mirv, MissMJ, Mkroh, Mmortal03, MontanNito,
Moronoman, Morris729, Moshe Constantine Hassan Al-Silverburg, Moyogo, Mpatel, Mporter, Myasuda, NPguy, Nakon, Narayanese, Naturespace, NawlinWiki, Ned Scott, Neilanderson,
Nergaal, Nestamachine, NewEnglandYankee, Newbyguesses, Nickslspride34, NithinBekal, No Guru, NoPetrol, Nobelprizewinner, Nomad, Nonsuch, Norm mit, Northfox,
Notpayingthepsychiatrist, NuclearFusion, NuclearWarfare, Nufy8, Nunquam Dormio, O, OlEnglish, Olaf Davis, Oli Filth, Olivier, Ollainen, Omegatron, Omgwaffels, Omnipaedista, OnePt618,

176

Article Sources and Contributors


Onorem, Opelio, Ophion, Ornamentalone, Oswaldo Zapata, Oxymoron83, Pallab1234, Panser Born, Panu, Paul August, Pavel Vozenilek, Pentasyllabic, Pepsidrinka, Perlman10s, Perspectival,
Peteryoung144, Philip Trueman, Phils, Phoenix2, Phys, PhysPhD, Physicistjedi, Phyte, Pigman, Pimpin101, Pip2andahalf, Pippo2001, Pjacobi, Plumbago, PoisonGM, Poli-Psy, Polyamorph,
Poshzombie, Pra1998, Pred, Pretty Green, ProGeek314, Prodego, Pwqn, Quadell, Quajafrie, Quaker phil, Quaqa, Qutt, Qweeveen, Qxz, R Calvete, R3m0t, Rabinzkaman, Rawling, RayAYang,
Raymond Hill, Razorflame, Rd232, Reaper Eternal, Reaperfromhell, Reconsider the static, Red Act, Rediahs, Remingtonhill1, Res2216firestar, RexNL, Reyk, Rhetoric Of A Sophist, Rhobite,
Rich Farmbrough, Rickstauduhar, Ringbang, Rjwilmsi, Rnricklefs, Robdunst, RobertM525, Ronhjones, Rory096, Roy Fultun, RoyBoy, Rpcappello, Rwmnau, RxS, Ryanjunk, Ryanmcdaniel,
S3000, SCZenz, SColombo, SDJ, SRoughsedge, SSJ 5, SWAdair, Saalstin, Salsa Shark, Sardinita, Sarvagnya, Sashhere, Savant13, SchfiftyThree, Schneelocke, SchnitzelMannGreek, Sciurin,
Scoobey, Scott1329m, Scourge of God, Sct72, Scwlong, Seahorseruler, Seaphoto, Seraphim, Sh4wz0r, Shamvil, Shanel, Shanes, Sheliak, Shidobu, Shirik, Shlomi Hillel, Sideswiper, Silence,
Silly rabbit, Simon_J_Kissane, Sjakkalle, SkepticVK, Skrewtape, SlashDot, Smamaret, Smokizzy, Sodium, Solancel, Solipsist, SpacePyjamas, Specs112, SpiderJon, Spiff, Spliffy, SpuriousQ,
Staffwaterboy, Stephenb, Steuard, Steven66s, Steveo27five, Stevertigo, Storm.sarup, StormbringerUK, Street Scholar, String4d, StringLove, StringyGuy, Stu181, Suisui, Sunoco, Superior IQ
Genius, Sverdrup, Swisstingle, Syebo, Synchronism, T-dot, T00g00d96, TERBAFAN, TKD, Taco Manipulator, Tallboyhoops1991, TallulahBelle, Tamaratrouts, Tarquin, Tassedethe, Taw,
Tayste, Tcturner2002, TeaDrinker, Teardrop onthefire, Teleprinter Sleuth, Terrifictriffid, Terrisknickers, Tgeorgescu, That Guy, From That Show!, The Thing That Should Not Be, The tooth,
TheRingess, TheTito, Theresa knott, Thesis4Eva, Thincat, Thingg, Three887, Thubsch, Thumperward, Tide rolls, Timeitsways, Timneu22, Timothy Clemans, TimothyRias, Timwi, Tobias
Bergemann, Tom Lougheed, Tom harrison, Tommy2010, Tong, Tony Fox, Torrazzo, Tothebarricades.tk, Trent, Trevor MacInnis, Truthnlove, Tschach, Turgan, Twri, Tycho, UU,
UberScienceNerd, Ucanlookitup, Ugha, Unique and proud of it, Uruk2008, Urvabara, Vald, Van helsing, VanishedUser314159, Vanka5, Vapour, Varrey280303, Vasiliy Faronov, Vastly,
Versus22, Verum, Vicarious, Visium, Visor, VolatileChemical, Vollrath2323, Vrkaul, WJBscribe, Waffleboy36, Wakabaloola, Waleswatcher, Watsup1313, Wavelength, Wayne Slam, Where,
White Cat, Wiggl3sLimited, Wiki incorp, WikiFew, WikiZorro, Wikipelli, Winner 42, Witten Is God, Wmahan, Woland37, Wolfmankurd, Wooba doob, XAXISx, Xgamer4, Xhaoz, Xiahou, Y.t.,
Yandman, Yevgeny Kats, You? Me? Us?, Ytomem, Yukiseaside, Zalgo, Zazaban, Zegoma beach, Zelos, Zidonuke, Zifnabxar, Zundark, Zunz, 1651 anonymous edits
Quantum Gravity Source: http://en.wikipedia.org/w/index.php?oldid=423140426 Contributors: 194.117.133.xxx, 200.191.188.xxx, 2over0, Acalamari, Allowgolf, Alvatros, Amareto2,
Anarchimede, Ancheta Wis, Andwor9, Apratim07, Apyule, Army1987, AstroNomer, Avono, B, Bdalevin, Ben.c.roberts, BenRG, Bender235, Bevo, Bobby D. Bryant, Brazmyth, Brews ohare,
Caco de vidro, Cal 1234, CalebNoble, CanadianLinuxUser, CardinalDan, CaseyPenk, Casimir9999, Charles Matthews, Christopher Thomas, Cjthellama, Ck lostsword, Clement Cherlin,
Complexica, Conversion script, Coren, Count Iblis, Craig Pemberton, Cthuljew, DAGwyn, DJIndica, Dabbler, Dac04, DancingPenguin, Daniel Arteaga, DanielBurnstein, David Schaich,
Davidclifford, Delaszk, Dicklyon, Dllahr, DonJStevens, DrArthurRubinPHD, Dude1818, Duduong, Ekwos, El C, Ems57fcva, Endlessnameless, Epbr123, Eric Drexler, ErkDemon, FT2, FayssalF,
Finn-Zoltan, Follyland, Fotoni, Francophile124, Frazzydee, Fropuff, Frosty726, Fullmetal2887, F, GarbagEcol, Garfield Salazar, Giftlite, Googledin!, Gravitophoton, GregorB, Harold f,
Hbackman, Headbomb, Hep thinker, Herbee, Hillman, Hirvenkrpa, Hmonroe, IRP, Igny, Ilya (usurped), IronGargoyle, Itinerant1, JHMM13, Jaganath, Jason Quinn, Jeffq, Jhmmok, JimJast,
JocK, Joke137, JonathanD, JorisvS, Joyous!, Jpod2, Jrasowsky, KasugaHuang, Keenan Pepper, Keraunos, Klasovsky, Knowandgive, Koeplinger, Korepin, KrisBogdanov, Kroggz, Kurtan, L
Kensington, Lagelspeil, Lambiam, Lambtron, LilHelpa, LorenzoB, Lumidek, MER-C, Malyctenar, Marcus2, Marek69, Markus Poessel, Masudr, Materialscientist, Matusz, Mcarling, Mhsb,
MichaelMaggs, Miguel, Modify, Mpatel, Mstuomel, Musictime4me, N0814444, NewEnglandYankee, Notburnt, NuclearWinner, Nucleophilic, Onebravemonkey, ParadiZio, Patrick O'Leary,
PhilHibbs, Phys, Piccor, Pie4all88, Pigetrational, Pjacobi, Pleasantville, Pmokeefe, Polyamorph, QFT, Quantity, RG2, Rabsmith, Raidr, Raul654, Reddi, ReluctantPhilosopher, Rettetast, Rholton,
Rjwilmsi, Roadrunner, Rolfguthmann, Roy Brumback, Rror, SCZenz, Saibod, Sam Staton, Samdutton, Scarykitty, Science writer, Sdedeo, Seattle Skier, Seth Ilys, Sheliak, Shushruth, Silly rabbit,
SirFozzie, Slightsmile, Smithfarm, Soosed, SpikeTorontoRCP, StaticG, StealthCopyEditor, StevenJohnston, Stevertigo, Susurrus, TAz69x, TVC 15, Tamaratrouts, Terra Novus, The Thing That
Should Not Be, Theanphibian, Theonlydavewilliams, ThomasWinwood, Throwmeaway, Tide rolls, Tim Shuba, Timwi, Titan1129, Tms9, ToddFincannon, TonyMath, Trifonov, TrueTeargem,
Truthnlove, Ttimespan, Twunchy, Tycho, Udifuchs, UncleBubba, VBGFscJUn3, Vacuunaut, Vald, Valeriy Pischenko, Vampus, Victor Blacus, Vincenzo.romano, Vitaleyes, Wereon, Wiki
Roxor, Wikiwikimoore, Wireader, Woohookitty, YapaTi, Yevgeny Kats, Yurik, Zunaid, 312 anonymous edits
Quantum Source: http://en.wikipedia.org/w/index.php?oldid=420584388 Contributors: 4C, Academic Challenger, Acalamari, AdjustShift, Ahoerstemeier, AlexiusHoratius, Alison,
Andrewpmk, Andypandy.UK, Anthony Appleyard, Ashmoo, Atlant, Bbq332, Bensaccount, Bicala, Bilbo1507, Bjankuloski06en, Booknotes, BryanD, Bupper, Burakburak, C. Trifle,
Capricorn42, Ceyockey, Charles Matthews, Chymicus, Complexica, Crunchy Numbers, DaGizza, Dan Granahan, Dannyeder, DavidBrahm, Db099221, Deasmumhain, Deathphoenix, Dennis
Estenson II, Djkrajnik, Dr Miles Long, Dragon's Blood, DragonHawk, Drakonicon, Dungodung, El C, Epbr123, EugeneZelenko, Falcon8765, Fdmt, Filelakeshoe, Finn-Zoltan, Fjjf,
FlamingSilmaril, Flightx52, Freakofnurture, Fresheneesz, Garion96, Geoffrey.landis, Gexmeansgecko, Graeme Bartlett, Gragox, Grapeguy7, Gratedparmesan, Guruspiritual, Harishng, HiDrNick,
Hosterweis, Hujaza, Jaan513, Jag123, Jalexsmith1991, Jeff02, JerrySteal, John Vandenberg, Jsjunkie, Juventas, K.zaman1710, Kilmer-san, Kinston eagle, KnowledgeBased, Kurt hueston,
Laplace's Demon, Laurascudder, Leebo, Lethe, Loggin12354, LoveEncounterFlow, LtBert44, Maedin, Marco Polo, Mariodivece, Masudr, Materialscientist, Maurice Carbonaro, Mebden,
Melchoir, Mental Blank, Mhocker, Mikeblas, NCDane, Nacefe, NawlinWiki, NerdyNSK, Ngexpert5, Nihiltres, Nsaa, Nutfortuna, Omoo, Ontarioboy, Openforbusiness, Oreo Priest, Orphan Wiki,
Orthografer, Oxymoron83, Parkyere, Pdcook, Pengo, Platypus222, Programming gecko, Pundit, Quantumpundit, Qwertzy2, RG2, Raidon Kane, Raylu, Ricky81682, RodC, RoyBoy, Sadi Carnot,
Salvatore Ingala, Sam Korn, Sampi, Scorpionman, Seanruiz, Smack, Smartcowboy, Spangineer, Stevenmitchell, Stevertigo, Stewartrfc, Sverdrup, TastyPoutine, The Anome, Tide rolls, Tiptoety,
Too Old, Topbanana, Truthflux, Truthnlove, Tsemii, Vald, Viriditas, Vortico, Wafulz, War sharks, Whodunit, WikiWikiPhil, Wimt, Yougotavirus, Yungjui, Zamb, , 209 anonymous edits
Quantum state Source: http://en.wikipedia.org/w/index.php?oldid=422116605 Contributors: ARTE, Agent Foxtrot, Alksentrs, Andres, B. Wolterding, Bambaiah, BenRG, Bevo, Bizzon,
Bkalafut, Bob K31416, CSTAR, CapitalR, Colin Watson, Cortonin, Dan East, Dragon's Blood, Dzordzm, Erwin, Freakofnurture, Freddy78, Fresheneesz, GeorgeLouis, Geschichte, Giftlite,
Harddk, Headbomb, Hidaspal, Hulten, J04n, JTXSeldeen, Julesd, Jutta234, Kbrose, Kinneytj, Laussy, Lotje, Machine Elf 1735, MathKnight, Mct mht, Michael Hardy, MichaelHaeckel, Mpatel,
Mschlindwein, Nathanielvirgo, Ott, Papadopc, Physis, Pierluigi.taddei, PoorLeno, Pvkeller, RTC, RealityDysfunction, RobinK, Rockfang, Rorro, Sbyrnes321, Sheliak, Slipstream, Stephen
Poppitt, Steve Quinn, StewartMH, The-tenth-zdog, Thurth, V81, Waxigloo, WaysToEscape, Woohookitty, 67 anonymous edits

177

Image Sources, Licenses and Contributors

Image Sources, Licenses and Contributors


Image:Bohr model 3.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Bohr_model_3.jpg License: GNU Free Documentation License Contributors: User:Sadi Carnot
Image:Hot metalwork.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Hot_metalwork.jpg License: unknown Contributors: Contributor, Fir0002, Jahobr, Wst, 1 anonymous edits
File:Max Planck Briefmarke 2008.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Max_Planck_Briefmarke_2008.jpg License: unknown Contributors: Deutsche Bundespost
File:Einstein.Painting.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Einstein.Painting.jpg License: Creative Commons Attribution-Sharealike 3.0 Contributors: GeorgeLouis
File:Niels.Bohr.Sketch.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Niels.Bohr.Sketch.jpg License: Creative Commons Attribution 3.0 Contributors: GeorgeLouis
Image:Emission spectrum-H.png Source: http://en.wikipedia.org/w/index.php?title=File:Emission_spectrum-H.png License: Public Domain Contributors: user:Merikanto
Image:Chicago Union Station 1943.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Chicago_Union_Station_1943.jpg License: Public Domain Contributors: Berrucomons, Bjh21,
Bmicomp, Caseyjonz, Davepape, Denniss, Duesentrieb, Grenavitar, Hailey C. Shannon, Infrogmation, Jon Harald Sby, MarkSweep, Red devil 666, Roberta F., Rocket000, SPUI, Shizhao,
Trialsanderrors, WikedKentaur,
File:Heisenberg 10.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Heisenberg_10.jpg License: Public Domain Contributors: JdH, Palamde, Quiris
File:Erwin Schroedinger.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Erwin_Schroedinger.jpg License: Creative Commons Attribution-Sharealike 3.0 Contributors:
GeorgeLouis
File:Paul.Dirac.monument.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Paul.Dirac.monument.jpg License: Creative Commons Attribution-Sharealike 3.0 Contributors:
GeorgeLouis, Ww2censor
File:10 Quantum Mechanics Masters.jpg Source: http://en.wikipedia.org/w/index.php?title=File:10_Quantum_Mechanics_Masters.jpg License: Creative Commons Attribution-Sharealike 3.0
Contributors: User:Patrick Edwin Moran
Image:Black body.svg Source: http://en.wikipedia.org/w/index.php?title=File:Black_body.svg License: Public Domain Contributors: User:Darth Kule
Image:Photoelectric effect.svg Source: http://en.wikipedia.org/w/index.php?title=File:Photoelectric_effect.svg License: unknown Contributors: Wolfmankurd
Image:Bohr atom model English.svg Source: http://en.wikipedia.org/w/index.php?title=File:Bohr_atom_model_English.svg License: Creative Commons Attribution-Sharealike 2.5
Contributors: User:Brighterorange
File:Young+Fringes.gif Source: http://en.wikipedia.org/w/index.php?title=File:Young+Fringes.gif License: GNU Free Documentation License Contributors: User:Mpfiz, User:Patrick Edwin
Moran
File:Single & double slit experiment.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Single_&_double_slit_experiment.jpg License: Creative Commons Attribution-Sharealike 3.0
Contributors: User:Patrick Edwin Moran
Image:neon orbitals.JPG Source: http://en.wikipedia.org/w/index.php?title=File:Neon_orbitals.JPG License: Public Domain Contributors: Benjah-bmm27, Mortadelo2005, 1 anonymous edits
File:Superposition.svg Source: http://en.wikipedia.org/w/index.php?title=File:Superposition.svg License: Creative Commons Attribution-Sharealike 3.0 Contributors: User:Patrick Edwin
Moran
File:QuantumHarmonicOscillatorAnimation.gif Source: http://en.wikipedia.org/w/index.php?title=File:QuantumHarmonicOscillatorAnimation.gif License: Creative Commons Zero
Contributors: User:Sbyrnes321
Image:HAtomOrbitals.png Source: http://en.wikipedia.org/w/index.php?title=File:HAtomOrbitals.png License: GNU Free Documentation License Contributors: Admrboltz, Benjah-bmm27,
Dbc334, Dbenbenn, Ejdzej, Falcorian, Kborland, MichaelDiederich, Mion, Saperaud, 6 anonymous edits
File:Rtd seq v3.gif Source: http://en.wikipedia.org/w/index.php?title=File:Rtd_seq_v3.gif License: Public Domain Contributors: Saumitra R Mehrotra & Gerhard Klimeck
File:QuantumDot wf.gif Source: http://en.wikipedia.org/w/index.php?title=File:QuantumDot_wf.gif License: Public Domain Contributors: Saumitra R Mehrotra & Gerhard Klimeck
File:Steppot.png Source: http://en.wikipedia.org/w/index.php?title=File:Steppot.png License: GNU Free Documentation License Contributors: Original uploader was Bamse at en.wikipedia
File:Infinite potential well.svg Source: http://en.wikipedia.org/w/index.php?title=File:Infinite_potential_well.svg License: unknown Contributors: User:Bdesham
Image:EPR-paradox-illus.png Source: http://en.wikipedia.org/w/index.php?title=File:EPR-paradox-illus.png License: GNU Free Documentation License Contributors: Original uploader was
CSTAR at en.wikipedia
Image:Bell's theorem.svg Source: http://en.wikipedia.org/w/index.php?title=File:Bell's_theorem.svg License: Creative Commons Attribution 3.0 Contributors: User:Manwesulimo2004
Image:Bells-thm.png Source: http://en.wikipedia.org/w/index.php?title=File:Bells-thm.png License: GNU Free Documentation License Contributors: Bdesham, Common Good, It Is Me Here,
Joshbaumgartner, Karelj, Maksim, Mdd, Pieter Kuiper, Tano4595
Image:Bell-test-photon-analyer.png Source: http://en.wikipedia.org/w/index.php?title=File:Bell-test-photon-analyer.png License: GNU Free Documentation License Contributors: Chetvorno,
Glenn, Joshbaumgartner, Karelj, Maksim, Mdd
File:Schrodingers cat.svg Source: http://en.wikipedia.org/w/index.php?title=File:Schrodingers_cat.svg License: Creative Commons Attribution-Sharealike 3.0 Contributors: User:Dhatfield
File:MWI Schrodingers cat.png Source: http://en.wikipedia.org/w/index.php?title=File:MWI_Schrodingers_cat.png License: Creative Commons Attribution-Sharealike 3.0 Contributors:
Dc987
Image:String theory.svg Source: http://en.wikipedia.org/w/index.php?title=File:String_theory.svg License: Creative Commons Attribution 3.0 Contributors: User:MissMJ
Image:World lines and world sheet.svg Source: http://en.wikipedia.org/w/index.php?title=File:World_lines_and_world_sheet.svg License: Public Domain Contributors: User:Actam,
User:Kurochka
Image:Calabi-Yau.png Source: http://en.wikipedia.org/w/index.php?title=File:Calabi-Yau.png License: Creative Commons Attribution-Sharealike 2.5 Contributors: D-Kuru, Darapti, German,
Kilom691, Lunch, Peter17, 1 anonymous edits
File:Quantum gravity.png Source: http://en.wikipedia.org/w/index.php?title=File:Quantum_gravity.png License: Public Domain Contributors: User:Raidr
File:Gravity Probe B.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Gravity_Probe_B.jpg License: Public Domain Contributors: Emesee, GDK, Pline
File:Point&string.png Source: http://en.wikipedia.org/w/index.php?title=File:Point&string.png License: unknown Contributors: Actam, Kilom691, Kurochka, 2 anonymous edits
File:Calabi-Yau.png Source: http://en.wikipedia.org/w/index.php?title=File:Calabi-Yau.png License: Creative Commons Attribution-Sharealike 2.5 Contributors: D-Kuru, Darapti, German,
Kilom691, Lunch, Peter17, 1 anonymous edits
File:Spinnetwork.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Spinnetwork.jpg License: Creative Commons Attribution-Sharealike 3.0 Contributors: User:Skoglund S
Image:Hydrogen Density Plots.png Source: http://en.wikipedia.org/w/index.php?title=File:Hydrogen_Density_Plots.png License: unknown Contributors: PoorLeno (talk)

178

License

License
Creative Commons Attribution-Share Alike 3.0 Unported
http:/ / creativecommons. org/ licenses/ by-sa/ 3. 0/

179

You might also like