Sie sind auf Seite 1von 711

ompound Chemical element chemical energy chemical potential chemical reaction chemical vapor deposition chemistry Chen Ning

Yang
Christopher Stasheff church bells circle circular detection probability CKM matrix clairvoyance classical electrodynamics classical
ics Bound state Bra-ket notation Breit equation
ory Claude Cohen-Tannoudji Claude E. Shannon Claus Jönsson Clay Math#The Millennium Prize problems Clebsch-Gordan coefficients
cients coherence (physics) coherent Coherent state colatitude Cold cathode cold emission collapse of the wavefunction collision Colloid
ute compact compact Lie group compact space compactification (physics) complementarity complementarity (physics) complementary
t Clebsch-Gordan coefficients Coherent state
mplex projective line Complex_numbers complexity Composite field composite particle Compton effect Compton suppression Compton
raction configuration space confinement conical intersection conjugate quantities conjugate transpose connected correlation function
bability consistent histories conspiracy theory constants of motion constraint Constructive quantum field theory context contextualism
tradiction Control theory convergent series convex hull convex set Cooper pair coordinate coordinate system Coordinates (mathematics)
ables Compton scattering Compton wavelength
mit correspondence principle cosmic inflation cosmic radiation cosmic ray cosmological constant cotangent bundle coulomb Coulomb
riant derivative CP violation CP-symmetry CPT invariance CPT symmetry CPT violation CPT-symmetry creation and annihilation operators
cal atom Curie curl Current (electricity) curvature form curvilinear Cyclotron#Mathematics_of_the_cyclotron cymbals d'Alembert operator
Politzer David Thouless David Wineland Davisson-Germer experiment de Broglie de Broglie hypothesis de Broglie wavelength de Rham
ion and annihilation operators Dark energy star
n degeneracy degeneracy (disambiguation) degeneracy pressure degenerate degenerate dwarf Degenerate energy level degenerate gases
onal theory Density matrix density of states density operator Density state Depleted uranium derivation derivative DESY detailed balance
tion diffraction pattern Diffusion diffusion equation Digital Devil Saga Digital object identifier dimension dimensional analysis dimensional
rac sea Dirac string Dirac's constant direct product direct sum direct sum#Direct sum of Hilbert spaces Dirichlet problem Dirk Gently's
energy level Degenerate matter Delayed choice
Displacement operator Dissociation (chemistry) distance distant anticipation Distribution (mathematics) divergence theorem divergent
experiment Douglas Adams down quark drag (physics) Driven harmonic motion drop dual space Duality (mathematics) Duncan MacInnes
don Edward Teller Edward Witten effective action effective field theory effective theory Effects of nuclear explosions Ehrenfest theorem
cal space Dirac operator Double-slit experiment
sambiguation) Einstein's summation convention Einstein-Podolsky-Rosen paradox elastic elastic collision electric Electric charge electric
g electrical network electrical potential electrical resistance electrical resonance electricity electrochemistry electrode electrodynamics
tic waves electromagnetically induced transparency electromagnetism Electron Electron affinity Electron capture electron charge electron
ctron spin electron subshell electron volt electron-degenerate matter electron-neutrino Electronic band structure electronic configuration
ctronic density Electronic Hamiltonian Electronic
interaction electroweak theory Elementary charge Elementary particle elementary particles Elitzur-Vaidman bomb-testing problem ellipse
rgy spectrum energy states Energy-dispersive_X-ray_spectroscopy Englert Englert-Greenberger duality Enriched uranium Enrico Fermi
ean epistemology eponym EPR paradox equation equation of state equations of state equilibrium equipartition theorem equivalence class
statistics Erwin Schroedinger Erwin Schrödinger Estermann Euclidean quantum gravity Euclidean space Eugene Wigner Euler-Lagrange
ess EP Quantum Mechanics Excited state Exotic
citons exemplar existence of God exotic (hadron) Exotic atom exotic baryon exotic meson expectation value expected value experiment
olator Fast breeder reactor Fast neutron reactor faster-than-light fate fault tolerant feminist science fiction Fermi energy Fermi gas Fermi
gnet Fertile material Feshbach resonance Feshbach-Fano partioning Feynman Feynman diagram Feynman path integral Feynman slash
nite potential well first-order phase transition flash memory Flatland Flatterland flavour (particle physics) Fluorescence flutter flux Flux
gas Fermi liquid Fermi's golden rule Fermi-Dirac
Fourier series Fourier transform Fowler-Nordheim equation fractal fractional statistics frames of reference Franck-Hertz experiment Frank
e free states Freeman Dyson freezing frequencies frequency frequency spectrum Frequency-resolved optical gating friction Fritz London
vative functional derivative operator functional integral functional integration functional integration (neurobiology) functional integration
ock matrix Fock space Fock state Franck-Hertz
amental particles fundamental representation Fusion power Futurama Future energy development G-parity G. Johnstone Stoney Galaxy
box Gas in a harmonic trap gauge anomaly gauge boson gauge field gauge fixing gauge group gauge invariance gauge invariant gauge
Marsden experiment Geissler tube General relativity General Semantics General Theory of Relativity generalized coordinate Generalized
geometrical optics geophysical George Alec Effinger George Chapline George Eugene Uhlenbeck George Gamow George Mackey George
ger-Marsden experiment Gibbs paradox
r_Schwerionenforschung GeV GHZ Gibbs paradox Gibbs state gigabyte Gilbert N. Lewis Glauber P representation global minimum global
tion theory gravastar gravitation gravitational coupling constant gravitational field gravitational singularities gravitational wave gravitino
ation group velocity Group_representation GURPS Gustav Ludwig Hertz Gyroscopic_precession#Torque-induced_precession Göttingen
all effect Hamel basis Hamilton-Jacobi equation Hamilton-Jacobi equations Hamiltonian hamiltonian (quantum mechanics) Hamiltonian
nics) Heisenberg picture Hilbert space Hydrogen
c oscillator harmonic series (music) harmonics Hartle-Hawking state Hartree product Hartree theory Hartree-Fock Hartree-Fock method
Principle Heisenberg uncertainty relations Heisenberg's microscope Heisenberg's uncertainty principle Heisenbug helium Hellsing HEMT
e Hermitian matrix hermitian operator Hermitian operator#Spectral theorem hertz heterostructure hidden variable hidden variable theories

Quantum Mechanics
ature superconductivity high-energy physics Hilary Putnam Hilbert space Hilbert-Schmidt hill potential Hindu History of nuclear weapons
plicate and Explicate Order according to David
lonomy Homodyne detection Homogeneity_(physics) homogeneous space homotopy group Hopf algebra Horst Ludwig Störmer Howard
ogen atom hydrogen molecule Hydrogen-2 hydrogen-like atom hypercube hypercylinder hyperfine structure Hyperion Cantos#Endymion
ntity function Igor Tamm image image (mathematics) imaginary number imaginary unit Immanuel Kant impact ionization Impedance Index
Internal conversion Interpretation of quantum
divisibility infinite potential well Infinitesimal influence information information entropy information retrieval Information theory infrared
titute for Theoretical Physics (Frankfurt) instrumentalism insulator integer integrable integral Integral Fast Reactor integrate Integration
nal conversion (chemistry) internal conversion coefficient internal symmetry International Phonetic Alphabet International Space Station
nt invariant (mathematics) invariant theory Inverse scattering inverse scattering problem Inverse scattering transform inverse-square law
ffect Klein-Gordon equation Ladder operators
y iron irreducible representation irreducible representations Is logic empirical? Isaac Newton Isidor Isaac Rabi Isidor Rabi Islam Islamic
er Jacobi identity Jahn-Teller effect Jain James Chadwick James Clerk Maxwell James Edward Zimmerman James Franck James Jeans
n Cramer John Dalton John F. Allen John Hasbrouck van Vleck John Searle John Stewart Bell John Strutt, 3rd Baron Rayleigh John von
h Willard Gibbs joule joule second Julian Schwinger June 15 June 5 K. K. Darrow Kanada kaon Karl K. Darrow Karl Popper Karl Pribram
d equation London moment Many-body problem
a formula Klystrode Klystron knowledge Kochen-Specker theorem Korteweg-de Vries equation Kristofer Straub Kronecker delta Ladder
flow Landau pole Lande interval rule Landé g-factor Laplace Laplace operator Laplacian Large Hadron Collider Larmor_precession laser
ity lens (optics) Lenz Leo Kadanoff Leon Lederman Leon Rosenfeld Leonid Mandelshtam Lepton leptons leptoquark Lester Germer Lester
t quantum light wave lightbulb limit (mathematics) limits to computing Lindblad equation line broadening line bundle Line integral linear
nics Matrix model Maxwell-Boltzmann statistics
ransformation Linewidth Linus Pauling Liouville equation Liouville's theorem (Hamiltonian) liquid liquid crystal Liquid drop model Liquid
t of isotopes by symbol list of mathematical topics in quantum theory list of noise topics List of nuclear tests List of optical topics list of
eories List of topics (scientific method) List_of_particles#Hypothetical_particles lithium local hidden variable theory local maximum local
Multiplicative quantum number Neutral particle
scattering Longitudinal wave loop expansion Lord Rayleigh Lords and Ladies Lords and Ladies (novel) Lorentz covariant Lorentz factor
is, 7th duc de Broglie Louis-Victor de Broglie lower bound Lp space LS coupling LS-coupling LSZ formalism Lucretius Ludwig Boltzmann
magnetic moment magnetic monopole magnetic potential Magnetic quantum number Magnetic Resonance Imaging magneto-optic effect
theory Many-minds interpretation Many-worlds interpretation many-worlds interpretation of quantum mechanics Marshall Stone Martin
physics Observable Oil-drop experiment Open
aster of Mosquiton material science mathematical Mathematical analysis mathematical formulation of quantum mechanics mathematical
x mechanics Matrix population models Matrix theory (physics) matrix_(mathematics) matter Max Born Max Planck Max Tegmark Maxwell
1 May 6 MCSCF mean measurable space measure measure (mathematics) measure space measure theory measurement measurement
Media:Stark splitting in hydrogen.png medical imaging Meissner effect Mellin transform memories MEMS Mendeleev mental model meson
ox Particle in Adiabatic
a one-dimensional lattice (periodic
invariant
black hole microscope microscopic microscopy microwave Middle Ages Mie theory Millennium (Hellsing)#HJ-Oberstammführer (Warrant
xy molecular geometry molecular Hamiltonian Molecular laser isotope separation molecular physics Molecular scattering Molecular term
MOSFET Mossbauer effect Mott insulator Multiple scattering multiplicative quantum number Multiverse (science) Muon muon neutrino
ademy of Sciences National Bureau of Standards Natural abundance natural satellite natural unit Nearly-free electron model Neil Gaiman
ic potential Path integral formulation Penrose
on matter Neutron star neutron-degenerate matter neutronium new age New Scientist New York Academy of Sciences New York Times
ogen dioxide nitrous oxide no cloning theorem Nobel Prize Nobel Prize for Physics Nobel Prize in Physics Nobel Prizes for Physics nodal

-
ng nondimensionalization nondimensionalization#Quantum harmonic oscillator nonholonomic mapping nonlinear sigma model nonlocal
uantum mechanics) Photoelectric effect Planck
r normalized wavefunction normalizing constant Norman F. Ramsey normed vector space Noumenon Nova (series) November 5 NP-hard
sonance Nuclear material Nuclear medicine Nuclear physics Nuclear power Nuclear power plant Nuclear proliferation Nuclear propulsion
Nuclear weapon design nuclei nucleon Nucleosynthesis nucleus Nukees NUMB3RS number operator numerical O.R. Lummer Observable
nd off shell One-dimensional periodic case one-loop Feynman diagram ontological Ontology open ball open quantum system operational
Plum pudding model Position
Zeeman operator Potential
effect
al phenomenon Optical Society of America optical theorem Optics Orbit orbital orbital angular momentum Orbital elements orbital motion
rthonormal orthonormal basis Osama bin Laden oscillation oscillator Oskar Klein Osterwalder-Schrader theorem Otto Stern outer product
partial derivative partial differential equation partial differential equations partial trace particle particle accelerator particle collider particle
cle physics Particle scattering particle statistics particle zoo particles partite partition function partition function (quantum field theory)
ility current Projective Hilbert space Pure gauge
ntegral Formulation patient pattern patterns Paul Adrien Maurice Dirac Paul Dirac Paul Ehrenfest Paul Gordan Paul Sophus Epstein Pauli
unction periodic table periodic table of elements periodicity permittivity permittivity of free space permutation perpendicular perturbation
leum Pfund series pharmaceutical Phase (matter) Phase (waves) phase distribution phase noise phase shift phase space phase transition
Philosophical interpretation of classical physics Philosophy Philosophy of science phonon Phosphorescence Photocurrent Photodiode
ology Quantum chaos Quantum Critical Point
mology physical limits to computing physical paradox physical phenomenon Physical property physical quantity Physical Review Letters
er Pieter Zeeman pilot wave pin group pion Pioneer plaque planck constant Planck Length Planck mass Planck time Planck units Planck's
ding model plum-pudding model Plutonium Poet Laureate Poincare group Poincare symmetry Poincaré symmetry poison Poisson bracket
on Quantum foam Quantum Hall effect Quantum
ded Pontryagin duality Portable Document Format position position manifold position operator positive definite positive linear functional
gy surface potential theory Potential well potential_energy#Graphical_representation Pound-Rebka falling photon experiment POVM power
Darkness (1987 film) Princeton University Princeton University Press Principal quantum number principle principle of complementarity
ity distribution probability flux probability measure probability theory Proca equation Process physics processes Product (mathematics)
Quantum level Quantum mechanics Quantum
ive space proof of the existence of God propagator property Prophecy (Stargate SG-1) proportionality constant Propositional calculus
ychology pudding pure mathematics pure state purification of quantum state Pyotr Leonidovich Kapitsa Pythagoreans Q factor QCD QCD
ation of gauge theories quantization_(physics) quantized Quantum quantum entanglement Quantum annealing quantum chaos Quantum
puting quantum cosmology Quantum cryptography quantum cybernetics quantum decoherence quantum dot Quantum electrochemistry
antum phase transition Quantum solid Quantum
n quantum gravity quantum gyroscope quantum Hall effect quantum harmonic oscillator Quantum harmonic oscillator#Ladder operator
ent quantum mechanic quantum mechanical Quantum Mechanics Quantum Mechanics - simplified Quantum mechanics#Description of the
quantum numbers quantum operation Quantum optics quantum particle Quantum Physics quantum process tomography quantum state
eory Quantum thermodynamics quantum trajectory representation theories of quantum mechanics quantum tunneling Quantum tunnelling
Quantum Theory Parallels to Consciousness
vel) Quark quark matter quark model quark star quark-degenerate matter Quark-gluon plasma quarks quartz quasar quasiparticle Qubit
io frequency radio wave radioactive Radioactive decay radioactive nuclei Radioactive waste Radioactivity radiobiology radiofrequency
k (mathematics) Rapid single flux quantum Rate distortion theory rationalism ray ray tracing Rayleigh Scattering Rayleigh-Jeans law real
uantum well Quantum Zeno effect Quasistability
m reference frame Reflection (physics) reflexive space refraction Reginald Cahill region Relationship between string theory and quantum
d theory relativistic wave equations relativity Relativity physics relativity theory Relic particles remote viewing Renaissance Renninger
kernel Hilbert space residue theorem resolution resolution of the identity resonance resonance (disambiguation) Resonant frequency
ction#Values at the integers Riemannian manifold Riesz representation theorem Rigged Hilbert space right triangle ring wave guide ripple
guide Ritz method Rutherford model Rutherford
t Millikan Robert Mills (physicist) Robert Oppenheimer Robert Serber Robertson-Schrödinger relation Rockefeller Foundation Rockefeller
cess rubidium Rudolf Grimm rule of thumb rumor Rutherford Rutherford backscattering Rutherford model Rutherford scattering Rydberg
c Satyendra Nath Bose scalar scalar field scalar potential scale anomaly scanning SQUID microscope scanning tunnelling microscope
equation Schrödinger picture Schrödinger's cat Schrödinger's cat in fiction Schrödinger's Cat trilogy Schrödinger's equation Schumann
ory Schrödinger equation Schrödinger picture
scientific model scientific notation scientific rigor scintillation screen screw dislocation sea level sea water second second law of
Self-energy semantics semiclassical semiclassical gravity semiconductor semiconductor devices semidefinite programming seminar
hannon entropy Shannon entropy#Formal definitions shape resonance Sheldon Lee Glashow Shell model Shelter Island (town), New York
L sine curve sine-Gordon equation singular spectrum singularity Sir Roger Penrose Skyrmion SLAC Slater determinant Slater-type orbital
riational principle Selection rule Semiclassical
olid solid angle solid helium solid state physics soliton Solvay Conference Sommerfeld-Wilson-Ishiwara quantization sound sound wave
oup special relativity special unitary group Special_relativity spectral line spectral measure spectral theorem spectral theory spectrometer
l harmonic spherical harmonics spin Spin (physics) spin (physics)#spin multiplets spin angular momentum spin quantum number spin
struction Slater determinant Spin-1/2 Spin-orbital
fission spontaneous parametric down conversion Spontaneous symmetry breaking spooky action at a distance square integrable square
tion standard model standard model (basic details) standing wave standing waves Stanford Encyclopedia of Philosophy star Star Trek
e statistical mechanics statistics Stefan-Boltzmann law Stellar mass black hole Stephen Donaldson Stephen Hawking Stephen Notley
ated scattering Stirling's approximation stochastic stochastic process Stokes theorem Stokes' law Stone's theorem Stone's theorem on
effect Stationary state Stern-Gerlach experiment
ong CP problem strong CP violation strong force strong interaction strong interactions strong nuclear force structures Stuart Kauffman
nyaev Zel'dovich effect super-consciousness Super-Kamiokande supercommutator Superconducting superconductive Superconductivity
n principle superpotential superselection superselection sector Supersolid superstring theory Supersymmetry supersymmetry breaking
ectic space symposium synchronicity synchrotron radiation synonym system T-symmetry T. D. Lee table of Clebsch-Gordan coefficients
upersymmetric quantum mechanics T-symmetry
eter-totter teleology telepathy teleportation temperature tensor tensor category tensor product tensor product#Tensor product of Hilbert
uantum Cats The Compass Rose The Elegant Universe The Elementary Particles The Feynman Lectures on Physics The Gap Cycle The
l chemistry theoretical physics theory Theory of Everything theory of relativity thermal de Broglie wavelength thermal equilibrium thermal
of thermodynamics Thomas Kuhn Thomas Young Thomas Young (scientist) Thomas-Fermi approximation Thomson scattering Thorium
antum number Transformation theory (quantum
Timeline of chemical element discovery Timeline of cosmic microwave background astronomy Timeline of quantum mechanics, molecular
n-Oppenheimer-Volkoff limit Tomography Tonks-Girardeau gas top quark topological defect topological dimension topological entropy
al angular momentum total angular momentum quantum number trace class trace-class trajectory transactional interpretation Transducer
To Scilesco
astrophe Uncertainty principle Unitarity Unitarity
or transition rate transition rule translational invariance translationally invariant transmission coefficient (physics) Transmission electron
diode tunnel_(quantum_mechanics) tunneling time Turbid media twentieth century Two Lumps two-body problem two-photon generation
operator Uncertainty Uncertainty principle uncertainty principle#One of the theorems uncertainty relation uncountable uncountable set
x unitary operator unitary representation unitary representation of a star Lie superalgebra unitary transformation United States National
rbation theory Wave packet Wave-particle duality
stin University of Tübingen University of Vienna Unobservables Unruh effect Unsolved problems in physics unstable unstable particle up
ntum-mechanical vacuum Vaisheshika valence shell Valentine Bargmann vapour pressure variance variational method Variational method
ons vector potential vector space Vector space dimension Vector_space velocity vernacular vertex renormalization Very high temperature
viscous visual system Vladimir Aleksandrovich Fock Vladimir Fock Voigt volt volume von Klitzing constant von Neumann von Neumann
istribution Wigner-Eckart theorem Work function
Ritz Walther Bothe Walther Gerlach Ward-Takahashi identity water wave wave equation wave function wave functions Wave interference
acket waves Wayne Itano weak decay weak force weak gauge boson Weak interaction weak interactions weak measurement weak nuclear
Dwarf white dwarf material white noise whole number Wick rotation Wiener measure Wiener process Wightman axioms Wigner 3-j symbol
ibson (novelist) William Rowan Hamilton Willis Lamb Willoughby Smith winding number wiretap WKB approximation Wojciech H. Zurek
eraction Z boson Zagreus (Doctor Who audio) Zeeman effect Zeno's paradoxes#The arrow paradox zero-point energy zig-zag zinc sulfide
FT
Quantum
Mechanics
Adiabatic invariant
A -
Zeeman effect
DR

compiled by To Scilesco
Quantum Mechanics
Compiled by: To Scilesco
Date: 14.07.2006
BookId: dcchcaruqoqeqfro

FT
All articles and pictures of this book were retrieved from the Wikipedia Project
(wikipedia.org) on 02.07.2006. The articles are free to use under the terms
of the GNU Free Documentation License. A copy of this license is included in
the section entitled "GNU Free Documentation License". Images in this book
have diverse licenses and you can find a list of figures and the corresponding
licenses in the section "List of Figures". The version history of all articles can
be retrieved from wikipedia.org. Each Article in this book has a reference to
the original article. The principal authors of articles are referenced at the end
of each article unless technical difficulties did not allow for a proper determi-
nation of the principal authors.

Logo design by Joerg Pelka


A
Printed by InstaBook Corporation (instabook.net)

Published by pediapress.com a service offered by brainbot technologies AG ,


Mainz, Germany
DR
Articles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
Adiabatic invariant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
Adiabatic theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
Aharonov-Bohm effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

FT
Atomic theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
Auger electron . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
Bargmann’s limit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
Bohr model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
Boltzmon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
Born probability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
Bose–Einstein condensate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
Bose-Einstein statistics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
Bose gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
Bound state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
Bra-ket notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
Breit equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
Canonical commutation relation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
Chladni’s law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
A
Classical limit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
Clebsch-Gordan coefficients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
Coherent state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
Complementarity (physics) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
Complete set of commuting observables . . . . . . . . . . . . . . . . . . . . . . . 67
Compton scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
Compton wavelength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
Conjugate variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
DR
Constraint algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
Coupling constant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
Creation and annihilation operators . . . . . . . . . . . . . . . . . . . . . . . . . . 79
Dark energy star . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
Davisson-Germer experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
De Broglie hypothesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
Degenerate energy level . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
Degenerate matter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
Delayed choice quantum eraser . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
Diabatic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
Dirac equation in the algebra of physical space . . . . . . . . . . . . . . . . . 98
Dirac operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
Double-slit experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
Duru-Kleinert transformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
Ehrenfest theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
Einselection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
Electronic density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
Electronic Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
Electronic state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117

FT
Elementary particle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
Energy level splitting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
Entanglement witness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
EP Quantum Mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
Excited state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
Exotic hadron . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
Faddeev equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
Fano resonance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
Fermi-Dirac statistics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
Fermi energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
Fermi gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
Fermi liquid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
Fermi’s golden rule . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
Field emission . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
A
Finite potential well . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Flux quantization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
146
149
Fock matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
Fock space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
Fock state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
Franck-Hertz experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
Free particle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
Functional integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
DR
Geiger-Marsden experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
Gibbs paradox . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
Greenberger-Horne-Zeilinger state . . . . . . . . . . . . . . . . . . . . . . . . . 167
Hamiltonian (quantum mechanics) . . . . . . . . . . . . . . . . . . . . . . . . . 168
Heisenberg picture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172
Hilbert space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
Hydrogen atom . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182
Hydrogen-like atom . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
Hyperfine structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191
Imaginary time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194
Implicate and Explicate Order according to David Bohm . . . . . . . 195
Incompleteness of quantum physics . . . . . . . . . . . . . . . . . . . . . . . . . 207
Interaction picture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 210
Internal conversion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211
Interpretation of quantum mechanics . . . . . . . . . . . . . . . . . . . . . . . 212
Introduction to quantum mechanics . . . . . . . . . . . . . . . . . . . . . . . . 224
Josephson effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253
Klein-Gordon equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 256
Ladder operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 259

FT
Laplace-Runge-Lenz vector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
Large Area Neutron Detector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265
Lindblad equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 266
London moment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 266
Many-body problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267
Mathematical formulation of quantum mechanics . . . . . . . . . . . . 268
Matrix mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
Matrix model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 285
Maxwell-Boltzmann statistics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 285
Measurement in quantum mechanics . . . . . . . . . . . . . . . . . . . . . . . 291
Molecular Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 299
Multiplicative quantum number . . . . . . . . . . . . . . . . . . . . . . . . . . . . 300
Neutral particle oscillations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 301
Normalisable wavefunction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303
A
Normal mode . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Nuclear physics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
306
311
Observable . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 314
Oil-drop experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 315
Open quantum system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 320
Optical theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 320
Parity (physics) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 322
Particle in a box . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 330
DR
Particle in a one-dimensional lattice (periodic potential) . . . . . . . 338
Particle in a ring . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 343
Particle in a spherically symmetric potential . . . . . . . . . . . . . . . . . 345
Path integral formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 348
Penrose Interpretation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 360
Peres-Horodecki criterion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 362
Perturbation theory (quantum mechanics) . . . . . . . . . . . . . . . . . . . 363
Photoelectric effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 371
Planck particle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 379
Planck postulate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 380
Planck’s law of black body radiation . . . . . . . . . . . . . . . . . . . . . . . . 380
Plum pudding model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 388
Position operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 390
Potential energy surface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 391
Potential well . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 391
POVM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 393
Probability amplitude . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 399
Probability current . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 400
Projective Hilbert space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 403

FT
Pure gauge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 404
Quantum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 404
Quantum 1/f noise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 409
Quantum acoustics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 410
Quantum biology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 410
Quantum chaos . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 411
Quantum Critical Point . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 413
Quantum entanglement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 414
Quantum field theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 421
Quantum fluctuation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 436
Quantum foam . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 438
Quantum Hall effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 440
Quantum harmonic oscillator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 442
Quantum indeterminacy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 451
A
Quantum leap . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Quantum level . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
458
459
Quantum mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 459
Quantum mechanics, philosophy and controversy . . . . . . . . . . . . 477
Quantum mineralogy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 485
Quantum phase transition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 486
Quantum solid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 486
Quantum state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 487
DR
Quantum statistical mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 490
Quantum superposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 494
Quantum Theory Parallels to Consciousness . . . . . . . . . . . . . . . . . 495
Quantum tomography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 497
Quantum tunnelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 497
Quantum vibration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 503
Quantum well . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 504
Quantum Zeno effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 505
Quasistability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 506
QWiki . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 507
Range criterion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 507
Relativistic particle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 508
Resonance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 509
Ring wave guide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 512
Ritz method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 514
Rutherford model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 516
Rutherford scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 517
Rydberg formula . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 518
Scattering channel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 521

FT
Scattering theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 522
Schrödinger equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 524
Schrödinger picture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 531
Schrödinger’s cat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 532
Schrödinger’s cat in fiction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 540
Schwinger’s variational principle . . . . . . . . . . . . . . . . . . . . . . . . . . . 545
Selection rule . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 546
Semiclassical . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 548
Separable states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 549
Shelter Island Conference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 552
Single particle reconstruction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 555
Slater determinant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 556
Spin-1/2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 557
Spin-orbital . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 559
A
Squashed entanglement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Squeezed coherent state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
560
564
SQUID . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 571
Stark effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 576
Stationary state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 581
Stern-Gerlach experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 582
Subatomic particle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 588
Superdense coding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 592
DR
Superselection sector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 594
Supersymmetric quantum mechanics . . . . . . . . . . . . . . . . . . . . . . . 596
Thermal de Broglie wavelength . . . . . . . . . . . . . . . . . . . . . . . . . . . . 601
Topological order . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 603
Topological quantum number . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 607
Transformation theory (quantum mechanics) . . . . . . . . . . . . . . . . 608
T-symmetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 609
Two interfering electron wave-packets . . . . . . . . . . . . . . . . . . . . . . 616
Ultraviolet catastrophe . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 617
Uncertainty principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 619
Unitarity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 630
Unitarity bound . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 630
Variational method (quantum mechanics) . . . . . . . . . . . . . . . . . . . 631
Variational perturbation theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . 633
Wavefunction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 634
Wave packet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 640
Wave-particle duality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 644
Wien’s displacement law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 652
Wigner-Eckart theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 656

FT
Wigner quasi-probability distribution . . . . . . . . . . . . . . . . . . . . . . . 656
Work function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 661
Zeeman effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 663

GNU Free Documentation License . . . . . . . . . . . . . . . . . . . . . . . . . . . . 669

List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 673

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 675

A
DR
1

Adiabatic invariant

An adiabatic invariant in general is a property of motion which is conserved

FT
to exponential accuracy in the small parameter representing the typical rate of
change of the gross properties of the body. For periodic motion, the adiabatic
H
invariants are the action integrals p dq taken over a period of the motion.
These are constants of the motion and remain so even when changes are made
in the system, as long as the changes are slow compared to the period of mo-
tion.
In plasma physics there are three adiabatic invariants of charged particle mo-
tion.

The first adiabatic invariant, µ


The magnetic moment of a gyrating particle,
1 2
2 mv⊥
µ= B ,
is a constant of the motion (as long as q/m does not change). In fact, it is in-
A
variant to all orders in an expansion in ω/ωc , so the magnetic moment remains
nearly constant even for changes at rates approaching the gyrofrequency.
There are some important situations in which the magnetic moment is not
invariant:

• Magnetic pumping: When µ is constant, the perpendicular particle energy


is proportional to B, so the particles can be heated by increasing B, but this
is a ’one shot’ deal because the field cannot be increased indefinitely. On
DR
the other hand, if the collision frequency is larger than the pump frequen-
cy, µ is no longer conserved. In particular, collisions allow net heating by
transferring some of the perpendicular energy to parallel energy.
• Cyclotron heating: If B is oscillated at the cyclotron frequency, the condi-
tion for adiabatic invariance is violated and heating is possible. In particu-
lar, the induced electric field rotates in phase with some of the particles and
continuously accelerates them.
• Magnetic cusps: The magnetic field at the center of a cusp vanishes, so the
cyclotron frequency is automatically smaller than the rate of any changes.
Thus the magnetic moment is not conserved and particles are scattered
relatively easily into the loss cone.

Adiabatic invariant
2

The second adiabatic invariant, J


The longitudinal invariant of a particle trapped in a magnetic mirror,
Rb
J = a v|| ds,

FT
where the integral is between the two turning points, is also an adiabatic in-
variant. This guarantees, for example, that a particle in the ionosphere moving
around the Earth will always return to the same line of force. The adiabatic
condition is violated in transit-time magnetic pumping, where the length of a
magnetic mirror is oscillated at the bounce frequency, resulting in net heating.

The third adiabatic invariant, Φ


The total magnetic flux Φ enclosed by a drift surface is the third adiabatic in-
variant, associated with the periodic motion of mirror-trapped particles drifting
around the axis of the system. Because this drift motion is relatively slow, Φ is
often not conserved in practical applications.

External links


A lecture notes on the second adiabatic invariant 1
lecture notes on the third adiabatic invariant 2

Source: http://en.wikipedia.org/wiki/Adiabatic_invariant

Principal Authors: Art Carlson, MathMartin, SimonP, Gurch, Linas


DR
Adiabatic theorem

The adiabatic theorem is an important theorem in quantum mechanics which


provides the foundation for perturbative quantum field theory.
There are different versions of this theorem. Max Born and V. A. Fock proved
the original version in 1928:
A physical system remains in its instantaneous eigenstate if a given per-
turbation is acting on it slowly enough and if there is a gap between the
eigenvalue and the rest of the Hamiltonian’s spectrum.

1 http://farside.ph.utexas.edu/teaching/plasma/lectures/node24.html
2 http://farside.ph.utexas.edu/teaching/plasma/lectures/node25.html

Adiabatic theorem
3

To be more precise, the adiabatic theorem does not tell us that there is any
finite lower bound for the duration over which we have to perform a pertur-
bation on the system in order to keep it in its instantaneous eigenstate. It just
tells that this is the case if the rate of change approaches zero!

FT
In 1990 J. E. Evron and A. Elgart found a new version of the adiabatic theorem
that does not require gaps.

External links and references


• J. E. Evron, A. Elgart: Adiabatic Theorem without a Gap Condition 3

Source: http://en.wikipedia.org/wiki/Adiabatic_theorem

Principal Authors: Artur adib, Deco, SeventyThree, Conscious, Charles Matthews, BeteNoir

Aharonov-Bohm effect
A
The Aharonov-Bohm effect, sometimes called the Ehrenberg-Siday-
Aharonov-Bohm effect, is a quantum mechanical phenomenon by which a
charged particle is affected by electromagnetic fields in regions from which the
particle is excluded. The earliest form of this effect was predicted by Werner
Ehrenberg and R.E. Siday in 1949, and similar effects were later rediscovered
by Aharonov and Bohm in 1959. Such effects are predicted to arise from both
magnetic fields and electric fields, but the magnetic version has been easier
DR
to observe. In general, the profound consequence of Aharonov-Bohm effects
is that knowledge of the classical electromagnetic field acting locally on a
particle is not sufficient to predict its quantum-mechanical behavior.
After the 1959 paper was published, Bohm was informed that the effect had
been predicted by Rory E. Siday and Werner Ehrenberg a decade earlier; Bohm
and Aharonov duly cited this in their second paper (Peat, 1997, p. 192).
The most commonly described case, sometimes called the Aharonov-Bohm
solenoid effect, is when the wave function of a charged particle passing
around a long solenoid experiences a phase shift as a result of the enclosed
magnetic field, despite the magnetic field being zero in the region through
which the particle passes. This phase shift has been observed experimentally
by its effect on interference fringes. (There are also magnetic Aharonov-Bohm

3 http://www.arxiv.org/abs/math-ph/9805022/

Aharonov-Bohm effect
4

effects on bound energies and scattering cross sections, but these cases have
not been experimentally tested.) An electric Aharonov-Bohm phenomenon was
also predicted, in which a charged particle is affected by regions with different
electrical potentials but zero electric field, and this has also seen experimental

FT
confirmation. A separate "molecular" Aharonov-Bohm effect was proposed for
nuclear motion in multiply-connected regions, but this has been argued to be
essentially different, depending only on local quantities along the nuclear path
(Sjöqvist, 2002).
A general review can be found in Peshkin and Tonomura (1989).

Magnetic Aharonov-Bohm effect


The magnetic Aharonov-Bohm effect can be seen as a result of the requirement
that quantum physics be invariant with respect to the gauge choice for the
vector potential A. This implies that a particle with charge q travelling along
some path P in a region with zero magnetic field (B = 0 = ∇ × A) must acquire
a phase φ; given in SI units by
φ = ~q P A · dx,
R
A
with a phase difference ∆φ between any two paths with the same endpoints
therefore determined by the magnetic flux Φ through the area between the
paths (via Stokes theorem and ∇ × A = B), and given by:

∆φ = ~ .

This phase difference can be observed by placing a shielded solenoid between


the slits of a double-slit experiment (or equivalent). A shielded solenoid en-
DR
closes a magnetic field B, but does not produce any magnetic field outside of
its cylinder, and thus the charged particle (e.g. an electron) passing outside
experiences no classical effect. However, there is a (curl-free) vector potential
outside the solenoid with an enclosed flux, and so the relative phase of particles
passing through one slit or the other is altered by whether the magnetically
shielded solenoid current is turned on. This corresponds to an observable shift
of the interference fringes on the observation plane.
The same phase effect is responsible for the quantized-flux requirement in su-
perconducting loops. This quantization is due to the fact that the supercon-
ducting wave function must be single valued: its phase difference ∆φ around
a closed loop must be an integer multiple of 2π (with the charge q=2e for the
electron Cooper pairs), and thus the flux Φ must be a multiple of h /2e. The
superconducting flux quantum was actually predicted prior to Aharonov and
Bohm, by London (1948) using a phenomenological model.

Aharonov-Bohm effect
5

A FT
Figure 1 Schematic of double-slit experiment in which Aharonov-Bohm effect can be observed:
electrons pass through two slits, interfering at an observation screen, with the interference pattern
shifted when a magnetic field B is turned on in the shielded cylindrical solenoid.

The magnetic Aharonov-Bohm effect is also closely related to Dirac’s argument


that the existence of a magnetic monopole necessarily implies that both electric
and magnetic charges are quantized. A magnetic monopole implies a mathe-
matical singularity in the vector potential, which can be expressed as an in-
finitely long Dirac string of infinitesimal diameter that contains the equivalent
DR
of all of the 4πg flux from a monopole "charge" g. Thus, assuming the absence
of an infinite-range scattering effect by this arbitrary choice of singularity, the
requirement of single-valued wave functions (as above) necessitates charge-
quantization: 2qg/c~ must be an integer (in cgs units) for any electric charge
q and magnetic charge g.
The magnetic Aharonov-Bohm effect was experimentally confirmed by Osak-
abe et al. (1986), following earlier work summarized in Olariu and Popèscu
(1984). Its scope and application continues to expand. Webb et al. (1985)
demonstrated Aharonov-Bohm oscillations in ordinary, non-superconducting
metallic rings; for a discussion, see Schwarzschild (1986) and Imry & Webb
(1989). Bachtold et al. (1999) detected the effect in carbon nanotubes; for a
discussion, see Kong et al. (2004).

Aharonov-Bohm effect
6

Electric Aharonov-Bohm effect


Just as the phase of the wave function depends upon the magnetic vector po-
tential, it also depends upon the scalar electric potential. By constructing a
situation in which the electrostatic potential varies for two paths of a particle,

FT
through regions of zero electric field, an observable Aharonov-Bohm interfer-
ence phenomenon from the phase shift has been predicted; again, the absence
of an electric field means that, classically, there would be no effect.
From the →Schrödinger equation, the phase of an eigenfunction with energy E
goes as exp(−iEt/~). The energy, however, will depend upon the electrostatic
potential V for a particle with charge q. In particular, for a region with constant
potential V (zero field), the electric potential energy qV is simply added to E,
resulting in a phase shift:
∆φ = − qV~ t ,

where t is the time spent in the potential.


The initial theoretical proposal for this effect suggested an experiment where
charges pass through conducting cylinders along two paths, which shield the
A
particles from external electric fields in the regions where they travel, but still
allow a varying potential to be applied by charging the cylinders. This proved
difficult to realize, however. Instead, a different experiment was proposed
involving a ring geometry interrupted by tunnel barriers, with a bias voltage V
relating the potentials of the two halves of the ring. This situation results in
an Aharonov-Bohm phase shift as above, and was observed experimentally in
1998.
DR
Mathematical interpretation
In the terms of modern differential geometry, the Aharonov-Bohm effect can be
understood to be the holonomy of the complex-valued line bundle representing
the electromagnetic field. The connection on the line bundle is given by the
electromagnetic potential A, and thus the electromagnetic field strength is the
curvature of the line bundle F =dA. The integral of A around a closed loop is
the holonomy, which, by Stokes theorem, is the magnetic field threading the
loop. Thus the wave function of the electron can be seen to be directly coupled
to the complex line bundle representing the electromagnetic field.
See also a related effect, the Berry phase.

Aharonov-Bohm effect
7

References
• Aharonov, Y. and D. Bohm, "Significance of electromagnetic potentials in
quantum theory," Phys. Rev. 115, 485–491 (1959).

FT
• Bachtold, A., C. Strunk, J. P. Salvetat, J. M. Bonard, L. Forro, T. Nussbaumer
and C. Schonenberger 4, “Aharonov-Bohm oscillations in carbon nanotubes”,
Nature 397, 673 (1999).
• Ehrenberg, W. and R. E. Siday, "The Refractive Index in Electron Optics and
the Principles of Dynamics," Proc. Phys. Soc. London Sect. B 62, 8–21
(1949).
• Imry, Y. and R. A. Webb, "Quantum Interference and the Aharonov-Bohm
Effect," Scientific American, 260(4), April 1989.
• Kong, J., L. Kouwenhoven, and C. Dekker, "Quantum change for nan-
otubes", Physics Web 5 (July 2004).
• London, F. "On the problem of the molecular theory of superconductivity,"
Phys. Rev. 74, 562–573 (1948).
• Murray, M. Line Bundles 6, (2002).
• Olariu, S. and I. Iovitzu Popèscu, "The quantum effects of electromagnetic


A fluxes," Rev. Mod. Phys. 57, 339–436 (1985).
Osakabe, N., T. Matsuda, T. Kawasaki, J. Endo, A. Tonomura, S. Yano, and
H. Yamada, "Experimental confirmation of Aharonov-Bohm effect using a
toroidal magnetic field confined by a superconductor." Phys Rev A. 34(2):
815-822 (1986). Abstract and full text. 7
• Peat, F. David 8, Infinite Potential: The Life and Times of David Bohm
(Addison-Wesley: Reading, MA, 1997). ISBN 0-201-40635-7.
• Peshkin, M. 9 and Tonomura, A., The Aharonov-Bohm effect (Springer-
DR
Verlag: Berlin, 1989). ISBN 3-540-51567-4.
• Schwarzschild, B. "Currents in Normal-Metal Rings Exhibit Aharonov-Bohm
Effect." Phys. Today 39, 17–20, Jan. 1986.
• Sjöqvist, E. "Locality and topology in the molecular Aharonov-Bohm effect,"
Phys. Rev. Lett. 89 (21), 210401/1–3 (2002).

4 http://pages.unibas.ch/phys-meso/
5 http://physicsweb.org/articles/world/17/7/3/1
6 http://www.maths.adelaide.edu.au/people/mmurray/dg99/line_bundles.pdf
7 http://prola.aps.org/abstract/PRA/v34/i2/p815_1
8 http://www.fdavidpeat.com/
9 http://www.phy.anl.gov/theory/staff/mp.html

Aharonov-Bohm effect
8

• van Oudenaarden, A., M. H. Devoret, Yu. V. Nazarov, and J. E. Mooij,


"Magneto-electric Aharonov-Bohm effect in metal rings," Nature 391, 768–
770 (1998).
• Webb, R., S. Washburn, C. Umbach, and R. Laibowitz. Phys. Rev. Lett. 54,

FT
2696 (1985).

Source: http://en.wikipedia.org/wiki/Aharonov-Bohm_effect

Principal Authors: Stevenj, Linas, Liontooth, Reddi, CYD

Atomic theory

In chemistry and physics, atomic theory is a theory of the nature of matter.


It states that all matter is composed of atoms. The philosophical background
of the atomic theory is called atomism. The theory applies to the common
phases of matter, namely solids, liquids and gases, as directly experienced on
A
Earth. Strictly speaking, it is not the appropriate theory for plasmas or neutron
stars where unusual environments such as extremes of temperature or density
prevent atoms from forming.

Importance
Arguably, the atomic theory is one of the most important theories in the history
of science, with wide-ranging implications for both pure and applied science.
The theory is largely credited to John Dalton, an 18th- and 19th century British
DR
chemist and physicist.
Modern chemistry (and biochemistry) is based upon the theory that all matter
is made up of atoms of different elements, which cannot be transmuted by
chemical means. In turn, chemistry has allowed for the development of the
pharmaceutical industry, the petrochemical industry, and many others.
Much of thermodynamics is understandable in terms of kinetic theory, whereby
gases are considered to be made up of either atoms or molecules, behaving in
accordance with Newton’s laws of motion. This was, in turn, a large driving
force behind the industrial revolution.
Indeed, many macroscopic properties of matter are best understood in terms
of atoms. Other examples include friction, material science and semiconductor
theory. The latter is particularly important, as it is the foundation of electronics.

Atomic theory
9

Historical precursors
Early atomism
Main article: Atomism

FT
From the 6th century BC, Hindu, Buddhist and Jaina philosophers in ancient
India developed the earliest atomic theories. The first philosopher who formu-
lated ideas about the atom in a systematic manner was Kanada who lived in
the 6th century BC. Another Indian philosopher, Pakudha Katyayana who also
lived in the 6th century BC and was a contemporary of Gautama Buddha, had
also propounded ideas about the atomic constitution of the material world. In-
dian atomists believed that an atom could be one of up to six elements, with
each element having up to 24 properties. They developed detailed theories of
how atoms could combine, react, vibrate, move, and perform other actions,
and had particularly elaborate theories of how atoms combine, which explains
how atoms first combine in pairs, and then group into trios of pairs, which are
the smallest visible units of matter. This parallels with the structure of mod-
A
ern atomic theory, in which pairs or triplets of supposedly fundamental quarks
combine to create most typical forms of matter. They had also suggested the
possibility of splitting an atom which, as we know today, is the source of atomic
energy. (See Indian atomism for more details.)
Democritus and Leucippus, Greek philosophers in the 5th century BC, present-
ed a theory of atoms. (See Atomism for more details.) The Greeks believed
that atoms were all made of the same material but had different shapes and
sizes, which determined the physical properties of the material. For instance,
DR
the atoms of a liquid were thought to be smooth, allowing them to slide over
each other. None of these ideas, however, were founded in scientific experi-
mentation.
During the Middle Ages (the Islamic Golden Age), Islamic atomists develop
atomic theories that represent a synthesis of both Greek and Indian atomism.
(See Islamic atomism for more details.) Older Greek and Indian ideas were fur-
ther developed by Islamic atomists, along with new Islamic ideas, such as the
possibility of there being particles smaller than an atom. As Islamic influence
began spreading through Europe, the ideas of Islamic atomism, along with the
older ideas of Greek and Indian atomism, spread throughout Europe by the
end of the Middle Ages, where modern atomic theories began taking shape.

Birth of modern atomic theory


In 1808, John Dalton proposed that an element is composed of atoms of a sin-
gle, unique type, and that although their shape and structure was immutable,

Atomic theory
10

atoms of different elements could combine to form more complex structures


(chemical compounds). He deduced this after the experimental discovery of
the law of multiple proportions — that is, if two elements form more than one
compound between them, then the ratios of the masses of the second element

FT
which combine with a fixed mass of the first element will be ratios of small
whole numbers.
The experiment in question involved combining nitrous oxide (NO) with oxy-
gen (O 2). In one combination, these gases formed dinitrogen trioxide (N 2O 3),
but when he repeated the combination with double the amount of oxygen (a
ratio of 1:2), they instead formed nitrogen dioxide (NO 2).
4NO + O 2 → 2N 2O 3
4NO + 2O 2 → 4NO 2
Atomic theory conflicted with the theory of infinite divisibility, which states
that matter can always be divided into smaller parts. In 1827, biologist Robert
Brown observed that pollen grains floating in water constantly jiggled about
for no apparent reason. In 1905, Albert Einstein theorised that this Brownian
motion was caused by the water molecules continuously knocking the grains
A
about, and developed a mathematical theory around it. This theory was vali-
dated experimentally in 1911 by French physicist Jean Perrin.

Discovery of subatomic particles


For much of this time, atoms were thought to be the smallest possible divi-
sion of matter. However, in 1897, J.J. Thomson published his work proving
that cathode rays are made of negatively charged particles (electrons). Since
cathode rays are emitted from matter, this proved that atoms are made up of
DR
subatomic particles and are therefore divisible, and not the indivisible atomos
postulated by Democritus. Physicists later invented a new term for such in-
divisible units, "elementary particles", since the word atom had come into its
common modern use.

Study of atomic structure


At first, it was believed that the light electrons were distributed in rings or oth-
er orbits in a more or less uniform sea or cloud of positive charge (the plum
pudding model). However, an experiment conducted in 1909 by colleagues
of Ernest Rutherford demonstrated that atoms have a most of their mass and
also their positive charge concentrated in a very small fraction of their vol-
ume, a region which Rutherford assumed to be at the very center of the atom.
In the gold foil experiment, alpha particles (emitted by polonium) were shot
through a sheet of gold (striking a fluorescent screen on the other side). The

Atomic theory
11

experimenters expected all the alpha particles to pass through without signif-
icant deflection, given the uniform distribution of positive charge in the plum
pudding model. On the contrary, about 1 in 8000 of the alpha particles were
heavily deflected (by more than 90 degrees). This led Rutherford to propose

FT
the planetary model of the atom in which pointlike electrons orbited in the
space around a massive compact nucleus like planets orbiting the Sun.
The nucleus was later discovered to contain protons, and further experimenta-
tion by Rutherford found that the nuclear mass of most atoms surpassed that
of the protons it possessed; this led him to postulate the existence of neutrons,
whose existence would be proven in 1932 by James Chadwick.
The planetary model of the atom still had shortcomings. First, a moving electric
charge emits electromagnetic waves; according to classical electromagnetism,
an orbiting charge would steadily lose energy and spiral towards the nucleus,
colliding with it in a tiny fraction of a second. Second, the model did not
explain why excited atoms emit light only in certain discrete spectra.
Quantum theory revolutionized physics at the beginning of the 20 th century
when Max Planck and Albert Einstein postulated that light energy is emitted or
A
absorbed in fixed amounts known as quanta. In 1913, Niels Bohr used this idea
in his →Bohr model of the atom, in which the electrons could only orbit the
nucleus in particular circular orbits with fixed angular momentum and energy.
They were not allowed to spiral into the nucleus, because they could not lose
energy in a continuous manner; they could only make quantum leaps between
fixed energy levels. Bohr’s model was extended by Arnold Sommerfeld in 1916
to include elliptical orbits, using a quantization of generalized momentum.
The ad hoc Bohr-Sommerfeld model was extremely difficult to use, but it made
DR
impressive predictions in agreement with certain spectral properties. However,
the model was unable to explain multielectron atoms, predict transition rates
or describe fine and hyperfine structure. In 1925, Erwin Schrödinger developed
a full theory of quantum mechanics, described by the →Schrödinger equation.
Together with Wolfgang Pauli’s exclusion principle, this allowed study of atoms
with great precision when digital computers became available. Even today,
these theories are used in the Hartree-Fock quantum chemical method to de-
termine the energy levels of atoms. Further refinements of quantum theory
such as the Dirac equation and quantum field theory made smaller impacts on
the theory of atoms.
Another model of historical interest, proposed by Gilbert N. Lewis in 1916, had
cubical atoms with electrons statically held at the corners. The cubes could
share edges or faces to form chemical bonds. This model was created to ac-
count for chemical phenomena such as bonding, rather than physical phenom-
ena such as atomic spectra.

Atomic theory
12

See also
• History of thermodynamics
• Kinetic theory

FT
• Development of Quantum Theory
• Quantum Chemistry
• John Dalton

Related lists

• Timeline of chemical element discovery


• Timeline of quantum mechanics, molecular physics, atomic physics, nuclear
physics, and particle physics
• Timeline of thermodynamics, statistical mechanics, and random processes

References

External links
A
• Ancient Atomism 10

Source: http://en.wikipedia.org/wiki/Atomic_theory

Principal Authors: Vsmith, Brighterorange, Ragesoss, Voyajer, Karol Langner, Linas, Rho, Timmy2,
Dustimagic, Eric Forste
DR
Auger electron

Auger emission (pronounced [o e]) is a phenomenon in physics in which the


emission of an electron from an atom causes the emission of a second electron.
This second ejected electron is called an Auger electron.
The name Auger electron comes from one of its discoverers, Pierre Victor
Auger. The name does not come from the similarly-named device, the auger.
When an electron is removed from a core level of an atom, leaving a vacancy,
an electron from a higher energy level may fall into the vacancy, resulting in a
release of energy. Although sometimes this energy is released in the form of an

10 http://plato.stanford.edu/entries/atomism-ancient/

Auger electron
13

emitted photon, the energy can also be transferred to another electron, which
is then ejected from the atom.
Upon ejection the kinetic energy of the Auger electron corresponds to the dif-
ference between the energy of the initial electronic transition and the ioniza-

FT
tion energy for the shell from which the Auger electron was ejected. These
energy levels depend on the type of atom and the chemical environment in
which the atom was located. Auger electron spectroscopy stimulates the emis-
sion of Auger electrons by bombarding a sample with either X-rays or energetic
electrons and measures the intensity of Auger electrons as a function of the
Auger electron energy. The resulting spectra can be used to determine the
identity of the emitting atoms and some information about their environment.
A similar Auger effect occurs in semiconductors. An electron and electron hole
can recombine giving up their energy to an electron in the conduction band,
increasing its energy.
The reverse effect is known as impact ionization.

History
A
The Auger emission process was discovered in the 1920s by Lise Meitner, an
Austrian physicist. Subsequently Pierre Victor Auger, a French Physicist, also
discovered the process. Auger reported the discovery in the journal Radium in
1925 and it was Auger that had the process named after him.

Source: http://en.wikipedia.org/wiki/Auger_electron

Principal Authors: AjAldous, Keenan Pepper, Srleffler, Stokerm, Tristanb


DR
Bargmann’s limit

In quantum mechanics, Bargmann’s limit, named for Valentine Bargmann,


provides an upper bound on the number N l of bound states in a system. It
takes the form
1 2m ∞
R
Nl ≤ 2l+1 ~2 0
r|V (r)|V <0 dr

Professor Hagen says, "The Bargmann limit provides, if not the best bound, a
pretty darn good one."
Note that the delta function potential attains this limit.

Bargmann’s limit
14

References
• Bargmann, Proc. Nat. Acad. Sci. 38 961 (1952)
• Schwinger, Proc. Nat. Acad. Sci. 47 122 (1961)

FT
Source: http://en.wikipedia.org/wiki/Bargmann%27s_limit

Principal Authors: TobinFricke, Covington, Amalas, Charles Matthews, Pjacobi

Bohr model

A
DR

Figure 2 The Bohr model of the atom

In atomic physics, the Bohr model depicts the atom as a small, positive-
ly charged nucleus surrounded by waves of electrons in orbit — similar in
structure to the solar system, but with electrostatic forces providing attraction,
rather than gravity, and with waves spread over entire orbit instead of localized
planets.

Bohr model
15

Introduced by Niels Bohr in 1913, the model’s key success was in explain-
ing the →Rydberg formula for the spectral emission lines of atomic hydrogen;
while the Rydberg formula had been known experimentally, it did not gain a
theoretical underpinning until the Bohr model was introduced.

FT
The Bohr model is a primitive model of the hydrogen atom which cannot ex-
plain the fine structure of the hydrogen atom nor any of the heavier atoms. As
a theory, it can be derived as a first-order approximation of the hydrogen atom
in the broader and much more accurate quantum mechanics, and thus may be
considered to be an obsolete scientific theory. However, because of its simplic-
ity, the Bohr model is still commonly taught to introduce students to quantum
mechanics.

History
In the early 20th century, experiments by Ernest Rutherford and others had
established that atoms consisted of a diffuse cloud of negatively charged elec-
trons surrounding a small, dense, positively charged nucleus. Given this ex-
perimental data, it is quite natural to consider a planetary model for the atom,
with electrons orbiting a sun-like nucleus. However, a naive planetary model
A
has several difficulties, the most serious of which is the loss of energy by syn-
chrotron radiation.That is, an accelerating electric charge emits electromag-
netic waves which carry energy; thus, with each orbit around the nucleus, the
electron would radiate away a bit of its orbital energy, gradually spiralling in-
wards to the nucleus until the atom was no more. A quick calculation shows
that this would happen almost instantly; thus, the naive planetary theory can-
not explain why atoms are extremely long-lived.
The naive planetary model also failed to explain atomic spectra, the observed
DR
discrete spectrum of light emitted by electrically excited atoms. Late 19th cen-
tury experiments with electric discharges through various low-pressure gasses
in evacuated glass tubes had shown that atoms will emit light (that is, electro-
magnetic radiation), but only at certain discrete frequencies. A naive planetary
model cannot explain this.
To overcome these difficulties, Niels Bohr proposed, in 1913, what is now
called the Bohr model of the atom. The key ideas were:

• The orbiting electrons existed in orbits that had discrete quantized energies.
That is, not every orbit is possible but only certain specific ones.
• The laws of classical mechanics do not apply when electrons make the jump
from one allowed orbit to another.
• When an electron makes a jump from one orbit to another the energy dif-
ference is carried off (or supplied) by a single quantum of light (called a

Bohr model
16

photon) which has an energy equal to the energy difference between the
two orbitals.
• The allowed orbits depend on quantized (discrete) values of orbital angular
momentum, L according to the equation

FT
h
L = n · ~ = n · 2π
Where n = 1,2,3,· · · and is called the principal quantum number, and h is
Planck’s constant.

Assumption (4) states that the lowest value of n is 1. This corresponds to a


smallest possible radius of 0.0529 nm. This is known as the Bohr radius. Once
an electron is in this lowest orbit, it can get no closer to the proton.
The Bohr model is sometimes known as the semiclassical model of the atom,
as it adds some primitive quantization conditions to what is otherwise a clas-
sical mechanics treatment. The Bohr model is certainly not a full quantum
mechanical description of the atom. Assumption 2) states that the laws of clas-
sical mechanics don’t apply during a quantum jump, but it doesn’t state what
laws should replace classical mechanics. Assumption 4) states that angular
momentum is quantised but does not explain why.
A
Refinements
Several enhancements to the Bohr model were proposed; most notably the
Sommerfeld model or Bohr-Sommerfeld model, which attempted to add
support for elliptical orbits to the Bohr model’s circular orbits. This model
supplemented condition (4) with an additional radial quantization condition,
the Sommerfeld-Wilson quantization condition
H
pdq = nh
DR
where p is the generalized momentum conjugate to the angular generalized
coordinate q; the integral is the action of action-angle coordinates.
The Bohr-Sommerfeld model proved to be extremely difficult and unwieldy
when its mathematical treatment was further fleshed out. In particular, the ap-
plication of traditional perturbation theory from classical planetary mechanics
led to further confusions and difficulties. In the end, the model was abandoned
in favour of the full quantum mechanical treatment of the hydrogen atom, in
1925, using Schrödinger’s wave mechanics.
However, this is not to say that the Bohr model was without its successes. Cal-
culations based on the Bohr-Sommerfeld model were able to accurately explain
a number of more complex atomic spectral effects. For example, up to first-
order perturbation, the Bohr model and quantum mechanics make the same
predictions for the spectral line splitting in the →Stark effect. At higher-order

Bohr model
17

perturbations, however, the Bohr model and quantum mechanics differ, and
measurements of the Stark effect under high field strengths helped confirm the
correctness of quantum mechanics over the Bohr model.
The Bohr-Sommerfeld quantization condition as first formulated can be viewed

FT
as a rough early draft of the more sophisticated condition that the symplectic
form of a classical phase space M be integral; that is, that it lie in the image of
Ȟ 2 (M, Z) → Ȟ 2 (M, R) → HDR
2
(M, R) , where the first map is the homo-
morphism of Čech cohomology groups induced by the inclusion of the integers
in the reals, and the second map is the natural isomorphism between the Čech
cohomology and the de Rham cohomology groups. This condition guarantees
that the symplectic form arise as the curvature form of a connection of a Her-
mitian line bundle. This line bundle is then called a prequantization in the
theory of geometric quantization.

Electron energy levels in hydrogen


The Bohr model is accurate only for one-electron systems such as the hydrogen
atom or singly-ionized helium. This section uses the Bohr model to derive the
A
energy levels of hydrogen.
The derivation starts with three simple assumptions:
1) All particles are wavelike, and an electron’s wavelength λ, is related to
its velocity v by:

h
λ= me v

where h is Planck’s Constant, and me is the mass of the electron. Bohr


DR
did not make this assumption (known as the de Broglie hypothesis) in
his original derivation, because it hadn’t been proposed at the time.
However it allows the following intuitive statement.

2) The circumference of the electron’s orbit must be an integer multiple of


its wavelength:

2πr = nλ

where r is the radius of the electron’s orbit, and n is a positive integer.

Bohr model
18

3) The electron is held in orbit by the coulomb force. That is, the coulomb
force is equal to the centripetal force:

kqe2 me v 2
r2
= r

FT
where k = 1/(4π0 ), and qe is the charge of the electron.

These are three equations with three unknowns: λ, r, v. After solving this
system of equations to find an equation for just v, it is placed into the equation
for the total energy of the electron:

Because of the virial theorem, the total energy simplifies to


E = − 12 me v 2

Substituting, one obtains the energy of the different levels of hydrogen:


A
Or, after plugging in values for the constants,

Thus, the lowest energy level of hydrogen (n = 1) is about -13.6 eV. The next
energy level (n = 2) is -3.4 eV. The third (n = 3) is -1.51 eV, and so on. Note
that these energies are less than zero, meaning that the electron is in a bound
state with the proton. Positive energy states correspond to the ionized atom
where the electron is no longer bound, but is in a scattering state.
DR
Energy in terms of other constants
Starting with what we found above,
−me qe4 1
En = 8h2 20 n2

We can multiply top and bottom by c2 , and we’ll arrive at


−me c2 qe4 1
En = 8h2 c2 20 n2

or re-grouping them to make it more clear:


 4 
q
En = − 21 me c2 4h2 ce2 2 n12
0

From here we can now write the energy level equation in terms of other con-
stants to:

Bohr model
19

−Er α2
En = 2n2

where,
En is the energy level

FT
Er is the rest energy of the electron

α is the fine structure constant

n is the principal quantum number.

Rydberg formula
The →Rydberg formula describes the transitions or quantum jumps between
one energy level and another. When the electron moves from one energy level
to another, a photon is given off. Using the derived formula for the different
’energy’ levels of hydrogen one may determine the ’wavelengths’ of light that a
A
hydrogen atom can give off.
The energy of photons that a hydrogen atom can give off are given by the
difference of two hydrogen energy levels:
 
me e4 1 1
E = Ei − Ef = 8h 2 2
 n2
− n2
0 f i

where nf means the final energy level, and ni means the initial energy
level. It is assumed that the final energy level is less than the initial energy
DR
level.

Since the energy of a photon is


hc
E= λ

the wavelength of the photon given off is


 
1 me e4 1 1
λ = 8ch3 2 n2
− n2
0 f i

The above is known as the →Rydberg formula. This formula was known in
the nineteenth century to scientists studying spectroscopy, but there was no
theoretical justification for the formula until Bohr derived it, more or less along
the lines above.

Bohr model
20

Shortcomings
The Bohr model gives an incorrect value L = ~ for the ground state orbital
angular momentum. The angular momentum in the true ground state is known
to be zero.

FT
The Bohr model also has difficulty with or fails to explain:

• The spectra of larger atoms. At best, it can make some approximate predic-
tions about the emission spectra for atoms with a single outer-shell electron
(atoms in the lithium group.)
• The relative intensities of spectral lines; although in some simple cases,
it was able to provide reasonable estimates (for example, calculations by
Kramers for the →Stark effect).
• The existence of fine structure and hyperfine structure in spectral lines.
• The →Zeeman effect - changes in spectral lines due to external magnetic
fields.

See also


A →Franck-Hertz experiment provided early support for the Bohr model.
Inert pair effect is adequately explained by means of the Bohr model.
• Lyman series
• →Schrödinger equation

References
Historical
DR
• Niels Bohr (1913). " On the Constitution of Atoms and Molecules (Part 1 of
3) 11". Philosophical Magazine 26: 1-25.
• Niels Bohr (1913). "On the Constitution of Atoms and Molecules, Part II
Systems Containing Only a Single Nucleus". Philosophical Magazine 26:
476-502.
• Niels Bohr (1913). "On the Constitution of Atoms and Molecules, Part III".
Philosophical Magazine 26: 857-875.
• Niels Bohr (1914). "The spectra of helium and hydrogen". Nature 92: 231-
232.

11 http://dbhs.wvusd.k12.ca.us/webdocs/Chem-History/Bohr/Bohr-1913a.html

Bohr model
21

• Niels Bohr (1921). " Atomic Structure 12". Nature.


• A. Einstein (1917). "Zum Quantensatz von Sommerfeld und Epstein". Ver-
handlungen der Deutschen Physikalischen Gesellschaft 19: 82-92. Reprint-
ed in The Collected Papers of Albert Einstein, A. Engel translator, (1997)

FT
Princeton University Press, Princeton. 6 p.434. (Provides an elegant re-
formulation of the Bohr-Sommerfeld quantization conditions, as well as an
important insight into the quantization of non-integrable (chaotic) dynam-
ical systems.)

Modern

• Paul Tipler and Ralph Llewellyn (2002). Modern Physics (4th ed.). W. H.
Freeman. ISBN 0716743450.

Source: http://en.wikipedia.org/wiki/Bohr_model

Principal Authors: Linas, JabberWok, Christopher Thomas, Tim Starling, MathKnight, Munchkinguy,
GoldenBoar, El C, Glenn
A
Boltzmon

A boltzmon (named after the nineteenth-century thermodynamicist Ludwig


Boltzmann) is a theoretical subatomic particle postulated to be created after
the explosion of a black hole.
DR
The boltzmon was proposed as a means of explaining what happens to the in-
formation of objects consumed by black holes while still preserving purity. One
theory, proposed by the Dutch researcher Gerard ’t Hooft, is that information is
contained in the particles that Hawking-radiate from the black hole. The other
theory includes the boltzmon particle.
This theory postulates that a black hole leaves behind a remnant when it
explodes—a single particle that has been dubbed the boltzmon. A boltzmon
would be about the size of the Planck-Wheeler area, or 10 -66 cm 2, which is
supposedly about as small as anything can be. It would contain the sum to-
tal of all the information ever consumed by the black hole, so each boltzmon
would be unique in the universe. While a typical particle has a few states (pos-
itive or negative electrical charge, integral or fractional spin, etc.), a boltzmon

12 http://dbhs.wvusd.k12.ca.us/webdocs/Chem-History/Bohr-Nature-1921.html

Boltzmon
22

would have an infinite number of states and as a result, would be highly un-
stable. If disturbed, it might make a hole in spacetime and vanish into it, thus
departing from our universe.

FT
References
• Ferris, Timothy. The Whole Shebang, 1997 Simon & Schuster.

Source: http://en.wikipedia.org/wiki/Boltzmon

Born probability

In quantum mechanics, the Born probability is a probability of an event cal-


culated from a wavefunction or more generally from the density matrix. The
probability (or its density) equals the squared modulus of the complex ampli-
A
tude an :
P = |an |2

The interpretation that the physical meaning of the wavefunction is proba-


bilistic was proposed by Max Born, and it became a pillar of the Copenhagen
interpretation of quantum mechanics.
DR
Source: http://en.wikipedia.org/wiki/Born_probability

Principal Authors: Jag123, Lumidek, Karol Langner, Conscious

Bose–Einstein condensate

A Bose–Einstein condensate is a phase of matter formed by bosons cooled to


temperatures very near to absolute zero. The first such condensate was pro-
duced by Eric Cornell and Carl Wieman in 1995 at the University of Colorado at
Boulder, using a gas of rubidium atoms cooled to 170 nanokelvins (nK). Under
such conditions, a large fraction of the atoms collapse into the lowest quantum
state, at which point quantum effects become apparent on a macroscopic scale.

Bose–Einstein condensate
23

FT
Figure 3 Velocity-distribution data confirming the discovery of a new phase of matter, the Bose–
Einstein condensate, out of a gas of rubidium atoms. The false colors indicate the number of atoms
at each velocity, with red being the fewest and white being the most. The areas appearing white
A
and light blue are at the lowest velocities. Left: just before the appearance of the Bose–Einstein
condensate. Center: just after the appearance of the condensate. Right: after further evaporation,
leaving a sample of nearly pure condensate. The peak is not infinitely narrow because of the
Heisenberg uncertainty principle: since the atoms are trapped in a particular region of space, their
velocity distribution necessarily possesses a certain minimum width.

Introduction
Condensates are extremely low temperature fluids with properties that are cur-
rently not completely understood, such as spontaneously flowing out of their
DR
container. The effect is the consequence of quantum mechanics, which states
that systems can only acquire energy in discrete steps. Now, if a system is at
such a low temperature that it is in the lowest energy state, it is no longer
possible for it to reduce its energy, not even by friction. Therefore, without
friction, the fluid will easily overcome gravity because of adhesion between the
fluid and the container wall, and it will take up the most favorable position,
i.e. all around the container.

Theory
The collapse of the atoms into a single quantum state is known as Bose con-
densation or Bose–Einstein condensation. This phenomenon was predicted
in 1925 by Albert Einstein, by generalizing Satyendra Nath Bose’s work on the
statistical mechanics of (massless) photons to (massive) atoms. (The Einstein
manuscript, believed to be lost, was found in a library at Leiden University in

Bose–Einstein condensate
24

2005.) The result of the efforts of Bose and Einstein is the concept of a →Bose
gas, governed by the →Bose-Einstein statistics, which describes the statisti-
cal distribution of identical particles with integer spin, now known as bosons.
Bosonic particles, which include the photon as well as atoms such as helium-4,

FT
are allowed to share quantum states with each other. Einstein speculated that
cooling bosonic atoms to a very low temperature would cause them to fall (or
"condense") into the lowest accessible quantum state, resulting in a new form
of matter.
This transition occurs below a critical temperature, which for a uniform three-
dimensional gas consisting of non-interacting particles with no apparent inter-
nal degrees of freedom is given by:
 2/3
n h2
Tc = ζ(3/2) 2πmkB

where:
<dl><dd>
Tc is the critical temperature,
n the particle density,
m
h
kB
A the mass per boson,
Planck’s constant,
the Boltzmann constant, and
ζ the Riemann zeta function; ζ(3/2) ≈ 2.6124.

</dd></dl>

Discovery
In 1938, Pyotr Kapitsa, John Allen and Don Misener discovered that helium-4
DR
became a new kind of fluid, now known as a superfluid, at temperatures below
2.17 kelvins (K) (lambda point). Superfluid helium has many unusual prop-
erties, including zero viscosity (the ability to flow without dissipating energy)
and the existence of quantized vortices. It was quickly realized that the super-
fluidity was due to Bose–Einstein condensation of the helium-4 atoms, which
are bosons. In fact, many of the properties of superfluid helium also appear in
the gaseous Bose–Einstein condensates created by Cornell, Wieman and Ket-
terle (see below). However, superfluid helium-4 is not commonly referred to
as a "Bose–Einstein condensate" because it is a liquid rather than a gas, which
means that the interactions between the atoms are relatively strong. The origi-
nal theory of Bose–Einstein condensation must be heavily modified in order to
describe it.
The first "true" Bose–Einstein condensate was created by Eric Cornell, Carl Wie-
man, and co-workers at JILA on June 5, 1995. They did this by cooling a dilute

Bose–Einstein condensate
25

vapor consisting of approximately 2000 rubidium-87 atoms to below 170 nK


using a combination of laser cooling (a technique that won its inventors Steven
Chu, Claude Cohen-Tannoudji, and William D. Phillips the 1997 Nobel Prize in
Physics) and magnetic evaporative cooling. About four months later, an in-

FT
dependent 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. Cornell,
Wieman and Ketterle won the 2001 Nobel Prize for their achievement.
The Bose–Einstein condensation also applies to quasiparticles in solids. A
magnon in an antiferromagnet carries spin 1 and thus obeys the Bose–Einstein
statistics. The density of magnons is controlled by an external magnetic field,
which plays the role of the magnon chemical potential. This technique pro-
vides access to a wide range of boson densities from the limit of a dilute Bose
gas to that of a strongly interacting Bose liquid. A magnetic ordering observed
at the point of condensation is the analog of superfluidity. In 1999 Bose con-
densation of magnons was demonstrated in the antiferromagnet TlCuCl3 by
Oosawa et al. The condensation was observed at temperatures as large as 14
A
K. Such a high transition temperature (relative to that of atomic gases) is due to
a greater density achievable with magnons and a smaller mass (roughly equal
to the mass of an electron).

Unusual characteristics
Further experimentation by the JILA team in 2000 uncovered a hitherto un-
known property of Bose–Einstein condensate. Cornell, Wieman, and their
coworkers originally used rubidium-87, an isotope whose atoms naturally repel
DR
each other making a more stable condensate. The JILA team instrumentation
now had better control over the condensate so experimentation was made on
naturally attracting atoms of another rubidium isotope, rubidium-85 (having
negative atom-atom scattering length). Through a process called Feshbach res-
onance involving a sweep of the magnetic field causing spin flip collisions, the
JILA researchers lowered the characteristic, discrete energies at which the ru-
bidium atoms bond into molecules making their Rb-85 atoms repulsive and
creating a stable condensate. The reversible flip from attraction to repulsion
stems from quantum interference among condensate atoms which behave as
waves.
When the scientists raised the magnetic field strength still further, the conden-
sate suddenly reverted back to attraction, imploded and shrank beyond detec-
tion, and then exploded, blowing off about two-thirds of its 10,000 or so atoms.
About half of the atoms in the condensate seemed to have disappeared from

Bose–Einstein condensate
26

the experiment altogether, not being seen either in the cold remnant or the
expanding gas cloud. Carl Wieman explained that under current atomic theory
this characteristic of Bose–Einstein condensate could not be explained because
the energy state of an atom near absolute zero should not be enough to cause

FT
an implosion; however, subsequent mean-field theories have been proposed to
explain it.
Due to the fact that supernovae explosions are implosions, the explosion of a
collapsing Bose–Einstein condensate was named "bosenova."
The atoms that seem to have disappeared are almost certainly still around in
some form, just not in a form that could be detected in that current experiment.
Two likely possibilities are that they have formed into molecules consisting of
two bonded rubidium atoms, or they received enough energy from somewhere
to fly away fast enough that they are out of the observation region before being
observed.

Current research
Compared to more commonly-encountered states of matter, Bose–Einstein con-
A
densates are extremely fragile. The slightest interaction with the outside world
can be enough to warm them past the condensation threshold, forming a nor-
mal gas and losing their interesting properties. It is likely to be some time
before any practical applications are developed.
Nevertheless, they have proved to be useful in exploring a wide range of ques-
tions in fundamental physics, and the years since the initial discoveries by the
JILA and MIT groups have seen an explosion in experimental and theoretical
activity. Examples include experiments that have demonstrated interference
DR
between condensates due to wave-particle duality 13, the study of superfluidity
and quantized vortices 14, and the slowing of light pulses to very low speeds
using electromagnetically induced transparency 15. Experimentalists have al-
so realized "optical lattices", where the interference pattern from overlapping
lasers provides a periodic potential for the condensate. These have been used
to explore the transition between a superfluid and a Mott insulator 16, and may
be useful in studying Bose–Einstein condensation in less than three dimensions,
for example the Tonks-Girardeau gas.
Bose–Einstein condensates composed of a wide range of isotopes have been
produced 17.

13 http://cua.mit.edu/ketterle_group/Projects_1997/Interference/Interference_BEC.htm
14 http://www.aip.org/pt/vol-53/iss-8/p19.html
15 http://www.europhysicsnews.com/full/26/article1/article1.html
16 http://qpt.physics.harvard.edu/qptsi.html
17 http://physicsweb.org/articles/world/18/6/1

Bose–Einstein condensate
27

Related experiments in cooling fermions rather than bosons to extremely low


temperatures have created degenerate gases, where the atoms do not congre-
gate in a single state due to the Pauli exclusion principle. To exhibit Bose–
Einstein condensate, the fermions must "pair up" to form compound particles

FT
(e.g. molecules or Cooper pairs) that are bosons. The first molecular Bose–
Einstein condensates were created in November 2003 by the groups of Rudolf
Grimm at the University of Innsbruck, Deborah S. Jin at the University of Col-
orado at Boulder and Wolfgang Ketterle at MIT. Jin quickly went on to create
the first fermionic condensate composed of Cooper pairs 18.

See also
• Atomic coherence
• →Bose gas
• Electromagnetically induced transparency
• Fermionic condensate
• Gas in a box
• Slow glass
A


Slow light
Superconductivity
• Superfluid
• Supersolid
• Super-heavy atom
• Tonks-Girardeau gas

External links
DR
• BEC Homepage 19 General introduction to Bose–Einstein condensation
• Nobel Prize in Physics 2001 20 - for the achievement of Bose–Einstein con-
densation in dilute gases of alkali atoms, and for early fundamental studies
of the properties of the condensates
• Physics Today: Cornell, Ketterle, and Wieman Share Nobel Prize for Bose–
Einstein Condensates 21
• Bose–Einstein Condensates at JILA 22
• The Bose–Einstein Condensate at Utrecht University, the Netherlands 23

18 http://physicsweb.org/articles/news/8/1/14/1
19 http://www.colorado.edu/physics/2000/bec/index.html
20 http://nobelprize.org/physics/laureates/2001/index.html
21 http://www.physicstoday.org/pt/vol-54/iss-12/p14.html
22 http://jilawww.colorado.edu/bec/
23 http://www.bec.phys.uu.nl/

Bose–Einstein condensate
28

• Alkali Quantum Gases at MIT 24


• Atom Optics at UQ 25
• Einstein’s manuscript on the Bose–Einstein condensate discovered at Leiden
University 26

FT
• The revolution that has not stopped 27 PhysicsWeb article from June 2005

References
• S. N. Bose, Z. Phys. 26, 178 (1924)
• A. Einstein, Sitz. Ber. Preuss. Akad. Wiss. (Berlin) 1, 3 (1925)
• L.D. Landau, J. Phys. USSR 5, 71 (1941)
• L. Landau (1941). "Theory of the Superfluidity of Helium II". Physical
Review 60: 356-358.
• M.H. Anderson, J.R. Ensher, M.R. Matthews, C.E. Wieman, and E.A. Cornell
(1995). "Observation of Bose–Einstein Condensation in a Dilute Atomic
Vapor" 28. Science 269: 198-201.
• K.B. Davis, M.-O. Mewes, M.R. Andrews, N.J. van Druten, D.S. Durfee, D.M.
Kurn, and W. Ketterle (1995). "Bose–Einstein condensation in a gas of sodi-
A

um atoms". Physical Review Letters 75: 3969-3973..
D. S. Jin, J. R. Ensher, M. R. Matthews, C. E. Wieman, and E. A. Cornell
(1996). "Collective Excitations of a Bose–Einstein Condensate in a Dilute
Gas". Physical Review Letters 77: 420-423.
• M. R. Andrews, C. G. Townsend, H.-J. Miesner, D. S. Durfee, D. M. Kurn,
and W. Ketterle (1997). "Observation of interference between two Bose
condensates". Science 275: 637-641..
• M. R. Matthews, B. P. Anderson, P. C. Haljan, D. S. Hall, C. E. Wieman, and
DR
E. A. Cornell (1999). "Vortices in a Bose–Einstein Condensate". Physical
Review Letters 83: 2498-2501.
• E.A. Donley, N.R. Claussen, S.L. Cornish, J.L. Roberts, E.A. Cornell, and
C.E. Wieman (2001). "Dynamics of collapsing and exploding Bose–Einstein
condensates". Nature 412: 295-299.
• M. Greiner, O. Mandel, T. Esslinger, T. W. Hänsch, I. Bloch (2002). "Quan-
tum phase transition from a superfluid to a Mott insulator in a gas of ultra-
cold atoms". Nature 415: 39-44..

24 http://cua.mit.edu/ketterle_group/home.htm
25 http://www.physics.uq.edu.au/atomoptics/
26 http://www.lorentz.leidenuniv.nl/history/Einstein_archive/
27 http://physicsweb.org/articles/world/18/6/8/1
28 http://links.jstor.org/sici?sici=0036-8075%2819950714%293%3A269%3A5221%3C198%3AOOBCIA

%3E2.0.CO%3B2-G

Bose–Einstein condensate
29

• S. Jochim, M. Bartenstein, A. Altmeyer, G. Hendl, S. Riedl, C. Chin, J.


Hecker Denschlag, and R. Grimm (2003). "Bose–Einstein Condensation of
Molecules". Science 302: 2101-2103.
• Markus Greiner, Cindy A. Regal and Deborah S. Jin (2003). "Emergence

FT
of a molecular Bose-Einstein condensate from a Fermi gas". Nature 426:
537-540.
• M. W. Zwierlein, C. A. Stan, C. H. Schunck, S. M. F. Raupach, S. Gupta,
Z. Hadzibabic, and W. Ketterle (2003). "Observation of Bose–Einstein Con-
densation of Molecules". Physical Review Letters 91: 250401.
• C. A. Regal, M. Greiner, and D. S. Jin (2004). "Observation of Reso-
nance Condensation of Fermionic Atom Pairs". Physical Review Letters 92:
040403.
• C. J. Pethick and H. Smith, "Bose–Einstein Condensation in Dilute Gases",
Cambridge University Press, Cambridge, 2001.
• Lev P. Pitaevskii and S. Stringari, "Bose–Einstein Condensation", Clarendon
Press, Oxford, 2003.
• Mackie M, Suominen KA, Javanainen J., "Mean-field theory of Feshbach-
resonant interactions in 85Rb condensates." Phys Rev Lett. 2002 Oct
A

28;89(18):180403.
Oxford Experimental BEC Group. http://www-matterwave.physics.ox.ac.
uk/bec/bec.html
• T. Nikuni, M. Oshikawa, A. Oosawa, and H. Tanaka, (1999). "Bose–Einstein
Condensation of Dilute Magnons in TlCuCl3" 29. Physical Review Letters 84:
5868.
DR
Source: http://en.wikipedia.org/wiki/Bose%E2%80%93Einstein_condensate

Principal Authors: Voyajer, CYD, Brian Jackson, Schneelocke, Michael Hardy, PAR, Matt Gies, Hfast-
edge, Fangz, R. Koot

29 http://dx.doi.org/10.1103/PhysRevLett.84.5868

Bose–Einstein condensate
30

Bose-Einstein statistics

For other topics related to Einstein see Einstein (disambiguation).

FT
In statistical mechanics, Bose-Einstein statistics (or more colloquially B-E
statistics) determines the statistical distribution of identical indistinguishable
bosons over the energy states in thermal equilibrium.
Fermi-Dirac and Bose-Einstein statistics apply when quantum effects have to
be taken into account and the particles are considered "indistinguishable". The
quantum effects appear if the concentration of particles (N/V) ≥ n q (where
n q is the quantum concentration). The quantum concentration is when the
interparticle distance is equal to the thermal de Broglie wavelength i.e. when
the wavefunctions of the particles are touching but not overlapping. As the
quantum concentration depends on temperature; high temperatures will put
most systems in the classical limit unless they have a very high density e.g.
a White dwarf. Fermi-Dirac statistics apply to fermions (particles that obey
A
the Pauli exclusion principle), Bose-Einstein statistics apply to bosons. Both
Fermi-Dirac and Bose-Einstein become Maxwell-Boltzmann statistics at high
temperatures or low concentrations.
Maxwell-Boltzmann statistics are often described as the statistics of "distin-
guishable" classical particles. In other words the configuration of particle A in
state 1 and particle B in state 2 is different from the case where particle B is
in state 1 and particle A is in state 2. When this idea is carried out fully, it
yields the proper (Boltzmann) distribution of particles in the energy states, but
DR
yields non-physical results for the entropy, as embodied in →Gibbs paradox.
These problems disappear when it is realized that all particles are in fact in-
distinguishable. Both of these distributions approach the Maxwell-Boltzmann
distribution in the limit of high temperature and low density, without the need
for any ad hoc assumptions. Maxwell-Boltzmann statistics are particularly use-
ful for studying gases F-D statistics are most often used for the study of elec-
trons in solids. As such, they form the basis of semiconductor device theory
and electronics.
Bosons, unlike fermions, are not subject to the Pauli exclusion principle: an
unlimited number of particles may occupy the same state at the same time.
This explains why, at low temperatures, bosons can behave very differently
than fermions; all the particles will tend to congregate together at the same
lowest-energy state, forming what is known as a Bose-Einstein condensate.

Bose-Einstein statistics
31

B-E statistics was introduced for photons in 1920 by Bose and generalized to
atoms by Einstein in 1924.
The expected number of particles in an energy state i for B-E statistics is:
gi
ni =

FT
e(i −µ)/kT −1

with i > µ and where:


n i is the number of particles in state i

g i is the degeneracy of state i

 i is the energy of the i -th state

µ is the chemical potential

k is Boltzmann’s constant
AT is absolute temperature

exp is the exponential function

This reduces to M-B statistics for energies (  i-µ ) » kT.

A Derivation of the Bose-Einstein distribution


DR
Suppose we have a number of energy levels, labelled by index i, each level
having energy  i and containing a total of n i particles. Suppose each level
contains g i distinct sublevels, all of which have the same energy, and which
are distinguishable. For example, two particles may have different momenta, in
which case they are distinguishable from each other, yet they can still have the
same energy. The value of g i associated with level i is called the "degeneracy"
of that energy level. Any number of bosons can occupy the same sublevel.
Let w(n,g) be the number of ways of distributing n particles among the g
sublevels of an energy level. There is only one way of distributing n particles
with one sublevel, therefore w(n,1) = 1. It’s easy to see that there are n + 1
ways of distributing n particles in two sublevels which we will write as:
(n+1)!
w(n, 2) = n!1! .

Bose-Einstein statistics
32

With a little thought it can be seen that the number of ways of distributing n
particles in three sublevels is w(n,3) = w(n,2) + w(n-1,2) + ... + w(0,2) so
that
(n−k+1)! (n+2)!
w(n, 3) = nk=0 w(n − k, 2) = nk=0 (n−k)!1! = n!2!
P P

FT
where we have used the following theorem involving binomial coefficients:
Pn (k+a)! (n+a+1)!
k=0 k!a! = n!(a+1)! .

Continuing this process, we can see that w(n,g) is just a binomial coefficient
(n+g−1)!
w(n, g) = n!(g−1)!
.

The number of ways that a set of occupation numbers n i can be realized is the
product of the ways that each individual energy level can be populated:
Q Q (n +g −1)! Q (n +g )!
W = i w(ni , gi ) = i n i!(g i−1)! ≈ i n i!(g i)!
i i i i

where the approximation assumes that gi >> 1. Following the same procedure
A
used in deriving the →Maxwell-Boltzmann statistics, we wish to find the set of
n i for which W is maximised, subject to the constraint that there be a fixed
number of particles, and a fixed energy. The maxima of W and ln(W ) occur
at the value of Ni and, since it is easier to accomplish mathematically, we will
maximise the latter function instead. We constrain our solution using Lagrange
multipliers forming the function:
P P
f (ni ) = ln(W ) + α(N − ni ) + β(E − ni i )
DR
Using the gi >> 1 approximation and using Stirling’s approximation for the
factorials (ln(x!) ≈ x ln(x) − x) gives:
P P
f (ni ) = i (ni + gi ) ln(ni + gi ) − ni ln(ni ) − gi ln(gi ) + α(N − ni ) + β(E −
P
ni i )

Taking the derivative with respect to n i, and setting the result to zero and
solving for n i yields the Bose-Einstein population numbers:
gi
ni = eα+βi −1

It can be shown thermodynamically that β = 1/kT where k is Boltzmann’s


constant and T is the temperature, and that α = -µ/kT where µ is the chemical
potential, so that finally:

Bose-Einstein statistics
33
gi
ni =
e(i −µ)/kT −1

Note that the above formula is sometimes written:


gi
ni =

FT
ei /kT /z−1

where z = exp(µ/kT ) is the absolute activity.

History
In the early 1920s Satyendra Nath Bose was intrigued by Einstein’s theory of
light waves being made of particles called photons. Bose was interested in
deriving Planck’s radiation formula, which Planck obtained largely by guessing.
In 1900 Max Planck had derived his formula by manipulating the math to
fit the empirical evidence. Using the particle picture of Einstein, Bose was
able to derive the radiation formula by systematically developing a statistics of
massless particles without the constraint of particle number conservation. Bose
derived Planck’s Law of Radiation by proposing different states for the photon.
Instead of statistical independence of particles, Bose put particles into cells and
A
described statistical independence of cells of phase space. Such systems allow
two polarization states, and exhibit totally symmetric wavefunctions.
He was quite successful in that he developed a statistical law governing the
behaviour pattern of photons. However he was not able to publish his work,
because no journals in Europe would accept his paper being unable to under-
stand it. Bose sent his paper to Einstein who saw the significance of it and he
used his influence to get it published.
DR
See also
• →Maxwell-Boltzmann statistics
• →Fermi-Dirac statistics
• Parastatistics
• →Planck’s law of black body radiation

Source: http://en.wikipedia.org/wiki/Bose-Einstein_statistics

Principal Authors: PAR, Michael Hardy, Stevenj, Voyajer, Mct mht, Salix alba, Phys, Youandme

Bose-Einstein statistics
34

Bose gas

An ideal Bose gas is a quantum-mechanical version of a classical ideal gas.

FT
It is composed of bosons, which have an integral value of spin, and obey
→Bose-Einstein statistics. The statistical mechanics of bosons were developed
by Satyendra Nath Bose for photons, and extended to massive particles by Al-
bert Einstein who realized that an ideal gas of bosons would form a condensate
at a low enough temperature, unlike a classical ideal gas. This condensate is
known as a Bose-Einstein condensate.

The Thomas-Fermi approximation


The thermodynamics of an ideal Bose gas is best calculated using the grand
partition function. The grand partition function for a Bose gas is given by:
Q −gi
Z(z, β, V ) = i 1 − ze−βi

where each term in the product corresponds to a particular energy  i , g i is the


A
number of states with energy  i , z is the absolute activity, which may also be
expressed in terms of the chemical potential µ by:
z(β, µ) ≡ eβµ

and β defined as:


1
β≡ kT
DR
where k is Boltzmann’s constant and T is the temperature. All thermodynamic
quantities may be derived from the grand partition function and we will con-
sider all thermodynamic quantities to be functions of only the three variables
z , β (or T ), and V . All partial derivatives are taken with respect to one of these
three variables while the other two are held constant. It is more convenient to
deal with the dimensionless
P grand potential
 defined as:
Ω = − ln(Z) = i gi ln 1 − ze−βi

Following the procedure described in the gas in a box article, we can apply the
Thomas-Fermi approximation which assumes that the average energy is large
compared to the energy difference between levels so that the above sum may
be replaced by an integral:
R∞
Ω ≈ 0 ln 1 − ze−βE dg


Bose gas
35

The degeneracy dg may be expressed for many different situations by the


general formula:
1 E α−1
dg = Γ(α) Ecα
dE

FT
where α is a constant, Ec is a "critical energy", and Γ is the Gamma function.
For example, for a massive Bose gas in a box, α=3/2 and the critical energy is
given by:
1 Vf
(βEc )α
= Λ3

where is the thermal wavelength. For a massive Bose gas in a harmonic trap
we will have α=3 and the critical energy is given by:
1 f
(βEc )α
= (~ωβ)3

where V(r) =mω 2r 2/2 is the harmonic potential. It is seen that E c is a function
of volume only.
We can solve the equation for the grand potential by integrating the Taylor
A
series of the integrand term by term, or by realizing that it is proportional to
the Mellin transform of the Li 1(z exp(-β E)) where Li s(x) is the polylogarithm
function. The solution is:
Ω ≈ − Li(βE
α+1 (z)
c)
α

The problem with this continuum approximation for a Bose gas is that the
ground state has been effectively ignored, giving a degeneracy of zero for zero
energy. This inaccuracy becomes serious when dealing with the Bose-Einstein
DR
condensate and will be dealt with in the next section.

Inclusion of the ground state


The total number of particles is found from the grand potential by
Liα (z)
N = −z ∂Ω
∂z ≈ (βEc )α

The polylogarithm term must remain real and positive, and the maximum value
it can possibly have is at z=1 where it is equal to ζ(α) where ζ is the Riemann
zeta function. For a fixed N , the largest possible value that β can have is a
critical value β c where
ζ(α)
N= (βc Ec )α

Bose gas
36

This corresponds to a critical temperature T c=1/kβ c below which the Thomas-


Fermi approximation breaks down. The above equation can be solved for the
critical temperature:
 1/α
N Ec
Tc = ζ(α)

FT
k

For example, for α = 3/2 and using the above noted value of Ec yields
 2/3 2
N h
Tc = V f ζ(3/2) 2πmk

Again, we are presently unable to calculate results below the critical tempera-
ture, because the particle numbers using the above equation become negative.
The problem here is that the Thomas-Fermi approximation has set the degen-
eracy of the ground state to zero, which is wrong. There is no ground state to
accept the condensate and so the equation breaks down. It turns out, howev-
er, that the above equation gives a rather accurate estimate of the number of
particles in the excited states, and it is not a bad approximation to simply "tack
on" a ground state term:
Liα (z)
A
N = N0 + (βEc )α

where N 0 is the number of particles in the ground state condensate:


g0 z
N0 = 1−z

This equation can now be solved down to absolute zero in temperature. Figure
1 shows the results of the solution to this equation for α=3/2, with k= c=1
DR
which corresponds to a gas of bosons in a box. The solid black line is the
fraction of excited states 1-N 0/N for N =10,000 and the dotted black line is
the solution for N =1000. The blue lines are the fraction of condensed particles
N 0/N The red lines plot values of the negative of the chemical potential µ and
the green lines plot the corresponding values of z . The horizontal axis is the
normalized temperature τ defined by
T
τ= Tc

It can be seen that each of these parameters become linear in τ α in the limit of
low temperature and, except for the chemical potential, linear in 1/τ α in the
limit of high temperature. As the number of particles increases, the condensed
and excited fractions tend towards a discontinuity at the critical temperature.
The equation for the number of particles can be written in terms of the nor-
malized temperature as:

Bose gas
37

N= g0 z
1−z +N ζ(α) τα
FT
Figure 4 Figure 1: Various Bose gas parameters as a function of normalized temperature τ . The
value of α is 3/2. Solid lines are for N=10,000, dotted lines are for N=1000. Black lines are
A
the fraction of excited particles, blue are the fraction of condensed particles. The negative of the
chemical potential µ is shown in red, and green lines are the values of z. It has been assumed that
k = c=1.

Liα (z)

For a given N and τ , this equation can be solved for τ α and then a series
solution for z can be found by the method of inversion of series, either in
powers of τ α or as an asymptotic expansion in inverse powers of τ α. From
DR
these expansions, we can find the behavior of the gas near T =0 and in the
Maxwell-Boltzmann as T approaches infinity. In particular, we are interested
in the limit as N approaches infinity, which can be easily determined from
these expansions.

Thermodynamics
Adding the ground state to the equation for the particle number corresponds
to adding the equivalent ground state term to the grand potential:
Liα+1 (z)
Ω = g0 ln(1 − z) − (βEc )α

All thermodynamic properties may now be computed from the grand poten-
tial. The following table lists various thermodynamic quantities calculated in

Bose gas
38

the limit of low temperature and high temperature, and in the limit of infi-
nite particle number. An equal sign (=) indicates an exact result, while an
approximation symbol indicates that only the first few terms of a series in τ α is
shown.

FT
Quantity General T  Tc T  Tc
z =1 ζ(α) ζ 2 (α)
≈ τ α − α 2α
2 τ
Vapor fraction Liα (z) = τα =1
= ζ(α) τα
1− N
N
0

Liα+1 (z) ζ(α+1) ζ(α)


Equation of state
= τα = ζ(α) τ α ≈ 1 − α+1 α
PV β Ω ζ(α) 2 τ
N = −N  
Gibbs Free Energy = ln(z) =0 ζ(α) ζ(α)
≈ ln τα
− 2α τ α
G = ln(z)

It is seen that all quantities approach the values for a classical ideal gas in the
limit of large temperature. The above values can be used to calculate other
thermodynamic quantities. For example, the relationship between internal en-
ergy and the product of pressure and volume is the same as that for a classical
ideal gas over all temperatures:
∂Ω
U= = αP V
A ∂β

A similar situation holds for the specific heat at constant volume


∂U
Cv = ∂T = k(α + 1) U β

The entropy is given by:


TS = U + PV − G
DR
Note that in the limit of high temperature, we have
 α 
τ
T S = (α + 1) + ln ζ(α)

which, for α=3/2 is simply a restatement of the Sackur-Tetrode equation.

See also
• Gas in a box
• Debye model
• Bose-Einstein condensate

Bose gas
39

References
• Huang, Kerson, "Statistical Mechanics", John Wiley and Sons, New York,
1967.

FT
• A. Isihara, "Statistical Physics", Academic Press, New York, 1971.
• L. D. Landau and E. M. Lifshitz, "Statistical Physics, 3rd Edition Part 1",
Butterworth-Heinemann, Oxford, 1996.
• C. J. Pethick and H. Smith, "Bose-Einstein Condensation in Dilute Gases",
Cambridge University Press, Cambridge, 2004.
• Zijun Yan, "General Thermal Wavelength and its Applications", Eur. J. Phys,
21 (2000), 625-631. online 30

Source: http://en.wikipedia.org/wiki/Bose_gas

Principal Authors: PAR, Schneelocke, Tom davis, SimonP

Bound state
A
In physics, a bound state is a composite of two or more building blocks (parti-
cles or bodies) that behaves as a single object. In quantum mechanics (where
the number of particles is conserved), a bound state is a state in the →Hilbert
space that corresponds to two or more particles whose interaction energy is
negative, and therefore these particles cannot be separated unless energy is
spent. The energy spectrum of a bound state is discrete, unlike the continuous
DR
spectrum of isolated particles. (Actually, it is possible to have unstable bound
states with a positive interaction energy provided that there is a "energy bar-
rier" that has to be tunnelled through in order to decay. This is true for some
radioactive nuclei.)
In general, a stable bound state is said to exist in a given potential of some
dimension if stationary wavefunctions exist (normalized in the range of the
potential). The energy of these wavefunctions is negative.
In relativistic quantum field theory, a stable bound state of n particles with
masses m 1, ..., m n shows up as a pole in the S-matrix with a center of mass
energy which is less than m 1+...+m n. An unstable bound state (see resonance)
shows up as a pole with a complex center of mass energy.

30 http://www.iop.org/EJ/article/0143-0807/21/6/314/ej0614.pdf

Bound state
40

Examples
• A proton and an electron can move separately; the total center-of-mass
energy is positive, and such a pair of particles can be described as an ionized

FT
atom. Once the electron starts to "orbit" the proton, the energy becomes
negative, and bound states - namely the hydrogen atom - is formed. Only
the lowest energy bound state, the ground state is stable. The other excited
states are unstable and will decay into bound states with less energy by
emitting a photon.
• A nucleus is a bound state of protons and neutrons (nucleons).
• A positronium "atom" is an unstable bound state of an electron and a
positron. It decays into photons.
• The proton itself is a bound state of three quarks (two up and one down;
one red, one green and one blue). However, unlike the case of the hydrogen
atom, the individual quarks can never be isolated. See confinement.

See also


AComposite field
→Resonance

Source: http://en.wikipedia.org/wiki/Bound_state

Principal Authors: Phys, SeventyThree, Dmr2, Conscious, Tony Sidaway


DR
Bra-ket notation

Bra-ket notation is the standard notation for describing quantum states in the
theory of quantum mechanics. It can also be used to denote abstract vectors
and linear functionals in pure mathematics. It is so called because the inner
product of two states is denoted by a bracket, hφ|ψi, consisting of a left part,
hφ|, called the bra, and a right part, |ψi, called the ket. The notation was
invented by Paul Dirac, and is also known as Dirac notation. It is also the
notation of choice in quantum computing.

Bra-ket notation
41

Bras and kets


In quantum mechanics, the state of a physical system is identified with a vector
in a complex →Hilbert space, H. Each vector is called a "ket", and written as

FT
|ψi

where ψ denotes the particular ket, read as "psi ket."


Every ket |ψi has a dual bra, written as
hψ|

This is a continuous linear function from H to the complex numbers C, defined


by:
 
hψ|ρi = |ψi , |ρi for all kets |ρi

where ( , ) denotes the inner product defined on the Hilbert space. The bra is
simply the conjugate transpose (also called the Hermitian conjugate) of the ket
and vice versa. The notation is justified by the Riesz representation theorem,
A
which states that a Hilbert space and its dual space are isometrically isomor-
phic. Thus, each bra corresponds to exactly one ket, and vice versa. This is not
always the case; on page 111 of Quantum Mechanics by Cohen-Tannoudji et al.
it is clarified that there is such a relationship between bras and kets, so long as
the defining functions used are square integrable. Consider a continuous basis
and a Dirac delta function or a sine or cosine wave as a wave function. Such
functions are not square integrable and therefore it arises that there are bras
that exist with no corresponding ket. This does not hinder quantum mechanics
DR
because all physically realistic wave functions are square integrable.
Bra-ket notation can be used even if the vector space is not a Hilbert space. In
any Banach space B, the vectors may be notated by kets and the continuous
linear functionals by bras. Over any vector space without topology, we may
also notate the vectors by kets and the linear functionals by bras. In these more
general contexts, the bracket does not have the meaning of an inner product,
because the Riesz representation theorem does not apply.
Applying the bra hφ| to the ket |ψi results in a complex number, called a "bra-
ket" or "bracket", which is written as
hφ|ψi.

In quantum mechanics, this is the probability amplitude for the state ψ to col-
lapse into the state φ.

Bra-ket notation
42

Properties
Because each ket is a vector in a complex →Hilbert space and each bra-ket is
an inner product, it follows directly that bras and kets can be manipulated in
the following ways:

FT
• Given any bra hφ|, kets |ψ1 i and |ψ2 i, and complex numbers c 1 and c 2,
then, since bras are linear functionals,
 
hφ| c1 |ψ1 i + c2 |ψ2 i = c1 hφ|ψ1 i + c2 hφ|ψ2 i.

• Given any ket |ψi, bras hφ1 | and hφ2 |, and complex numbers c 1 and c 2, then,
by the definition of addition and scalar multiplication of linear functionals,
 
c1 hφ1 | + c2 hφ2 | |ψi = c1 hφ1 |ψi + c2 hφ2 |ψi.

• Given any kets |ψ1 i and |ψ2 i, and complex numbers c 1 and c 2, from the
Aproperties of the inner product (with c* denoting the complex conjugate of
c),

c1 |ψ1 i + c2 |ψ2 i is dual to c∗1 hψ1 | + c∗2 hψ2 |.

• Given any bra hφ| and ket |ψi, an axiomatic property of the inner product
gives
DR
hφ|ψi = hψ|φi∗ .

Linear operators
If A : H → H is a linear operator, we can apply A to the ket |ψi to obtain
the ket (A|ψi). Linear operators are ubiquitous in the theory of quantum me-
chanics. For example, hermitian operators are used to represent observable
physical quantities, such as energy or momentum, whereas unitary linear op-
erators represent transformative processes such as rotation or the progression
of time.
Operators can also be viewed as acting on bras from the right hand side. Com-
posing the bra hφ| with the operator A results in the bra (hφ|A), defined as a
linear functional on H by the rule

Bra-ket notation
43
   
hφ|A |ψi = hφ| A|ψi .

This expression is commonly written as

FT
hφ|A|ψi.

A convenient way to define linear operators on H is given by the outer product:


if hφ| is a bra and |ψi is a ket, the outer product
|φihψ|

denotes the rank one operator that maps the ket |ρi to the ket |φihψ|ρi (where
hψ|ρi is a scalar multiplying the vector |φi). One of the uses of the outer prod-
uct is to construct projection operators. Given a ket |ψi of norm 1, the orthog-
onal projection onto the subspace spanned by |ψi is
|ψihψ|.

Composite bras and kets


A
Two Hilbert spaces V and W may form a third space V ⊗W by a tensor product.
In quantum mechanics, this is used for describing composite systems. If a
system is composed of two subsystems described by V and W respectively, then
the Hilbert space of the entire system is the tensor product of the two spaces.
(The exception to this is if the subsystems are actually identical particles. In
that case, the situation is a little more complicated.)
If |ψi is a ket in V and |φi is a ket in W, the tensor product of the two kets is a
DR
ket in V ⊗ W . This is written variously as
|ψi|φi or |ψi ⊗ |φi or |ψφi or |ψ, φi.

Representations in terms of bras and kets


In quantum mechanics, it is often convenient to work with the projections of
state vectors onto a particular basis, rather than the vectors themselves. The
reason is that the former are simply complex numbers, and can be formulated
in terms of partial differential equations (see, for example, the derivation of
the position-basis →Schrödinger equation). This process is very similar to the
use of coordinate vectors in linear algebra.
For instance, the Hilbert space of a zero-spin point particle is spanned by a
position basis {|xi}, where the label x extends over the set of position vectors.

Bra-ket notation
44

Starting from any ket |ψi in this Hilbert space, we can define a complex scalar
function of x, known as a wavefunction:
ψ(x) ≡ hx|ψi.

FT
It is then customary to define linear operators acting on wavefunctions in terms
of linear operators acting on kets, by
Aψ(x) ≡ hx|A|ψi.

For instance, the momentum operator p has the following form:


pψ(x) ≡ hx|p|ψi = −i~∇ψ(x).

One occasionally encounters an expression like


−i~∇|ψi.

This is something of an abuse of notation, though a fairly common one. The


differential operator must be understood to be an abstract operator, acting on
A
kets, that has the effect of differentiating wavefunctions once the expression is
projected into the position basis:
−i~∇hx|ψi.

For further details, see rigged Hilbert space.

Further reading
DR
• Feynman, Leighton and Sands (1965). The Feynman Lectures on Physics
Vol. III. Addison-Wesley. ISBN 0201021153.

Source: http://en.wikipedia.org/wiki/Bra-ket_notation

Principal Authors: CYD, MathKnight, AxelBoldt, Trewornan, Laurascudder, Mct mht, AugPi, Theresa
knott, Ancheta Wis

Bra-ket notation
45

Breit equation

The Breit equation is a relativistic wave equation derived by Gregory Breit in

FT
1929 based on the Dirac equation, which formally describes two or more mas-
sive spin-1/2 particles (electrons, for example) interacting electromagnetically
to the first order in perturbation theory. It accounts for magnetic interactions
and retardation effects to the order of 1/c 2. When other quantum electrody-
namic effects are negligible, this equation has been shown to give results in
good agreement with experiment.

Introduction
The Breit equation is not only an approximation in terms of quantum mechan-
ics, but also in terms of relativity theory as it is not completely invariant with
respect to the Lorentz transformation. Just as does the Dirac equation, it treats
nuclei as point sources of an external field for the particles it describes. For N
particles, the Breit equation has the form (r ij is the distance between particle i
and j ):
A nP
i ĤD (i) +
P 1
i>j rij −
P o
i>j B̂ij Ψ = EΨ,

where
h i
ĤD (i) = qi φ(ri ) + c s=x,y,z αs (i)πs (I) + α0 (I)m0 c2
P

is the Dirac hamiltonian (see Dirac equation) for particle i at position r i and
φ( r i) is the scalar potential at that position; q i is the charge of the particle,
DR
thus for electrons q i = - e.
The one-electron Dirac hamiltonians of the particles, along with their instan-
taneous Coulomb interactions 1/r ij, form the Dirac-Coulomb operator. To this,
Breit added the operator (now known as the Breit operator):
 
(a(i)·rij )(a(j)·rij )
B̂ij = 2r1ij a(i) · a(j) + r2
,
ij

where the Dirac matrices for electron i : a(i) = [α x(i),α y(i),α z(i)]. The two
terms in the Breit operator account for retardation effects to the first order.
The wave function Ψ in the Breit equation is a spinor with 4 N elements, since
each electron is described by a Dirac bispinor with 4 elements as in the Dirac
equation and total wave function is the cartesian product of these.

Breit equation
46

Breit hamiltonians
The total hamiltonian of the Breit equation, sometimes called the Dirac-
Coulomb-Breit hamiltonian (H DCB) can be decomposed into the following
practical energy operators for electrons in electric and magnetic fields (also

FT
called the Breit-Pauli hamiltonian) 1, which have well-defined meanings in
the interaction of molecules with magnetic fields (for instance for nuclear mag-
netic resonance):
B̂ij = Ĥ0 + Ĥ1 + ... + Ĥ6 ,

in which the consequitive partial operators are:


P p̂2
• Ĥ0 = i 2mi i + V is the nonrelativistic hamiltonian (m_{i} is the stationary
mass of particle i ).
p̂4i
Ĥ1 = − 8c12
P
• i m3 is connected to the dependence of mass on velocity:
2
 i
2 2
Ekin − m0 c = m2 v 2 c2 .


AĤ2 = −
P qi qj
i>j 2rij mi mj c2

p̂i · p̂j +
rij (rij p̂i )·p̂j
2
rij


counts for retardation and can be described as the interaction between the
is a correction that partly ac-

magnetic dipole moments of the particles, which arise from the orbital mo-
tion of charges (also called orbit-orbit interaction).
 
• Ĥ3 = µcB i m1i si · F(rij ) × p̂i + j>i 2q
P P
3 rij × p̂j is the classical interac-
i
r ij
tion between the orbital magnetic moments (from the orbital motion of
DR
charge) and spin magnetic moments (also called spin-orbit interaction).
The first term describes the interaction of a particles spin with its own or-
bital moment (F ( r i) is the electric field at the particles position), and the
second term between two different particles.
ih P qi
• Ĥ4 = 8πc2 i m2 p̂i · F(ri ) is a nonclassical term characteristic for Dirac
i
theory, sometimes called the Darwin term.
  
(si ·rij )(sj ·rij )
• Ĥ5 = 4µ2B i>j − 8π 1
P
3 (s i · s j )δ(rij + r 3 si · sj − 2
rij
is the mag-
ij
netic moment spin-spin interaction. The first term is called the contact
interaction, because it is nonzero only when the particles are at the same
position; the second term is the interaction of the classical dipole-dipole
type.

Breit equation
47
P h qi
i
• Ĥ6 = 2µB i H(ri ) · si + mi c A(ri ) · p̂i is the interaction between spin and
orbital magnetic moments with an external magnetic field H.

Notes

FT
• Note 1: H.A. Bethe, E.E. Salpeter, Quantum Mechanics of One- and Two-
Electron Atoms, Plenum Press, New York 1977, pg.181

Source: http://en.wikipedia.org/wiki/Breit_equation

Principal Authors: Karol Langner, Neilc, CambridgeBayWeather

Canonical commutation relation

In physics, the canonical commutation relation is the relation


[x, p] = i~
A
among the position x and momentum p of a point particle in one dimension,
where [x, p] = xp−px is the so-called commutator of x and p, i is the imaginary
unit and ~ is the reduced Planck’s constant h/2π. This relation is attributed to
Heisenberg, and it implies his uncertainty principle.

Relation to classical mechanics


DR
By contrast, in classical physics all observables commute and the commutator
would be zero; however, an analogous relation exists, which is obtained by
replacing the commutator with the Poisson bracket and the constant i~ with 1:
{x, p} = 1

This observation led Dirac to postulate that, in general, the quantum counter-
parts fˆ, ĝ of classical observables f, g should satisfy
[fˆ, ĝ] = i~{f,
\ g}.

In 1927, Hermann Weyl showed that a literal correspondence between a quan-


tum operator and a classical distribution in phase space could not hold. How-
ever, he did propose a mechanism, Weyl quantization, that underlies a mathe-
matical approach to quantization known as deformation quantization.

Canonical commutation relation


48

Representations
According to the standard mathematical formulation of quantum mechanics,
quantum observables such as x and p should be represented as self-adjoint
operators on some →Hilbert space. It is relatively easy to see that two oper-

FT
ators satisfying the canonical commutation relations cannot both be bounded.
The canonical commutation relations can be made tamer by writing them in
terms of the (bounded) unitary operators e−ikx and e−iap . The result is the so-
called Weyl relations. The uniqueness of the canonical commutation relations
between position and momentum is guaranteed by the Stone-von Neumann
theorem. The group associated with the commutation relations is called the
Heisenberg group.

Generalizations
The simple formula
[x, p] = i~,

valid for the quantization of the simplest classical system, can be generalized
A
to the case of an arbitrary Lagrangian L . We identify canonical coordinates
(such as x in the example above, or a field φ(x) in the case of quantum field
theory) and canonical momenta πx (in the example above it is p, or more
generally, some functions involving the derivatives of the canonical coordinates
with respect to time).
∂L
πi ≡ ∂(∂xi /∂t)
DR
This definition of the canonical momentum ensures that one of the Euler-
Lagrange equations has the form
∂ ∂L
∂t πi = ∂xi

The canonical commutation relations then say


[xi , πj ] = i~δij

where δij is the Kronecker delta.

Gauge invariance
Canonical quantization is performed, by definition, on canonical coordinates.
However, in the presence of an electromagnetic field, the canonical momentum

Canonical commutation relation


49

p is not gauge invariant. The correct gauge-invariant momentum (or "kinetic


momentum") is
p − eA/c

FT
where e is the quantum of electric charge, and A is the vector potential and c
is the speed of light. Although this quantity is the "physical momentum" in that
it is the quantity to be identified with momentum in laboratory experiments, it
does not satisfy the canonical commutation relations; only the kinetic momen-
tum does that. This can be seen as follows.
The non-relativistic Hamiltonian for a quantized charged particle of mass m in
a classical electromagnetic field is
 2
1
H = 2m p − eA
c + eφ

where A is the three-vector potential and φ is the scalar potential. This form
of the Hamiltonian, as well as the Schroedinger equation Hψ = i~∂ψ/∂t, the
Maxwell equations and the Lorentz force law are invariant under the gauge
transformation
AA → A0 = A + ∇Λ

φ → φ0 − 1 ∂Λ
c ∂t

ψ → ψ0 = U ψ

H → H 0 = U HU †
DR
where
 
ieΛ
U = exp ~c

and Λ = Λ(x, t) is the gauge function.


The canonical angular momentum is
L=r×p

and obeys the canonical quantization relations


[Li , Lj ] = i~ijk Lk

Canonical commutation relation


50

defining the Lie algebra for so(3), where ijk is the Levi-Civita symbol. Under
gauge transformations, the angular momentum transforms as
hψ|L|ψi → hψ 0 |L0 |ψ 0 i = hψ|L|ψi + e
~c hψ|r × ∇Λ|ψi

FT
The gauge-invariant angular momentum (or "kinetic angular momentum") is
given by
 
K = r × p − eAc

which has the commutation relations


 
[Ki , Kj ] = i~ij k Kk + e~
c xk (x · B)

where
B =∇×A

is the magnetic field. The inequivalence of these two formulations shows up in


the →Zeeman effect and the →Aharonov-Bohm effect.
A
See also
• canonical quantization
• CCR algebra
• Lie derivative
DR
Source: http://en.wikipedia.org/wiki/Canonical_commutation_relation

Principal Authors: Miguel, Ancheta Wis, Linas, Lumidek, Michael Hardy

Chladni’s law

Chladni’s law, named after Ernst Chladni, relates the frequency of modes of
vibration for flat circular surfaces with fixed center as a function of the numbers
m of diametric (linear) nodes and n of radial (circular) nodes. It is stated as
the equation
f = C(m + 2n)p

where C and p are coefficients which depend on the properties of the plate.

Chladni’s law
51

For flat circular plates, p is roughly 2, but Chladni’s law can also be used to
describe the vibrations of cymbals, handbells, and church bells in which case p
can vary from 1.4 to 2.4. In fact, p can even vary for a single object, depending
on which family of modes is being examined.

FT
External links
• A Study of Vibrating Plates 31 by Derek Kverno and Jim Nolen

Source: http://en.wikipedia.org/wiki/Chladni%27s_law

Principal Authors: Laurascudder, Michael Hardy, Choster

Classical limit

The classical limit is the ability of a physical theory to approximate or "recover"


A
classical mechanics when considered over special values of its parameters. The
classical limit is used with physical theories that predict non-classical behavior.
A postulate called the correspondence principle was introduced to quantum
theory by Niels Bohr; it states that, in effect, some kind of continuity argument
should apply to the classical limit of quantum systems as the value of Planck’s
constant tends to zero.
In quantum mechanics, due to the Heisenberg’s uncertainty principle, an elec-
tron can never be at rest; it must always have a non-zero kinetic energy, a result
DR
not found in classical mechanics. For example, if we consider something very
large relative to an electron, like a baseball, the uncertainty principle predicts
that it cannot have zero kinetic energy, but the uncertainty in kinetic energy
is so small that the baseball can appear to be at rest, and hence appears to
obey classical mechanics. In general, if large energies and large objects (rel-
ative to the size and energy levels of an electron) are considered in quantum
mechanics, the result will appear to obey classical mechanics.
In general and special relativity, if we consider flat space, small masses, and
small speeds (in comparison to the speed of light), we find that objects once
again appear to obey classical mechanics.

Source: http://en.wikipedia.org/wiki/Classical_limit

31 http://www.phy.davidson.edu/StuHome/derekk/Chladni/pages/menu.htm

Classical limit
52

Principal Authors: Cyan, Michael Hardy, Charles Matthews, Salsb

Clebsch-Gordan coefficients

FT
In physics, the Clebsch-Gordan coefficients are sets of numbers that arise
in calculations involving addition of angular momentum under the laws of
quantum mechanics.
In more mathematical terms, the CG coefficients are used in representation
theory, particularly of compact Lie groups, to perform the explicit direct sum
decomposition of the tensor product of two irreducible representations into ir-
reducible representations, in cases where the numbers and types of irreducible
components are already known abstractly. The name derives from the German
mathematicians Alfred Clebsch (1833-1872) and Paul Gordan (1837-1912),
who encountered an equivalent problem in invariant theory.
In terms of classical mathematics, the CG coefficients, or at least those asso-
ciated to the group SO(3), may be defined much more directly, by means of
A
formulae for the multiplication of spherical harmonics. The addition of spins
in quantum-mechanical terms can be read directly from this approach. The
formulas below use Dirac’s bra-ket notation.

Formal definition and some results


The Clebsch-Gordan coefficients are the numerical constants that express the
probability amplitude for the spins j1 , j2 with z-projections m1 , m2 to add to j
with z projection m
DR
|j1 j2 ; jmi = jm1 1 =−j1 jm2 2 =−j2 |j1 j2 ; m1 m2 ihj1 j2 ; m1 m2 |j1 j2 ; jmi
P P

where hj1 j2 ; m1 m2 |j1 j2 ; jmi are the CG coefficients.


m = m1 + m2 if hj1 j2 ; m1 m2 |j1 j2 ; jmi =
6 0.

The following holds:


Pj1 Pj2
|j1 j2 ; jmi = m 1 =−j1 m2 =−j2 |j1 j2 ; m1 m2 ihj1 j2 ; m1 m2 |j1 j2 ; jmi

renaming m1 to m01 and m2 to m02 and applying the J± operator


J± |j1 j2 ; jmi = (J1± +J2± ) jm1 0 =−j
Pj2
|j1 j2 ; m01 m02 ihj1 j2 ; m01 m02 |j1 j2 ; jmi
P
1 m0 =−j
1 2 2

We get a some-what long equation:

Clebsch-Gordan coefficients
53
p
(j ∓ m)(j ± m + 1)|j1 j2 ; j, m ± 1i

(j1 ∓ m01 )(j1 ± m01 + 1)|j1 j2 ; m01 ± 1, m02 i + (j2 ∓ m02 )(j2 ±
P P p p
= m01 m02
hj1 j2 ; m01 m02 |j1 j2 ; jmi

FT
and arbitrarily choosing one particular m1 and m2 and multiplying both sides
by a bra on the left (note that with the new m1 and m2 , m1 + m2 = m ± 1 when
the coefficients are not 0)
p
(j ∓ m)(j ± m + 1)hj1 j2 ; m1 m2 |j1 j2 ; j, m ± 1i

(j1 ∓ m01 )(j1 ± m01 + 1)hj1 j2 ; m1 m2 |j1 j2 ; m01 ± 1, m02 i +


P P p p
= m01 m02 (j2 ∓ m
hj1 j2 ; m01 m02 |j1 j2 ; jmi

and summing, now that most elements of the sum are 0 (note that
hj1 j2 ; m1 m2 |j1 j2 ; m01 ± 1, m02 i ≥ 0 if m01 6= m1 ∓ 1 or m02 6= m2 ) etc...
p
(j ∓ m)(j ± m + 1)hj1 j2 ; m1 m2 |j1 j2 ; j, m ± 1i
A p p
= (j1 ∓ m1 + 1)(j1 ± m1 )hj1 j2 ; m1 ∓1, m2 |j1 j2 ; jmi+ (j2 ∓ m2 + 1)(j2 ± m2 )h
1|j1 j2 ; jmi

replacing m with m ∓ 1 so that again m1 + m2 = m when the coefficients are


not 0
p
(j ∓ m + 1)(j ± m)hj1 j2 ; m1 m2 |j1 j2 ; jmi
DR
p
=
p (j1 ∓ m1 + 1)(j1 ± m1 )hj1 j2 ; m1 ∓ 1, m2 |j1 j2 ; j, m ∓ 1i +
(j2 ∓ m2 + 1)(j2 ± m2 )hj1 j2 ; m1 , m2 ∓ 1|j1 j2 ; j, m ∓ 1i

The above formula is useful for finding the last Clebsch-Gordan coefficients,
when the other one or two coefficients in the formula are known. Note that
there are sometimes only two coefficients in the formula, the third being both
invalid (j < |m|) and multiplied by 0. Guessing one of the coefficients, using
the formula to find the rest and normalising so that the first coefficient becomes
correct, one can find, for example (up to sign)

5 1
5 7 3
q 10
2 1; m1 = 2 , m2 = 1 2 1; j = 2 , m = 2 =

21

For more coefficients, see table of Clebsch-Gordan coefficients.

Clebsch-Gordan coefficients
54

See also
• Wigner 3-j symbol
• 6-j symbol

FT
• Spherical harmonics
• Associated Legendre polynomials
• Angular momentum
• Angular momentum coupling

External links
• Java TM Clebsch-Gordan Coefficient Calculator 32

References
• A.R. Edmonds, Angular Momentum in Quantum Mechanics, (1957) Prince-
ton University Press, ISBN 0-691-07912-9.
• E. U. Condon and G. H. Shortley, The Theory of Atomic Spectra, (1970)
A Cambridge at the University Press, ISBN 521-09209-4 See chapter 3.
• Albert Messiah, Quantum Mechanics (Volume II), (1966) North Holland
Publishing, ISBN ???? (something that looks the same but doesn’t mention
1966 is ISBN 0720400457)

Source: http://en.wikipedia.org/wiki/Clebsch-Gordan_coefficients
DR
Principal Authors: Cyp, Charles Matthews, Linas, ArnoldReinhold, JabberWok

Coherent state

In quantum mechanics a coherent state is a specific kind of quantum state of


the quantum harmonic oscillator whose dynamics most closely resemble the
oscillating behaviour of a classical harmonic oscillator system. It was the first
example of quantum dynamics when Erwin Schrödinger derived it in 1926
while searching for solutions of the →Schrödinger equation that satisfy the
correspondence principle. The quantum harmonic oscillator and hence, the co-
herent state, arise in the quantum theory of a wide range of physical systems.
For instance, a coherent state describes the oscillating motion of the particle

32 http://www.gleet.org.uk/cleb/cgjava.html

Coherent state
55

in a quadratic potential well. In the quantum theory of light (quantum elec-


trodynamics) and other bosonic quantum field theories they were introduced
by the work of Roy J. Glauber. Here the coherent state of a field describes an
oscillating field, the closest quantum state to a classical sinusoidal wave such

FT
as a continuous laser wave.

A
DR
Figure 5 Figure 1: The electric field, measured by optical homodyne detec-
tion, as a function of phase for three coherent states emitted by a Nd:YAG laser.
The amount of quantum noise in the electric field is completely independent of
the phase. As the field strength, i.e. the oscillation amplitude α of the coherent
state is increased, the quantum noise or uncertainty is constant at 1/2, and so
becomes less and less significant. In the limit of large field the state becomes
a good approximation of a noiseless stable classical wave. The average pho-
ton numbers of the three states from top to bottom are <n>=4.2, 25.2, 924.5
(source: link 1 and ref. 2)

Coherent states in quantum optics


In classical optics light is thought of as electromagnetic waves radiating from a
source. Specifically, coherent light is thought of as light that is emitted by many
such sources that are in phase. For instance a light bulb radiates light that is
the result of waves being emitted at all the points along the filament. Such
light is incoherent because the process is highly random in space and time (see

Coherent state
56

A
width.
FT
Figure 6 Figure 2: The oscillating wave packet corresponding to the second coherent state de-
picted in Figure 1. At each phase of the light field, the distribution is a Gaussian of constant

thermal light). In a laser, however, light is emitted by a carefully controlled


system in processes that are not random but interconnected by stimulation and
the resulting light is highly ordered, or coherent. Therefore a coherent state
corresponds closely to the quantum state of light emitted by an ideal laser.
Semi-classically we describe such a state by an electric field oscillating as a
DR
stable wave.
Contrary to the coherent state, which is the most wave-like quantum state, the
→Fock state (e.g. a single photon) is the most particle-like state. It is indivisible
and contains only one quanta of energy. These two states are examples of the
opposite extremes in the concept of wave-particle duality. A coherent state
distributes its quantum-mechanical uncertainty equally, which means that the
phase and amplitude uncertainty are approximately equal. Conversely, in a
single-particle state the phase is completely uncertain.

Quantum mechanical definition


Mathematically, the coherent state |αi is defined to be the eigenstate of the
annihilation operator a. Formally, this reads:
a|αi = α|αi

Coherent state
57

A
experimental errors.

Since a is not hermitian,


α = |α|eiθ
FT
Figure 7 Figure 3: Wigner function of the coherent state depicted in Figure 2. The distribution
is centered on state’s amplitude α and is symmetric around this point. The ripples are due to

is complex. Here |α| and θ are called the amplitude and phase of the state.
Physically, this formula means that a coherent state is left unchanged by the
DR
detection (or annihilation) of a particle. Consequently, in a coherent state,
one has exactly the same probability to detect a second particle. Note, this
condition is necessary for the coherent state’s Poissonian detection statistics, as
discussed below. Compare this to a single-particle state (→Fock state): Once
one particle is detected, we have zero probability of detecting another.
For the following discussion we need to define the dimensionless X and P
quadratures. For a harmonic oscillator, x = (mωπ/h) -1/2X is the oscillating
particle’s position and p = (mωh/π) -1/2P is its momentum. For an optical
field, E R = (hω/π 0V) 1/2cosθX ; and E I = (hω/π 0V) 1/2sinθP ; are the real and
imaginary components of the electric field.
Erwin Schrödinger was searching for the most classical-like states when he first
introduced coherent states. He described them as the quantum state of the
harmonic oscillator which minimizes the uncertainty relation with uncertainty
equally distributed in both X and P quadratures (ie. ∆X = ∆Y = 1/2). From

Coherent state
58

A FT
Figure 8 Figure 4: The probability of detecting n photons, the photon number distribution, of the
coherent state in Figure 3. As is necessary for a Poissonian distribution the mean photon number is
equal to the variance of the photon number distribution. Bars refer to theory, dots to experimental
values.

the generalized uncertainty relation, it is shown that such a state |α> must
obey the equation
(P − hP i)|αi = i(X − hXi)|αi

In the general case, if the uncertainty is not equally distributed in the X and P
DR
component, the state is called a squeezed coherent state.
If this formula is written back in terms of a and a †, it becomes:
a|αi = (hXi + ihP i)|αi

The coherent state’s location in the complex plane (phase space) is centered
at the position and momentum of a classical oscillator of the same phase θ
and amplitude (or the same complex electric field value for an electromagnetic
wave). As shown in Figure 2, the uncertainty, equally spread in all directions, is
represented by a disk with diameter 1/2. As the phase increases the coherent
state circles the origin and the disk neither distorts nor spreads. This is the
most similar a quantum state can be to a single point in phase space.
Since the uncertainty (and hence measurement noise) stays constant at 1/2 as
the amplitude of the oscillation increases, the state behaves more and more

Coherent state
59

A FT
Figure 9 Figure 5: Phase space plot of a coherent state. This shows that the uncertainty
in a coherent state is equally distributed in all directions. The horizontal and vertical axes
are the X and P quadratures, respectively (see text). The red dots on the x-axis trace out
the boundaries of the quantum noise in Figure 1.

like a sinusoidal wave, as shown in Figure 1. Conversely, since the vacuum


DR
state |0> is just the coherent state with α=0, all coherent states have the same
uncertainty as the vacuum. Therefore one can interpret the quantum noise of
a coherent state as being due to the vacuum fluctuations.
Furthermore, it is sometimes useful to define a coherent state simply as the
vacuum state displaced to a location α in phase space. Mathematically this is
done by the action of the displacement operator D(α):

|αi = eαa −α∗ a |0i = D(α)|0i

This can be easily obtained, as can virtually all results involving coherent states,
using the representation of the coherent state in the basis of Fock states:
|α|2 P
∞ √ αn
|αi = e− 2 n=0 |ni.
n!

Coherent state
60

A stable classical wave has a constant intensity. Consequently, the probability


of detecting n photons in a given amount of time is constant with time. This
condition ensures there will be shot noise in our detection. Specificially, the
probability of detecting n photons is Poissonian:

FT
hnin
P (n) = e−hni n!

Similarly, the average photon number in coherent state <n>=<a †a>=|α| 2


and the variance (∆n) 2= Var(a †a)=|α| 2, identical to the variance of the Pois-
sonian distribution. Not only does a coherent state go to a classical sinusoidal
wave in the limit of large α but the detection statistics of it are equal to that of
a classical stable wave for all values of α.
This also follows from the fact that for the prediction of the detection results
at a single detector (and time) any state of light can always be modelled as a
collection of classical waves (see degree of coherence). However, for the pre-
diction of higher-order measurement like intensity correlations (which mea-
sure the degree of nth-order coherence) this is not true. The coherent state is
unique in the fact that all n-orders of coherence are equal to 1. It is perfectly
A
coherent to all orders.
There are other reasons why a coherent state can be considered the most clas-
sical state. Roy J. Glauber coined the term "coherent state" and proved they are
produced when a classical electrical current interacts with the electromagnetic
field. In the process he introduced the coherent state to quantum optics. In
general when a quantum state of light is split at a beamsplitter, the two output
modes are entangled. Aharonov proved that coherent states are the only pure
states of light that remain unentangled (and thus classical) when split into two
DR
states.
From Figure 5, simple geometry gives ∆θ=1/2|α|. From this we can see that
there is a tradeoff between number uncertainty and phase uncertainty ∆θ∆n
= 1/2, the number-phase uncertainty relation. This is not a formal uncertainty
relation: there is no uniquely defined phase operator in quantum mechanics.

Mathematical characteristics
The coherent state does not display all the nice mathematical features of a
→Fock state; for instance two different coherent states are not orthogonal:
1
hβ|αi = e− 2 (|β|
2
+|α|2 −2β ∗ α) 6= δ(α − β)

so that if the oscillator is in the quantum state |α> it is also with nonzero
probability in the other quantum state |β> (but the farther apart the states are
situated in phase space, the lower the probability is). However, since they obey

Coherent state
61

a closure relation, any state can be decomposed on the set of coherent states.
They hence form an overcomplete basis in which one can diagonally decom-
pose any state. This is the premise for the Sudarshan-Glauber P representation.
Another difficulty is that a † has no eigenket (and a has no eigenbra). The fol-

FT
lowing formal equality is the closest substitute and turns out to be very useful
for technical computations:
 ∗

a† |αi = ∂α∂
+ α2 |αi

Coherent states of Bose-Einstein condensates


• A Bose-Einstein condensate (BEC) is a collection of boson atoms that are
all in the same quantum state. An approximate theoretical description of
its properties can be derived by assuming the BEC is in a coherent state.
However, unlike photons atoms interact with each other so it now appears
that it more likely to be one of the squeezed coherent states mentioned
above.
A
Generalizations
• In quantum field theory and string theory, a generalization of coherent
states to the case of infinitely many degrees of freedom is used to define
a vacuum state with a different vacuum expectation value from the original
vacuum.

See also
DR
• →Quantum field theory
• Quantum optics
• Electromagnetic field
• degree of coherence
• quantum coherence

External links
• Quantum states of the light field 33

33 http://gerdbreitenbach.de/gallery

Coherent state
62

References
• E. Schrödinger, Naturwissenschaften 14 (1926) 664.
• R.J. Glauber, Phys. Rev. 131 (1963) 2766.

FT
• Loudon, Rodney, The Quantum Theory of Light (Oxford University Press,
2000), [ISBN 0198501773]
• G. Breitenbach, S. Schiller, and J. Mlynek, "Measurement of the quantum
states of squeezed light", Nature, 387, 471 (1997) 34

Source: http://en.wikipedia.org/wiki/Coherent_state

Principal Authors: J S Lundeen, Gerd Breitenbach, Laussy, Charles Matthews, JerryFriedman

Complementarity (physics)

In physics, complementarity is a basic principle of quantum theory, and refers


A
to effects such as the wave-particle duality, in which different measurements
made on a system reveal it to have either particle-like or wave-like properties.
Niels Bohr is usually associated with this concept; in the orthodox form, it is
stated that a quantum mechanical system consisting of a boson or fermion can
either behave as a particle or as wave, but never simultaneously as both. A less
orthodox interpretation is the "duality condition," described by the inequality
proven by Jaeger, Shimony, and Vaidman (G. Jaeger, A. Shimony, and L. Vaid-
man, Phys. Rev. A, Vol. 51, 54 (1995)), and later by Englert (B. Englert, Phys.
DR
Rev. Lett., Vol. 77, 2154 (1996)), which allows wave and particle attributes
to co-exist, but postulates that a stronger manifestation of the particle nature
leads to a weaker manifestation of the wave nature and vice versa.
→Wave-particle duality is considered to be one of the distinguishing character-
istics of quantum mechanics, whose theoretical and experimental development
has been honoured by more than a few Nobel Prizes for Physics. It has been
discussed by prominent physicists for the last 100 years, from the time of Al-
bert Einstein, Niels Bohr and Werner Heisenberg, onwards. On the basis of
Bohr’s principle of complementarity, it is indeed universally accepted that the
observation of two complementary properties, such as position and momen-
tum, requires mutually exclusive experimental measurements.
The emergence of complementarity in a system occurs when one considers the
circumstances under which one attempts to measure its properties; as Bohr

34 http://www.exphy.uni-duesseldorf.de/Publikationen/1997/N387/471z.htm

Complementarity (physics)
63

noted, the principle of complementarity "implies the impossibility of any sharp


separation between the behaviour of atomic objects and the interaction with
the measuring instruments which serve to define the conditions under which
the phenomena appear." It is important to distinguish, as did Bohr in his origi-

FT
nal statements, the principle of complementarity from a statement of the uncer-
tainty principle. For a technical discussion of contemporary issues surrounding
complementarity in physics, see, e.g., 35 (from which parts of this discussion
were drawn.)
Various neutron interferometry experiments demonstrate the subtleness of the
notions of duality and complementarity in an interesting way. In order to pass
through the interferometer, the neutron must act as a wave. Yet upon passage,
the neutron is subject to gravitation, which one would think only particles, and
not waves, are subject to. As the neutron interferometer is rotated through
Earth’s gravitational field a phase change between the two arms of the inter-
ferometer is created, resulting in a change in the constructive and destructive
interference of the neutron waves on exit from the interferometer. Note that
in order to understand the interference effect one must concede that a single
neutron takes both paths through the interferometer at the same time: a sin-
A
gle neutron must "be in two places at once", as it were. Since the two paths
through a neutron interferometer can be as far as five to 15 cm apart, the effect
is hardly microscopic. This is not in contrast to traditional double-slit experi-
ments (or mirror interferometer) where the slits (or mirrors) can be arbitrary
far apart. So in interference and diffraction experiments neutron behaves the
same way as a photon (or an electron) of corresponding wavelength.

The mathematics of two-slit diffraction


DR
This section reviews the mathematical formulation of the double-aperture ex-
periment (see Fig.1). The formulation is in terms of the diffraction and in-
terference of waves. The culmination of the develpment is a presentation of
two numbers that characterizes the visibility of the interference fringes in the
experiment, linked together as the Englert-Greenberger duality relation.The
next will then discuss the orthodox quantum mechanical interpretation of the
duality relation in terms of wave-particle duality. Of this experiment, Richard
Feynman once said that it “has in it the heart of quantum mechanics. In reality
it contains the only mystery´´.
The wave function in the Young double-aperture experiment can be written as
ΨTotal (x) = ΨA (x) + ΨB (x).

35 http://citebase.eprints.org/cgi-bin/citations?id=oai:arXiv.org:quant-ph/0003073

Complementarity (physics)
64

The function
ΨA (x) = CA Ψ0 (x − xA )

is the wave function associated with the pinhole at A centered on xA ; a similar

FT
relation holds for pinhole B. The variable x is a position in space downstream
of the slits. The constants CA and CB are proportionality factors for the cor-
responding wave amplitudes, and Ψ0 (x) is the single hole wave function for
an aperture centered on the origin. The single-hole wave-function is taken to
be that of Fraunhofer diffraction; the pinhole shape is irrelevant, and the pin-
holes are considred to be idealized. The wave is taken to have a fixed incident
momentum p0 = h/λ:
eip0 ·|x|/~
Ψ0 (x) ∝ |x|

where |x| is the radial distance from the pinhole.


To distinguish which pinhole a photon passed through, one needs some mea-
sure of the distinguishability between pinholes. Such a measure is given by

|C |2 −|C |2
D = |CA |2 +|CB |2 = |PA − PB |,
A A B

where
|CA |2
PA = |CA |2 +|CB |2

and
|CB |2
PB = |CA |2 +|CB |2
DR
are the probabilities of finding that the particle passed through aperture A or
aperture B.
We have in particular D = 0 for two symmetric holes and D = 1 for a single
aperture (perfect distinguishability). In the far-field of the two pinholes the two
waves interfere and produce fringes. The intensity of the interference pattern
at a point y in the focal plane (denoted by (F) in the image) is given by
I(y) ∝ 1 + V cos (py d/~ + φ)

where py = h/λ · sin(α) is the momentum of the particle along the y direc-
tion, φ = Arg(CA ) − Arg(CB ) is a fixed phase shift, and d is the separa-
tion between the two pinholes. The angle α from the horizontal is given by
sin(α) ' tan(α) = y/L where L is the distance between the aperture screen
and the far field analysis plane. If a lens is used to observe the fringes in the

Complementarity (physics)
65

rear focal plane (F), the angle is given by sin(α) ' tan(α) = y/f where f is
the focal length of the lens.
The visibility of the fringes is defined by
Imax −Imin
V =

FT
Imax +Imin

where max and min denote the maximum and minimum intensity of the
fringes. Equivalently, this can be written as
|CA ·CB∗ |
V = 2 |C 2 2.
A | +|CB |

In a single hole experiment, the fringe visibility will be zero (as there are no
fringes); that is, V = 0. On the other hand, for a two slit configuration, where
the two slits are indistinguishable, one has perfect visibility; that is, V = 1 or

|CA | = |CB | = 1/ 2. It is straighforward to see that the duality relation
V 2 + D2 = 1

is always true.
A
The above presentation was limited to a pure quantum state. More generally,
for a mixture of quantum states, one will have
V 2 + D2 ≤ 1.

For the remainder of the development, the light source will be assumed to
be a laser, so that V 2 + D2 = 1 can be assumed to hold, following from the
coherence properties of laser light.
DR
Complementarity
The mathematical discussion presented above does not require quantum me-
chanics at its heart. In particular, the derivation is essentially valid for waves of
any sort. With slight modifications to account for the squaring of amplitudes,
the derivation could be applied to, for example, sound waves or water waves
in a ripple tank.
In order for the relation to be a precise formulation of Bohr complementarity
one must introduce wave-particle duality in the discussion. This means one
must consider both wave and particle behavior of light on an equal footing.
Wave-particle duality implies that one must A) use the unitary evolution of
the wave before the observation and B) consider the particle aspect after the
detection (this is called the Heisenberg-von Neumann collapse postulate). In-
deed since one could only observe the photon in one point of space (a photon
can not be absorbed twice) this implies that the meaning of the wave function

Complementarity (physics)
66

is essentially statistical and can not be confused with a classical wave (like it
exists in air or water).
In this context the direct observation of a photon in the aperture plane pre-
cludes the following recording of the same photon in (F). Reciprocally the ob-

FT
servation in (F) means that we did not absorb the photon before. If both holes
are open this implies that we don’t know where we would have detected the
photon in the aperture plane. D defines thus the distinguishability of the two
holes A and B.
A maximal value of distinguishability D = 1 means that only one hole (say A )
is open. If now we detect the photon at (F), we know that that photon would
have been detected in A necessarily. Conversely, D = 0 means that both holes
are open and play a symmetric role. If we detect the photon at (F), we don’t
know where the photon would have been detected in the aperture plane and
D = 0 characterizes our ignorance.
Similarly, if D = 0 then V = 1 and this means that a statistical accumulation
of photons at (F) will build up an interference pattern with maximal visibility.
Conversely, D = 1 implies V = 0 and thus, no fringes will appear after a
A
statistical recording of several photons.
The above treatment formalizes wave particle duality for the double-slit exper-
iment.

See also
• Afshar experiment
• →Wave-particle duality
• →Quantum entanglement
DR
• →Quantum indeterminacy

Further Reading
• Bethold-Georg Englert, Marlan O Scully & Herbert Walther, Quantum Op-
tical Tests of Complementarity , Nature, Vol 351, pp 111-116 (9 May
1991). Demonstrates that quantum interference effects are destroyed by
irreversible object-apparatus correlations ("measurement"), not by Heisen-
berg’s uncertainty principle itself. See also The Duality in Matter and Light
Scientific American, (December 1994)

Source: http://en.wikipedia.org/wiki/Complementarity_%28physics%29

Principal Authors: Danko Georgiev MD, Linas, Dewain Belgard, Enormousdude, Arcturus

Complementarity (physics)
67

Complete set of commuting observables

In quantum mechanics, a complete set of commuting observables (CSCO) is

FT
a set of commuting operators whose eigenvalues completely specify the state
of a system.
For example, in the case of the hydrogen atom, the Hamiltonian H, the angular
momentum L and its projection L z along any arbitrary z axis form a CSCO (if
one ignores the spin of the proton and of the electron as well as the movement
of the proton).

Source: http://en.wikipedia.org/wiki/Complete_set_of_commuting_observables

Principal Authors: TobinFricke, Conscious, Gazpacho

Compton scattering
A
In physics, Compton scattering or the Compton effect, is the decrease in
energy (increase in wavelength) of an X-ray or gamma ray photon, when it
interacts with matter. The amount the wavelength increases by is called the
Compton shift. Although nuclear compton scattering exists, what is meant by
Compton scattering usually is the interaction involving only the electrons of
an atom. Compton effect was observed by Arthur Holly Compton in 1923, for
which he earned the 1927 Nobel Prize in Physics.
DR
The effect is important because it demonstrates that light cannot be explained
purely as a wave phenomenon. Thomson scattering, the classical theory of
charged particles scattered by an electromagnetic wave, cannot explain any
shift in wavelength. Light must behave as if it consists of particles in order to
explain the Compton scattering. Compton’s experiment convinced physicists
that light can behave as a stream of particles whose energy is proportional to
the frequency.
The interaction between high energy photons and electrons results in the elec-
tron being given part of the energy (making it recoil), and a photon containing
the remaining energy being emitted in a different direction from the original,
so that the overall momentum of the system is conserved. If the photon still
has enough energy left, the process may be repeated.

Compton scattering
68

Compton scattering occurs in all materials and predominantly with photons


of medium energy, i.e. about 0.5 to 3.5 MeV. It is also observed that high-
energy photons; photons of visible light of higher frequency, for example, have
sufficient energy to even eject the bound electrons from the atom (photoelectric

FT
effect).

The Compton shift formula


For differential cross section of Compton scattering, see Klein-Nishina formula.

A
Figure 10 Compton Scattering (in the rest frame of the target)

Compton used a combination of three fundamental formulas representing the


various aspects of classical and modern physics, combining them to describe
DR
the quantum behavior of light.

• Light as a particle, as noted previously in the photoelectric effect.


• Relativistic dynamics Special Theory of Relativity
• Trigonometry - Law of cosines

The final result gives us the Compton scattering equation:


h
λ2 = me c (1 − cos θ) + λ1

where
λ1 is the wavelength of the photon before scattering,

λ2 is the wavelength of the photon after scattering,

Compton scattering
69

m e is the mass of the electron,

h/(m ec) is known as the →Compton wavelength,

FT
θ is the angle by which the photon’s heading changes,

h is Planck’s constant, and

c is the speed of light.

Collectively, the Compton wavelength is 2.43×10 -12 meter.

Derivation
We use that:
Eγ + Ee = Eγ 0 + Ee0
A
(Conservation of energy, where Eγ is the energy of a photon before the collision
and Ee is the energy of an electron before collision — its rest mass). The
variables with a prime are used for those after the collision.
And:
p~γ + p~e = p~γ 0 + p~e0

(Conservation of momentum, with the pe = 0 because we assume that the


electron is at rest.)
DR
We then use E = hf = pc:
p~e0 = p~γ − p~γ 0

p~e0 2 = (p~γ − p~γ 0 )2

p~e0 2 = p~γ 2 − 2 · p~γ · p~γ 0 + p~γ 0 2

p~e0 · p~e0 = p~γ · p~γ − 2 · p~γ · p~γ 0 + p~γ 0 · p~γ 0

pe0 2 · cos(0) = p2γ · cos(0) − 2 · pγ · pγ 0 · cos(θ) + p2γ 0 · cos(0)

The cos(θ) term appears because the momenta are spatial vectors, all of which
lie in a single 2D plane, thus their inner product is the product of their norms

Compton scattering
70

multiplied by the cosine of the angle between them.


hf hf 0
substituting pγ with c and pγ 0 with c , we derive
h2 f 2 h2 f 02 2h2 f f 0 cos θ
p2e0 = c2
+ c2
− c2

FT
Now we fill in for the energy part:

Eγ + Ee = Eγ 0 + Ee0

hf + mc2 = hf 0 +
p
(pe0 c)2 + (mc2 )2

We solve this for p e’:


(hf + mc2 − hf 0 )2 = (pe0 c)2 + (mc2 )2

(hf +mc2 −hf 0 )2 −m2 c4


c2
= p2e0

Then we have two equations for p2e0 , which we equate:


A(hf +mc2 −hf 0 )2 −m2 c4
c2
= h2 f 2
c2
+ h2 f 02
c2
− 2h2 f f 0 cos θ
c2

Now it’s just a question of rewriting:


h2 f 2 + h2 f 02 − 2h2 f f 0 + 2h(f − f 0 )mc2 = h2 f 2 + h2 f 02 − 2h2 f f 0 cos θ

−2h2 f f 0 + 2h(f − f 0 )mc2 = −2h2 f f 0 cos θ


DR
hf f 0 − (f − f 0 )mc2 = hf f 0 cos θ

hf f 0 (1 − cos θ) = (f − f 0 )mc2

h λc0 λc (1 − cos θ) = c c
mc2

λ − λ0
 
cλ0
h λc0 λc (1 − cos θ) = λλ0
− cλ
λ0 λ
mc2
 
λ0 λ cλ0 cλ
h(1 − cos θ) = c c λ0 λ
− λλ0
mc2
 
λ0 λ
h(1 − cos θ) = c − c mc2

Compton scattering
71

h
mc (1 − cos θ) = λ0 − λ

Applications

FT
Compton scattering is of prime importance to radiobiology, as it happens to be
the most probable interaction of high energy X rays with atomic nuclei in living
beings and is applied in radiation therapy.
Compton scattering has on occasion been proposed as an alternative explana-
tion for the phenomenon of the redshift by opponents of the Big Bang theory,
although this is not generally accepted because the influence of the Compton
scattering would be noticeable in the spectral lines of distant objects and this
is not observed.
In material physics, Compton scattering can be used to probe the wave function
of the electrons in matter in the momentum representation.
Compton Scatter is an important effect in Gamma spectroscopy, as it is possible
for the gamma rays to scatter out of the detectors used. Compton suppression
is used to detect stray scatter gamma rays to counteract this effect.
A
See also
• Thomson scattering
• →Photoelectric effect
• Timeline of cosmic microwave background astronomy
• Peter Debye
• Sunyaev Zel’dovich effect
DR
• Walther Bothe
• List of astronomical topics
• List of physics topics

External links
• Compton Effect 36 (PDF file) by Michael Brandl for Project PHYSNET 37.

Source: http://en.wikipedia.org/wiki/Compton_scattering

Principal Authors: Pt, Eteq, Fresheneesz, Eleassar777, Omegatron, Silenced, Reddi, Pfalstad, Niven,
AugPi

36 http://35.9.69.219/home/modules/pdf_modules/m219.pdf
37 http://physnet2.pa.msu.edu/

Compton scattering
72

Compton wavelength

The Compton wavelength λ of a particle is given by

FT
h ~
λ= mc = 2π mc ,

where
h is the Planck constant,

m is the particle’s mass,

c is the speed of light.

The Compton wavelength of the electron is approximately 2.4 × 10 -12 meters.


The Compton wavelength can be thought of as a fundamental limitation on
measuring the position of a particle, taking quantum mechanics and special
A
relativity into account. This depends on the mass m of the particle. To see
this, note that we can measure the position of a particle by bouncing light off
it - but measuring the position accurately requires light of short wavelength.
Light with a short wavelength consists of photons of high energy. If the energy
of these photons exceeds mc2 , when one hits the particle whose position is
being measured the collision may have enough energy to create a new particle
of the same type. This renders moot the question of the original particle’s
location.
DR
This argument also shows that the Compton wavelength is the cutoff below
which quantum field theory– which can describe particle creation and annihi-
lation – becomes important.
We can make the above argument a bit more precise as follows. Suppose we
wish to measure the position of a particle to within an accuracy ∆x . Then
the uncertainty relation for position and momentum says that ∆x ∆p ≥ ~/2
~
so the uncertainty in the particle’s momentum satisfies ∆p ≥ 2∆x Using the
relativistic relation between momentum and energy, when ∆p exceeds mc then
the uncertainty in energy is greater than mc2 , which is enough energy to
create another particle of the same type. So, with a little algebra, we see
~
there is a fundamental limitation ∆x ≥ 2mc So, at least to within an order
of magnitude, the uncertainty in position must be greater than the Compton
wavelength h/mc .

Compton wavelength
73

The Compton wavelength can be contrasted with the de Broglie wavelength,


which depends on the momentum of a particle and determines the cutoff be-
tween particle and wave behavior in quantum mechanics.
For fermions, the Compton wavelength sets the cross-section of interactions.

FT
For example, the cross-section for Thomson scattering of a photon from an
electron is equal to (8π/3)α2 λ2e , where α is the fine-structure constant and
λe is the Compton wavelength of the electron. For gauge bosons, the Compton
wavelength sets the effective range of the Yukawa interaction: since the photon
is massless, electromagnetism has infinite range.
The Compton wavelength of the electron is one of a trio of related units of
length, the other two being the Bohr radius a0 and the classical electron radius
re . The Compton wavelength is built from the electron mass me , Planck’s con-
stant h and the speed of light c. The Bohr radius is built from me , h and the
electron charge e. The classical electron radius is built from me , c and e. Any
one of these three lengths can be written in terms of any other using the fine
structure constant α:
αλe
re = 2π = α2 a0
A
The Planck mass is special because ignoring factors of 2π and the like, the
Compton wavelength for this mass is equal to its Schwarzschild radius. This
special distance is called the Planck length. This is a simple case of dimensional
analysis: the Schwarzschild radius is proportional to the mass, whereas the
Compton wavelength is proportional to the inverse of the mass.

External links
DR
• Length Scales in Physics: the Compton Wavelength 38

Source: http://en.wikipedia.org/wiki/Compton_wavelength

Principal Authors: Joke137, John Baez, MagnaMopus, Rmrfstar, Wigie

38 http://math.ucr.edu/home/baez/lengths.html#compton_wavelength

Compton wavelength
74

Conjugate variables

In physics, especially in quantum mechanics, conjugate variables are pairs

FT
of variables that share an uncertainty relation. The terminology comes from
classical Hamiltonian mechanics, but also appears in quantum mechanics and
engineering.
Examples of canonically conjugate variables include the following:

• Time and frequency: the longer a musical note is sustained, the more pre-
cise we know its frequency (but it spans more time). Conversely, a very
short musical note becomes just a click, and so one can’t know its frequen-
cy very accurately.
• Position and momentum: precise measurements of position lead to ambi-
guity of momentum, and v.v.
• Doppler and range: the more we know about how far away a radar target
is, the less we can know about the exact velocity of approach or retreat, and
vice versa. In this case, the two dimensional function of doppler and range
A
is known as a radar ambiguity function or radar ambiguity diagram.

A pair of conjugate variables are often Fourier transform duals of one-another,


or more generally are related through Pontryagin duality. The duality relations
lead naturally to an uncertainty relation between them.
A more precise mathematical definition, in the context of Hamiltonian mechan-
ics, is given in the article canonical coordinates.
DR
Source: http://en.wikipedia.org/wiki/Conjugate_variables

Principal Authors: Linas, Charles Matthews, Glogger, Demoscn, Arthur Rubin

Constraint algebra

In theoretical physics, a constraint algebra is a linear space of all constraints


and all of their polynomial functions or functionals whose action on the physi-
cal vectors of the →Hilbert space should be equal to zero.
For example, in electromagnetism, the equation for the Gauss’ law
∇·E ~ =ρ

Constraint algebra
75

is an equation of motion that does not include any time derivatives. This is why
it is counted as a constraint, not a dynamical equation of motion. In quantum
electrodynamics, one first constructs a Hilbert space in which Gauss’ law does
not hold automatically. The true Hilbert space of physical states is constructed

FT
as a subspace of the original Hilbert space of vectors that satisfy
~
(∇ · E(x) − ρ(x))|ψi = 0.

In more general theories, the constraint algebra may be a noncommutative


algebra.

Source: http://en.wikipedia.org/wiki/Constraint_algebra

Coupling constant

In physics, a coupling constant, usually denoted g, is a number that deter-


A
mines the strength of an interaction. Usually the Lagrangian or the Hamiltoni-
an of a system can be separated into a kinetic part and an interaction part. The
coupling constant determines the strength of the interaction part with respect
to the kinetic part, or between two sectors of the interaction part. For example,
the electric charge of a particle is a coupling constant.
A coupling constant plays an important role in dynamics. For example, one
often sets up hierarchies of approximation based on the importance of various
coupling constants. In the motion of a large lump of magnetized iron, the
DR
gravitational forces are more important than the magnetic forces because of
the relative coupling constants. However, in classical mechanics one usually
makes these decisions directly by comparing forces.

Fine structure constant


The coupling constant comes into its own in a quantum field theory. A special
role is played in relativistic quantum theories by coupling constants which are
dimensionless, ie, are pure numbers. For example, the fine-structure constant,
e2
α= 4π0 ~c

(where e is the charge of an electron,  0 is the permittivity of free space,


is Dirac’s constant and c is the speed of light) is such a dimensionless cou-
pling constant that determines the strength of the electromagnetic force on an
electron.

Coupling constant
76

Gauge coupling
In a non-Abelian gauge theory, the gauge coupling parameter, g, appears in
the Lagrangian as
1
Tr Gµν Gµν

FT
4g 2

(where G is the gauge field tensor) in some conventions. In another widely


used convention, G is rescaled so that the coefficient of the kinetic term is 1/4
and g appears in the covariant derivative. This should be understood to be
similar to a dimensionless version of the electric charge defined as

4π0 α.

Weak and strong coupling


In a quantum field theory with a dimensionless coupling constant, g, if it is
(much) smaller than one, then one says that the theory is weakly coupled. In
this case it is well described by an expansion in powers of g, called perturbation
theory. If the coupling constant is of order one or larger, the theory is said to
A
be strongly coupled. An example of the latter is the hadronic theory of strong
interactions (which is why it is called strong in the first place). In such a case
non-perturbative methods have to be used to investigate the theory.

Running coupling
DR

Figure 11 Virtual particles renormalize the coupling

Coupling constant
77

One can probe a quantum field theory at short times or distances by changing
the wavelength or momentum, k of the probe one uses. With a high frequency,
ie, short time probe, one sees virtual particles taking part in every process. The
reason this can happen, seemingly violating the conservation of energy is the

FT
uncertainty relation
∆E∆t ≥ ~

which allows such violations at short times. The previous remark only applies
to some formulations of QFT, in particular, canonical quantization in the inter-
action picture. In other formulations, the same event is described by "virtual"
particles going off the mass shell. Such processes renormalize the coupling
and make it dependent on the scale, k at which one observes the coupling.
The phenomenon of scale dependence of the coupling, g(k) is called running
coupling in a quantum field theory.

Beta-function
The beta function β(g) of a quantum field theory measures the running of a
A
coupling parameter. It is defined by the relation:
β(g) = k ∂g
∂k = ∂g
∂ ln k .

For most theories the beta function is positive, so that the coupling is increasing
in k (equivalently, the coupling rises as the scale at which the theory is observed
becomes shorter). This is also the case in quantum electrodynamics (QED). At
low energy, i.e. long distances, α ≈ 1/137. At the scale of the Z boson, about
90 GeV, α ≈ 1/127.
DR
In a classical field theory in which a scale change is an invariance (symmetry)
of the theory, the beta function breaks this scale invariance. Since this is a
quantum effect arising directly from the uncertainty principle, a non-zero beta
function implies a scale anomaly in such a quantum field theory.

Landau pole and asymptotic freedom


We noted that QED is weakly coupled at long distances, but the coupling in-
creases at short distances. This increase was first noticed by Lev Landau who
showed that QED becomes strongly coupled at high energy, and in fact the
coupling becomes infinite at asympototically high energy. This phenomenon is
called the Landau pole.
In non-Abelian gauge theories, the beta function is negative, as first found
by Frank Wilczek, David Politzer and David Gross. As a result the coupling

Coupling constant
78

decreases at short distances. Furthermore, the coupling decreases logarithmi-


cally, a phenomenon known as asymptotic freedom. The coupling decreases
approximately as
gs2 (k 2 ) 1
αs (k 2 ) ≡ 4π ≈ β0 ln(k 2 /Λ2 )

FT
where β 0 is a constant computed by Wilczek, Gross and Politzer.

QCD scale
In quantum chromodynamics (QCD), the quantity is called the QCD scale.
The value is
ΛM S = 217+25
−23 MeV

This value is to be used at a scale above the bottom quark mass of about 5 GeV.
The meaning of MS is given in the article on dimensional regularization.

Charge, colour charge, etc


A
In quantum field theory, since the size of the interaction term is absorbed into
the notion of the coupling constant (more correctly coupling parameter, since
it runs), the word charge is freed up for another use. One says, for example,
that the electrical charge of an electron is -1 and that of any observable particle
is an integer multiple of this. The notion of charge is now exactly the same as
the representation of the gauge group to which the particle belongs. Thus the
colour charge of a quark is fixed at 4/3 since it belongs to the fundamental
representation of SU(3), and the colour charge of a gluon is 8 since it belongs
DR
to the adjoint representation.
This difference in the notion of charge in classical and quantum field theory is
alluded to in a shorthand phrase that is sometimes used: "charge in units of
the positron charge".

String theory
A remarkably different situation exists in string theory. Each perturbative de-
scription of string theory depends on a string coupling constant. However, in
the case of string theory, these coupling constants are not pre-determined, ad-
justable, or universal parameters; rather they are dynamical scalar fields that
can depend on the position in space and time and whose values are determined
dynamically.

Coupling constant
79

See also
• →Quantum field theory, especially quantum electrodynamics and quantum
chromodynamics

FT
• Canonical quantization, renormalization and dimensional regularization
• fine structure constant
• gravitational coupling constant

References and external links


• An introduction to quantum field theory, by M.E.Peskin and H.D.Schroeder,
ISBN 0201503972
• The Nobel Prize in Physics 2004 – Information for the Public 39

Source: http://en.wikipedia.org/wiki/Coupling_constant

Principal Authors: Bambaiah, Phys, Xerxes314, Ricky81682, Cmdrjameson


A
Creation and annihilation operators

In physics, an annihilation operator is an operator that lowers the number


of particles in a given state by one. A creation operator is an operator that
increases the number of particles in a given state by one, and it is the adjoint
of the annihilation operator. Depending on the context, the identity of the
DR
particles in question varies; for example, in quantum chemistry and many-
body theory the creation and annihilation operators often act on electrons.
Annihilation and creation operators can also refer specifically to the ladder
operators for the quantum harmonic oscillator. In the latter case, the raising
operator is interpreted as a creation operator, adding a quantum of energy to
the oscillator system. Vice versa for the lowering operator. In many subfields of
physics and chemistry, using these operators instead of a wavefunction picture,
is known as second quantization.
The mathematics behind the creation and the annihilation operators is identical
as the formulae for ladder operators that appear in the quantum harmonic
oscillator. For example, the commutator of the annihilation and the creation
operator associated with the same state equals one; all other commutators
vanish.

39 http://nobelprize.org/physics/laureates/2004/public.html

Creation and annihilation operators


80

While the concept of creation and annihilation operators is well defined for free
field theories, in interacting QFTs, they can only be defined in the interaction
picture, which does not exist according to Haag’s theorem.

FT
Derivation of bosonic creation and annihilation op-
erators
In the context of the quantum harmonic oscillator, we reinterpret the ladder
operators as creation and annihilation operators, adding or subtracting fixed
quanta of energy to the oscillator system. Creation/annihilation operators are
different for bosons (integer spin) and fermions (half-integer spin). This is
because their wavefunctions have different symmetry properties.
Suppose the wavefunctions are dependent on N properties. Then
For bosons: ψ(1,2,3,4,...N ) = ψ(2,1,3,4,...N )

For fermions: ψ(1,2,3,4,...N ) = -ψ(2,1,3,4,...N )


A
For now let’s just consider the case of bosons because fermions are more com-
plicated.
Start with the →Schrödinger equation for the one dimensional time indepen-
dent quantum harmonic oscillator
 2 2 
~ d 1 2 2 ψ(x) = Eψ(x)
− 2m dx 2 + 2 mω x

Make a coordinate substitution to nondimensionalize the differential equation


q
DR
~
x ≡ mω q.

and the Schrödinger equation for the oscillator becomes


 
~ω d2 2 ψ(q) = Eψ(q).
2 − dq 2 +q

Notice that the quantity hω = h ν is the same energy as that found for light
quanta and that the parenthesis in the Hamiltonian can be written as
h ih i
d2 2 d d d d
− dq 2 + q = − dq + q dq + q + dq q − q dq

The last two terms in that equation form the commutator of q with its deriva-
tive. So let’s calculate that commutator [ q, ∂/∂q ]
 
d d df d
q dq − dq q f (q) = q dq − dq (qf (q)) = −f (q)

Creation and annihilation operators


81

In other words [ q, d /dq ] = - 1 or [ d /dq, q ] = 1.


Therefore
  h ih i
1 d2 2 = ~ω −d/dq+q d/dq+q
+ 21 ~ω
2 ~ω − dq 2 + q
√ √
2 2

FT
If we define
 
a† ≡ √1 d
− dq + q as the "creation operator" or the "raising operator" and
2
 
a≡ √1 d
+ dq + q as the "annihilation operator" or the "lowering operator"
2

the Hamiltonian becomes


H = ~ω a† a + 12 .


This Hamiltonian is significantly simpler than the original form. Further sim-
plifications of this equation enables one to derive all the properties listed above
thus far.
Letting p = - i d /dq, where p is the nondimensionalized momentum operator
Aa† = √1 (q
2
− ip)

a= √1 (q + ip).
2

Substituting backwards, the laddering operators are recovered.

Mathematical details
DR
The operators derived above are actually a specific instance of a more gener-
alized class of creation and annihilation operators. The more abstract (and
hence more applicable) form of the operators satisfy the properties below.
Let H be the one-particle →Hilbert space. To get the bosonic CCR algebra, look
at the algebra generated by a(f ) for any f in H. The operator a(f ) is called an
annihilation operator and the map a(.) is antilinear. Its adjoint is a †(f ) which
is linear in H.
For a boson,
[a(f ), a(g)] = [a† (f ), a† (g)] = 0

[a(f ), a† (g)] = hf |gi,

where we are using bra-ket notation.

Creation and annihilation operators


82

For a fermion, the anticommutators are


{a(f ), a(g)} = {a ∗ (f ), a ∗ (g)} = 0

{a(f ), a ∗ (g)} = hf |gi.

FT
A CAR algebra.
Physically speaking, a(f ) removes (i.e. annihilates) a particle in the state |f >
wheareas a †(f ) creates a particle in the state |f >.
The free field vacuum state is the state with no particles. In other words,
a(f )|0i = 0

where |0> is the vacuum state.


If |f > is normalized so that <f |f >=1, then a †(f ) a(f ) gives the number of
particles in the state |f >.
Note that the creation and annihilation operators are "generalized complex
conjugates" of each other. Usually, the notation is chosen in such a way that
A
the a †(f ) is the creation operator, and a(f ) is the annihilation operator. The †
reminds us that something "extra" is being added to the system. The topic can
be misleadingly confusing if this is not done.

Notational caveats and considerations


In quantum mechanics, Dirac bra-ket notation is often used. However, there is
some ambiguity in this notation, particularly when there is the need to differ-
entiate between these things:
DR
• The lowest energy state
• The zero state
• The vacuum state
• The zero ket

Often, these are all interchangeably notated as |0>, or even | >. As a result, it
is necessary to read carefully, and consider the context in which the notation is
used.
For example, in the quantum harmonic oscillator, the ground state has the
property that when the annihilation operator b is applied to it, it satisfies b|0>
= 0| > = 0
The intermediate step is rarely indicated as it is considered necessary only
when more conceptual/mathematical rigour is needed.

Creation and annihilation operators


83

In this example, the lowest energy state is denoted as |0>. It is labeled as the
"zero state", but it is important to emphasize that any state can be labeled as the
"zero" state. The zero state is often used as a reference state to other quantum
states. Therefore, the |0> state need not be the state with the absolutely lowest

FT
energy. In the case of the harmonic oscillator, it is due to the particulars of the
mathematics that the ground state is chosen to be |0>. The vacuum state is
the state where no quanta is available to be extracted. This special null state
is denoted by | >. This vacuum state is also known as the "zero ket" because
there are zero particles in the state. Unfortunately, the lowest energy state |0>
is also known as the "zero ket" for the different reason that the state is labeled
as "zero". Care must be taken that the four concepts listed above are not mixed
together.
Sometimes, the terms "null state" and "empty state" are used interchangeably
for |> and |0>. The meaning for this usage is again dependent on the context.

The vacuum state


The vacuum state is a conceptual state which has no particles. The state is
usually denoted as |0>, not the "empty ket" | >. Interestingly enough, no
A
actual function actually represents the |0> state, but for notational purposes,
we define the vacuum state as being normalized such that <0|0> = 1 and that
|0> is orthogonal to all other states of the form |N >, where N is any indexing
of quantum states for a particular system.

Energy spectra
In a quantum mechanical system, the range of discrete energies allowed in a
system may be either finite, or infinite, or "semi-infinite".
DR
In a system where the energies are confined to be semi-infinite on the interval
[constant, ∞) such as the quantum harmonic oscillator, the vacuum state | >,
(different from |0>) needs to be introduced in order to make the theory of
creation and annihilation operators consistent. The lowest energy allowed in a
semi-infinite energy system is known as the ground state. Since it is often used
as the reference state, it is denoted by |0>. However, this state is not empty -
the vacuum state | > is introduced to disambiguate these two states.
In a system where the range of energies is (-∞, ∞), the vacuum state is almost
always denoted by |0>. There is no need for the "null" state | > as |0> already
is sufficient to denote "emptiness". There is also no "ground state" present,
which is why the notational ambiguity arises. This interpretation arises directly
from the relativistic formalism of quantum mechanics by Paul Dirac, which
later became one of the foundations for quantum field theory. One of the
shortcomings of quantum field theory however, is its allowance of energy states

Creation and annihilation operators


84

tending to infinity. The attempt to resolve this problem is very much an active
part of quantum mechanical research today.
In summary for infinite and semi infinite systems

FT
COMMON NOTATION FOR STATES
infinite semi-infinite

ground state none |0>

vacuum state |0> | >

There is no | > state needed for infinite-ranged-energy-systems in quantum


mechanics.

See also
• →Ladder operators
• Bogolibov transformations - arises in the theory of quantum optics. Also
transliterated as Bogolubov transformations’
A
Source: http://en.wikipedia.org/wiki/Creation_and_annihilation_operators

Principal Authors: HappyCamper, Laurascudder, Phys, Salsb, Sunev

Dark energy star


DR
A dark-energy star is a hypothetical compact astrophysical object which a
minority of physicists feel might constitute an alternative explanation for ob-
servations of astronomical black hole candidates.
The concept was proposed by physicist George Chapline. The theory states that
infalling matter is converted into vacuum energy, or dark energy as the matter
falls through the event horizon. The space within the event horizon would end
up with a large value for the cosmological constant, and have negative pressure
to exert against gravity. There would be no information destroying singularity.

Theory
In March 2005, physicist George Chapline claimed that quantum mechanics
makes it a "near certainty" black holes do not exist and are instead dark energy
stars. The dark energy star is a different concept than that of a gravastar.

Dark energy star


85

Dark-energy stars were first proposed due to the fact that in quantum physics,
absolute time is required, however in general relativity, an object falling to-
wards a black hole would to an outside observer seem to have time pass in-
finitely slowly at the event horizon. The object itself would feel as if time

FT
flowed normally.
In order to reconcile quantum mechanics with black holes, Chapline theorized
that a phase transition in the phase of space occurs at the event horizon. He
based his ideas on the physics of superfluids. As a column of superfluid grows
taller, at some point, density increases slowing down the speed of sound so
that it approaches zero. However, at that point, quantum physics makes sound
waves dissipate their energy into the superfluid, so that the zero sound speed
condition is never encountered.
In the dark-energy star hypothesis, infalling matter approaching the event hori-
zon decay into successively lighter particles. Nearing the event horizon, envi-
ronmental effects accelerate proton decay. This may account for high energy
cosmic ray sources and positron sources in the sky. When the matter falls
through the event horizon, the energy equivalent of some or all of that matter
is converted into dark energy. This negative pressure counteracts the mass the
A
star gains, avoiding a singularity.
The negative pressure also gives a very high number for the cosmological con-
stant. As there is no singularity to evaporate, Hawking radiation may not exist
in this model of black holes.
Furthermore ’primordial’ dark-energy stars could form by fluctuations of space-
time itself which is analogous to "blobs of liquid condensing spontaneously out
of a cooling gas." This not only alters the understanding of black holes but has
DR
the potential to explain the dark energy and dark matter that are indirectly
observed.

See also
• Gravastar
• acoustic metric
• analog model of gravity
• Stellar mass black hole

External links
• Dark Energy Stars 40
• MPIE Galactic Center Research 41

40 http://www.llnl.gov/tid/lof/documents/pdf/317506.pdf

Dark energy star


86

Source: http://en.wikipedia.org/wiki/Dark_energy_star

Principal Authors: RoyBoy, Hillman, Phys, Algri, Christopher Thomas

FT
Davisson-Germer experiment

In 1927 at Bell Labs, Clinton Davisson and Lester Germer fired slow moving
electrons at a crystalline nickel target. The angular dependence of the reflect-
ed electron intensity was measured, and was determined to have the same
diffraction pattern as those predicted by Bragg for X-Rays.
This experiment, like Arthur Compton’s proof of the particle-like nature of
light, proved the wave-like nature of matter and completed the wave-particle
duality hypothesis, which was a fundamental step in quantum theory.

Source: http://en.wikipedia.org/wiki/Davisson-Germer_experiment
A
Principal Authors: Linas, Gershwinrb, Tone, Charles Matthews, Wiccan Quagga

De Broglie hypothesis

In physics, the de Broglie hypothesis is the statement that all matter has a
wave-like nature (wave-particle duality). The de Broglie relations show that
DR
the wavelength is inversely proportional to the momentum of a particle and
that the frequency is directly proportional to the particle’s kinetic energy. The
hypothesis was advanced by Louis de Broglie in 1923 in his PhD thesis 42; he
was awarded the Nobel Prize for Physics in 1929 for this work, which made
him the first person to receive a Nobel Prize on a PhD thesis.

The de Broglie relations


The first de Broglie equation relates the wavelength to the particle momentum
as q
h h v2
λ= p = mv 1− c2

41 http://www.mpe.mpg.de/ir/GC/index.php
42 L. de Broglie, PhD thesis, reprinted in Ann. Found. Louis de Broglie 17 (1992) p. 22.

De Broglie hypothesis
87

where λ is the particle’s wavelength, h is Planck’s constant, p is the particle’s


momentum, m is the particle’s rest mass, v is the particle’s velocity, and c is
the speed of light in a vacuum.
The greater the energy, the larger the frequency and the shorter (smaller) the

FT
wavelength. Given the relationship between wavelength and frequency, it fol-
lows that short wavelengths are more energetic than long wavelengths. The
second de Broglie equation relates the frequency of a particle to the kinetic
energy such that
 
Ek mc2 q 1
f= h = h − 1
2
1− v2
c

where f is the frequency and Ek is the kinetic energy. The two equations are
often written as
p = ~k

Ek = ~ω
A
where ~ is Dirac’s constant, k is the wavenumber, and ω is the angular frequen-
cy.
See the article on group velocity for detail on the argument and derivation of
the de Broglie relations.

Experimental confirmation
In 1927 at Bell Labs, Clinton Davisson and Lester Germer fired slow-moving
DR
electrons at a crystalline nickel target. The angular dependence of the reflected
electron intensity was measured, and was determined to have the same diffrac-
tion pattern as those predicted by Bragg for X-Rays. Before the acceptance of
the De Broglie hypothesis, diffraction was a property that was only exhibited
by waves. Therefore, the presence of any diffraction effects by matter, demon-
strated the wave-like nature of matter. When the De Broglie wavelength was
inserted into the Bragg condition, the observed diffraction pattern was predict-
ed, thereby experimentally confirming the De Broglie hypothesis for electrons.
This was a pivotal result in the development of quantum mechanics. Just as
Arthur Compton demonstrated the particle nature of light, the →Davisson-
Germer experiment showed the wave-nature of matter, and completed the the-
ory of wave-particle duality. For physicists this idea was important because it
means that not only can any particle exhibit wave characteristics, but that one
can use wave equations to describe phenomena in matter if one uses the de
Broglie wavelength.

De Broglie hypothesis
88

Since the original Davisson-Germer experiment for electrons, the De Broglie


hypothesis has been confirmed for other elementary particles. Recent experi-
ments even confirm the relations for macromolecules, which are normally con-
sidered too large to undergo quantum mechanical effects. In 1999, a research

FT
team in Vienna demonstrated diffraction for molecules as large as fullerenes. 43

Wavelength of large objects


Theoretically, all objects, not just sub-atomic particles, exhibit wave properties
according to the De Broglie Hypothesis.
Consider the following example:
A baseball has a mass of 0.15 kg and is thrown by a professional baseball player
at 40 m/s. The de Broglie wavelength of the baseball is given by:
m = 0.15kg
v = 40m/s (about 90 mph)
h
λ= p where p = mv
6.626×10−34 kgm2 /s
λ=
A
λ=
0.15kg×40m/s
1.10 × 10−34 m
This wavelength is considerably smaller than the diameter of a proton (about
10−15 m) and is approaching the Planck Length. As such, the wave-like proper-
ties of this baseball are so small as to be unobservable.

See also
• →Bohr model
DR
References
• Steven S. Zumdahl, Chemical Principles 5th Edition, (2005) Houghton Mif-
flin Company.
• Tipler, Paul A. and Ralph A. Llewellyn (2003). Modern Physics. 4th ed.
New York; W. H. Freeman and Company. ISBN 0-7167-4345-0. pp. 203-4,
222-3, 236.

Source: http://en.wikipedia.org/wiki/De_Broglie_hypothesis

43 Arndt, M., O. Nairz, J. Voss-Andreae, C. Keller, G. van der Zouw, A. Zeilinger (14 October 1999).

"Wave-particle duality of C60". Nature 401: 680-682.

De Broglie hypothesis
89

Principal Authors: Linas, Teply, Cdang, Blinken, Dataphile

Degenerate energy level

FT
This article refers to physical states having the same energy. For other uses
of the word degeneracy, see degeneracy (disambiguation).

In physics two or more physical states are said to be degenerate if they are
both at the same energy level; the energy level is said to be degenerate if it
contains two or more such states. The number of occupation states available
at a particular energy level is called the level’s degeneracy.
In quantum theory, this usually pertains to electronic configurations and the
electron energy levels, where different possible occupation states for particles
may be related by symmetry. The usage comes from the fact that degener-
ate eigenstates correspond to identical eigenvalues of the Hamiltonian. Since
eigenvalues correspond to roots of the characteristic equation, degeneracy here
A
has the same meaning as the common mathematical usage of the word.
If the symmetry is broken by a perturbation, such as applying an external elec-
tric field, this can change the energies of the states, causing energy level split-
ting.

See also
• Density of states
DR
Source: http://en.wikipedia.org/wiki/Degenerate_energy_level

Principal Authors: Jheald, Tim Starling, Linas, Karol Langner

Degenerate energy level


90

Degenerate matter

Degenerate matter is matter which has sufficiently high density that the dom-

FT
inant contribution to its pressure arises from the Pauli exclusion principle. The
pressure maintained by a body of degenerate matter is called the degeneracy
pressure, and arises because the Pauli principle forbids the constituent parti-
cles from occupying identical quantum states. Therefore, reducing the volume
requires forcing the particles into higher-energy quantum states. The species of
fermion are sometimes identified, so that we may speak of electron degeneracy
pressure, neutron degeneracy pressure, and so forth.
Unlike a classical ideal gas, whose pressure is proportional to its temperature
(P = nkT , where P is pressure, n is particles per unit volume, k is Boltzmann’s
constant, and T is temperature), the pressure exerted by degenerate matter
depends only weakly on its temperature. In particular, the pressure remains
nonzero even at absolute zero temperature. At relatively low densities, the
pressure of a fully degenerate gas is given by P = Kn5/3 , where K depends on
the properties of the particles making up the gas. At very high densities, where
A
most of the particles are forced into quantum states with relativistic energies,
the pressure is given by P = K 0 n4/3 , where K 0 again depends on the properties
of the particles making up the gas.
Exotic examples of degenerate matter include neutronium, strange matter,
metallic hydrogen and white dwarf matter. Degeneracy pressure contributes
to the pressure of conventional solids, but these are not usually considered to
be degenerate matter as a significant contribution to their pressure is provid-
ed by the interplay between the electrical repulsion of atomic nuclei and the
DR
screening of nuclei from each other by electrons allocated among the quantum
states determined by the nuclear electrical potentials. In metals it is useful to
treat the conduction electrons alone as a degenerate, free electron gas while
the majority of the electrons are regarded as occupying bound quantum states.
This contrasts with the case of the degenerate matter that forms the body of a
white dwarf where all the electrons would be treated as occupying free particle
momentum states.

Degenerate gases
Degenerate gases are gases composed of fermions that have a particular config-
uration which usually forms at high densities. Fermions are subatomic particles
with half-integer spin. Their behaviour is regulated by a set of quantum me-
chanical rules called the →Fermi-Dirac statistics. One particular rule is the

Degenerate matter
91

Pauli exclusion principle that states that there can be only one fermion occu-
pying each quantum state which also applies to electrons that are not bound
to a nucleus but merely confined to a fixed volume, such as the deep interior
of a star. Such particles as electrons, protons, neutrons, and neutrinos are all

FT
fermions and obey Fermi-Dirac statistics.
A fermion gas in which all the energy states below a critical value, designated
Fermi energy, are filled is called a fully degenerate fermion gas. The electron
gas in ordinary metals and in the interior of white dwarf stars constitute two
examples of a degenerate electron gas. Most stars are supported against their
own gravitation by normal gas pressure. White dwarf stars are supported by
the degeneracy pressure of the electron gas in their interior. For white dwarfs
the degenerate particles are the electrons while for neutron stars the degener-
ate particles are neutrons.

Electron degeneracy
In ordinary gas, most of the energy levels called n-spheres of only certain dis-
crete energy states available to electrons are unfilled and the electrons are free
to move about. As particle density is increased in a fixed volume, electrons
A
progressively fill the lower energy states and additional electrons are forced
to occupy states of higher energy. Therefore, degenerate gases strongly resist
further compression because the electrons cannot move to lower energy levels
which are already filled due to the Pauli Exclusion Principle. The degenerate
electrons are locked into place because all of the lower energy shells are filled
up so they no longer move freely as in a normal gas. Even though thermal en-
ergy may be extracted from the gas, it still may not cool down, since electrons
cannot give up energy by moving to a lower energy state. This increases the
DR
pressure of the fermion gas termed degeneracy pressure. In a degenerate gas,
the average pressure is high enough to keep material from being compressed
by gravity.
Under high densities the matter becomes a degenerate gas when the electrons
are all stripped from their parent atoms. In the core of a star once hydrogen
burning in nuclear fusion reactions stops, it becomes a collection of positively
charged ions, largely helium and carbon nuclei, floating in a sea of electrons
which have been stripped from the nuclei. Degenerate gas is an almost perfect
conductor of heat and does not obey the ordinary gas laws. White dwarfs are
luminous not because they are generating any energy but rather because they
have trapped a large amount of heat. Normal gas exerts higher pressure when
it is heated and expands, but the pressure in a degenerate gas does not depend
on the temperature. When gas become super-compressed, particles position
right up against each other to produce degenerate gas that behaves more like

Degenerate matter
92

a solid. In degenerate gases the kinetic energies of electrons are quite high and
the rate of collision between electrons and other particles is quite low, there-
fore degenerate electrons can travel great distances at velocities that approach
the speed of light. Instead of temperature, the pressure in a degenerate gas

FT
depends only on the speed of the degenerate particles; however, adding heat
does not increase the speed. Pressure is only increased by the mass of the parti-
cles which increases the gravitational force pulling the particles closer together.
Therefore, the phenomenon is opposite of that normally found in matter where
if the mass of the matter is increased, the object becomes bigger. In degenerate
gas, when the mass is increased, the pressure is increased, and the particles
become spaced closer together, so the object becomes smaller. Degenerate gas
can be compressed to very high densities, typical values being in the range of
10 7 grams per cubic centimetre.
There is an upper limit to the mass of an electron-degenerate object, the Chan-
drasekhar limit, beyond which electron degeneracy pressure cannot support
the object against collapse. The limit is approximately 1.44 solar masses for
objects with compositions similar to the sun. The mass cutoff changes with the
chemical composition of the object, as this affects the ratio of mass to num-
A
ber of electrons present. Celestial objects below this limit are white dwarf
stars, formed by the collapse of the cores of stars which run out of fuel. Dur-
ing collapse, an electron-degenerate gas forms in the core, providing sufficient
degeneracy pressure as it is compressed to resist further collapse. Above this
mass limit, a neutron star (supported by neutron degeneracy pressure) or a
black hole may be formed instead.

Proton degeneracy
DR
Sufficiently dense matter containing protons experiences proton degeneracy
pressure, in a manner similar to the electron degeneracy pressure in electron-
degenerate matter. As protons and electrons occur in roughly equal numbers
in most forms of matter, proton degeneracy is usually modelled as a correction
to the equations of state of electron-degenerate matter, as opposed to the dom-
inant source of degeneracy pressure (which would require proton-degenerate
matter that was free of electrons).

Neutron degeneracy
Neutron degeneracy is analogous to electron degeneracy and is demonstrat-
ed in neutron stars, which are supported by the pressure from a degenerate
neutron gas. This happens when a stellar core above 1.44 solar masses (the
Chandrasekhar limit) collapses and is not halted by the degenerate electrons.
As the star collapses, the →Fermi energy of the electrons increases to the point
where it is energetically favorable for them to combine with protons to produce

Degenerate matter
93

neutrons (via inverse beta decay, also termed "neutralization"). The result of
this collapse, for the simplest models of neutron-degenerate matter, is an ex-
tremely compact star composed almost entirely of a degenerate neutron gas.
Neutrons in a degenerate neutron gas are spaced much more closely than elec-

FT
trons in an electron-degenerate gas, because the more massive neutron has a
much shorter wavelength at a given energy. In the case of neutron stars and
white dwarf stars, this is compounded by the fact that the pressures within neu-
tron stars are much higher than those in white dwarfs. The pressure increase
is caused by the fact that the compactness of neutron stars causes gravitation-
al forces to be much higher than in a less compact body with similar mass,
resulting in a star on the order of a thousand times smaller than a white dwarf.
There is an upper limit to the mass of a neutron-degenerate object, the Tolman-
Oppenheimer-Volkoff limit, which is analogous to the Chandrasekhar limit for
electron-degenerate objects. The precise limit is unknown, as it depends on the
equations of state of neutron-degenerate matter, for which a highly accurate
model is not yet available. Above this limit, a neutron star may collapse into
a black hole, or into other, denser forms of degenerate matter (such as quark
matter) if these forms exist and have suitable properties (mainly related to
A
degree of compressibility, or "stiffness", described by the equations of state).

Quark degeneracy
At densities greater than those supported by neutron degeneracy, quark matter
is expected to occur. Several variations of this have been proposed that rep-
resent quark-degenerate states. Strange matter is a degenerate gas of quarks
that is often assumed to contain strange quarks in addition to the usual up and
down quarks. Color superconductor materials are degenerate gases of quarks
DR
in which quarks pair up in a manner similar to Cooper pairing in electrical
superconductors. The equations of state for the various proposed forms of
quark-degenerate matter vary widely, and are usually also poorly defined, due
to the difficulty modelling strong force interactions.
Quark-degenerate matter may occur in the cores of neutron stars, depending
on the equations of state of neutron-degenerate matter. It may also occur in
quark stars, formed by the collapse of objects above the Tolman-Oppenheimer-
Volkoff mass limit for neutron-degenerate objects. Whether quark-degenerate
matter forms at all in these situations depends on the equations of state of
both neutron-degenerate matter and quark-degenerate matter, both of which
are poorly known.

Degenerate matter
94

Speculative types of degenerate matter


Preon degeneracy
Preons are subatomic particles proposed to be the constituents of quarks, which

FT
become composite particles in preon-based models. If preons exist, preon-
degenerate matter might occur at densities greater than that which can be
supported by quark-degenerate matter. The properties of preon-degenerate
matter depend very strongly on the model chosen to describe preons, and the
existence of preons is not assumed by the majority of the scientific community,
due to conflicts between the preon models originally proposed and experimen-
tal data from particle accelerators.

See also
• Degenerate star
• White dwarf (degenerate dwarf) - white dwarf material - electron-
degenerate matter
A



Neutron star - neutron matter - neutron-degenerate matter
Quark star - quark matter - quark-degenerate matter
Preon star - preon matter - preon-degenerate matter
• Pauli Exclusion Principle
• Uncertainty Principle

External links
Detailed mathematical explanation of degenerate gases 44
DR

Source: http://en.wikipedia.org/wiki/Degenerate_matter

Principal Authors: Voyajer, Christopher Thomas, Alan Peakall, Eyu100, Roadrunner

44 http://ircamera.as.arizona.edu/astr_250/Lectures/Lec17_sml.htm

Degenerate matter
95

Delayed choice quantum eraser

A delayed choice quantum eraser is a combination between a quantum eraser

FT
experiment and Wheeler’s delayed choice experiment. This experiment has
actually been performed and published by Yoon-Ho Kim, R. Yu, S.P. Kulik, Y.H.
Shih, and Marlon O. Scully 45 Phys.Rev.Lett. 84 1-5 (2000). This experiment
was designed to investigate a very peculiar result of the well known double slit
experiment of quantum mechanics, the dual wave particle nature of light, and
in fact all matter.

Introduction
In the double slit experiment, a photon passes through a double slit apparatus,
in which the photon must pass either through one or the other of two slits, and
then registers on a detector, which can determine where the photon reaches
the detector, like an image projected on a screen. If one allows many photons
to individually pass through either slit A or slit B and doesn’t know which slit
they passed through, an interference pattern emerges on the detector. The
A
interference pattern indicates that the light beam is in fact made up of waves.
However, if one somehow observes which of the two slits each photon actually
passes through, a different result will be obtained. In this case, each photon
hits the detector after going through only one slit and a single concentration
of hits in the middle of the detection field. This result is consistent with light
behaving as individual particles, like tiny bullets. The very odd thing about
this is that a different outcome results based on whether or not the photon is
observed after it goes through the slit but before it hits the detector.
DR
In a quantum eraser experiment, one arranges to detect which one of the slits
the photon passes through, but also construct the experiment in such a way that
this information can be "erased" after the fact. It turns out that if one observes
which slit the photon passes through, the "no interference" or particle behavior
will result, which is what quantum mechanics predicts, but if the quantum
information is "erased" regarding which slit the photon passed through, the
photons revert to behaving like waves.
However, Kim, et al. have shown that it is possible to delay the choice to erase
the quantum information until after the photon has actually hit the target. But,
again, if the information is "erased," the photons revert to behaving like waves,
even if the information is erased after the photons have hit the detector.

45 http://xxx.lanl.gov/pdf/quant-ph/9903047

Delayed choice quantum eraser


96

The experiment
The experimental setup, described in much more detail at 46, is as follows. First,
generate a photon and pass it through a double slit apparatus. After the photon
goes through slit A or B, a special crystal (one at each slit) uses spontaneous

FT
parametric down conversion (SPDC) to convert the photon into two identical
entangled photons with 1/2 the frequency of the original photon. One of these
photons continues to the target detector, while the other entangled photon is
deflected by a prism to bounce off a mirror some distance away. Now, if the
second photon (coming from slit A or slit B) is observed, it is known which
slit the original photon went through, so the photon behaves like a particle. If
the second photon’s paths from slit A and B are combined, the which-way path
is not observed, and the first photon behaves like a wave. The experimenter
can choose to observe or not observe the which-way information by erasing (or
detecting) information about the second photon’s path.
The results from Kim, et al. have shown that, in fact, observing the second pho-
ton’s path will determine the particle or wavelike behavior of the first photon
at the detector, even if the second photon is not observed until after the first
A
photon arrives at the detector. In other words, the delayed choice to observe
or not observe the second photon will change the outcome of an event in the
past.

Discussion
In terms of the conventional way of viewing the physical universe, this result
seems to be a paradox. This experiment demonstrates the possibility of ob-
serving both particle-like and wave-like behavior of a photon using quantum
DR
entanglement. Furthermore, the behavior of the photon at the primary detec-
tor can be changed even after the registration of the event by the detector.
In their paper, Kim, et al. 47 explain that the concept of complementarity is one
of the most basic principles of quantum mechanics. According to the Heisen-
berg Uncertainty Principle, it is not possible to measure both precise position
and momentum of a quantum particle at the same time. In other words, posi-
tion and momentum are complementary. In 1927, Niels Bohr described com-
plementarity as “wave-like” and “particle-like” behavior of a quantum particle.
this has come to be known as the wave-particle duality of quantum mechan-
ics. The double-slit experiment is a good example of this concept. Feynman
believed that this was the basic mystery of quantum mechanics. The actual

46 http://xxx.lanl.gov/pdf/quant-ph/9903047
47 http://xxx.lanl.gov/PS_cache/quant-ph/pdf/9903/9903047.pdf

Delayed choice quantum eraser


97

mechanisms that enforce complementarity vary from one experimental situa-


tion to another. In the double-slit experiment, the common wisdom is that the
Heisenberg Uncertainty Principle makes it impossible to determine which slit
the photon passes through without at the same time disturbing it enough to

FT
destroy the interference pattern. However, in 1982, Scully and Druhl found
a way around the position-momentum uncertainty obstacle and proposed a
quantum eraser to obtain which-path or particle-like information without in-
troducing large uncontrolled phase factors to disturb the interference. They
found that the interference pattern disappears when which-path information
is obtained, even if this information was obtained without directly observing
the original photon. Even more surprising was that, if you somehow "erase"
the which-path information, the interference pattern reappears! And, perhaps
most provocative of all, you can delay the "choice" to "erase" or "observe" the
which-path information and still restore the interference pattern, even after
the original photon has been "observed" at the primary detector!
How can this be? It would seem that the "choice" to observe or erase the which-
path information can change the position where the photon is recorded on the
detector, even after it should have already been recorded.
A
One explanation of this paradox would be that this is a kind of time travel.
In other words, the delayed "choice" to "erase" or "observe" the which-path
information of the original photon can change the outcome of an event in the
past. Another explanation would be that in fact both outcomes occur. The
universe itself exists in a superposition of states in which either the original
photon goes through slit A or slit B and in which the which-path information
either "observed" or "erased". This is described in detail in the Everett many-
worlds interpretation of quantum mechanics.
DR
External links
• basic delayed choice experiment 48
• delayed choice quantum eraser 49
• the notebook of philosophy and physics 50

Source: http://en.wikipedia.org/wiki/Delayed_choice_quantum_eraser

48 http://www.bottomlayer.com/bottom/basic_delayed_choice.htm
49 http://www.bottomlayer.com/bottom/kim-scully/kim-scully-web.htm
50 http://www.bottomlayer.com/

Delayed choice quantum eraser


98

Diabatic

In quantum chemistry, the potential energy surfaces are obtained within the

FT
adiabatic or Born-Oppenheimer approximation. This corresponds to a repre-
sentation of the molecular wave function where the variables corresponding to
the molecular geometry and the electronic degrees of freedom are separated.
The non separable terms are due to the nuclear kinetic energy terms in the
molecular Hamiltonian and are said to couple the potential energy surfaces. In
the neighbourhood of an avoided crossing or conical intersection, these terms
cannot be neglected. One therefore usually performs one unitary transforma-
tion from the adiabatic representation to the so-called diabatic representation
in which the nuclear kinetic energy operator is diagonal. In this representation,
the coupling is due to the electronic energy and is a scalar quantity which is
much more easy to estimate numerically.

Source: http://en.wikipedia.org/wiki/Diabatic
A
Principal Authors: TimBentley, Elfguy

Dirac equation in the algebra of physical


space

The Dirac equation, as the relativistic equation that describes spin 1/2 particles
DR
in quantum mechanics can be written in terms of the Algebra of physical space
(APS), which is a case of a Clifford algebra or geometric algebra that is based
in the use of paravectors.
The Dirac equation in APS, including the electromagnetic interaction, reads
¯ 3 + eĀΨ = mΨ̄†
i∂Ψe

Another form of the Dirac equation in terms of the Space time algebra was
given earlier by D. Hestenes.
In general, the Dirac equation in the formalism of geometric algebra has the
advantage of providing a direct geometric interpretation.

Dirac equation in the algebra of physical space


99

Relation with the standard form


The Dirac equation can be written also as
i∂ Ψ̄† e3 + eAΨ̄† = mΨ

FT
The standard form is obtained multiplying the spinor with the projector P =
1
2 (1+e3 ). Without electromagnetic interaction, the following matrix is obtained
from
 the two  previous
 equivalent forms
 of the Dirac equation
i∂¯ Ψ̄† P3 Ψ̄† P3

0
=m
i∂ 0 ΨP3 ΨP3

This colum of projected spinors are related to the spinors in the Weyl represen-
tation. This is more evident identifying the right and left Weyl spinors as
Ψ̄† P3 = ψL

ΨP3 = ψR

so
A
 that
0 i∂0 + i∇
 
ψL
=m
 
ψL
i∂0 − i∇ 0 ψR ψR

In the matrix representation each expression is replaced by a 2 by 2 matrix,


including ψL and ψR . The nabla operator can be written in terms of the Pauli
matrices as
DR
∇ → σ · ∇.

On the other hand, only the firs column of each spinor is taken
ψL,R → F irstColum(ψL,R ),

so
the Dirac
 equation
 becomes
    
0 1 0 σ ψL ψL
∂ + ·∇ =m ,
1 0 0 −σ 0 ψR ψR

from which, the standard relativistic covariant form of the Dirac equation is
found
iγ µ ∂µ ψ = mψ.

Dirac equation in the algebra of physical space


100

Spinor expansion in a null basis


The spinor Ψ can be expanded in a null basis as follows
Ψ = ψ1∗ P̄3 − ψ2∗ P3 e1 + ψ3 P3 + ψ4 e1 P3 ,

FT
where each coefficient is extracted from the spinor such that
ψ1 = 2hP3 Ψ̄† P3 iS

ψ2 = 2he1 P̄3 Ψ̄† P3 iS

ψ3 = 2hP3 Ψ P3 iS

ψ4 = 2he1 P̄3 ΨP3 iS

This expansion is cleanly related with the colum spinor in the Weyl represen-
tation 
so that
 the Weyl spinor components are
ψ1
Ψ→
A
ψ2 
ψ3 

ψ4

Similarly, the spinor components in the Pauli-Dirac representation are calculat-


ed as  
ψ3 + ψ1
ψ4 + ψ2 
DR
Ψ→
ψ3 − ψ1 

ψ4 − ψ2

Current
The current is defined as
J = ΨΨ† ,

which satisfies the continuity equation


¯


∂J S
=0

Dirac equation in the algebra of physical space


101

Second order Dirac equation


An application of the Dirac equation on itself leads to the second order Dirac
equation
(−∂ ∂¯ + AĀ)Ψ − i(2e A∂¯ + eF )Ψe3 = m2 Ψ

FT
S

Free particle solutions


Positive energy solutions
A solution for the free particle with momentum p = p0 + p and positive energy
p0 > 0 is
q
p
Ψ= m R(0) exp(−i hpx̄iS e3 ).

This solution is unimodular


ΨΨ̄ = 1
A
and the current resembles the classical proper velocity
u= p
m

p
J = ΨΨ† = m

Negative energy solutions


A solution for the free particle with negative energy and momentum p =
DR
−|p0 | − p = −p0 is
q
p0
Ψ=i m R(0) exp(i hp0 x̄iS e3 ),

This solution is anti-unimodular


ΨΨ̄ = −1

p
and the current resembles the classical proper velocity u = m
p
J = ΨΨ† = −m ,

but with a remarkable feature: "the time runs backwards"


dt

p
dτ = m S < 0

Dirac equation in the algebra of physical space


102

Dirac Lagrangian
The Dirac Lagrangian is
L = hi∂ Ψ̄† e3 Ψ̄ − eAΨ̄† Ψ̄ − mΨΨ̄i0

FT
See also
• Paravector
• Algebra of physical space
• Geometric algebra

References
• Baylis, William (2002). Classical eigenspinors and the Dirac equation, Phys.
Rev. A 45, 4293–4302 (1992)

• Hestenes D., Observables, operators, and complex numbers in the Dirac


theory, J. Math. Phys. 16, 556 (1975)
A
Source: http://en.wikipedia.org/wiki/Dirac_equation_in_the_algebra_of_physical_space

Dirac operator
DR
In mathematics and quantum mechanics, a Dirac operator is a differential
operator that is a formal square root, or half-iterate, of a second-order operator
such as a Laplacian. The original case which concerned Paul Dirac was to
factorise formally an operator for Minkowski space, to get a form of quantum
theory compatible with special relativity; to get the relevant Laplacian as a
product of first-order operators he introduced spinors.
In general, let D be a first-order differential operator acting on a vector bundle
V over a Riemannian manifold M .
If
D2 = ∆,

with ∆ being the Laplacian of V , D is called a Dirac operator.


In high-energy physics, this requirement is often relaxed: only the second-order
part of D2 must equal the Laplacian.

Dirac operator
103

Examples
1. −i∂x is a Dirac operator on the tangent bundle over a line.
2: We now consider a simple bundle of importance in physics: The configura-
tion space of a particle with spin 1/ 2 confined to a plane, which is also the base

FT
manifold. Physicists generally think of wavefunctions ψ: R 2 → C 2 which they
write
 
χ(x, y)
η(x, y)

where x and y are the usual coordinate functions on R 2. χ specifies the prob-
ability amplitude for the particle to be in the spin-up state, and similarly for η.
The so-called spin-Dirac operator can then be written
D = −iσx ∂x − iσy ∂y ,

where σ i are the Pauli matrices. Note that the anticommutation relations for
the Pauli matrices make the proof of the above defining property trivial. Those
A
relations define the notion of a Clifford algebra.
3: The most famous Dirac operator describes the propagation of a free electron
in three dimensions and is elegantly written
D = γ µ ∂µ

using Einstein’s summation convention and even more elegantly as


D = ∂/
DR
using the Feynman slash notation.

See also
• Clifford algebra
• Connection
• Dolbeault operator
• Heat kernel.

Source: http://en.wikipedia.org/wiki/Dirac_operator

Principal Authors: Phys, Charles Matthews, Dreamy224, Dcarlson, Rex the first

Dirac operator
104

Double-slit experiment

The double-slit experiment or two-slit experiment consists of letting light

FT
diffract through two slits producing fringes on a screen. These fringes or inter-
ference patterns have light and dark regions corresponding to where the light
waves have constructively and destructively interfered. The experiment can
also be performed with a beam of electrons or atoms, showing similar interfer-
ence patterns; this is taken as evidence of the "wave-particle duality" predicted
by quantum physics. Note, however, that a double-slit experiment can also be
performed with water waves in a ripple tank; the explanation of the observed
wave phenomena does not require quantum mechanics in any way. The phe-
nomenon is quantum mechanical only when quantum particles - such as atoms,
electrons, or photons - manifest as waves.

Importance to physics
Although the double-slit experiment is now often referred to in the context
of quantum mechanics, it was originally performed by the English scientist
A
Thomas Young some time around 1805 in an attempt to resolve the question of
whether light was composed of particles (the "corpuscular" theory), or rather
consisted of waves traveling through some aether, just as sound waves travel
in air.
The interference patterns observed in the experiment seemed to discredit the
corpuscular theory, and the wave theory of light remained well accepted un-
til the early 20th century, when evidence began to accumulate which seemed
instead to confirm the particle theory of light.
DR
The double-slit experiment, and its variations, then became a classic Gedanken-
experiment (thought experiment) for its clarity in expressing the central puz-
zles of quantum mechanics; although in this form the experiment was not ac-
tually performed with anything other than light until 1961, when Claus Jöns-
son of the University of Tübingen performed it with electrons. (C Jönsson,
Zeitschrift für Physik 161, 454; C. Jönsson 1974 "Electron diffraction at mul-
tiple slits", American Journal of Physics 42 4-11), and not until 1974 in the
form of "one electron at a time", in a laboratory at the University of Milan, by
researchers led by Pier Giorgio Merli, of LAMEL-CNR Bologna.
The results of the 1974 experiment were published and even made into a short
film, but did not receive wide attention. The experiment was repeated in 1989
by Tonomura et al at Hitachi in Japan. Their equipment was better, reflecting
15 years of advances in electronics and a dedicated development effort by the
Hitachi team. Their methodology was more precise and elegant, and their

Double-slit experiment
105

results agreed with the results of Merli’s team. Although Tonomura asserted
that the Italian experiment had not detected electrons one at a time - a key to
demonstrating the wave-particle paradox - single electron detection is clearly
visible in the photos and film taken by Merli and his group.

FT
In September 2002, the double-slit experiment of Claus Jönsson was voted "the
most beautiful experiment" by readers of Physics World.

Explanation of experiment

A
In Young’s original experiment, sunlight passes first through a single slit, and
then through two thin vertical slits in otherwise solid barriers, and is then
viewed on a rear screen.
When either slit is covered, a single peak is observed on the screen from the
DR
light passing through the other slit.
But when both slits are open, instead of the sum of these two singular peaks
that would be expected if light were made of particles, a pattern of light and
dark fringes is observed.
This pattern of fringes was best explained as the interference of the light waves
as they recombined after passing through the slits, much as waves in water re-
combine to create peaks and swells. In the brighter spots, there is "constructive
interference", where two "peaks" in the light wave coincide as they reach the
screen. In the darker spots, "destructive interference" occurs where a peak and
a trough occur together.

Double-slit experiment
106

FT
Figure 12

Replicating Young’s experiment


This experiment can easily be demonstrated in just the way that Young demon-
strated it to the Royal Society of London. An assistant outside used mirrors
to direct sunlight at a pinhole opening. The beam from the opening was then
bisected by "a slip of card". To make things easier, a modern experimenter
could replace the sunlight and mirrors with a laser pointer covered, except for
A
a pinhole, by black paper. Splitting the beam with a small strip of notecard will
produce a visible interference pattern when the beam is projected across the
room. 51

Quantum version of experiment


By the 1920s, various other experiments (such as the photoelectric effect) had
demonstrated that light interacts with matter only in discrete, "quantum"-sized
packets called photons.
DR
If sunlight is replaced with a light source that is capable of producing just one
photon at a time, and the screen is sensitive enough to detect a single photon,
Young’s experiment can, in theory, be performed one photon at a time – with
identical results.
If either slit is covered, the individual photons hitting the screen, over time,
create a pattern with a single peak. But if both slits are left open, the pattern of
photons hitting the screen, over time, again becomes a series of light and dark
fringes. This result seems to both confirm and contradict the wave theory. On
the one hand, the interference pattern confirms that light still behaves much
like a wave, even though we send it one particle at a time. On the other hand,
each time a photon with a certain energy is emitted, the screen detects a pho-
ton with the same energy. Under the Copenhagen Interpretation of quantum

51 http://www.cavendishscience.org/phys/tyoung/tyoung.htm

Double-slit experiment
107

theory, an individual photon is seen as passing through both slits at once, and
interfering with itself, producing the interference pattern.
A remarkable refinement of the double-slit experiment consists of putting a
detector at each of the two slits, to determine which slit the photon passes

FT
through on its way to the screen (If the photon or electron passes through
only one slit - which it must do, as, by definition, a photon or an electron is
a quantum, or "packet" of energy which cannot be subdivided - then logically
it cannot interfere with itself and produce an interference pattern). When the
experiment is arranged in this way, the fringes disappear.
The Copenhagen interpretation posits the existence of probability waves which
describe the likelihood of finding the particle at a given location. Until the
particle is detected at any location along this probability wave, it effectively
exists at every point. Thus, when the particle could be passing through either of
the two slits, it will actually pass through both, and so an interference pattern
results. But if the particle is detected at one of the two slits, then it can no
longer be passing through both - it must exist at one or the other, and so no
interference pattern appears.
A
The many worlds interpretation states that the particle not only goes through
both slits but that it is detected at every possible final location as well – but in
different, mutually unobservable worlds.
This is similar to the path integral formulation of quantum mechanics provided
by Richard Feynman (although Feynman stresses that this is merely a math-
ematical description, not an attempt to describe some "real" process that we
cannot see), in which a particle such as a photon takes every possible path
through space-time to get from point A to point B. In the double-slit experi-
DR
ment, point A might be the emitter, and point B the screen upon which the
interference pattern appears, and a particle takes every possible path - through
both slits at once - to get from A to B. When a detector is placed at one of
the slits, the situation changes, and we now have a different point B at the
detector, and a new path between the detector and the screen - upon which
the interference pattern no longer appears).

Conditions for interference


A necessary condition for obtaining an interference pattern in a double-slit
experiment concerns the difference in pathlength between two paths that light
can take to reach a zone of constructive interference on the viewing screen.
This difference must be the wavelength of the light that is used, or a multiple
of this wavelength. (See illustration. 52) If a beam of sunlight is let in, and that

52 http://schools.matter.org.uk/Content/Interference/formula.html

Double-slit experiment
108

beam is allowed to fall immediately on the double slit, then the fact that the
Sun is not a point source degrades the interference pattern. The light from a
source that is not a point source behaves like the light of many point sources
side by side. Each can create an interference pattern, but the interference

FT
patterns of each of the many-side-by-side sources does not coincide on the
screen, so they average each other out, and no interference pattern is seen.
The presence of the first slit is necessary to ensure that the light reaching the
double slit is light from a single point source. The path length from the single
slit to the double slit is equally important for obtaining the interference pattern
as the path from the double slit to the screen.
Newton’s rings show that light does not have to be coherent in order to produce
an interference pattern. Newton’s rings can be readily obtained with plain
sunlight. 53 More rings are discernible if for example light from a Sodium lamp
is used, since Sodium lamp light is only a narrow band of the spectrum. Light
from a Sodium lamp is incoherent. Other examples of interference patterns
from incoherent light are the colours of soap bubbles and of oil films on water.
In general, interference patterns are clearer when monochromatic or near-
A
monochromatic light is used. Laserlight is as monochromatic as light can be
made, therefore laserlight is used to obtain an interference pattern.
If the two slits are illuminated by coherent waves, but with polarizations per-
pendicular with respect to each other, the interference pattern disappears.

Results observed
The bright bands observed on the screen happen when the light has interfered
constructively – where a crest of a wave meets a crest. The dark regions show
DR
destructive interference – a crest meets a trough.
nλ x
d = L

xd
nλ = L

where

is the wavelength of the light

d is the separation of the slits

53 Newton’s rings. Newton’s Rings from Eric Weisstein’s World of Physics(http://scienceworld.wolfram.

com/physics/NewtonsRings.html)

Double-slit experiment
109

x is the distance between the bands of light (also called fringe distance)

L is the distance from the slits to the screen

FT
n is the order of maxima observed (Central Maximum is n=1)

This is only an approximation and depends on certain conditions.


It is possible to work out the wavelength of light using this equation and the
above apparatus. If d and L are known and x is observed, then can be easily
calculated.
A detailed treatment of the mathematics of double-slit interference in the con-
text of quantum mechanics is given in the article on Englert-Greenberger dual-
ity.

Shape of interference fringes


The theoretical shapes of the interference fringes observed in Young’s double
slit experiment are straight lines which is easily proved.
A
In case two pinholes are used instead of slits, as in the original Young’s exper-
iment, hyperbolic fringes are observed. This is because the difference in paths
travelled by the light from the two sources is a constant for a fringe which is
the property of a hyperbola.
If the two sources are placed on a line perpendicular to the screen, the shape
of the interference fringes is circular as the individual paths travelled by light
from the two sources are always equal for a given fringe. This can be done in
DR
simpler way by placing a mirror parallel to a screen at a distance and a source
of light just above the mirror. (Note the extra phase difference of π due to
reflection at the interface of a denser medium)

See also
• Elitzur-Vaidman bomb-testing problem
• Quantum eraser experiment
• Quantum coherence

Video Demonstration
• From Movie "Down the Rabbit Hole" (sequel to What the Bleep Do We
Know!?) Clip of Double Split Experiment

Double-slit experiment
110

http://www.whatthebleep.com/trailer/doubleslit.wm.low.html (The down-


load time over a modem connection is very slow, and playback is interrupted
so frequently that it may be impossible to understand.)

FT
References
• Tipler, Paul (2004). Physics for Scientists and Engineers: Electricity, Mag-
netism, Light, and Elementary Modern Physics, 5th ed., W. H. Freeman.
ISBN 0716708108.
• Gribbin, John (1999). Q is for Quantum: Particle Physics from A to Z.
Weidenfeld & Nicholson. ISBN 0753806851.
• Feynman, Richard P. (1988). QED: The Strange Theory of Light and Matter.
Princeton University Press. ISBN 0691024170.

External links
• Simple Derivation of Interference Conditions 54
• The Double Slit Experiment 55
A


Double-Slit in Time 56
Keith Mayes explains the Double Slit Experiment in plain English 57

• Carnegie Mellon department of physics, photo images of Newton’s rings 58


• Java demonstration of double slit experiment 59
• Double-slit experiment animation 60
• Double-slit experiment cartoon animation 61
DR
Source: http://en.wikipedia.org/wiki/Double-slit_experiment

Principal Authors: Eequor, Linas, The Anome, Reddi, Samboy, Pfalstad, Laurascudder, Afshar, Cleonis,
Lumidek

54 http://schools.matter.org.uk/Content/Interference/formula.html
55 http://physicsweb.org/article/world/15/9/1
56 http://physicsweb.org/articles/news/9/3/1/1?rss=2.0
57 http://www.thekeyboard.org.uk/Quantum%20mechanics.htm
58 http://physdemo.phys.cmu.edu/newton_rings.htm
59 http://www.falstad.com/ripple/ex-2slit.html
60 http://homepage.univie.ac.at/Franz.Embacher/KinderUni2005/waves.gif
61 http://video.google.com/videoplay?docid=-4237751840526284618&q=quantum

Double-slit experiment
111

Duru-Kleinert transformation

The Duru-Kleinert transformation, named after H. Duru and Hagen Klein-

FT
ert, is a mathematical method for solving path integrals of physical systems
with singular potentials, which is necessary for the solution of all atomic path
integrals due to the presence of Coulomb potentials (singular like 1/r). The
Duru-Kleinert transformation replaces the diverging time-sliced path integral
of Richard Feynman (which thus does not exist) by a well-defined convergent
one.

Papers
• H. Duru and H. Kleinert, Solution of the Path Integral for the H-Atom, Phys.
Letters B 84, 185 (1979) 62

• H. Duru and H. Kleinert, Quantum Mechanics of H-Atom from Path Inte-


grals, Fortschr. d. Phys. 30, 401 (1982) 63
A
• H. Kleinert, Path Integrals in Quantum Mechanics, Statistics, Polymer
Physics, and Financial Markets 3. ed., World Scientific (Singapore, 2004) 64
( read book here 65)

Source: http://en.wikipedia.org/wiki/Duru-Kleinert_transformation

Principal Authors: Tobias Bergemann, Michael Hardy, Phys, Conscious


DR

62 http://www.physik.fu-berlin.de/~kleinert/65/65.pdf
63 http://www.physik.fu-berlin.de/~kleinert/83/83.pdf
64 http://www.worldscibooks.com/physics/5057.html
65 http://www.physik.fu-berlin.de/~kleinert/b5

Duru-Kleinert transformation
112

Ehrenfest theorem

The Ehrenfest theorem, named after Paul Ehrenfest, relates the time deriva-

FT
tive of the expectation value for a quantum mechanical operator to the com-
mutator of that operator with the Hamiltonian of the system. It is
d
dt hAi = ~i h[H, A]i

where A is some QM operator and <A> is its expectation value. Notice how
neatly Ehrenfest’s theorem fits into the →Heisenberg picture of quantum me-
chanics.
Ehrenfest’s theorem is closely related to Liouville’s theorem from Hamiltonian
mechanics, which involves the Poisson bracket instead of a commutator. In
fact, it is a general rule of thumb that a theorem in quantum mechanics which
contains a commutator can be turned into a theorem in Classical mechanics by
changing the commutator into a Poisson bracket and multiplying by i~.
The theorem can be shown to follow from the →Lindblad equation, a master
A
equation for the time evolution of a mixed state.

Derivation
Suppose some system is presently in a quantum state Φ. If we want to know
the instantaneous time derivative of the expectation value of A, that is, by
definition
R ∗ R  dΦ∗   
d d
AΦ dV + Φ∗ dA
R
dt hAi = dt Φ AΦ dV = dt dt Φ dV +
DR
R ∗  dΦ 
Φ A dt dV

R  dΦ∗   
Φ∗ A dΦ
R
= dt AΦ dV + dt dV,

where we are integrating over all space, and we have assumed the operator A is
time independent, so that its derivative is zero. If we apply the →Schrödinger
equation, we find that
dΦ 1
dt = i~ HΦ

and
dΦ∗ −1 ∗ ∗ −1 ∗
dt = i~ Φ H = i~ Φ H.

Ehrenfest theorem
113

Notice H = H ∗ because the Hamiltonian is hermitian. Placing this into the


above equation we have
d 1
R ∗ 1 i
dt hAi = i~ Φ (AH − HA)Φ dV = i~ h[A, H]i = ~ h[H, A]i.

FT
General example
For the very general example of a massive particle moving in a potential, the
Hamiltonian is simply
p2
H= 2m + V (r)
where r is just the location of the particle. Suppose we wanted to know the
instantaneous change in momentum p. Using Ehrenfest’s theorem, we have
d 1 1
dt hpi = i~ h[p, H]i = i~ h[p, V (r)]i

since p commutes with itself. When represented in coordinate space, the mo-
mentum operator p = −i~∇, so
d
R ∗ R ∗
dt hpi = Φ V (r)∇Φ dV − Φ ∇(V (r)Φ) dV.
A
After applying a product rule, we have
d
dt hpi = h−∇V (r)i = hF i,

but we recognize this as Newton’s second law. This is an example of the corre-
spondence principle, the result manifests as Newton’s second law in the case of
having so many particles that the net motion is given exactly by the expectation
value of a single particle.
DR
Notes
• ↑ In →Bra-ket notation
d −1 −1 −1 ∗
dt hφ|xi = i~ hφ|Ĥ|xi = i~ hφ|xiH = i~ Φ H

where Ĥ is the Hamiltonian operator, and H is the Hamiltonian represent-


ed in coordinate space (as is the case in the derivation above). In other
words, we applied the adjoint operation to the entire Schrödinger equation,
which flipped the order of operations for H and Φ.

Source: http://en.wikipedia.org/wiki/Ehrenfest_theorem

Ehrenfest theorem
114

Principal Authors: Michael Hardy, Hyandat, WMDickson, Pathfinder, Brienanni

Einselection

FT
Einselection is short for environmentally-induced superselection, a nickname
coined by Wojciech H. Zurek. Einselection is the quantum process whereby the
environment persistently monitors a quantum system, causing decoherence be-
tween its states. The decoherence process selects a certain subset of states from
the enormous →Hilbert space. These ’pointer states’ are stable despite environ-
mental interaction, which explains the emergence of a preferred basis in quan-
tum measurement. The einselected states lack coherence, and therefore do not
exhibit the quantum behaviours of entanglement and superposition. Since only
quasi-local, essentially classical states survive the decoherence process, einse-
lection can in many ways explain the emergence of a (seemingly) classical
reality in a fundamentally quantum universe (at least to local observers).
A
Source: http://en.wikipedia.org/wiki/Einselection

Principal Authors: Conscious, Mo0, Icairns, ShaneKing, Jag123

Electronic density

In quantum mechanics, and in particular in quantum chemistry, the electron-


DR
ic density ρ corresponding to an N -electron wavefunction Ψ(N ) is the one-
electron function given by
ρ(x) = dx2 ... dxN |Ψ(N ) (x, x2 , ..., xN )|2
R

In the case Ψ(N ) is a →Slater determinant made of N spin orbitals ϕk :


PN
ρ(x) = N1 k=1 |ϕk (x)|
2

The two-electron electronic density is given by


ρ(x, x0 ) = dx3 ... dxN |Ψ(N ) (x, x0 , x3 , ..., xN )|2
R

Those quantities are particularly important in the context of density functional


theory.
The coordinates x used here are the spin-spatial coordinates.

Electronic density
115

Source: http://en.wikipedia.org/wiki/Electronic_density

Principal Authors: Vb, Gsp

FT
Electronic Hamiltonian

The Electronic Hamiltonian is an operator in quantum mechanics (and in par-


ticular quantum chemistry) which describes the motions of electrons and nuclei
in a polyatomic molecule. The terminology is sometimes used interchange-
ably to mean either the Electronic molecular Hamiltonian or the full electronic
Hamiltonian. The latter includes a kinetic energy operator corresponding to
the contributions from the nuclei.
There are a number of interrelated concepts associated with the term "Elec-
tronic Hamiltonian". These include the following:

• Full electronic Hamiltonian




A Electronic Hamiltonian
Nuclear Hamiltonian
• Clampled Hamiltonian

Depending on the context, "full electronic Hamiltonian" may be used inter-


changeably with "electronic Hamiltonian". Furthermore, the "Clamped Hamil-
tonian" or may be used interchangeably with "Electronic Hamiltonian". The
latter is usually used only when discussing various methods associated with
DR
the Born-Oppenheimer approximation.

Full electronic Hamiltonian


Let R denote the vector of nuclear coordinates, and r the vector of electronic
coordinates.
The full electronic Hamiltonian consists of 5 terms. They are

• The kinetic energy operators for each nuclei in the system;


• The kinetic energy operators for each electron in the system;
• The potential energy between the electrons and nuclei - the total electron-
nucleus Coulombic attraction in the system;
• The potential energy arising from Coulombic electron-electron repulsions
• The potential energy arising from Coulombic nuclei-nuclei repulsions - al-
so known as the nuclear repulsion energy. See electric potential for more
details.

Electronic Hamiltonian
116

Hence, the full electronic Hamiltonian is


Ĥel = T̂ n + T̂ e + Û en + Û ee + Û nn

FT
Electronic Hamiltonian and potential energy surfaces
Very often, the electronic Hamiltonian is defined to be
Ĥ e = T̂ e + Û en + Û nn + Û ee

so that the full Hamiltonian would be written as


Ĥel = Ĥ e + T̂ n .

In this case, the kinetic energy operator T̂ n would be known as the nuclear
Hamiltonian. The electronic Hamiltonian contains all three potential terms
because their sum
V̂ (r, R) = Û en + Û nn + Û ee
A
is the expression which gives rise to the physically meaningful potential energy
surfaces ubiquitous in chemistry, for fixed nuclear geometry R.
Written in atomic units, the electronic Hamiltonian becomes:
P P P Za P P
Ĥel = i − 21 ∇2i − i
1
a |ri −da | + 2 i
1
j6=i |ri −rj | +
1 P P Za Zb
2 a b6=a |da −db |

where

• ri is the vector position of electron i with vector components in Bohr radii,


DR
• Za is the charge of fixed nucleus a in units of the elementary charge,
• da is the vector position of nucleus a with vector components in Bohr radii.

Electronic molecular Hamiltonian


The electronic molecular Hamiltonian is the term of the molecular Hamilto-
nian obtained when the molecular geometry is frozen. This is also known as
the clamped Hamiltonian or clamped Hamiltonian approximation. With-
in the Born-Oppenheimer approximation, the electronic Hamiltonian is said to
depend adiabatically on the molecular geometry. Its discrete eigenvalues are
called potential energy surfaces and the corresponding eigenstates the elec-
tronic states of the molecule. The electronic states are labelled according to
their group representation and spin multiplicity.

Electronic Hamiltonian
117

Adiabatic and diabatic states


By definition, the adiabatic states are diagonal in the electronic Hamiltonian.
A consequence is that it is not diagonal in the kinetic energy operator. The off
diagonal terms of this operator are known as the nonadiabatic operator.

FT
Strictly diabatic states do not exist in general, although in the ideal case, it
is diagonal in the kinetic energy operator, and off diagonal in the electronic
Hamiltonian. In other words, the diabatic states minimize the magnitude of
the contributions of the nonadiabatic operator.

Source: http://en.wikipedia.org/wiki/Electronic_Hamiltonian

Principal Authors: HappyCamper, Ian Pitchford

Electronic state

Electronic state is a quantum state of a system consisting of electrons (usually


A
orbitals or chemical bonds in crystals or molecules). The state with lowest
energy is called ground state, states with higher energy are excited states.

Source: http://en.wikipedia.org/wiki/Electronic_state

Elementary particle
DR
For the novel by Michel Houellebecq, see The Elementary Particles.

In particle physics, an elementary particle or fundamental particle is a particle


not known to have substructure; that is, it is not made up of smaller particles.
If an elementary particle truly has no substructure, then it is one of the basic
particles of the universe from which all larger particles are made. In the mod-
ern theory of particle physics, the Standard Model, the quarks, leptons, and
gauge bosons are elementary particles. 6667 Historically, the hadrons (mesons
and baryons such as the proton and neutron) and even whole atoms were once
regarded as elementary particles.

66 Gribbon, John (2000). Q is for Quantum - An Encyclopedia of Particle Physics. Simon & Schuster.

ISBN 068485578X.
67 Clark, John, E.O. (2004). The Essential Dictionary of Science. Barnes & Noble. ISBN 0760746168.

Elementary particle
118

Elementary particles are grouped according to spin: particles normally associ-


ated with matter are fermions, having spin 1/2; they are divided into twelve
flavours. Particles associated with fundamental forces are bosons, having spin
1 (or 2 for gravity). 68

FT
• Fermions:

Quarks — up, down, strange, charm, bottom, top

Leptons — electron, muon, tau, electron neutrino, muon neutrino, tau


neutrino

• Bosons:

Gauge bosons – gluon, W and Z bosons, photon

Other bosons — Higgs boson, graviton


A
Standard Model
Main article: Standard Model

The Standard Model of particle physics contains 12 flavours of elementary


fermions ("matter particles"), plus their corresponding antiparticles, as well
as elementary bosons that mediate the forces and the still undiscovered Higgs
DR
boson. However, the Standard Model is widely considered to be a provisional
theory rather than a truly fundamental one, since it is fundamentally incom-
patible with Einstein’s general relativity. There are likely to be hypothetical el-
ementary particles not described by the Standard Model, such as the graviton,
the particle that would carry the gravitational force or the sparticles, supersym-
metric partners of the ordinary particles.

Fundamental fermions
Main article: fermion

The 12 fundamental fermionic flavours are divided into three generations of


four particles each. Six of the particles are quarks. The remaining six are

68 Veltman, Martinus (2003). Facts and Mysteries in Elementary Particle Physics. World Scientific. ISBN

981238149X.

Elementary particle
119

leptons, three of which are neutrinos, and the remaining three of which have
an electric charge of -1: the electron and its two cousins, the muon and the tau
lepton.

First generation Second generation Third generation

FT
• electron: e - • muon: µ - • tau lepton: τ -
• electron-neutrino: ν e • muon-neutrino: ν µ • tau-neutrino: ν τ
• up quark: u • charm quark: c • top quark: t
• down quark: d • strange quark: s • bottom quark: b

Table 1 Particle Generations

Antiparticles
Main article: antimatter

There are also 12 fundamental fermionic antiparticles which correspond to


these 12 particles. The positron e + corresponds to the electron and has an
electric charge of +1 and so on:
A First generation



positron: e +
Second generation


electron-antineutrino: ν̄e •
positive muon: µ +
muon-antineutrino: ν̄µ
Third generation



positive tau lepton: τ +
tau-antineutrino: ν̄τ
• up antiquark: ū • charm antiquark: c̄ • top antiquark: t̄
• down antiquark: d¯ • strange antiquark: s̄ • bottom antiquark: b̄

Table 2 Antiparticles

Quarks
Main article: quark
DR
Quarks and antiquarks have never been detected to be isolated, a fact explained
by confinement. Every quark carries one of three color charges of the strong in-
teraction; antiquarks similarly carry anticolor. Color charged particles interact
via gluon exchange in the same way that charged particles interact via photon
exchange. However, gluons are themselves color charged, resulting in an am-
plification of the strong force as color charged particles are separated. Unlike
the electromagnetic force which diminishes as charged particles separate, col-
or charged particles feel increasing force; effectively, they can never separate
from one another.
However, color charged particles may combine to form color neutral compos-
ite particles called hadrons. A quark may pair up to an antiquark: the quark
has a color and the antiquark has the corresponding anticolor. The color and

Elementary particle
120

anticolor cancel out, forming a color neutral meson. Or three quarks can ex-
ist together: one quark is "red", another "blue", another "green". These three
colored quarks together form a color neutral baryon. Or three antiquarks can
exist together: one antiquark is "antired", another "antiblue", another "anti-

FT
green". These three anticolored antiquarks form a color neutral antibaryon.
Quarks also carry fractional electric charges, but since they are confined within
hadrons whose charges are all integral, fractional charges have never been
isolated. Note that quarks have electric charges of either +2/3 or -1/3, whereas
antiquarks have corresponding electric charges of either -2/3 or +1/3.
Evidence for the existence of quarks comes from deep inelastic scattering: fir-
ing electrons at nuclei to determine the distribution of charge within nucleons
(which are baryons). If the charge is uniform, the electric field around the pro-
ton should be uniform and the electron should scatter elastically. Low-energy
electrons do scatter in this way, but above a particular energy, the protons de-
flect some electrons through large angles. The recoiling electron has much
less energy and a jet of particles is emitted. This inelastic scattering suggests
that the charge in the proton is not uniform but split among smaller charged
particles: quarks.
A
Fundamental bosons
Main article: boson

In the Standard Model, vector (spin-1) bosons (gluons, photons, and the W
and Z bosons) mediate forces, while the Higgs boson (spin-0) is responsible for
particles having intrinsic mass.
DR
Gluons
Main article: gluon

Gluons are the mediators of the strong interaction and carry both colour and
anticolour. Although gluons are massless, they are never observed in detectors
due to colour confinement; rather, they produce jets of hadrons, similar to sin-
gle quarks. The first evidence for gluons came from annihilations of electrons
and positrons at high energies which sometimes produced three jets — a quark,
an antiquark, and a gluon.

Electroweak bosons
Main article: W and Z bosons

Elementary particle
121

There are three weak gauge bosons: W +, W -, and Z 0; these mediate the weak
interaction. The massless photon mediates the electromagnetic interaction.

Higgs boson

FT
Main article: higgs boson

Although the weak and electromagnetic forces appear quite different to us at


everyday energies, the two forces are theorized to unify as a single electroweak
force at high energies. This prediction was clearly confirmed by measurements
of cross-sections for high-energy electron-proton scattering at the HERA col-
lider at DESY. The differences at low energies is a consequence of the high
masses of the W and Z bosons, which in turn are a consequence of the Higgs
mechanism. Through the process of spontaneous symmetry breaking, the Hig-
gs selects a special direction in electroweak space that causes three electroweak
particles to become very heavy (the weak bosons) and one to remain massless
(the photon). Although the Higgs mechanism has become an accepted part of
the Standard Model, the Higgs boson itself has not yet been observed in detec-
tors. Indirect evidence for the Higgs boson suggests its mass lies below about
A
200 GeV. In this case, the LHC experiments will be able to discover this last
missing piece of the Standard Model.

Beyond the Standard Model


Although all experimental evidence confirms the predictions of the Standard
Model, many physicists find this model to be unsatisfactory due to its many
undetermined parameters, many fundamental particles, the non-observation
DR
of the Higgs boson and other more theoretical considerations such as the hi-
erarchy problem. There are many speculative theories beyond the Standard
Model which attempt to rectify these deficiencies.

Grand unification
Main article: grand unification theory

One extension of the Standard Model attempts to combine the electroweak in-
teraction with the strong interaction into a single ’grand unified theory’ (GUT).
Such a force would be spontaneously broken into the three forces by a Higgs-
like mechanism. The most dramatic prediction of grand unification is the ex-
istence of X bosons, which cause proton decay. However, the non-observation
of proton decay at Super-Kamiokande rules out the simplest GUTs, including
SU(5) and SO(10).

Elementary particle
122

Supersymmetry
Main article: supersymmetry

Supersymmetry extends the Standard Model by adding an additional class of

FT
symmetries to the Lagrangian. These symmetries exchange fermionic particles
with bosonic ones. Such a symmetry predicts the existence of supersymmet-
ric particles, abbreviated as sparticles, which include the sleptons, squarks,
neutralinos and charginos. Each particle in the Standard Model would have a
superpartner whose spin differs by 1/2 from the ordinary particle. Due to the
breaking of supersymmetry, the sparticles are much heavier than their ordi-
nary counterparts; they are so heavy that existing particle colliders would not
be powerful enough to produce them. However, some physicists believe that
sparticles will be detected when the Large Hadron Collider at CERN begins
running.

String theory
Main article: string theory
A
According to string theorists, each kind of fundamental particle corresponds to
a different pattern of fundamental string. All strings are essentially the same,
although they may be open (lines) or closed (loops). Different particles differ
in the coordination of their strings. Modern string theories include supersym-
metry, making them superstring theories.
One particular prediction of string theory is the existence of extremely massive
counterparts of ordinary particles due to vibrational excitations of the funda-
DR
mental string. Another important prediction of string theory is the existence
of a massless spin-2 particle behaving like the graviton. By predicting gravity,
string theory unifies quantum mechanics with general relativity, making it the
first consistent theory of quantum gravity.
One problem with string theory is that it predicts that the number of dimen-
sions for spacetime much greater than 4 (the number of observed dimensions).
These extra dimensions are supposedly compactified or rolled-up. Other relat-
ed theories such as brane theories contain extended extra dimensions, which
are hidden from us by our confinement to a brane.

Preon theory
Main article: preon

Elementary particle
123

According to preon theory there are one or more orders of particles more fun-
damental than those (or most of those) found in the Standard Model. The most
fundamental of these are normally called preons, which is derived from "pre-
quarks". In essence, preon theory tries to do for the Standard Model what the

FT
Standard Model did for the particle zoo that came before it. Most models as-
sume that almost everything in the Standard Model can be explained in terms
of three to half a dozen more fundamental particles and the rules that govern
their interactions. Interest in preons has waned since the simplest models were
experimentally ruled out in the 1980’s.

See also
• →Subatomic particle
• Particle physics
• List of particles

References
A
• Brian Greene, The Elegant Universe, W.W.Norton & Company, 1999, ISBN
0-393-05858-1.

External links
• Greene, Brian, " Elementary particles 69". The Elegant Universe, NOVA (PBS)
• particleadventure.org: The Standard Model 70, * Unsolved Mysteries. Be-
yond The Standard Model 71, * What is the World Made of? The Naming of
DR
Quarks 72
• University of California: Particle Data Group 73
• particleadventure.org: Particle chart 74
• CERNCourier: Season of Higgs and melodrama 75
• Pentaquark information page 76

Source: http://en.wikipedia.org/wiki/Elementary_particle

69 http://www.pbs.org/wgbh/nova/elegant/part-flash.html
70 http://particleadventure.org/particleadventure/frameless/standard_model.html
71 http://particleadventure.org/particleadventure/frameless/beyond_start.html
72 http://particleadventure.org/particleadventure/frameless/quarknaming.html
73 http://pdg.lbl.gov/
74 http://particleadventure.org/particleadventure/frameless/chart.html
75 http://www.cerncourier.com/main/article/41/2/17
76 http://plato.phy.ohiou.edu/~hicks/thplus.htm

Elementary particle
124

Principal Authors: AugPi, Xerxes314, Glenn, Sadi Carnot, Reddi

Energy level splitting

FT
Energy level splitting occurs in physics when the degenerate energy levels of
two or more states are split because of external fields or other effects. The
term is most commonly used in quantum theory in reference to the electron
configuration in atoms or molecules.

Examples
• The →Zeeman effect - the splitting of electronic levels in an atom because
of an external magnetic field.
• The →Stark effect - splitting because of an external electric field.
• The Jahn-Teller effect - splitting of electronic levels in a molecule because
breaking the symmetry lowers the energy when the degenerate orbitals are
A
partially filled.

See also
• Energy level

Source: http://en.wikipedia.org/wiki/Energy_level_splitting
DR
Principal Authors: AnonUser, GangofOne, Donbert, Karol Langner, Itub

Entanglement witness

In quantum information theory, an entanglement witness is an object of geo-


metric nature which distinguishes an entangled state from separable ones.

Details
We first recall a few preliminary facts before giving the main result which shows
the existence of entanglement witnesses.
Let a composite quantum system have state space HA ⊗ HB . A mixed state ρ
is then a trace-class positive operator on the state space which has trace 1. We
can view the family of states as a subset of the real Banach space generated

Entanglement witness
125

by the Hermitian trace-class operators, with the trace norm. A mixed state ρ is
separable if it can be approximated, in the trace norm, by states of the form
ξ = ki=1 pi ρA B
P
i ⊗ ρi

FT
,where ρA B
i ’s and ρi ’s are pure states on the subsystems A and B respectively. So
the family of separable states is the closed convex hull of pure product states.
We will make use of the following variant of Hahn-Banach theorem:
Theorem Let S1 and S2 be convex closed sets in a real Banach space and one
of them is compact, then there exists a bounded functional f separating the
two sets.
This is a generalization of the fact that, in real Euclidean space, given a convex
set and a point outside, there always exists an affine subspace separating the
two. The affine subspace manifests itself as the functional f. In the present
context, the family of separable states is a convex set in the space of trace
class operators. If ρ is an entangled state (thus lying outside the convex set),
then by theorem above, there is a functional f separating ρ from the separable
states. It is this functional f, or its identification as an operator, that we call
A
an entanglement witness. There are more than one hyperplane separating a
closed convex set and a point lying outside of it. So for an entangled state there
are more than one entanglement witnesses. Recall the fact that the dual space
of the Banach space of trace-class operators is isomorphic to the set of bounded
operators. Therefore we can identify f with a Hermitian operator A. Therefore,
modulo a few details, we have shown the existence of an entanglement witness
given an entangled state:
Theorem For every entangled state ρ, there exists a Hermitian operator A such
DR
that T r(A ρ) < 0 and T r(Aσ) ≥ 0, for all separable state σ.
When both HA and HB have finite dimensions, there is no difference between
trace-class and Hilbert-Schmidt operators. So in that case A can be given by
Riesz representation theorem. As an immediate corollary, we have:
Theorem A mixed state σ is separable if and only if
T r(Aσ) ≥ 0

for any bounded operator A satisfying T r(A · P ⊗ Q) ≥ 0, for all product pure
state P ⊗ Q.
If a state is separable, clearly the desired implication from the theorem must
hold. On the other hand, given an entangled state, one of its entanglement
witnesses will violate the given condition.

Entanglement witness
126

Thus if a bounded functional f of the trace-class Banach space and f is positive


on the product pure states, then f, or its identification as a Hermitian operator,
is an entanglement witness. Such a f indicates the entanglement of some state.
Using the isomorphism between entanglement witnessnes and non-completlely

FT
positive maps, it was shown (by the Horodecki’s) that
Theorem A mixed state σ ∈ L(HA ) ⊗ L(HB ) is separable if for every positive
map from bounded operators on HB to bounded operators on HA , the op-
erator (IA ⊗ Λ)(σ) is positive, where IA is the identity map on L(HA ), the
bounded operators on HA .

References
• R.B. Holmes. Geometric Functional Analysis and Its Applications, Springer-
Verlag, 1975.

• M. Horodecki, P. Horodecki, R. Horodecki, Separability of Mixed States:


Necessary and Sufficient Conditions, Physics Letters A 210, 1996.
A
Source: http://en.wikipedia.org/wiki/Entanglement_witness

EP Quantum Mechanics

In physics, EP quantum mechanics is a theory of motion of point particles,


partly included in the framework of quantum trajectory representation the-
DR
ories of quantum mechanics, based upon an equivalence postulate similar
in content to the equivalence principle of general relativity, rather than on
the traditional Copenhagen interpretation of quantum mechanics. The equiv-
alence postulate states that all one-particle systems can be connected by a
non-degenerate coordinate transformation, more precisely by a map over the
cotangent bundle of the position manifold, so that there exists a quantum ac-
tion function S(q) transforms as a scalar field. Here, the action is defined as
dS(q) = pi (q)dq i

is the canonical one-form. This property is the heart of the EP formulation


of quantum mechanics. An immediate consequence of the EP is the removal
of the rest frame. The theory is based on symmetry properties of Schwarzian
derivative and on the quantum stationary Hamilton-Jacobi equation (QSHJE),
which is a partial differential equation for the quantum action function S(q),

EP Quantum Mechanics
127

the quantum version of the Hamilton-Jacobi equations differing from the clas-
sical one for the presence of a quantum potential term
~2
Q(q) = 4m {S(q), q}

FT
with {, } denoting the Schwarzian derivative. The QSHJE can be demonstrat-
ed to imply the →Schrödinger equation with square-summability of the wave
function, and thus quantization of energy, due to continuity conditions of the
quantum potential, without any assumption on the probabilistic interpretation
of the wave function. The theory, which is a work in progress, may or may
not include probabilistic interpretation as a consequence OR a hidden variable
description of trajectories.

References
• Alon E. Faraggi, M. Matone (2000) "The Equivalence Postulate of Quantum
Mechanics", International Journal of Modern Physics A, Volume 15, Issue
13, pp. 1869-2017. arXiv hep-th/9809127 77
A
• G. Bertoldi, Alon E. Faraggi, M. Matone (2000) "Equivalence principle,
higher dimensional Mobius group and the hidden antisymmetric tensor of
Quantum Mechanics", Class. Quantum Grav. 17 (2000) 3965–4005. arXiv
hep-th/9909201 78

Source: http://en.wikipedia.org/wiki/EP_Quantum_Mechanics

Principal Authors: Matteoeo, Pjacobi, Enochlau


DR
Excited state

In quantum mechanicsan excited state of a system (such as an atom, molecule


or nucleus) is any quantum state of the system that has a higher energy than
the ground state (that is, more energy than the absolute minimum).
The lifetime (see resonance) of a system in an excited state is usually short:
spontaneous or induced emission of a quantum of energy (such as a photon or
a phonon) usually occurs shortly after the system is promoted to the excited

77 http://arxiv.org/abs/hep-th/9809127
78 http://arxiv.org/abs/hep-th/9909201

Excited state
128

state, returning the system to a state with lower energy (a less excited state or
the ground state).

Example: the hydrogen atom

FT
A simple example of this concept comes by considering the hydrogen atom.
The ground state of the hydrogen atom corresponds to having the atom’s single
electron in the lowest possible orbit (that is, the spherically symmetric "1s"
wavefunction, which has the lowest possible quantum numbers). By giving
the atom additional energy (for example, by the absorption of a photon of an
appropriate energy), the electron is able to move into an excited state (one
with one or more quantum numbers greater than the minimum possible). If
the photon has too much energy, the electron will cease to be bound to the
atom, and the atom will become ionised.
Once the electron is in its excited state, we deem the hydrogen atom to be in its
excited state. The atom may return to a lower excited state, or the ground state,
by emitting a photon with a characteristic energy. Emission of photons from
atoms in various excited states leads to an electromagnetic spectrum showing
A
a series of characteristic emission lines (including, in the case of the hydrogen
atom, the Lyman series, the Balmer series, the Paschen series, the Brackett
series.)

External links
• Picture of a hydrogen atom changing from ground state to an excited state 79

See also
DR
• →Rydberg formula
• →Quantum state

Source: http://en.wikipedia.org/wiki/Excited_state

Principal Authors: Andrewwall, Bensaccount, ALoan, Vogon, Conscious

79 http://www.klimaforschung.net/kernreaktion/Orbital01.gif

Excited state
129

Exotic hadron

Exotic hadrons are subatomic particles made of quarks and bound by the

FT
strong interaction that are not predicted by the simple quark model. That is,
they do not have the same quark content as ordinary hadrons: exotic baryons
have more than just the three quarks of ordinary baryons, and exotic mesons
do not have one quark and one antiquark like ordinary mesons. Experimental
signatures for exotic hadrons have been seen recently but remain a topic of
controversy in particle physics.

Source: http://en.wikipedia.org/wiki/Exotic_hadron

Faddeev equations

The Faddeev equations are equations that describes, at once, all the possible
A
exchanges/interactions in a system of three particles in a fully quantum me-
chanical formulation. They can be solved iteratively with powerful computer
codes.
In general, Faddeev equations need as input a potential that describes the in-
teraction between two individual particles. It is also possible to introduce a
term in the equation in order to take also three-body forces into account.
The Faddeev equations are the most often used non-perturbative formulation of
DR
the quantum-mechanical three-body problem. Unlike the three body problem
in classical mechanics, the quantum three body problem is uniformly soluble.
In nuclear physics, the off the energy shell nucleon-nucleon interaction has
been studied by analyzing (n,2n) and (p,2p) reactions on deuterium targets,
using the Faddeev Equations. The nucleon-nucleon interaction is expanded
(approximated) as a series of separable potentials. The Coulomb interaction
between two protons is a special problem, in that its expansion in separable
potentials does not converge, but this is handled by matching the Faddeev
solutions to long range coulomb solutions, instead of to plain waves.
Separable potentials are interactions that do not preserve a particle’s location.
Ordinary local potentials can be expressed as sums of separable potentials.
The physical nucleon-nucleon interaction, which involves exchange of mesons,
is not expected to be either local or separable.

Faddeev equations
130

Source: http://en.wikipedia.org/wiki/Faddeev_equations

Principal Authors: David R. Ingham, Charles Matthews, Freakofnurture, Conscious, Philipum

FT
Fano resonance

In physics, a Fano resonance, in contrast with a Breit-Wigner resonance, is


a resonance for which the corresponding profile in the cross-section has the
so-called Fano shape, i.e. it can be fitted with a function proportional to:
(qΓres /2+E−Eres )2
(E−Eres )2 +(Γres /2)2
.

The Eres and Γres parameters are the standard Breit-Wigner parameters (po-
sition and width of the resonance, respectively). The q parameter is the so-
called Fano parameter. It is interpreted (within the Feshbach-Fano partioning
theory) as the ratio between the resonant and direct (background) scattering
probability. In the case the direct scattering probability is vanishing, the q pa-
rameter becomes infinite and the Fano formula is boiling down to the usual
A
Breit-Wigner (Lorentzian) formula:
1
(E−Eres )2 +(Γres /2)2 .

The classical reference is U. Fano, Phys. Rev. 124, 1866 (1961).

Source: http://en.wikipedia.org/wiki/Fano_resonance
DR
Principal Authors: Michael Hardy, Vipul

Fermi-Dirac statistics

In statistical mechanics, Fermi-Dirac statistics is a particular case of particle


statistics developed by Enrico Fermi and Paul Dirac that determines the sta-
tistical distribution of fermions over the energy states for a system in thermal
equilibrium. In other words, it is a probability of a given energy level to be
occupied by a fermion. Fermions are particles which are indistinguishable and
obey the Pauli exclusion principle, i.e., no more than one particle may occu-
py the same quantum state at the same time. Statistical thermodynamics is
used to describe the behaviour of large numbers of particles. A collection of
non-interacting fermions is called a →Fermi gas.

Fermi-Dirac statistics
131

A FT
Figure 13 Fermi-Dirac distribution as a function of /µ plotted for 4 different tem-
peratures. Occupancy transitions are smoother at higher temperatures.

F-D statistics was introduced in 1926 by Enrico Fermi and Paul Dirac and ap-
plied in 1927 by Arnold Sommerfeld to electrons in metals.
For F-D statistics, the expected number of particles in states with energy i is
DR
gi
ni =
e(i −µ)/kT +1

where:
ni is the number of particles in state i,

i is the energy of state i,

gi is the degeneracy of state i (the number of states with energy i ),

µ is the chemical potential (Sometimes the →Fermi energy EF is used


instead, as a low-temperature approximation),

Fermi-Dirac statistics
132

k is Boltzmann’s constant, and

T is absolute temperature.

FT
In the case where µ is the →Fermi energy EF and gi = 1 , the function is called
the Fermi function:
1
F (E) =
e(i −EF )/kT +1

A
DR

Figure 14 Fermi-Dirac distribution as a function of temperature. More states are


occupied at higher temperatures.

Which distribution to use


Fermi-Dirac and Bose-Einstein statistics apply when quantum effects have to
be taken into account and the particles are considered "indistinguishable". The
quantum effects appear if the concentration of particles (N/V) ≥ n q (where
n q is the quantum concentration). The quantum concentration is when the
interparticle distance is equal to the thermal de Broglie wavelength i.e. when
the wavefunctions of the particles are touching but not overlapping. As the
quantum concentration depends on temperature; high temperatures will put

Fermi-Dirac statistics
133

most systems in the classical limit unless they have a very high density e.g.
a White dwarf. Fermi-Dirac statistics apply to fermions (particles that obey
the Pauli exclusion principle), Bose-Einstein statistics apply to bosons. Both
Fermi-Dirac and Bose-Einstein become Maxwell-Boltzmann statistics at high

FT
temperatures or low concentrations.
Maxwell-Boltzmann statistics are often described as the statistics of "distin-
guishable" classical particles. In other words the configuration of particle A in
state 1 and particle B in state 2 is different from the case where particle B is
in state 1 and particle A is in state 2. When this idea is carried out fully, it
yields the proper (Boltzmann) distribution of particles in the energy states, but
yields non-physical results for the entropy, as embodied in →Gibbs paradox.
These problems disappear when it is realized that all particles are in fact in-
distinguishable. Both of these distributions approach the Maxwell-Boltzmann
distribution in the limit of high temperature and low density, without the need
for any ad hoc assumptions. Maxwell-Boltzmann statistics are particularly use-
ful for studying gases F-D statistics are most often used for the study of elec-
trons in solids. As such, they form the basis of semiconductor device theory
and electronics.
A
A derivation
Consider a single-particle state of a multiparticle system, whose energy is .
For example, if our system is some quantum gas in a box, then a state might
be a particular single-particle wave function. Recall that, for a grand canonical
ensemble in general, the grand partition function is
Z = s e−(E(s)−µN (s))/kT
P
DR
where
E(s) is the energy of a state s,

N (s) is the number of particles possessed by the system when in the state s,

µ denotes the chemical potential, and

s is an index that runs through all possible microstates of the system.

In the present context, we take our system to be a fixed single-particle state


(not a particle). So our system has energy n ·  when the state is occupied by
n particles, and 0 if it is unoccupied. Consider the balance of single-particle
states to be the reservoir. Since the system and the reservoir occupy the same

Fermi-Dirac statistics
134

A FT
Figure 15 Fermi-Dirac distribution as a function of . High energy states are less
probable. Or, low energy states are more probable.

physical space, there is clearly exchange of particles between the two (indeed,
this is the very phenomenon we are investigating). This is why we use the
grand partition function, which, via chemical potential, takes into considera-
DR
tion the flow of particles between a system and its thermal reservoir.
For fermions, a state can only be either occupied by a single particle or unoccu-
pied. Therefore our system has multiplicity two: occupied by one particle, or
unoccupied, called s1 and s2 respectively. We see that E(s1 ) = , N (s1 ) = 1,
and E(s2 ) = 0, N (s2 ) = 0. The partition function is therefore
Z = 2i=1 e−(E(si )−µN (si ))/kT = e−(−µ)/kT + 1.
P

For a grand canonical ensemble, probability of a system being in the microstate


sα is given by
e−(E(sα )−µN (sα )
P (sα ) = Z .

Our state being occupied by a particle means the system is in microstate s1 ,


whose probability is

Fermi-Dirac statistics
135

e−(E(s1 )−µN (s1 ))/kT e−(−µ)/kT 1


n̄ = P (s1 ) = Z = = .
e−(−µ)/kT +1 e(−µ)/kT +1

n̄ is called the Fermi-Dirac distribution. For a fixed temperature T, n̄() is the


probability that a state with energy  will be occupied by a fermion. Notice n̄ is

FT
a decreasing function in . This is consistent with our expectation that higher
energy states are less likely to be occupied.
Note that if the energy level  has degeneracy g , then we would make the
simple modification:
1
n̄ = g · .
e(−µ)/kT +1

This number is then the expected number of particles in the totality of the
states with energy .
For all temperature T, n̄(µ) = 21 , that is, the states whose energy is µ will
always have equal probability of being occupied or unoccupied.
In the limit T → 0, n̄ becomes a step function (see graph above). All states
whose energy is below the chemical potential will be occupied with probability
1 and those states with energy above µ will be unoccupied. The chemical
A
potential at zero temperature is called →Fermi energy, denoted by EF , i.e.
EF = µ(T = 0).
It may be of interest here to note that, in general the chemical potential is
temperature-dependent. However, for systems well below the Fermi tempera-
ture TF = EkF , it is often sufficient to use the approximation µ ≈ EF .

Another derivation
DR
In the previous derivation, we have made use of the grand partition function
(or Gibbs sum over states). Equivalently, the same result can be achieved by
directly analysing the multiplicities of the system.
Suppose there are two fermions placed in a system with four energy levels.
There are six possible arrangements of such a system, which are shown in the
diagram below.

1 2 3 4
A * *
B * *
C * *
D * *
E * *
F * *

Fermi-Dirac statistics
136

Each of these arrangements is called a microstate of the system. Assume that,


at thermal equilibrium, each of these microstates will be equally likely, subject
to the constraints that there be a fixed total energy and a fixed number of
particles.

FT
Depending on the values of the energy for each state, it may be that total
energy for some of these six combinations is the same as others. Indeed, if we
assume that the energies are multiples of some fixed value , the energies of
each of the microstates become:
A: 3

B: 4

C: 5

D: 5

E: 6
A
F: 7

So if we know that the system has an energy of 5, we can conclude that it
will be equally likely that it is in state C or state D. Note that if the particles
were distinguishable (the classical case), there would be twelve microstates
altogether, rather than six.
DR
Now suppose we have a number of energy levels, labelled by index i , each
level having energy  i and containing a total of n i particles. Suppose each level
contains g i distinct sublevels, all of which have the same energy, and which
are distinguishable. For example, two particles may have different momenta, in
which case they are distinguishable from each other, yet they can still have the
same energy. The value of g i associated with level i is called the "degeneracy"
of that energy level. The Pauli exclusion principle states that only one fermion
can occupy any such sublevel.
Let w(n, g) be the number of ways of distributing n particles among the g
sublevels of an energy level. Its clear that there are g ways of putting one
particle into a level with g sublevels, so that w(1, g) = g which we will write
as:
g!
w(1, g) = 1!(g−1)!

Fermi-Dirac statistics
137

We can distribute 2 particles in g sublevels by putting one in the first sublev-


el and then distributing the remaining n - 1 particles in the remaining g - 1
sublevels, or we could put one in the second sublevel and then distribute the
remaining n - 1 particles in the remaining g - 2 sublevels, etc. so that w’(2, g)

FT
= w(1, g - 1) + w(1,g - 2) + ... + w(1, 1) or
Pg−1 (g−k)!
w(1, g − k) = g−1 g!
P
w(2, g) = k=1 k=1 1!(g−k−1)! = 2!(g−2)!

where we have used the following theorem involving binomial coefficients:


Pg k! (k+1)!
k=n n!(k−n)! = (n+1)!(k−n)!

Continuing this process, we can see that w(n, g) is just a binomial coefficient
g!
w(n, g) = n!(g−n)!

The number of ways that a set of occupation numbers n i can be realized is the
product of the ways that each individual energy level can be populated:
W = i w(ni , gi ) = i n !(ggi−n
!
Q Q
i i)! i
A
Following the same procedure used in deriving the Maxwell-Boltzmann distri-
bution, we wish to find the set of n i for which W is maximised, subject to the
constraint that there be a fixed number of particles, and a fixed energy. We
constrain our solution using Lagrange multipliers forming the function:
P P
f (ni ) = ln(W ) + α(N − ni ) + β(E − ni i )

Again, using Stirling’s approximation for the factorials and taking the deriva-
DR
tive with respect to n i, and setting the result to zero and solving for n i yields
the Fermi-Dirac population numbers:
gi
ni = eα+βi +1

It can be shown thermodynamically that β = 1/kT where k is Boltzmann’s


constant and T is the temperature, and that α = -µ/kT where µ is the chemical
potential, so that finally:
gi
ni =
e(i −µ)/kT +1

Note that the above formula is sometimes written:


gi
ni =
ei /kT /z+1

where z = exp(µ/kT ) is the absolute activity.

Fermi-Dirac statistics
138

See also
• →Maxwell-Boltzmann statistics
• →Bose-Einstein statistics

FT
• Parastatistics

Source: http://en.wikipedia.org/wiki/Fermi-Dirac_statistics

Principal Authors: Mct mht, PAR, Unc.hbar, Fresheneesz, Tim Starling

Fermi energy

In physics and →Fermi-Dirac statistics, the Fermi energy (E F) of a system of


non-interacting fermions is the smallest possible increase in the ground state
energy when exactly one particle is added to the system. It is equivalent to the
chemical potential of the system in its ground state at absolute zero. It can also
A
be interpreted as the maximum energy of an individual fermion in this ground
state. The Fermi energy is one of the central concepts of condensed matter
physics.

Fermi level
The Fermi level is the top of the collection of electron energy levels at abso-
lute zero temperature. Since fermions cannot exist in identical energy states
(see the exclusion principle), at absolute zero, electrons pack into the lowest
DR
available energy states and build up a "Fermi sea" of electron energy states. 80
In this state (at 0 K), the average energy of an electron is given by:
Eav = 35 Ef

where Ef is the fermi energy.


The Fermi momentum is the momentum of fermions at the Fermi surface. The
Fermi momentum is given by:
p
pF = 2me Ef

where me is the mass of the electron.

80 http://hyperphysics.phy-astr.gsu.edu/hbase/solids/fermi.html

Fermi energy
139

This concept is usually applied in the case of dispersion relations between the
energy and momentum that do not depend on the direction. In more general
cases, one must consider the Fermi energy.
The Fermi velocity is the average velocity of an electron in an atom at absolute

FT
zero. This average velocity corresponds to the average energy given above. The
Fermi velocity is defined by:
q
2Ef
Vf = me

where me is the mass of the electron.


Below the Fermi temperature, a substance gradually expresses more and more
quantum effects of cooling. The Fermi temperature is defined by:
Ef
T.f = k

where k is the Boltzmann constant.

Quantum mechanics
A
According to quantum mechanics, fermions – particles with a half-integer spin,
usually 1/2, such as electrons – follow the Pauli exclusion principle, which
states that multiple particles may not occupy the same quantum state. Con-
sequently, fermions obey →Fermi-Dirac statistics. The ground state of a non-
interacting fermion system is constructed by starting with an empty system
and adding particles one at a time, consecutively filling up the lowest-energy
unoccupied quantum states. When the desired number of particles has been
reached, the Fermi energy is the energy of the highest occupied molecular or-
DR
bital (HOMO). Within conductive materials, this is equivalent to the lowest
unoccupied molecular orbital (LUMO), however within other materials there
will be a significant gap between the HOMO and LUMO on the order of 2-3 eV.
This gap does exist in conductors, however it is infinitesimally small.

Free electron gas


In the free electron gas, the quantum mechanical version of an ideal gas of
fermions, the quantum states can be labelled according to their momentum.
Something similar can be done for periodic systems, such as electrons moving
in the atomic lattice of a metal, using something called the "quasi-momentum"
(see Bloch wave). In either case, the Fermi energy states reside on a surface in
momentum space known as the Fermi surface. For the free electron gas, the
Fermi surface is the surface of a sphere; for periodic systems, it generally has a
contorted shape (see Brillouin zones). The volume enclosed by the Fermi sur-
face defines the number of electrons in the system, and the topology is directly

Fermi energy
140

related to the transport properties of metals, such as electrical conductivity.


The study of the Fermi surface is sometimes called Fermiology. The Fermi
surfaces of most metals are well studied both theoretically and experimentally.
The Fermi energy of the free electron gas is related to the chemical potential

FT
by the equation
  2  4 
2
π 4 kT
µ = EF 1 − π12 EkT
F
+ 80 E F
+ · · ·

where E F is the Fermi energy, k is the Boltzmann constant and T is tempera-


ture. Hence, the chemical potential is approximately equal to the Fermi energy
at temperatures of much less than the characteristic Fermi temperature E F/k.
The characteristic temperature is on the order of 10 5 K for a metal, hence at
room temperature (300 K), the Fermi energy and chemical potential are essen-
tially equivalent. This is significant since it is the chemical potential, not the
Fermi energy, which appears in →Fermi-Dirac statistics.

See also
A


fermi gas
semiconductors
• electrical engineering
• electronics
• thermodynamics

References
DR
• Table of fermi energies, velocities, and temperatures for various elements 81.
• a discussion of fermi gases and fermi temperatures 82.

Source: http://en.wikipedia.org/wiki/Fermi_energy

Principal Authors: Fresheneesz, Tim Starling, The Anome, CYD, Tantalate

81 http://hyperphysics.phy-astr.gsu.edu/hbase/tables/fermi.html
82 http://physicsweb.org/articles/world/15/4/7

Fermi energy
141

Fermi gas

A Fermi gas, or Free electron gas, is a collection of non-interacting fermions.

FT
It is the quantum mechanical version of an ideal gas, for the case of fermionic
particles. Electrons in metals and semiconductors and neutrons in a neutron
star can be approximately considered Fermi gases. The energy distribution
of the fermions in a Fermi gas in thermal equilibrium is determined by their
density, the temperature and the set of available energy states, via →Fermi-
Dirac statistics. By the Pauli principle, no quantum state can be occupied by
more than one fermion, so the total energy of the Fermi gas at zero temperature
is larger than the product of the number of particles and the single-particle
ground state energy. For this reason, the pressure of a Fermi gas is nonzero
even at zero temperature, in contrast to that of a classical ideal gas. This so-
called degeneracy pressure stabilizes a neutron star (a Fermi gas of neutrons)
or a White Dwarf star (a Fermi gas of electrons) against the inward pull of
gravity.
It is possible to define a Fermi temperature below which the gas can be consid-
A
ered degenerate. This temperature depends on the mass of the fermions and
the energy density of states. For metals, the electron gas’s Fermi temperature
is generally many thousands of kelvins, so they can be considered degenerate.
The maximum energy of the fermions at zero temperature is called the →Fermi
energy. The Fermi energy surface in momentum space is known as the Fermi
surface.
Since interactions are neglected by definition, the problem of treating the equi-
librium properties and dynamical behaviour of a Fermi gas reduces to the study
DR
of the behaviour of single independent particles. As such, it is still relatively
tractable and forms the starting point for more advanced theories (such as
Fermi liquid theory or perturbation theory in the interaction) which take into
account interactions to some degree of accuracy.

See also
• Gas in a box
• →Bose gas

Source: http://en.wikipedia.org/wiki/Fermi_gas

Principal Authors: PAR, SimonP, Tantalate, Schneelocke, Tom davis

Fermi gas
142

Fermi liquid

A Fermi liquid is a generic term for a quantum mechanical liquid of fermions

FT
that arises under certain physical conditions—when the temperature is suffi-
ciently low, and when the system is translationally invariant. The interaction
between the particles of the many-body system does not need to be small (see
e.g. electrons in a metal). The phenomenological theory of Fermi liquids,
which was introduced by the Russian physicist Lev Davidovich Landau in 1956,
explains why some of the properties of an interacting fermion system are very
similar to those of the →Fermi gas (i.e. non-interacting fermions), and why
other properties differ.
Liquid He-3 is a Fermi liquid at low temperatures (but not low enough to be in
its superfluid phase.) He-3 is an isotope of Helium, with 2 protons, 1 neutron
and 2 electrons per atom; because there is an odd number of fermions inside
the atom, the atom itself is also a fermion. The electrons in a normal (non-
superconducting) metal also form a Fermi liquid.
A
The Fermi liquid is qualitatively analogous to the non-interacting Fermi gas, in
the following sense: The system’s dynamics and thermodynamics at low ex-
citation energies and temperatures may be described by substituting the non-
interacting fermions with so-called quasiparticles, each of which carries the
same spin, charge and momentum as the original particles. Physically these
may be thought of as being particles whose motion is disturbed by the sur-
rounding particles and which themselves perturb the particles in their vicinity.
Each many-particle excited state of the interacting system may be described by
listing all occupied momentum states, just as in the non-interacting system. As
DR
a consequence, quantities such as the heat capacity of the Fermi liquid behave
qualitatively in the same way as in the Fermi gas (e.g. the heat capacity rises
linearly with temperature).
However, the following differences to the non-interacting Fermi gas arise:

• The energy of a many-particle state is not simply a sum of the single-particle


energies of all occupied states. Instead, the change in energy for a giv-
en change δnk in occupation of states k contains terms both linear and
quadratic in δnk (for the Fermi gas, it would only be linear, δnk k , where k
denotes the single-particle energies). The linear contribution corresponds
to renormalized single-particle energies, which involve, e.g., a change in
the effective mass of particles. The quadratic terms correspond to a sort of
"mean-field" interaction between quasiparticles, which is parameterized by
so-called Landau Fermi liquid parameters and determines the behaviour of

Fermi liquid
143

density oscillations (and spin-density oscillations) in the Fermi liquid. Still,


these mean-field interactions do not lead to a scattering of quasi-particles
with a transfer of particles between different momentum states.

• Specific heat, compressibility, spin-susceptibility and other quantities show

FT
the same qualitative behaviour (e.g. dependence on temperature) as in the
Fermi gas, but the magnitude is (sometimes strongly) changed.

• In addition to the mean-field interactions, some weak interactions between


quasiparticles remain, which lead to scattering of quasiparticles off each
other. Therefore, quasiparticles acquire a finite lifetime. However, at low
enough energies above the Fermi surface, this lifetime becomes very long,
such that the product of excitation energy (expressed in frequency) and
lifetime is much larger than one. In this sense, the quasiparticle energy is
still well-defined (in the opposite limit, Heisenberg’s uncertainty relation
would prevent an accurate definition of the energy).

• Green’s function and momentum distribution of quasiparticles behave as


for the fermions in the Fermi gas (apart from the broadening of the delta


A peak in the Green’s function by the finite lifetime).

The structure of the "bare" particles (as opposed to quasiparticle) Green’s


function is similar to that in the Fermi gas (where, for a given momen-
tum, the Green’s function in frequency space is a delta peak at the re-
spective single-particle energy). The delta peak in the density-of-states
is broadened (with a width given by the quasiparticle lifetime). In ad-
dition (and in contrast to the quasiparticle Green’s function), its weight
(integral over frequency) is suppressed by a quasiparticle weight factor
DR
0 < Z < 1. The remainder of the total weight is in a broad "incoherent
background", corresponding to the strong effects of interactions on the
fermions at short time-scales.

• The distribution of particles (as opposed to quasiparticles) over momentum


states at zero temperature still shows a discontinuous jump at the Fermi
surface (as in the Fermi gas), but it does not drop from 1 to 0: the step is
only of size Z.

• In a metal the resistance at low temperatures is dominated by electron-


electron scattering in combination with Umklapp scattering. For a Fermi
liquid, the resistance from this mechanism varies as T 2 , which is often taken
as an experimental check for Fermi liquid behaviour (in addition to the
linear temperature-dependence of the specific heat), although it only arises
in combination with the lattice.

Fermi liquid
144

Source: http://en.wikipedia.org/wiki/Fermi_liquid

Principal Authors: CYD, Dschwen, RedWolf, Itai, Jofox

FT
Fermi’s golden rule

In quantum physics, Fermi’s golden rule is a way to calculate the transition


rate between two eigenstates of a quantum system using time-dependent per-
turbation theory, which means it’s an approximation.
We consider the system to begin in an eigenstate |ii of a given Hamiltonian H0 .
We consider the effect of a time-independent perturbing Hamiltonian H 0 .
The one-to-many transition probability per unit of time from the state |ii to a
set of states |f i is given, to first order in the perturbation, by:
Ti→f = 2π
~ |hf |H 0 |ii|2 ρ
A
where ρ is the density of final states, and < f | H’ | i > is the matrix element (in
bra-ket notation) of the perturbation, H’, between the final and initial states.
Fermi’s golden rule is valid when H 0 is time-independent, |ii is an eigenstate of
the unperturbed Hamiltonian, the states |f i form a continuum, and the initial
state has not been significantly depleted (eg, by scattering into the final states).
The most common way to derive the equation is to start with time-dependent
perturbation theory and to take the limit for absorption under the assumption
that the time of the measurement is much larger than the time needed for the
DR
transition.
Although named after Fermi, most of the work leading to the Golden Rule was
done by Dirac.

External links
• More information on Fermi’s golden rule 83
• Derivation using time-dependent perturbation theory 84

Source: http://en.wikipedia.org/wiki/Fermi%27s_golden_rule

83 http://hyperphysics.phy-astr.gsu.edu/hbase/quantum/fermi.html
84 http://www.ph.utexas.edu/~schwitte/PHY362L/QMnote.pdf

Fermi’s golden rule


145

Principal Authors: Donvinzk, Laurascudder, Jheald, MikeMorley, Alfredo.correa

Field emission

FT
Also known as Fowler-Nordheim tunneling, field emission is a form of quan-
tum tunneling in which electrons pass through a barrier in the presence of a
high electric field. This phenomenon is highly dependent on both the prop-
erties of the material and the shape of the particular cathode, so that higher
aspect ratios produce higher field emission currents. The current density pro-
duced by a given electric field is governed by the Fowler-Nordheim equation.
Applications of field emission include its use as an electron source in flash
memory, electron microscopy, MEMS systems, and field emission displays.
In the field of vacuum electronics, field emission is seen as an alternative to
thermionic emission, with advantages such as dramatically higher efficiency,
less scatter of emitted electrons, faster turn-on times, compactness, and, in
many cases, redundancy. Some disadvantages include lower current per emis-
A
sion source and, often, lower overall current density. Field emission limits the
maximum operating voltage for high voltage vacuum devices such as vacuum
capacitors and vacuum switches.
Vacuum tubes based on thermionic emission require several minutes to warm
up before they can be used; by contrast, the function of field emission devices
is effectively instantaneous, allowing switching times of many megahertz. The
ability to modulate the electron source, rather than modifying a stream of elec-
trons from a constant source (i.e., by velocity modulation), has allowed many
DR
vacuum devices to be greatly simplified. For instance, the Klystrode functions
much like the two-chamber Klystron, without the need for a first chamber.

See also
• Cold cathode

External links
• Field emission - Fowler-Nordheim tunneling, Principles of Semiconductor
Devices, Bart Van Zeghbroeck, 1997 85

85 http://ece-www.colorado.edu/~bart/book/msfield.htm

Field emission
146

Source: http://en.wikipedia.org/wiki/Field_emission

Principal Authors: Ajdecon, Bert Hickman, Pjacobi, Heron, Adoarns

FT
Finite potential well

The finite potential well (also known as the finite square well) is a simple
problem from quantum mechanics. It is an extension of the infinite potential
well, in which a particle is confined to a box, but one which has finite - not
infinite - potential walls. This means unlike the infinite potential well, there
is a probability associated with the particle being found outside of the box.
The quantum mechanical interpretation is unlike the classical interpretation,
where if the total energy of the particle is less than potential energy barrier of
the walls it cannot be found outside the box. In the quantum interpretation,
there is a probability of the particle being outside the box even when the energy
of the particle is less than the potential energy barrier of the walls.
A
The particle in a 1-dimensional box
For the 1-dimensional case on the x-axis, the time-independent Schrödinger
equation can be written as:

~2 d2 ψ + V (x)ψ = Eψ
− 2m (1) where
dx 2
h
~ = 2π
h is Planck’s constant
m is the mass of the particle
ψ is the (complex valued) wavefunction that we want to find
DR
V (x) is a function describing the potential at each point x and
E is the energy, a real number.

For the case of the particle in a 1-dimensional box of length L, the potential
is zero inside the box, but rises abruptly to a value Γ at x = -L/2 and x =
L/2. The wavefunction is considered to be made up of different wavefuctions
at different ranges of x, depending on whether x is inside or outside of the box.
Therefore
 the wavefunction is defined such that:
ψ1 , if x < −L/2 (the region outside the box)

ψ = ψ2 , if − L/2 < x < L/2 (the region inside the box)

ψ3 if x > L/2 (the region outside the box)

Finite potential well


147

Inside the box


For the region inside the box V (x) = 0 and Equation 1 reduces to:
2 2
~ d ψ2
− 2m dx2
= Eψ2

FT
Letting

2mE
k= ~

the equation becomes


d2 ψ2
dx2
= −k 2 ψ2

This is a well studied differential equation and eigenvalue problem with a gen-
eral solution of:
ψ2 = A sin(kx) + B cos(kx)

Hence:
k 2 ~2
A E= 2m

Here, A and B can be any complex numbers, and k can be any real number (k
must be real because E is real).

Outside the box


For the region outside of the box, V (x) = Γ and Equation 1 becomes:
2 2
~ d ψ1
− 2m dx2
= (E − Γ)ψ1
DR
There are two possible families of solutions, depending on whether E is less
than Γ (the particle is bound in the potential) or E is greater than Γ (the
particle is free).
For a free particle, E > Γ, and letting

2m(E−Γ)
κ= ~

produces
d2 ψ1
dx2
= −κ2 ψ1

with the same solution form as the inside-well case:


ψ1 = C sin(κx) + D cos(κx)

Finite potential well


148

This analysis will first focus on the bound state, where Γ > E. Letting

2m(Γ−E)
α= ~

produces

FT
d2 ψ1
dx2
= α 2 ψ1

where the general solution is exponential:


ψ1 = F e−αx + Geαx

Similarly, for the other region outside the box:


ψ3 = He−αx + Ieαx

Now in order find the specific solution for the problem at hand, we must specify
the appropriate boundary conditions and find the values for A , B , F , G , H
and I that satisfy those conditions.
A
Finding wavefunctions for the bound state
Solutions to the Schrödinger equation must be continuous, and continuously
differentiable. These requirements are boundary conditions on the differential
equations previously derived.
In this case, the finite potential well is symmetrical, so symmetry can be ex-
ploited to reduce the necessary calculations.
Summarizing
 the previous section:
DR
ψ1 , if x < −L/2 (the region outside the box)

ψ = ψ2 , if − L/2 < x < L/2 (the region inside the box)

ψ3 if x > L/2 (the region outside the box)

where we found ψ1 , ψ2 and ψ3 to be:


ψ1 = F e−αx + Geαx

ψ2 = A sin(kx) + B cos(kx)

ψ3 = He−αx + Ieαx

Finite potential well


149

We see that as x goes to −∞, the F term goes to infinity. Likewise, as x goes to
+∞, the I term goes to infinity. As the wave function must be finite for all x,
this means we must set F = I = 0, and we have:
ψ1 = Geαx and ψ3 = He−αx

FT
Next, we know that the overall ψ function must be continuous and differen-
tiable. In other words the values of the functions and their derivatives must
match up at the dividing points:

See also
• →Potential well
• Infinite potential well
• →Quantum tunnelling

References
• Griffiths, David J. (2005). Introduction to Quantum Mechanics, 2 nd ed.,
Prentice Hall. ISBN 0131118927.
A
Source: http://en.wikipedia.org/wiki/Finite_potential_well

Flux quantization
DR
Flux quantization is a quantum phenomenon in which the magnetic field is
quantized in the unit of h/2e, also known variously as flux quanta, fluxoids,
vortices or fluxons.
Flux quantization occurs in Type II superconductors subjected to a magnetic
field. Below a critical field H c1, all magnetic flux is expulsed according to the
Meissner effect and perfect diamagnetism is observed, exactly as in a Type I
superconductor. Up to a second critical field value, H c2, flux penetrates in
discrete units while the bulk of the material remains superconducting. Both
critical fields are temperature dependent, and tabulated values are the zero-
temperature extrapolation unless otherwise noted.

Flux quantization
150

See also
• Flux pinning
• Magnetic flux quantum

FT
Source: http://en.wikipedia.org/wiki/Flux_quantization

Principal Authors: Charles Matthews, Stevelihn, Eldereft, Art LaPella, Zowie

Fock matrix

In quantum mechanics, the Fock matrix is a matrix approximating the single-


electron energy operator of a given quantum system in a given set of basis
vectors.
It is most often formed in computational chemistry when attempting to solve
the Roothaan equations for an atomic or molecular system. The Fock matrix
A
is actually an approximation to the true Hamiltonian operator of the quantum
system. It includes the effects of electron-electron repulsion only in an average
way. Importantly, because the Fock operator is a one-electron operator, it does
not include the electron correlation energy.
The Fock matrix is defined by the Fock operator. For the restricted case which
assumes closed-shell orbitals and single-determinantal wavefunctions, the Fock
operator for the first electron is given by:
F̂ (1) = Ĥ core (1) + n [2Jˆj (1) − K̂j (1)]
DR
P
j=1

where:
F̂ (i)

is the Fock operator for the i -th electron in the system,


Ĥ core (i)

is the core Hamiltonian for the i -th electron,


n

is the total number of orbitals in the system (equal to half the number of elec-
trons),

Fock matrix
151

Jˆj (i)

is the Coulomb operator, defining the repulsive force between the j -th and i -th
electrons in the system,

FT
K̂j (i)

is the exchange operator, defining the effect of exchanging the two electrons.

See also
• Roothaan equations
• Hartree-Fock

Source: http://en.wikipedia.org/wiki/Fock_matrix

Principal Authors: Edsanville, Agentsoo, Charles Matthews, Remuel, Njerseyguy


A
Fock space

The Fock space is an algebraic system (→Hilbert space) used in quantum me-
chanics to describe quantum states with a variable or unknown number of
particles. It is named for V. A. Fock.
Technically, the Fock space is the Hilbert space made from the direct sum of
DR
tensor products of single-particle Hilbert spaces:
Fν (H) = ∞ ⊗n
L
n=0 Sν H

where S ν is the operator which symmetrizes or antisymmetrizes the space,


depending on whether the Hilbert space describes particles obeying bosonic (ν
= +) or fermionic (ν = -) statistics respectively. H is the single particle Hilbert
space. It describes the quantum states for a single particle, and to describe the
quantum states of systems with n particles, or superpositions of such states,
one must use a larger Hilbert space, the Fock space, which contains states for
unlimited and variable number of particles. →Fock states are the natural basis
of this space. (See also the →Slater determinant.)

Fock space
152

Example
An example of a state of the Fock space is
|Ψiν = |φ1 , φ2 , · · · , φn iν

FT
describing n particles, one of which has wavefunction φ 1, another φ 2 and so on
up to the n th particle, where each φ i is any wavefunction from the single parti-
cle Hilbert space H. When we speak of one particle in state φ i it must be borne
in mind that in quantum mechanics identical particles are indistinguishable,
and in a same Fock space all particles are identical (to describe many species
of particles, make the tensor products of as many different Fock spaces). It
is one of the most powerful features of this formalism that states are intrin-
sically properly symmetrized. So that for instance, if the above state |Ψ> - is
fermionic, it will be 0 if two (or more) of the φ i are equal, because by the Pauli
exclusion principle no two (or more) fermions can be in the same quantum
state. Also, the states are properly normalized, by construction.
A useful and convenient basis for this space is the occupancy number basis. If
|ψ i> is a basis of H, then we can agree to denote the state with n 0 particles in
A
state |ψ 0>, n 1 particles in state |ψ 1>, ..., n k particles in state |ψ k> by
|n0 , n1 , · · · , nk iν ,

with each n i taking the value 0 or 1 for fermionic particles and 0,1,2,... for
bosonic particles.
Such a state is called a →Fock state. Since |ψ i> are understood as the steady
states of the free field, i.e., a definite number of particles, a Fock state describes
DR
an assembly of non-interacting particles in definite numbers. The most general
pure state is the linear superposition of Fock states.
Two operators of paramount importance are the creation and annihilation op-
erators, which upon acting on a Fock state respectively remove and add a par-
ticle, in the ascribed quantum state. They are denoted a† (φi ) and a(φi ) re-
spectively, with φi referring to the quantum state |φi i in which the particle is
removed or added. It is often convenient to work with states of the basis of H
so that these operators remove and add exactly one particle in the given state.
These operators also serve as a basis for more general operators acting on the
Fock space, for instance the number operator giving the number of particles in
a specific state |φi i is a† (φi )a(φi ).

Source: http://en.wikipedia.org/wiki/Fock_space

Principal Authors: Phys, Stupidmoron, Laussy, Maury Markowitz, Charles Matthews

Fock space
153

Fock state

A Fock state, in quantum mechanics, is any state of the →Fock space with a

FT
well-defined number of particles in each state. The name is for V. A. Fock.
If we limit to a single mode for simplicity (doing so we formally describe a mere
harmonic oscillator), a Fock state is of the type |n> with n an integer value.
This means that there are n quanta of excitation in the mode. |0> corresponds
to the ground state (no excitation). It is different from 0 which is the null
vector.
Fock states form the most convenient basis of the Fock space. They are defined
to obey the following relations in the bosonic algebra:

a† |ni = n + 1|n + 1i


a|ni = n|n − 1i

|ni = √1 (a† )n |0i


A n!

with a (resp. a †) the annihilation (resp. creation) bose operator. Similar


relations hold for fermionic algebra.
This allows to check that <a †a>=n and Var(a †a)=0, i.e., that measuring the
number of particles a †a in a Fock state returns always a definite value with no
fluctuation.
DR
Source: http://en.wikipedia.org/wiki/Fock_state

Principal Authors: Digitalme, Phys, Unyoyega, Charles Matthews, Laussy

Franck-Hertz experiment

In physics, the Franck-Hertz experiment was an early physics experiment that


provided support for the Bohr model of the atom, a precursor to quantum me-
chanics. In 1914, physicists James Franck and Gustav Ludwig Hertz sought to
experimentally probe the energy levels of the atom. The now-famous Franck-
Hertz experiment elegantly supported Niels Bohr’s model of the atom, with
electrons orbiting the nucleus with specific, discrete energies. Franck and Hertz
were awarded the Nobel Prize in Physics in 1925 for this work.

Franck-Hertz experiment
154

The experiment
The classic experiment involved a tube containing low pressure gas and bound-
ed at each end by an electrode, and containing a mesh accelerating grid near
the ground electrode. (This ’ground’ was actually held very slightly negative, so

FT
that electrons had to have a small amount of kinetic energy to reach it.) Instru-
ments were fitted to measure the current passing between the electrodes, and
to adjust the potential difference (the voltage) between the cathode (negative
electrode) and the accelerating grid.

• At low voltages–up to 7 volts when the tube contained mercury vapour–


the current through the tube increased steadily with increasing potential
difference. The higher voltage increased the electric field in the tube and
electrons were drawn more forcefully towards and through the accelerating
grid.
• At 7 volts the current drops sharply, almost back to zero.
• The current increases steadily once again if the voltage is increased further,
until 11.9 volts is reached (exactly 7+4.9 volts).
• At 11.9 volts a similar sharp drop is observed.

AThis series of dips in current at 4.9 volt increments will visibly continue to
potentials of at least 100 volts.

Franck and Hertz were able to explain their experiment in terms of elastic
and inelastic collisions. At low potentials, electrons acquired only a modest
amount of kinetic energy. When they encountered mercury atoms in the tube,
they participated in elastic collisions. The total amount of kinetic energy in
the system remained the same. (Since electrons are significantly less massive
DR
than mercury atoms, this meant that the electrons held on to the vast majority
of that energy, too.) Higher potentials served to drive more electrons to the
ground and increase the observed current.
The lowest energy electronic excitation a mercury atom can participate in re-
quires 4.9 electron volts (eV). When the accelerating potential reached 4.9
volts, each electron possessed exactly that amount of energy when it reached
the anode grid. Consequently, a collision between a mercury atom and an elec-
tron at that point could be inelastic. Its kinetic energy could be converted into
potential energy, and used to excite the mercury atom. With the loss of all
of its kinetic energy, the electron can no longer overcome the slight negative
potential at the ground electrode, and the measured current drops sharply.
As the voltage is increased, electrons will participate in one inelastic collision,
lose their 4.9 eV, but then continue to be accelerated. In this manner, the
current rises again after the accelerating potential exceeds 4.9 V. At 9.8 V, the

Franck-Hertz experiment
155

situation changes again. There, each electron now has just enough energy to
participate in two inelastic collisions, excite two mercury atoms, and then be
left with no kinetic energy. Once again, the observed current drops. At intervals
of 4.9 volts this process will repeat; each time the electrons will undergo one

FT
additional inelastic collision.
A similar pattern is observed with neon gas, but at intervals of approximately
19 volts. The process is identical, just with a much different threshold. One
additional difference is that a glow will appear near the accelerating grid at 19
volts–one of the transitions of relaxing neon atoms emits red-orange light. This
glow will move closer to the cathode with increasing accelerating potential, to
whatever point in the tube the electrons acquire the 19 eV required to excite
a neon atom. At 38 volts two distinct glows will be visible: one between the
cathode and grid, and one right at the accelerating grid. Higher potentials will
result in additional glowing regions in the tube, spaced at 19 volt intervals.
The Franck-Hertz experiment confirmed Bohr’s quantized model of the atom
by demonstrating that atoms could indeed only absorb (and be excited by)
specific amounts of energy (quanta).
A
References
• The Franck-Hertz Experiment at Hyperphysics 86
• Up-to-date literature on the Franck-Hertz Experiment 87

Source: http://en.wikipedia.org/wiki/Franck-Hertz_experiment
DR
Principal Authors: Icairns, TenOfAllTrades, Mac Davis, Linas, Pnicolet

Free particle

In physics, a free particle is a particle that, in some sense, is not bound. In the
classical case, this is represented with the particle not being influenced by any
external force.

86 http://hyperphysics.phy-astr.gsu.edu/hbase/FrHz.html
87 http://users.skynet.be/P.Nicoletopoulos/references.html

Free particle
156

Classical Free Particle


The classical free particle is characterized simply by a fixed velocity. The mo-
mentum is given by

FT
p = mv

and the energy by


E = 12 mv 2

where m is the mass of the particle and v is the vector velocity of the particle.

Non-Relativistic Quantum Free Particle


The →Schrödinger equation for a free particle is:
2

~
− 2m ∇2 ψ(r, t) = i~ ∂t ψ(r, t)

The solution for a particular momentum is given by a plane wave:


ψ(r, t) = ei(k·r−ωt)
A
with the constraint
~2 k 2
2m = ~ω

where r is the position vector, t is time, k is the wave vector, and ω is the
angular frequency. Since the integral of ψψ * over all space must be unity, there
will be a problem normalizing this momentum eigenfunction. This will not be a
DR
problem for a general free particle which is somewhat localized in momentum
and position. (See particle in a box for a further discussion.)
The expectation value of the momentum p is
hpi = hψ| − i~∇|ψi = ~k

The expectation value of the energy E is



hEi = hψ|i~ ∂t |ψi = ~ω

Solving for k and ω and substituting into the constraint equation yields the fa-
miliar relationship between energy and momentum for non-relativistic massive
particles
hpi2
hEi = 2m

Free particle
157

where p=|p|. The group velocity of the wave is defined as


vg = dω/dk = dE/dp = v

where v is the classical velocity of the particle. The phase velocity of the wave

FT
is defined as
vp = ω/k = E/p = p/2m = v/2

A general free particle need not have a specific momentum or energy. In this
case, the free particle wavefunction may be represented by a superposition of
free particle momentum eigenfunctions:
ψ(r, t) = A(k)ei(k·r−ωt) dk
R

where the integral is over all k-space.

Relativistic free particle


There are a number of equations describing relativistic particles. For a descrip-


A
tion of the free particle solutions, see the individual articles.

The →Klein-Gordon equation describes charge-neutral, spinless, relativistic


quantum particles

• The Dirac equation describes the relativistic electron (charged, spin 1/2)

Source: http://en.wikipedia.org/wiki/Free_particle
DR
Principal Authors: PAR, Linas, Lupin, Sverdrup, Karol Langner

Functional integration

This article’s topic is not that of functional integration (neurobiology) or


functional integration (sociology).

Functional integration is a collection of results in mathematics and physics


where the domain of an integral is no longer a region of space, but a space
of functions. Functional integrals arise in probability, in the study of partial
differential equations and in Feynman’s approach to the quantum mechanics
of particles and fields.

Functional integration
158

In an ordinary integral there is a function to be integrated—the integrand—and


a region of space over which to integrate the function—the domain of integra-
tion. The process of integration consists of adding the values of the integrand
at each point of the domain of integration. Making this procedure rigorous

FT
requires a limiting procedure, where the domain of integration is divided into
smaller and smaller regions. For each small region the value of the integrand
cannot vary much so it may be replaced by a single value. In a functional
integral the domain of integration is a space of functions. For each function
the integrand returns a value to add up. Making this procedure rigorous poses
challenges that are the topic of research in the beginning of the 21st century.
Functional integration was introduced by Wiener in 1921 in his studies of
Brownian motion. He developed a rigorous method —now known as the
Wiener measure— for assigning a probability to a particle’s random path. Feyn-
man developed another functional integral, the path integral, useful for com-
puting the quantum properties of systems. In Feynman’s path integral, the clas-
sical notion of a unique trajectory for a particle is replaced by an infinite sum
of classical paths, each weighed differently according to its classical properties.
Functional integration is central to quantization techniques in theoretical
A
physics. The algebraic properties of functional integrals are used to develop
series used to calculate properties in quantum electrodynamics and the stan-
dard model.

The problem of functional integration


Integration of functions is a summation. If the domain of integration is the
square [0,&nbsp;1] &times; [0,&nbsp;1], the integral is computed by breaking
the region into small rectangles. Each rectangle serves as the base of a prism
DR
whose height is any value of the function within the rectangle. The integral is
the sum of the volumes (base × height) of all the prisms. If the rectangles are
small enough and the function smooth, the process converges.
A functional is a function that associates a number to a function. This is in
distinction to common functions that associate numbers to other numbers. Ex-
amples of functionals include the functional that is one for any function, or the
functional the returns the integral of the function over a domain.
By analogy with integration of functions, functional integration is a summa-
tion procedure where the domain of integration is a space of functions and the
functional integral an addition of cylinders (just like the prisms in ordinary in-
tegration) with the functional as the height and some amount (or measure) of
function space as the base. The “area” of the functions can be represented by

Functional integration
159

Dω , the functional by F [ω] (square brackets are often used to distinguish func-
tionals from common functions), the space of functions by I and the functional
integral by

FT
R
I Dω F [ω] . (1)

The result of integrating a functional is to be a number. Developing a defini-


tion to equation (1) that has properties similar to ordinary integration is the
problem of the definition of functional integration. There is no general the-
ory to make sense of this formal expression as there is for the conventional
integration.
In functional integration there are different spaces to consider:

• The functions ω take values from a ν-dimensional space called space-time.


This is how many dimensions are used to specify a point in the domain of
ω. Space-time in often assumed to be a subset of R ν , the Euclidean space
with ν dimensions. In applications of functional integration, the functions
ω represent particle paths (in which case ν = 1) or a physical field such as
the vector potential (in which case ν = 4).
A
• The space for the range of the functions ω also varies depending on appli-
cations. This space is locally a subset of R κ. In applications it can be R 3, as
in the quantum mechanics; or a more complicated space, such as a tangent
bundle as in the case of quantum chromodynamics.
• The domain of integration is a function space, and likely to be infinite di-
mensional.

A definition of functional integration, by analogy with common integration, is


DR
expected to satisfy certain properties. Functional integration should be itself
R R R
a linear functional, such that Dω (F +α G ) = Dω F + α Dω G. Volumes
in functional space should be invariant under translation. A ball in functional
space centered around a function f or that same function plus a constant should
result in the same value for the functional integral. Also, if the functional
space happens to be finite dimensional, then the functional integral should be
related to the ordinary integration. These conditions are impossible to satisfy
for functional integrals.
Attempting to directly generalize the notion of volume in functional space has
not led to a useful theory of functional integration. Discretization of the func-
tional integral in equation (1) could be an approach towards its definition. For
the case of one-dimensional paths (ν=1 and κ=1), the functional integration
is replaced by an n-dimensional integral and the functional is computed from
the value of the path ω and n points. The functional integral would then be
the value of the n-dimensional integral in the limit of n going to infinity:

Functional integration
160

R R R R
Dω F (ω) = lim dω1 dω2 . . . dωn F (ω1 , . . . , ωn ) (2)
n→∞

The “size” of a function space can be computed from this expression by using
the simple functional F [ω]=1. Choosing a function space where each ω i varies

FT
over limited range of length W, the n-dimensional integral is equal to W n.
This will diverge to infinity or converge to zero in the limit. Building a theory
of integration when the value of the integral can only be zero or infinity is not
very interesting.
Most of the cylinders that contribute to the functional integral (2) correspond
to discontinuous functions. In Brownian motion or in the path integral formu-
lation of quantum mechanics, the paths are continuous. Both applications did
not generalize the notion of volume to functional spaces, as in equation (1),
but rather generalized the notion of a Gaussian integral. In applications, the
functional being integrated is related to an action functional S arising from
classical mechanics. Action functionals can be written as the sum of two terms,
S 0 + S i , with S 0 involving the derivative of the function squared. For exam-
ple, for the case of one-dimensional paths
A 2
S0 [ω] = dτ ω̇2 .
R

Smooth paths lead to small values of the functional, and large variations of
the path (as if almost discontinuous) lead to large values of the functional.
Introducing a term exp(-S 0) into the functional integral should dampen the
effects of discontinuous paths. This leads to Gaussian functional integrals.

Gaussian integration
DR
Instead of generalizing the notion of volume to infinite dimensions, the Gaus-
sian integral can be generalized. If M is a positive n×n symmetric matrix, and
x and J are n-dimensional vectors, the basic Gaussian integral can be used to
show that
1 1 1
Dx e− 2 x·M ·x+x·J = √ 1 e 2 J· M ·J
R
| det M |

The integration variables have been abbreviated with Dx=(2π ) -n/2dx 1...dx n.
The determinant and the matrix operations in the result of the many-
dimensional Gaussian can be interpreted in terms of infinite-dimensional ob-
jects. This result then can be used as the basis for the definition of a functional
integral. The action functional S of a path x(t) can be approximated by a
discretization of the time domain of the path into n+1 pieces of length a. At
these time values the path assumes the values x 0, ..., x n+1. The two end points

Functional integration
161

x 0 and x n+1 remain fixed and are not part of the n-fold integration. The dis-
cretized action S can be put in the form x∆x + xJ. The matrix ∆ is similar to
a discretized Laplace operator, the vector J will have only two non-zero entries
and the vector x runs from x 1 to x n . The determinant of ∆ and the inverse

FT
matrix ∆ -1 can be evaluated as a function of the number of discretization steps
n. The determinant is n+1. The inverse matrix is a matrix with entries of order
one divided by the determinant.
Using a limit of iterated Gaussians it is possible to define a functional integral
for paths (ν=1). With the notation O (1) to indicate a matrix with entries of
order one, the iterated Gaussian integral
1
 n+1 1/2 1
O(1)
Dx e− 2 x·M ·x+x·J = an+1
R
e 2a(n+1)

will converge if the entire expression is divided by a (n+1)/2.


In the case of a spacetime of more than one dimension, generalizations of
the Gaussian integral will not converge. This is at the root of many of the
difficulties in quantum field theory, as the Gaussian integral corresponds to the
quantum field theory with no interactions. The difficulty in generalizing the
A
Gaussian integral is that the growth of the determinant in the result cannot
be removed by a rescaling. (The determinant of the operator in ν spacetime
dimensions grows as 2nν.) In physics it is common to refer to the divergence
due to the determinant as an infrared divergence and the divergence due to
the limit of the discretization a going to zero as an ultraviolet divergence.
Due to the behavior of Gaussian integrals, in one spacetime dimension (ν=1)
it becomes possible to define a functional integral of the form
Dxe−S[x] .
R
DR
In the applications of functional integration to quantum mechanics the action
S[x] is pure imaginary and the exponential is an oscillating function. This leads
to further difficulties relating to conditional convergence. In spacetime dimen-
sions greater than one (ν ≥ 2), it is not known how to define a functional inte-
gral without resorting to regularization procedures that do away with many of
the properties of integration.

Approaches to path integrals


Functional integrals where the space of integration are paths (ν = 1) can be
defined in many different ways. The definitions fall in two different classes: the
constructions derived from Wiener’s theory yield a integral based on a measure;
whereas the constructions following Feynman’s path integral do not. Even

Functional integration
162

within these two broad divisions, the integrals are not identical, that is, they
are defined for different classes of functions.

The Wiener integral

FT
In the Wiener integral a probability is assigned to a class of Brownian motion
paths. The class consists of the paths w that are known to go through a small
region of space at a given time. The passage through different regions of space
is assumed independent of each other and the distance between any two points
of the Brownian path is assumed to be Gaussian distributed with a variance that
depends on the time t and on a diffusion constant D :
||w(s+t)−w(s)||2
 
Prob(w(s + t), t|w(s), s) = √ 1 exp − 2Dt
2πDt

The probability for the class of paths can be found by multiplying the probabil-
ities of starting in one region and then being at the next. The Wiener measure
can be developed by considering the limit of many small regions.

• Ito and Stratanovich calculus


A
The Feynman integral

• Trotter formula
• The Kac idea of Wick rotations.
• Using x-dot-dot-squared or i S[x] + x-dot-squared.
• The Cartier DeWitt-Morette relies on integrators rather than measures

See also
DR
• Feynman path integral
• Partition function

Reference and external links

Source: http://en.wikipedia.org/wiki/Functional_integration

Principal Authors: Michael Hardy, Miguel, XaosBits, Phys, The Anome

Functional integration
163

Geiger-Marsden experiment

The Geiger-Marsden experiment (also called the Gold foil experiment or the

FT
Rutherford experiment) was an experiment done by Hans Geiger and Ernest
Marsden in 1909, under the direction of Ernest Rutherford at the Physical Lab-
oratories of the University of Manchester which led to the downfall of the plum
pudding model of the atom.
They measured the deflection of alpha particles directed normally onto a sheet
of very thin gold foil. Under the prevailing plum pudding model, the alpha
particles should all have been deflected by, at most, a few degrees. However
they observed that a very small percentage of particles were deflected through
angles much larger than 90 degrees. From this observation Rutherford con-
cluded that the atom contained a very small positive charge which could repel
the alpha particles if they came close enough, subsequently developed into the
→Bohr model.

Methodology
A
Geiger and Marsden bombarded a number of different metal foils with alpha
particles generated from a tube of radium bromide gas. A low power micro-
scope was used to count the scattering of these particles, a procedure requiring
many hours in a darkened room watching for tiny flashes of light as the scat-
tered particles struck a zinc sulfide scintillant screen.
A variety of different foils were used such as aluminium, iron, gold and lead
along with different thicknesses of gold foil made by packing several pieces of
DR
very thin foil together. Given the very high mass and momentum of an alpha
particle, the expectation was that the particles would pass through having been
deflected by a tiny angle at most, with the number of particles penetrating
falling off as the thickness of foil (and the atomic weight of its material) was
increased; the remainder being absorbed.
However they were astonished to find that although this was generally true,
around 1 in 8000 particles were reflected through more than 90 degrees even
with a single sheet of extremely thin, 6x10 -8 metre (or about 200 atoms) thick,
gold foil, an observation completely at odds with the predictions of the plum
pudding model.

Geiger-Marsden experiment
164

Conclusions
The result was completely unpredicted, prompting Rutherford to later com-
ment "It was almost as incredible as if you fired a fifteen-inch shell at a piece
of tissue paper and it came back and hit you".

FT
Early in 1911 Rutherford published a revised model of the atom, known as the
Rutherford atom. The observations indicated that a model of the atom with a
diffuse positive charge was incorrect and that it was instead concentrated. He
concluded that the atom is mostly empty space, with most of the atom’s mass
concentrated in a tiny center, the nucleus, and electrons being held in orbit
around it by electrostatic attraction. The nucleus was around 10 -15 meters in
diameter, in the centre of a 10 -10 metre diameter atom. Those alpha particles
that had come into proximity with the nucleus had been strongly deflected
whereas the majority had passed at a relatively great distance to it.
Rutherford’s model was developed by Niels Bohr into the →Bohr model pro-
posed in 1913. The Rutherford atom had a number of problems, in particular
electrons should radiate electromagnetic energy and rapidly spiral into the nu-
cleus.
A
See also
• List of famous experiments

References
• Geiger H. & Marsden E. (1909). "On a Diffuse Reflection of the α-
DR
Particles" 88. Proceedings of the Royal Society, Series A 82: 495–500.
• Rutherford E. (1911). "The Scattering of α and β Particles by Matter and the
Structure of the Atom" 89. Philosophical Magazine, Series 6 21: 669–688.
• JPEG images of Rutherford’s 1911 paper 90
• Description of the experiment, from the New Mexico Institute of Mining
and Technology 91
• Short biography of Ernest Rutherford 92

Source: http://en.wikipedia.org/wiki/Geiger-Marsden_experiment

88 http://dbhs.wvusd.k12.ca.us/webdocs/Chem-History/GM-1909.html
89 http://dbhs.wvusd.k12.ca.us/webdocs/Chem-History/Rutherford-1911/Rutherford-1911.html
90 http://www.math.ubc.ca/~cass/rutherford/rutherford.html
91 http://www.physics.nmt.edu/~raymond/classes/ph13xbook/node193.html
92 http://www.phy.hr/~dpaar/fizicari/xrutherf.html

Geiger-Marsden experiment
165

Principal Authors: Jll, Fastfission, Sjakkalle, Raul654, Malcolm Farmer

Gibbs paradox

FT
This article is about Gibbs paradox on the extensivity of entropy; for Gibbs
mixing paradox, see mixing paradox.

In statistical mechanics, a simple derivation of the entropy of an ideal gas,


based on the Boltzmann distribution yields an expression for the entropy which
is not an extensive variable as it must be, leading to an apparent paradox
known as the Gibbs paradox. The difficulty is resolved by requiring the par-
ticles be indistinguishable which results in "correct Boltzmann counting". The
resulting equation for the entropy of a classical ideal gas is extensive, and is
known as the Sackur-Tetrode equation.
If you have a fixed volume of an ideal gas, the measurement of the total en-
tropy within the volume should not change if you were to divide the volume
A
into two (or more) equal partitions, then remove the partitions. However, if
you calculate the total entropy by measuring (and adding up) the position and
momentum of each individual particle inside the volume (and keep track of
each particle separately), then the entropy you calculate will change depend-
ing on whether you add and remove partitions inside the fixed volume. This
discrepancy is called the Gibb’s Paradox. The paradox is resolved by concluding
that every particle is indistinguishable from every other particle in the volume,
thus you cannot measure entropy by measuring particles as if each were indi-
DR
vidually identifiable.

Calculating the Gibbs paradox


If we have an ideal gas of energy U, volume V and with N particles, then we
can represent the state of the gas by specifying the 3D momentum vector and
the 3D position vector for each of the N particles. This can be thought of as
specifying the coordinates of a point in a 6N-dimensional phase space, where
each of the axes corresponds to one of the momentum or position coordinates
of one of the particles. The set of all possible points that the gas could ever
occupy in phase space is specified by the constraint that the gas will have a
particular energy:
1 PN P3 2
U = 2m i=1 j=1 pij

Gibbs paradox
166

and be contained inside of the volume V (let’s say V is a box of side X so that
X 3=V ):
0 ≤ xij ≤ X

FT
where [p i1, p i2, p i3] and [x i1, x i2, x i3] are the vector momentum and position
of particle i. The first constraint defines the surface of a 3N-dimensional hy-
persphere of radius (2mU ) 1/2 and the second is a 3N-dimensional hypercube
of volume V N. These combine to form a 6N-dimensional "hypercylinder". Just
as the area of the wall of a cylinder is the circumference of the base times the
height, so the area φ of the wall of this hypercylinder is:
3N −1
 3N 
2π 2 (2mU ) 2
φ(U, V, N ) = V N Γ(3N/2)
(1)

The entropy is proportional to the logarithm of the number of states that


the gas could have while satisfying these constraints. Another way of stating
Heisenberg’s uncertainty principle is to say that we cannot specify a volume in
phase space smaller than h 3N where h is Planck’s constant. The above "area"
must really be a shell of thickness h, so we therefore write the entropy as:
A
S = k log(φh/h3N )

where k is the constant of proportionality, Boltzmann’s constant. Using Stir-


ling’s approximation for the Gamma function, and keeping only terms of order
N the entropy becomes:
  3   
U 2
S(=?) kN log V N + 32 kN 53 + log 4πm
3h2
DR
This quantity is not extensive as can be seen by considering two identical vol-
umes with the same particle number and the same energy. Suppose the two
volumes are separated by a barrier in the beginning. Removing or reinserting
the wall is reversible, but the entropy difference after removing the barrier is
δS = k [2N log(2V ) − N log V − N log V ] = 2kN log 2 > 0

which is in contradiction to thermodynamics. This is the Gibbs paradox. It was


resolved by J.W. Gibbs himself, by postulating that the gas particles are in fact
indistinguishable. This means that all states that differ only by a permutation
of particles should be considered as the same point. For example, if we have
a 2-particle gas and we specify AB as a state of the gas where the first particle
(A ) has momentum p 1 and the second particle (B ) has momentum p 2, then
this point as well as the BA point where the B particle has momentum p 1 and
the A particle has momentum p 2 should be counted as the same point. It can

Gibbs paradox
167

be seen that for an N -particle gas, there are N! points which are identical in
this sense, and so to calculate the volume of phase space occupied by the gas
we must divide Equation 1 by N!. This will give for the entropy:
    3   
V U 2
S = kN log N + 32 kN 53 + log 4πm

FT
N 3h2

which can be easily shown to be extensive. This is the Sackur-Tetrode equation.

Source: http://en.wikipedia.org/wiki/Gibbs_paradox

Principal Authors: PAR, Marco Krohn, Charles Matthews, Kjkolb, Linas

Greenberger-Horne-Zeilinger state

In physics, in the area of quantum information theory, a Greenberger-Horne-


Zeilinger state is a certain type of entangled quantum state.
A
Definition
The GHZ state is an entangled quantum state in an M dimensions:
|0i⊗M√
+|1i⊗M
|GHZi = .
2

|000i+|111i
Most notably the 3-qubit GHZ state is: |GHZi = √ .
2

Properties
DR
Apparently there is no standard measure of multi-partite entanglement, but
many measures define the GHZ to be maximally entangled.
Another important property of the GHZ state is that when we trace over one of
the three systems
T r3 ((|000i + |111i)(h000| + h111|)) = |00ih00| + |11ih11|

which is a mixed state.


On the other hand, if we were to measure any of subsystems we will leave
behind either |00i or |11i which are not entangled. Thus, we say that the
GHZ is maximally entangled. This is unlike the W state which leaves bipartite
entanglements even when we measure one of its subsystems.

Source: http://en.wikipedia.org/wiki/Greenberger-Horne-Zeilinger_state

Greenberger-Horne-Zeilinger state
168

Hamiltonian (quantum mechanics)

The quantum Hamiltonian is the physical state of a system, which may be

FT
characterized as a ray in an abstract →Hilbert space (or, in the case of ensem-
bles, as a trace class operator with trace 1). (See →Mathematical formulation
of quantum mechanics)
Physically observable quantities are described by self-adjoint operators acting
on the Hilbert space. For example, the Hilbert space associated with the spin
degrees of freedom of a spin-1/2 particle is C 2, while the Hilbert space associat-
ed to a spinless particle moving on a line is L 2(R), the space of complex-valued
and square-integrable (with respect to the Lebesgue measure) functions on the
real line.
The quantum Hamiltonian H is the observable corresponding to the total en-
ergy of the system. If the state space is finite dimensional, then it is of course
bounded. In the infinite dimensional case, it is almost always unbounded,
therefore not defined everywhere.
A
In introductory physics literature, the following is considered either as a defi-
nition or an axiom:
The eigenkets (eigenvectors) of H, denoted |ai

(using Dirac Bra-ket notation), provide an orthonormal basis for

the Hilbert space. The spectrum of allowed energy levels of the system
is given by the set of eigenvalues, denoted {E a}, solving the equation:
DR
H |ai = Ea |ai.

Since H is a Hermitian operator, the energy is always a real number.

From a mathematically rigorous point of view, care must be taken with the
above axiom. Operators on infinite-dimensional Hilbert spaces need not have
eigenvalues (notice in the infinite-dimensional case, the set of eigenvalues
need not coincide with the spectrum, which can always safely assumed to
be nonempty). However, all routine quantum mechanical calculations can be
done using the physical formulation.
As with all observables, the spectrum of the Hamiltonian are the possible out-
comes when one measures the total energy of a system. Like any other self

Hamiltonian (quantum mechanics)


169

adjoint operator, the spectrum of the Hamiltonian can be decomposed, via its
spectral measures, into discrete, absolutely continuous, and singular parts. The
discrete spectrum can be associated to eigenvectors, which in turn usually are
the bound states of the system. The absolutely continuous spectrum corre-

FT
spond to the free states. The singular spectrum, interesting enough, are phys-
ically impossible outcomes. For example, consider the finite potential well,
which admits bound states with discrete negative energies and free states with
continuous positive energies.
The Hamiltonian generates the time evolution of quantum states. If |ψ(t)i is
the state of the system at time t, then

H |ψ(t)i = i~ ∂t |ψ(t)i.

where ~ is h-bar. This equation is known as the →Schrödinger equation. (It


takes the same form as the Hamilton-Jacobi equation, which is one of the rea-
sons H is also called the Hamiltonian.) Given the state at some initial time (t =
0), we can integrate it to obtain the state at any subsequent time. In particular,
if H is independent of time, then
A 
|ψ(t)i = exp − iHt~

|ψ(0)i.

In physical literature, the exponential operator on the right hand side is defined
by the power series. One might note that taking polynomials of unbounded
and not everywhere defined operators makes no mathematical sense, much
less power series. Rigorously, to take functions of unbounded operators, one
requires a functional calculus. In the case of the exponential function, the con-
tinuous, or just the holomorphic functional calculus suffices. We note again,
DR
however, that for common calculations the physicist’s formulation is quite pro-
ficient.
By the *-homomorphism property of the Borel functional calculus, the operator
 
U = exp − iHt
~

is an unitary operator, and is a common form of the time evolution operator


(also called the propagator). If the Hamiltonian is time-independent, {U(t)}
form a one parameter unitary group (more than a semigroup); this gives rise
to the physical principle of detailed balance.

Hamiltonian (quantum mechanics)


170

Energy eigenket degeneracy, symmetry, and conser-


vation laws
In many systems, two or more energy eigenstates have the same energy. A

FT
simple example of this is a free particle, whose energy eigenstates have wave-
functions that are propagating plane waves. The energy of each of these plane
waves is inversely proportional to the square of its wavelength. A wave prop-
agating in the x direction is a different state from one propagating in the y
direction, but if they have the same wavelength, then their energies will be the
same. When this happens, the states are said to be degenerate.
It turns out that degeneracy occurs whenever a nontrivial unitary operator U
commutes with the Hamiltonian. To see this, suppose that |a> is an energy
eigenket. Then U |a> is an energy eigenket with the same eigenvalue, since
U H|ai = U Ea |ai = Ea (U |ai) = H (U |ai).

Since U is nontrivial, at least one pair of |ai and U |ai must represent distinct
states. Therefore, H has at least one pair of degenerate energy eigenkets. In
A
the case of the free particle, the unitary operator which produces the symmetry
is the rotation operator, which rotates the wavefunctions by some angle while
otherwise preserving their shape.
The existence of a symmetry operator implies the existence of a conserved
observable. Let G be the Hermitian generator of U :
U = I − iG + O(2 )

It is straightforward to show that if U commutes with H, then so does G :


DR
[H, G] = 0

Therefore,
∂ 1
∂t hψ(t)|G|ψ(t)i = i~ hψ(t)|[G, H]|ψ(t)i =0

In obtaining this result, we have used the Schrödinger equation, as well as its
dual,

hψ(t)|H = −i~ ∂t hψ(t)|.

Thus, the expected value of the observable G is conserved for any state of the
system. In the case of the free particle, the conserved quantity is the angular
momentum.

Hamiltonian (quantum mechanics)


171

Hamilton’s equations
Hamilton’s equations in classical Hamiltonian mechanics have a direct analogy
in quantum mechanics. Suppose we have a set of basis states {|ni}, which need
not necessarily be eigenstates of the energy. For simplicity, we assume that they

FT
are discrete, and that they are orthonormal, i.e.,
hn0 |ni = δnn0

Note that these basis states are assumed to be independent of time. We will
assume that the Hamiltonian is also independent of time.
The instantaneous state of the system at time t, |ψ (t)i, can be expanded in
terms of these basis states:
P
|ψ(t)i = n an (t)|ni

where
an (t) = hn|ψ(t)i
A
The coefficients a n(t) are complex variables. We can treat them as coordinates
which specify the state of the system, like the position and momentum coor-
dinates which specify a classical system. Like classical coordinates, they are
generally not constant in time, and their time dependence gives rise to the
time dependence of the system as a whole.
The expectation value of the Hamiltonian of this state, which is also the mean
energy, is
hH(t)i ≡ hψ(t)|H|ψ(t)i = nn0 a∗n0 an hn0 |H|ni
P
DR
where the last step was obtained by expanding |ψ (t)i in terms of the basis
states.
Each of the a n(t) s actually corresponds to two independent degrees of
freedom, since the variable has a real part and an imaginary part. We
now perform the following trick: instead of using the real and imaginary
parts as the independent variables, we use a n(t) and its complex conju-
gate a n*(t). With this choice of independent variables, we can calculate
the partial derivative
∂hHi P 0 0
∂a∗ = n an hn |H|ni = hn |H|ψi
n0

By applying Schrödinger’s equation and using the orthonormality of the basis


states, this further reduces to

Hamiltonian (quantum mechanics)


172

∂hHi ∂an0
∂a∗n0 = i~ ∂t

Similarly, one can show that


∂hHi ∂a∗n
= −i~

FT
∂an ∂t

If we define "conjugate momentum" variables π n by


πn (t) = i~a∗n (t)

then the above equations become


∂hHi ∂an ∂hHi
∂πn = ∂t , ∂an = − ∂π
∂t
n

which is precisely the form of Hamilton’s equations, with the an s as the gen-
eralized coordinates, the πn s as the conjugate momenta, and hHi taking the
place of the classical Hamiltonian.

See also

AHamiltonian mechanics

Source: http://en.wikipedia.org/wiki/Hamiltonian_%28quantum_mechanics%29

Principal Authors: CYD, Lethe, Mct mht, Enormousdude, Oleg Alexandrov, Looxix, SeventyThree

Heisenberg picture
DR
In physics, the Heisenberg picture is that formulation of quantum mechanics
where the operators (observables and others) are time-dependent and the state
vectors are time-independent. It stands in contrast to the →Schrödinger picture
in which operators are constant and the states evolve in time.
The "Heisenberg Picture" is not to be confused with matrix mechanics which
is sometimes called Heisenberg quantum mechanics.

Mathematical details
In quantum mechanics in the Heisenberg picture the state vector, |ψ> does
not change with time, and an observable A satisfies
 
d −1 ∂A
dt A = (i~) [A, H] + ∂t .
classical

Heisenberg picture
173

In some sense, the Heisenberg picture is more natural and fundamental than
the →Schrödinger picture, especially for relativistic theories. Lorentz invari-
ance is manifest in the Heisenberg picture.
Moreover, the similarity to classical physics is easily seen: by replacing the

FT
commutator above by the Poisson bracket, the Heisenberg equation becomes
an equation in Hamiltonian mechanics.
By the Stone-von Neumann theorem, the Heisenberg picture and the
→Schrödinger picture are unitarily equivalent.
See also →Schrödinger picture.

Deriving Heisenberg’s equation


Suppose we have an observable A (which is a Hermitian linear operator). The
expectation value of A for a given state |ψ(t)> is given by:
hAit = hψ(t)|A|ψ(t)i

or if we write following the →Schrödinger equation


A|ψ(t)i = e−iHt/~ |ψ(0)i

(where H is the Hamiltonian and hbar is Planck’s constant divided by 2·π) we


get
hAit = hψ(0)|eiHt/~ Ae−iHt/~ |ψ(0)i,

and so we define
A(t) := eiHt/~ Ae−iHt/~ .
DR
Now,
 
d
dt A(t) = ~i HeiHt/~ Ae−iHt/~ + ∂A
∂t classical + ~i eiHt/~ A · (−H)e−iHt/~

(differentiating according to the product rule),


 
= ~i eiHt/~ (HA − AH) e−iHt/~ + ∂A ∂t classical =
i
~ (HA(t) − A(t)H) +
 
∂A
∂t classical

(the last passage is valid since exp(-iHt/hbar) commutes with H )


 
= ~i [H, A(t)] + ∂A∂t classical

Heisenberg picture
174

(where [X, Y ] is the commutator of two operators and defined as [X, Y ] :=


XY - YX )
So we get
 
d
dt A(t) = ~i [H, A(t)] + ∂A
∂t classical .

FT
for a Time independent observable A and using Jacobi identity and integration
by parts respect to time we get:
2
A = C + i~t [A, H] + ~t 2 [A, [A, H]] where C is a constant operator, due to the
relationship between Poisson Bracket and Commutators this relation also
holds for classical mechanics

See also
• →Interaction picture
• →Quantum mechanics
• →Schrödinger equation


A→Bra-ket notation
→Matrix mechanics

Source: http://en.wikipedia.org/wiki/Heisenberg_picture

Principal Authors: Lethe, Pjacobi, -Ril-, Karl-H, MathKnight, MarSch, GangofOne, GregorB, Naar

Hilbert space
DR
In mathematics, a Hilbert space is a generalization of Euclidean space that is
not restricted to finite dimensions. Thus it is an inner product space, which
means that it has notions of distance and of angle (especially the notion of or-
thogonality or perpendicularity). Moreover, it satisfies a more technical com-
pleteness requirement which ensures that limits exist when expected, which
facilitates various definitions from calculus. Hilbert spaces allow geometric in-
tuition to be applied to certain infinite dimensional functional spaces. They
provide a context with which to formalize and generalize the concepts of the
Fourier series in terms of arbitrary orthogonal polynomials and of the Fourier
transform, which are central concepts from functional analysis. Hilbert spaces
are of crucial importance in the mathematical formulation of quantum me-
chanics.

Hilbert space
175

Introduction
Hilbert spaces were named after David Hilbert, who studied them in the con-
text of integral equations. John von Neumann originated the designation "der
abstrakte Hilbertsche Raum" in his famous work on unbounded Hermitian op-

FT
erators published in 1929. Von Neumann was perhaps the mathematician who
most clearly recognized their importance as a result of his seminal work on the
foundations of quantum mechanics begun with Hilbert and Lothar (Wolfgang)
Nordheim and continued with Eugene Wigner. The name "Hilbert space" was
soon adopted by others, for example by Hermann Weyl in his book The The-
ory of Groups and Quantum Mechanics published in 1931 (English language
paperback ISBN 0486602699).
The elements of an abstract Hilbert space are sometimes called "vectors". In
applications, they are typically sequences of complex numbers or functions. In
quantum mechanics for example, a physical system is described by a complex
Hilbert space which contains the "wavefunctions" that stand for the possible
states of the system. See mathematical formulation of quantum mechanics for
details. The Hilbert space of plane waves and bound states commonly used in
A
quantum mechanics is known more formally as the rigged Hilbert space.

Definition
Every inner product <·,·> on a real or complex vector space H gives rise to a
norm ||·|| as follows:
p
||x|| = hx, xi.

In any normed space, the open balls constitute a compatible topology; any
DR
normed vector space is a topological vector space (and even a uniform struc-
ture) and therefore so is any inner product space.
The Cauchy criterion may be defined for sequences in this space (as it can
in any uniform space): a sequence {x n } n is a Cauchy sequence if for every
natural number N there is a real number  such that for all m, n > N, ||x n –
x m || < . We call H a Hilbert space if it is complete with respect to this norm, that is if every Cauchy sequence
All finite-dimensional inner product spaces (such as Euclidean space with the
ordinary dot product) are Hilbert spaces. However, the infinite-dimensional ex-
amples are much more important in applications. These applications include:

• The theory of unitary group representations


• The theory of square integrable stochastic processes
• The Hilbert space theory of partial differential equations, in particular for-
mulations of the Dirichlet problem

Hilbert space
176

• Spectral analysis of functions, including theories of wavelets


• Mathematical formulations of quantum mechanics

The inner product allows one to adopt a "geometrical" view and use geomet-
rical language familiar from finite dimensional spaces. Of all the infinite-

FT
dimensional topological vector spaces, the Hilbert spaces are the most "well-
behaved" and the closest to the finite-dimensional spaces.
One goal of Fourier analysis is to write a given function as a (possibly infi-
nite) sum of multiples of given base functions. This problem can be studied
abstractly in Hilbert spaces: every Hilbert space has an orthonormal basis, and
every element of the Hilbert space can be written in a unique way as a sum of
multiples of these base elements.

Examples
In these examples, we will assume the underlying field of scalars is C, although
the definitions apply to the case in which the underlying field of scalars is R.

Euclidean spaces
A
C n with the inner product definition
hx, yi = nk=1 xk yk
P

where the bar over a complex number denotes its complex conjugate.

Sequence spaces
Much more typical are the infinite dimensional Hilbert spaces however. If B is
DR
any set, we define the sequence space little l 2 over B, denoted by
 
x : B → C b∈B |x (b)|2 < ∞
P
`2 (B) =

This space becomes a Hilbert space with the inner product


P
hx, yi = b∈B x(b)y(b)

for all x and y in l 2(B ). B does not have to be a countable set in this definition,
although if B is not countable, the resulting Hilbert space is not separable. In
a sense made more precise below, every Hilbert space is isomorphic to one the
form l 2(B ) for a suitable set B. If B =N, we write simply l 2.

Hilbert space
177

Lebesgue spaces
These are function spaces associated to measure spaces (X, M, µ), where M is
a σ-algebra of subsets of X and µ is a countably additive measure on M. Let
L 2 µ(X ) be the space of complex-valued square-integrable measurable func-

FT
tions on X, modulo equality almost everywhere. Square integrable means the
integral of the square of its absolute value is finite. Modulo equality almost
everywhere means functions are identified if and only if they are equal outside
of a set of measure 0.
The inner product of functions f and g is here given by
R
hf, gi = X f (t)g(t) dµ(t)

One needs to show:

• That this integral indeed makes sense;


• The resulting space is complete.

These are technically easy facts, and the interested reader should consult the
A
Halmos reference below, Section 42. Note that the use of the Lebesgue integral
ensures that the space will be complete. See L p space for further discussion of
this example.

Sobolev spaces
Sobolev spaces, denoted by H s or W s,2 , are another example of Hilbert spaces,
and are used very often in the field of Partial differential equations.

Operations on Hilbert spaces


DR
Given two (or more) Hilbert spaces, we can combine them into a single Hilbert
space by taking their direct sum or their tensor product.

Bases
An important concept is that of an orthonormal basis of a Hilbert space H :
this is a family {e k } k ∈ B of H satisfying:

• Elements are normalized: Every element of the family has norm 1: ||e k || =
1 for all k in B
• Elements are orthogonal: Every two different elements of B are orthogonal:
<e k , e j > = 0 for all k, j in B with k 6= j.
• Dense span: The linear span of B is dense in H.

We also use the expressions orthonormal sequence and orthonormal set.

Hilbert space
178

Examples of orthonormal bases include:

• the set {(1,0,0),(0,1,0),(0,0,1)} forms an orthonormal basis of R 3


• the sequence {f n : n ∈ Z} with f n (x) = exp(2πinx) forms an orthonormal
basis of the complex space L 2([0,1])

FT
• the family {e b : b ∈ B } with e b (c) = 1 if b =c and 0 otherwise forms an
orthonormal basis of l 2(B ).

Note that in the infinite-dimensional case, an orthonormal basis will not be a


basis in the sense of linear algebra; to distinguish the two, the latter basis is
also called a Hamel basis. That the span of the basis vectors is dense means
that every vector in the space can be written as the limit of an infinite series
and the orthogonality implies that this decomposition is unique.
Using Zorn’s lemma, one can show that every Hilbert space admits an orthonor-
mal basis; furthermore, any two orthonormal bases of the same space have the
same cardinality. A Hilbert space is separable if and only if it admits a count-
able orthonormal basis.
Since all infinite-dimensional separable Hilbert spaces are isomorphic, and
A
since almost all Hilbert spaces used in physics are separable, when physicists
talk about the Hilbert space they mean any separable one.
If {e k } k ∈ B is an orthonormal basis of H, then every element x of H may be
written as
P
x = k∈B hek , xiek

Even if B is uncountable, only countably many terms in this sum will be non-
zero, and the expression is therefore well-defined. This sum is also called the
DR
Fourier expansion of x.
If {e k } k ∈ B is an orthonormal basis of H, then H is isomorphic to l 2(B ) in the
following sense: there exists a bijective linear map Φ : H → l 2(B ) such that
hΦ (x) , Φ (y)i = hx, yi

for all x and y in H.

Orthogonal complements and projections


If S is a subset of a Hilbert space H, we define the set of vectors orthogonal to
S
S perp = {x ∈ H : hx, si = 0 ∀s ∈ S}

Hilbert space
179

S perp is a closed subspace of H and so forms itself a Hilbert space. If V is a


closed subspace of H, then V perp is called the orthogonal complement of V. In
fact, every x in H can then be written uniquely as x = v + w, with v in V
and w in V perp. Therefore, H is the internal Hilbert direct sum of V and V perp.

FT
The linear operator P V : H → H which maps x to v is called the orthogonal
projection onto V.
Theorem. The orthogonal projection P V is a self-adjoint linear operator on H
of norm ≤ 1 with the property P V 2 = P V . Moreover, any self-adjoint linear
operator E such that E 2 = E is of the form P V , where V is the range of E. For
every x in H, P V (x) is the unique element v of V which minimizes the distance
||x - v||.
This provides the geometrical interpretation of P V (x): it is the best approxima-
tion to x by elements of V.

Reflexivity
An important property of any Hilbert space is its reflexivity. In fact, more is
true: one has a complete and convenient description of its dual space (the
A
space of all continuous linear functions from the space H into the base field),
which is itself a Hilbert space. Indeed, the Riesz representation theorem states
that to every element φ of the dual H there exists one and only one u in H
such that
φ (x) = hu, xi

for all x in H and the association φ ↔ u provides an antilinear isomorphism


between H and H . This correspondence is exploited by the bra-ket notation
DR
popular in physics but frowned upon by mathematicians.

Bounded operators
For a Hilbert space H, the continuous linear operators A : H → H are of partic-
ular interest. Such a continuous operator is bounded in the sense that it maps
bounded sets to bounded sets. This allows to define its norm as
kAk = sup { kAxk : kxk ≤ 1 } .

The sum and the composition of two continuous linear operators is again con-
tinuous and linear. For y in H, the map that sends x to <y, Ax> is linear and
continuous, and according to the Riesz representation theorem can therefore
be represented in the form
hA∗ y, xi = hy, Axi.

Hilbert space
180

This defines another continuous linear operator A * : H → H, the adjoint of A.


The set L(H ) of all continuous linear operators on H, together with the addition
and composition operations, the norm and the adjoint operation, forms a C *-
algebra; in fact, this is the motivating prototype and most important example

FT
of a C *-algebra.
An element A of L(H ) is called self-adjoint or Hermitian if A * = A. These
operators share many features of the real numbers and are sometimes seen as
generalizations of them.
An element U of L(H ) is called unitary if U is invertible and its inverse is given
by U *. This can also be expressed by requiring that <Ux, Uy> = <x, y> for
all x and y in H. The unitary operators form a group under composition, which
can be viewed as the automorphism group of H.

Unbounded operators
If a linear operator has a closed graph and is defined on all of a Hilbert space,
then, by the closed graph theorem in Banach space theory, it is necessarily
bounded. However, if we allow ourselves to define a linear map that is defined
A
on a proper subspace of the Hilbert space, then we can obtain unbounded
operators.
In quantum mechanics, several interesting unbounded operators are defined on
a dense subspace of Hilbert space. It is possible to define self-adjoint unbound-
ed operators, and these play the role of the observables in the mathematical
formulation of quantum mechanics.
Examples of self-adjoint unbounded operator on the Hilbert space L 2(R) are:
DR
• A suitable extension of the differential operator
d
[Af ](x) = i dx f (x),

where i is the imaginary unit and f is a differentiable function of compact


support.

• The multiplication by x operator:

[Bf ](x) = xf (x).

These correspond to the momentum and position observables, respectively.


Note that neither A nor B is defined on all of H, since in the case of A the
derivative need not exist, and in the case of B the product function need not

Hilbert space
181

be square integrable. In both cases, the set of possible arguments form dense
subspaces of L 2(R).

See also

FT
• Topologies on the set of operators on a Hilbert space
• Operator algebra
• Reproducing kernel Hilbert space
• Rigged Hilbert space

• Mathematical analysis
• Functional analysis
• Harmonic analysis

References
• Jean Dieudonné, Foundations of Modern Analysis, Academic Press, 1960.
• Paul Halmos, Measure Theory, D. van Nostrand Co, 1950.

A David Hilbert, Lothar Nordheim, and John von Neumann, "Über die Grund-
lagen der Quantenmechanik," Mathematische Annalen, volume 98, pages
1-30, 1927.
• John von Neumann, "Allgemeine Eigenwerttheorie Hermitescher Funktion-
aloperatoren," Mathematische Annalen, volume 102, pages 49-131, 1929.
• Hermann Weyl, The Theory of Groups and Quantum Mechanics, Dover
Press, 1950. This book was originally published in German in 1931.
DR
Source: http://en.wikipedia.org/wiki/Hilbert_space

Principal Authors: CSTAR, Jitse Niesen, Michael Hardy, Vilemiasma, Lupin, Lethe, Paul August, Zun-
dark

Hilbert space
182

Hydrogen atom

An hydrogen atom is an atom of the chemical element hydrogen. It is

FT
composed of a single negatively-charged electron circling a single positively-
charged proton which is the nucleus of the hydrogen atom. The electron is
bound to the proton by the Coulomb force.
The hydrogen atom has special significance in quantum mechanics and quan-
tum field theory as a simple physical system for which the solution to the
→Schrödinger equation is analytical, from which the positions of energy levels
(thus, the frequencies of the hydrogen spectral lines) can be calculated.
In 1913, Niels Bohr had deduced the spectral frequencies of the hydrogen atom
making several assumptions (see The Bohr Model). The results of Bohr for the
frequencies and underlying energy values are confirmed by the full quantum-
mechanical analysis which uses the Schrödinger equation, as was shown in
1925/26. The solution of the Schrödinger equation goes much further, because
it also yields the shape of the electron’s wave function ("orbital") for the various
A
possible quantum-mechanical states - thus explaining the anisotropic character
of atomic bonds. The Schrödinger equation also applies to more complicated
atoms and molecules, however, in most cases the solution is not analytical and
either computer calculations are necessary or some simplifying assumptions
must be made.

Solution of Schrödinger equation: Overview of re-


sults
DR
The solution of the Schrödinger equation for the hydrogen atom uses the fact
that the Coulomb potential produced by the nucleus is isotropic (it only de-
pends on the distance to the nucleus). Although the resulting energy eigen-
functions (the "orbitals") are not necessarily isotropic themselves, their de-
pendence on the angular coordinates follows completely generally from this
isotropy of the underlying potential: The states are not only eigenstates of
the Hamiltonian, but also eigenstates of the angular momentum operator (so
called spherical harmonics). This corresponds to the fact that angular mo-
mentum is conserved in the orbital motion of the electron around the nucleus.
Therefore, the energy eigenstates may be classified by two angular momen-
tum quantum numbers, l and m (integer numbers). The "angular momentum"
quantum number l = 0, 1, 2, ... determines the magnitude of the angular
momentum. The "magnetic" quantum number m = -l, .., +l determines the
projection of the angular momentum on the (arbitrarily chosen) z-axis.

Hydrogen atom
183

In addition, the radial dependence of the wave functions has to be found. It


is only here that the details of the 1/r Coulomb potential enter (leading to
Laguerre polynomials in r). This leads to a third quantum number, the principal
quantum number n = 1, 2, 3, ... Note that the angular momentum quantum

FT
number can run only up to n - 1, i.e. l = 0, 1, ..., n - 1.
Due to angular momentum conservation, states of the same l but different m
have the same energy (this holds for all problems with rotational symmetry).
In addition, for the hydrogen atom, the states of the same n are also degenerate
(i.e. they have the same energy); but this is a specialty and it is no longer true
for more complicated atoms which have a (effective) potential differing from
the form 1/r (due to the presence of the inner electrons shielding the nucleus
potential).
Taking into account the spin of the electron adds a last quantum number, the
projection of the electrons spin along the z axis, which can take on two values.
Therefore, any eigenstate of the electron in the hydrogen atom is described
fully by four quantum numbers. According to the usual rules of quantum me-
chanics, the actual state of the electron may be any superposition of these
states. This explains also why the choice of z-axis for the quantization of angu-
A
lar momentum is immaterial: An orbital of given l and m’ obtained for another
preferred axis z’ can always be represented as a suitable superposition of the
various states of different m (but same l) that have been obtained for z.

Mathematical summary of eigenstates of hydrogen


atom
Main article: hydrogen-like atom
DR
The normalized position wavefunctions, given in spherical coordinates are:
r 
3 (n−l−1)!
ψnlm (θ, φ, r) = 2
na0 2n[(n+l)!]
e−ρ/2 ρl L2l+1
n−l−1 (ρ) · Yl,m (θ, φ)

where:
2r
• ρ= na0 and a 0 is the Bohr radius.
• L2l+1
n−l−1 (ρ) are the Generalized Laguerre polynomials of degree n-l-1.
• Yl,m (θ, φ) is a spherical harmonic.

The eigenvalues are:

• For Angular momentum operator:

Hydrogen atom
184

L2 |n, l, mi = ~2 l(l + 1)|n, l, mi

Lz |n, l, mi = ~m|n, l, mi

FT
• For the Hamiltonian:

H|n, l, mi = En |n, l, mi

where

2 2 2
 2
Ze2
En = − mc2·n
Z α
2
m
= − 2~ 2 4π0
1
n2
and α is the fine structure constant
and in hydrogen atom Z =1.

Visualizing the hydrogen electron orbitals


A
DR

The image to the right shows the first few hydrogen atom orbitals (energy
eigenfunctions). These are cross-sections of the probability density that are

Hydrogen atom
185

color-coded (black=zero density, white=highest density). The angular mo-


mentum quantum number l is denoted in each column, using the usual spec-
troscopic letter code ("s" means l = 0; "p": l = 1; "d ": l = 2). The main
quantum number n (= 1, 2, 3, ...) is marked to the right of each row. For

FT
all pictures the magnetic quantum number m has been set to 0, and the cross-
sectional plane is the xz-plane (z is the vertical axis). The probability density
in three-dimensional space is obtained by rotating the one shown here around
the z-axis.
The "ground state", i.e. the state of lowest energy, in which the electron is
usually found, is the first one, the "1s" state (n = 1, l = 0).
An image with more orbitals is also available (up to higher numbers n and l).
Note the number of black lines that occur in each but the first orbital. These
are "nodal lines" (which are actually nodal surfaces in three dimensions). Their
total number is always equal to n - 1, which is the sum of the number of radial
nodes (equal to n - l - 1) and the number of angular nodes (equal to l ).

Features going beyond the Schrödinger solution


A
There are several important effects that are neglected by the Schrödinger equa-
tion and which are responsible for certain small but measurable deviations of
the real spectral lines from the predicted ones:

• Although the maximum effective speed of the electron in hydrogen is only


1/137th of the speed of light there is an increase in the electron’s mass, as
predicted by special relativity. For heavier elements, this is more significant
(see 93).
DR
• The spin of the electron has a magnetic moment attached to it. Even when
there is no external magnetic field, within the inertial frame of the moving
electron the electric field of the nucleus partly acts like a magnetic field.
This effect is also explained by special relativity, and it leads to the so-called
spin-orbit coupling, i.e. an influence of the electron’s orbital motion around
the nucleus onto its spin.

Both of these features (and more) are incorporated in the relativistic Dirac
equation, whose predictions come still closer to experiment. It can still be
solved exactly for the hydrogen atom. The resulting states now must be clas-
sified by the total angular momentum number j (arising through the coupling
between electron spin and orbital angular momentum). States of the same j
and the same n are still degenerate.

93 http://www.chem1.com/acad/webtut/atomic/qprimer/#Q26

Hydrogen atom
186

• There are always vacuum fluctuations of the electromagnetic field, accord-


ing to quantum mechanics. This means in particular that the electron un-
dergoes a kind of "jitter" motion. As a consequence, the degeneracy be-
tween states of the same j but different l is lifted. This has been demonstrat-

FT
ed in the famous Lamb-Rutherford experiment and was the starting point
for the development of the theory of Quantum electrodynamics (which is
able to deal with these vacuum fluctuations and employs the famous Feyn-
man diagrams for approximations using perturbation theory). This effect is
now called Lamb shift.

For these developments, it was essential that the solution of the Dirac equation
for the hydrogen atom could be worked out exactly, such that any experimen-
tally observed deviation had to be taken seriously as a signal of failure of the
theory.

See also
• Hydrogen
• quantum mechanics


A quantum chemistry
quantum field theory
• quantum state

References
• Griffiths, David (1995). Introduction to Quantum Mechanics. Upper Saddle
River, NJ: Prentice Hall. ISBN 0-13-111892-7.
DR
Section 4.2 deals with the hydrogen atom specifically, but all of Chapter 4 is
relevant.

• Bransden, B.H.; C.J. Joachain (1983). Physics of Atoms and Molecules.


London: Longman. ISBN 0-582-44401-2.

External links
• Physics of hydrogen atom on Scienceworld 94
• Interactive graphical representation of orbitals 95
• Applet which allows viewing of all sorts of hydrogenic orbitals 96

94 http://scienceworld.wolfram.com/physics/HydrogenAtom.html
95 http://webphysics.davidson.edu/faculty/dmb/hydrogen/
96 http://www.falstad.com/qmatom/

Hydrogen atom
187

Source: http://en.wikipedia.org/wiki/Hydrogen_atom

Principal Authors: Oo64eva, DÅ‚ugosz, Edsanville, The Anome, Anville, Michael Hardy, John C PI,
WAS 4.250, Jsalazar, Karol Langner

FT
Hydrogen-like atom

Hydrogen-like atoms (or hydrogenic atoms) are atoms with one single elec-
tron. Like the hydrogen atom, hydrogen-like atoms are one of the few quan-
tum mechanical problems which can be exactly solved. Atoms or ions whose
valence shell is made of one single electron (e.g. alkali metals) have similar
chemical bonding or spectroscopic propreties to hydrogen-like atoms.
The simplest atomic orbitals are those that occur in an atom with a single
electron, such as the hydrogen atom. In this case the atomic orbitals are the
eigenstates of the hydrogen Hamiltonian. They can be obtained analytically
(see →Hydrogen atom). An atom of any other element ionized down to a
A
single electron is very similar to hydrogen, and the orbitals take the same form.
For atoms with two or more electrons, the governing equations can only be
solved with the use of methods of iterative approximation. Orbitals of multi-
electron atoms are qualitatively similar to those of hydrogen, and in the sim-
plest models, they are taken to have the same form. For more rigorous and
precise analysis, numerical approximations must be used. Atomic orbitals
are often expanded in a basis set of Slater-type orbitals which are orbitals of
hydrogen-like atoms with arbitrary nuclear charge Z.
DR
A given (hydrogen-like) atomic orbital is identified by unique values of three
quantum numbers: n, l, and m l. The rules restricting the values of the quantum
numbers, and their energies (see below), explain the electron configuration of
the atoms and the periodic table.
The stationary states (quantum states) of the hydrogen-like atoms are its atom-
ic orbital. However, in general, an electron’s behavior is not fully described by
a single orbital. Electron states are best represented by time-depending "mix-
tures" (linear combinations) of multiple orbitals. See Linear combination of
atomic orbitals molecular orbital method.
The quantum number n first appeared in the →Bohr model. It determines,
among other things, the distance of the electron from the nucleus; all electrons
with the same value of n lay at the same distance. Modern quantum mechanics
confirms that these orbitals are closely related. For this reason, orbitals with

Hydrogen-like atom
188

the same value of n are said to comprise a "shell". Orbitals with the same value
of n and also the same value of l are even more closely related, and are said to
comprise a "subshell".

Mathematical characterization

FT
Derivation
Atomic orbitals are solutions to the →Schrödinger equation. In this case, the
potential term is the potential given by Coulomb’s law:
1 Ze2
V = − 4π0 r

where

• The first term is a constant, usually abbreviated by the letter k,


• Z is the atomic number,
• e is the elementary charge,
• r is the magnitude of the distance from the nucleus.
A
The wavefunction is a function of three spatial variables, so that after removing
the time-dependence, the Schrödinger equation is a partial differential equa-
tion in three variables (see separation of variables). However, since the poten-
tial is spherically symmetric, it is profitable to write the equation in spherical
coordinates. In this form, any individual eigenfunction ψ can be written as a
product of three single-variable functions, often denoted as follows:
ψ(r, θ, φ) = R(r)f (θ)g(φ)
DR
(where θ represents the polar angle (colatitude) and φ the azimuthal angle.) It
can further be reduced to three separate equations, each in one variable.
Two separations are required, resulting in two separation constants. A third
arbitrary constant results from the application of boundary conditions to R.
The equations given below use a form of the separation constants that seems
arbitrary, but it simplifies matters later on.
 
1 d
r 2 dR + 2µr2 (E − V (r)) = l(l + 1)
R(r) dr dr ~2

2
1 d g(φ)
g(φ) dφ2
= −m2
h i
sin(θ) d df
l(l + 1) sin2 (θ) + f (θ) dθ
sin(θ) dθ = m2

where:

Hydrogen-like atom
189

h
• h is the reduced Planck constant ( 2π ), and
• µ is the reduced mass of the electron vis-à-vis the nucleus.

Results

FT
In addition to ` and m, a third arbitrary integer, called n, emerges from the
boundary conditions placed on R. The functions R, f and g that solve the equa-
tions above depend on the values of these integers, called quantum numbers.
As a result, it is customary to subscript the functions with the values of the
quantum numbers they depend on. The forms of the functions are:
ψ = Cnlm Rnl (r) flm (θ) gm (φ)

Zr l
− na
  
2Zr 2l+1 2Zr
Rnl (r) = e µ
naµ Ln−l−1 naµ

il+|m|
(sin θ)|m|
h
d
flm (θ) = 2l l! d(cos θ)
(cos2 (θ) − 1)l (the associated Legendre
functions)

gm (φ) = eimφ
A
n, l, m are integer numbers, with the restrictions:
n = 1, 2, 3, ...

l = 0, 1, 2, ..., n − 1

m = −l, −(l − 1), ..., 0, ..., l − 1, l


DR
where:

• L2l+1
n−l−1 are the generalized Laguerre polynomials.
• aµ is defined by:
4πε0 ~2
aµ = µe2

Note that aµ is approximately equal to a0 (the Bohr radius)

• i is the imaginary number.

• Cnlm is a normalization constant. Since the wavefunction must be normal-


R
ized ( R3
|Ψ|2 d3~r = 1 ), by calculating this integral we get:

Hydrogen-like atom
190
 3 1/2
2Z (n−l−1)!
Cnlm = naµ 2n(n+l)!

The functions f and g are usually consolidated into the function Y (θ, φ) =

FT
f (θ)g(φ). This function is a spherical harmonic.
Thus, the complete expression for the normalized wavefunctions is:
r 3
(n−l−1)! −Zr/naµ 2Zr l 2l+1
   
2Z
ψnlm (θ, φ, r) = naµ 2n[(n+l)!]
e naµ Ln−l−1 2Zr
naµ ·
Yl,m (θ, φ)

Angular momentum
Each atomic orbital is associated with an angular momentum L. It is a vector,
and its magnitude is given by:
p
|L| = ~ l(l + 1)

The projection of this vector onto any arbitrary direction is quantized. If the
A
arbitrary direction is called z, the quantization is given by:
Lz = ml ~

where ml is restricted as described above. This value is always less than the
total angular momentum. Thus, if the L-vector is measured in some direction,
it will not lie entirely in that direction; part of it will lie in perpendicular direc-
tions. This allows the uncertainty principle to stand. It mandates that no two
components of L may be known at once. If one component were known to be
DR
equal to the total L, the other two would necessarily be zero.
These two relations do not give the total angular momentum of the electron.
For that, electron spin must be included.
This quantization of angular momentum closely parallels that proposed by
Niels Bohr (see →Bohr model) in 1913, with no knowledge of wavefunctions.

See also
• Positronium
• Exotic atom

Hydrogen-like atom
191

References
• Tipler, Paul & Ralph Llewellyn (2003). Modern Physics (4th ed.). New
York: W. H. Freeman and Company. ISBN 0-7167-4345-0

FT
Source: http://en.wikipedia.org/wiki/Hydrogen-like_atom

Principal Authors: Pfalstad, John C PI, Eequor, JimR, Achoo5000

Hyperfine structure

In atomic physics, hyperfine structure is a small perturbation in the energy


levels (or spectra) of atoms or molecules due to the magnetic dipole-dipole
interaction, arising from the interaction of the nuclear magnetic dipole with
the magnetic field of the electron.
A
Theory
According to classical thinking, the electron moving around the nucleus has a
magnetic dipole moment, because it is charged. The interaction of this mag-
netic dipole moment with the magnetic moment of the nucleus (due to its spin)
leads to hyperfine splitting.
However, due to the electron’s spin, there is also hyperfine splitting for s-shell
electrons, which have zero orbital angular momentum. In this case, the mag-
netic dipole interaction is even stronger, as the electron probability density
DR
does not vanish inside the nucleus (r = 0).
The amount of correction to the Bohr energy levels due to hyperfine splitting
of the hydrogen atom is on the order of:
m 4 2
mp α mc

where
m is the mass of an electron

m p is the mass of a proton

α is the fine structure constant (1/137.036 )

Hyperfine structure
192

c is the speed of light.

For atoms other than hydrogen, the nuclear spin I~ and the total electron angu-
lar momentum J~ = L ~ +S~ get coupled, giving rise to the total angular momen-

FT
~ ~ ~
tum F = J + I. The hyperfine splitting is then
∆Ehf s = −~
µI B~ J = a [F (F + 1) − I(I + 1) − J(J + 1)],
2

where
~
a = √gI µ~ N BJ ,
J(J+1)

with µ
~ N the magnetic dipole moment of the nucleus.
This interaction obeys the Lande interval rule: The energy level is split into
(J + I) − |J − I| + 1 energy levels, where J denotes the total electron angular
momentum and I denotes the nuclear spin.
With ∆Ehf s ≈ ~, the hyperfine splitting is a much smaller perturbation than
the fine structure.
A
In a more advanced treatment, one also has to take the nuclear magnetic
quadrupole moment into account. This is sometimes (?) referred to as "hy-
perfine structure anomaly".

History
The optical hyperfine structure was already observed in 1881 by Albert Abra-
ham Michelson. It could, however, only be explained in terms of quantum
mechanics in the 1920s. Wolfgang Pauli proposed the existence of a small
DR
nuclear magnetic moment in 1924.
In 1935, M. Schiiler and T. Schmidt proposed the existence of a nuclear
quadrupole moment in order to explain anomalies in the hyperfine structure.

Applications
Astrophysics
As the hyperfine splitting is very small, the transition frequencies usually are
not optical, but in the range of radio- or microwave frequencies.
Hyperfine structure gives the 21 cm line observed in HI region in interstellar
medium.
Carl Sagan and Frank Drake considered the hyperfine transition of hydrogen to
be a sufficiently universal phenomenon so as to be used as a base unit of time
and length on the Pioneer plaque and later Voyager Golden Record.

Hyperfine structure
193

Nuclear technology
The AVLIS and MLIS processes use hyperfine splitting caused by differences be-
tween the mass of atomic nucleus for uranium-235 and uranium-238 to selec-
tively photoionize only the uranium-235 atoms and then separate the ionized

FT
particles from the non-ionized ones. Precisely tuned dye lasers are used as the
sources of the necessary exact wavelength radiation.

Use in defining the SI second and meter


The hyperfine structure transition can be used to make a microwave notch
filter with very high stability, repeatability and Q factor, which can thus be
used as a basis for very precise atomic clocks. Typically, the hyperfine structure
transition frequency of a particular isotope of caesium or rubidium atoms is
used as a basis for these clocks.
Due to the accuracy of hyperfine structure transition-based atomic clocks, they
are now used as the basis for the definition of the second. One second is now
defined to be exactly 9,192,631,770 cycles of the hyperfine structure transition
frequency of caesium-133 atoms.
A
Since 1983, the meter is defined by declaring the speed of light in a vacuum to
be exactly 299,792,458 metres per second. Thus:
The metre is the length of the path travelled by light in vacuum during a time
interval of 1/299 792 458 of a second.

Qubit in ion-trap quantum computing


The hyperfine states of a trapped ion are commonly used for storing qubits in
ion-trap quantum computing. They have the advantage of having a very long
DR
lifetimes, experimentally exceeding 10 min (compared to 1 s for metastable
electronic levels).
The frequency associated with the states’ energy separation is in the microwave
region, making it possible to drive hyperfine transitions using microwave radia-
tion. However, at present no emitter is available that can be focused to address
a particular ion from a sequence. Instead, a pair of laser pulses can be used
to drive the transition, by having their frequency difference (detuning) equal
to the required transition’s frequency. This is essentially a stimulated Raman
transition.

See also
• Energy levels
• Quantum numbers

Hyperfine structure
194

Source: http://en.wikipedia.org/wiki/Hyperfine_structure

Principal Authors: The Anome, LAk loho, ArnoldReinhold, Icairns, Linas

FT
Imaginary time

Imaginary time is a concept derived from quantum mechanics. It is used


to describe models of the universe in physical cosmology. Stephen Hawking
popularized the concept of imaginary time in his book A Brief History of Time.
Imaginary time is difficult to visualize. If we imagine "regular time" as a hor-
izontal line with "past" on one side and "future" on the other, then imaginary
time would run perpendicular to this line as the imaginary numbers run per-
pendicular to the real numbers in the complex plane. However, imaginary time
is not imaginary in the sense that it is unreal or made-up—it simply runs in a
direction different from the type of time we experience. In essence, imaginary
time is a way of looking at the time dimension as if it were a dimension of
A
space: you can move forward and backward along imaginary time, just like
you can move right and left in space.
The concept is useful in cosmology because it can help smooth out gravitational
singularities in models of the universe (see Hartle-Hawking state. Singularities
pose a problem for physicists because they are areas where known physical
laws do not apply. The Big Bang, for example, appears as a singularity in "reg-
ular time." But when visualized with imaginary time, the singularity is removed
and the Big Bang functions like any other point in spacetime.
DR
The No-Boundary Universe and imaginary time
To further illustrate this concept, imagine spacetime as the surface of the Earth,
with all three space dimensions combined into the "east-west" axis, and the
imaginary time dimension running along the "north-south" axis. In this model,
spacetime is both finite and boundless—like the surface of the Earth, it has a
finite area, but lacks any edge or boundary.
The "North Pole" in this model would be analogous to the Big Bang. It is the
"northernmost" point on the surface of the earth, just as it is the "earliest" point
of time in the universe. But in the imaginary time/space cosmology, the Big
Bang/north pole is not a singularity; it is a point of spacetime just like any
other. If you travel "north" or "before" the Big Bang, physics would not break
down, you would simply encounter more of the universe.

Imaginary time
195

It must be noted that this globe visualization of the No-Boundary Universe is


not without faults. The visualization suggests that space will expand until it
reaches an "equator" and then contract, due to gravity, until it becomes the
South Pole, or the Big Crunch. However, current models of the universe sug-

FT
gest that space will continue to expand faster and faster. Imaginary time, as a
quantum mechanics concept, is perhaps more accurately integrated with cos-
mology via the wavefunction model of the universe.

See also
• Wick rotation
• Euclidean quantum gravity

External links
• The Beginning of Time 97 — Lecture by Stephen Hawking which discusses
imaginary time.
• Stephen Hawking’s Universe: Strange Stuff Explained 98 — PBS site on
A imaginary time

Source: http://en.wikipedia.org/wiki/Imaginary_time

Implicate and Explicate Order according to


David Bohm
DR
The physicist David Bohm proposed a conception of order which is radically
different from most conceptions of order, and in doing so made a distinction
between the Implicate and Explicate Order, which he characterised as fol-
lows:
In the enfolded [or Implicate] order, space and time are no longer the dom-
inant factors determining the relationships of dependence or independence
of different elements. Rather, an entirely different sort of basic connection
of elements is possible, from which our ordinary notions of space and time,
along with those of separately existent material particles, are abstracted as
forms derived from the deeper order. These ordinary notions in fact appear

97 http://www.hawking.org.uk/text/public/bot.html
98 http://www.pbs.org/wnet/hawking/strange/html/imaginary.html

Implicate and Explicate Order according to David Bohm


196

in what is called the "explicate" or "unfolded" order, which is a special and


distinguished form contained within the general totality of all the Implicate
Orders (Bohm, 1980, p. xv).

FT
David Bohm’s challenges to some generally prevail-
ing views
In proposing this new notion of order, Bohm explicitly challenged a number of
tenets that are fundamental to much scientific work. The tenets challenged by
Bohm include:

• That phenomena are reducible to fundamental particles and laws describing


the behaviour of particles, or more generally to any static (i.e. unchanging)
entities, whether separate events in space-time, quantum states, or static
entities of some other nature.
• Related to (1), that human knowledge is most fundamentally concerned
with mathematical prediction of statistical aggregates of particles.
• That an analysis or description of any aspect of reality (e.g. quantum theory,


Athe speed of light) can be unlimited in its domain of relevance.
That the Cartesian coordinate system, or its extension to a curvilinear sys-
tem, is the deepest conception of underlying order as a basis for analysis
and description of the world.
• That there is ultimately a sustainable distinction between reality and
thought, and that there is a corresponding distinction between the observer
and observed in an experiment or any other situation (other than a dis-
tinction between relatively separate entities valid in the sense of Explicate
DR
Order).
• That it is, in principle, possible to formulate a final notion concerning the
nature of reality; e.g. a Theory of Everything.

Bohm’s proposals have at times been ’dismissed’ largely on the basis of such
tenets, without due consideration necessarily given to the fact that they had
been challenged by Bohm.
Bohm’s paradigm is inherently antithetical to reductionism, in most forms, and
accordingly can be regarded as a form of ontological holism. On this, Bohm
noted of prevailing views among physicists: "the world is assumed to be consti-
tuted of a set of separately existent, indivisible and unchangeable ’elementary
particles’, which are the fundamental ’building blocks’ of the entire universe · · ·
there seems to be an unshakable faith among physicists that either such parti-
cles, or some other kind yet to be discovered, will eventually make possible a
complete and coherent explanation of everything" (Bohm, 1980, p. 173).

Implicate and Explicate Order according to David Bohm


197

verse

FT
Figure 16 A Helium Atom and its constituent particles: an exam-
ple of a small collection of the posited building blocks of the uni-

In Bohm’s conception of order, then, primacy is given to the undivided whole,


A
and the Implicate Order inherent within the whole, rather than to parts of the
whole, such as particles, quantum states, and continua. For Bohm, the whole
encompasses all things, structures, abstractions and processes, including pro-
cesses that result in (relatively) stable structures as well as those that involve
metamorphosis of structures or things. In this view, parts may be entities nor-
mally regarded as physical, such as atoms or sub-atomic particles, but they
may also be abstract entities, such as quantum states. Whatever their nature
and character, according to Bohm, these parts are considered in terms of the
DR
whole, and in such terms, they constitute relatively autonomous and indepen-
dent "sub-totalities". The implication of the view is, therefore, that nothing is
entirely separate or autonomous.
Bohm (1980, p. 11) said: "The new form of insight can perhaps best be called
Undivided Wholeness in Flowing Movement. This view implies that flow is, in
some sense, prior to that of the ‘things’ that can be seen to form and dissolve
in this flow". According to Bohm, a vivid image of this sense of analysis of
the whole is afforded by vortex structures in a flowing stream. Such vortices
can be relatively stable patterns within a continuous flow, but such an analysis
does not imply that the flow patterns have any sharp division, or that they
are literally separate and independently existent entities; rather, they are most
fundamentally undivided. Thus, according to Bohm’s view, the whole is in
continuous flux, and hence is referred to as the holomovement (movement of
the whole).

Implicate and Explicate Order according to David Bohm


198

A
Quantum theory and relativity theory
FT
Figure 17 A vortex in a stream - a relatively stable pat-
tern which occurs within a continuous flow of liquid

A key motivation for Bohm in proposing a new notion of order was what he
saw as the incompatibility of quantum theory with relativity theory, in terms of
certain features of the theoris as observed in relevant experimental contexts.
Bohm (1980, p. xv) summarised the state of affairs he perceived to exist in the
DR
following terms:
· · ·in relativity, movement is continuous, causally determinate and well
defined, while in quantum mechanics it is discontinuous, not causally
determinate and not well-defined. Each theory is committed to its own
notions of essentially static and fragmentary modes of existence (relativity
to that of separate events connectible by signals, and quantum mechanics
to a well-defined quantum state). One thus sees that a new kind of theory
is needed which drops these basic commitments and at most recovers some
essential features of the older theories as abstract forms derived from a
deeper reality in which what prevails is unbroken wholeness.

Bohm maintained that relativity and quantum theory are in basic contradic-
tion in these essential respects, and that a new notion of order should begin
with that which both point toward: undivided wholeness. This should not be

Implicate and Explicate Order according to David Bohm


199

taken, however, to imply that he considered such powerful theories should be


discarded. Nevertheless, he argued that each was relevant in a certain con-
text - i.e. a set of interrelated conditions within the Explicate Order - rather
than having unlimited relevance, and that apparent contradictions stem from

FT
attempts to overgeneralize by superposing the theories on one another, imply-
ing greater generality or broader relevance than is ultimately warranted. Thus,
Bohm (1980, pp. 156-167) argued: "... in sufficiently broad contexts such
analytic descriptions cease to be adequate ... ’the law of the whole’ will gen-
erally include the possibility of describing the ’loosening’ of aspects from each
other, so that they will be relatively autonomous in limited contexts ... how-
ever, any form of relative autonomy (and heteronomy) is ultimately limited by
holonomy, so that in a broad enough context such forms are seen to be mere-
ly aspects, relevated in the holomovement, rather than disjoint and separately
existent things in interaction".

Hidden variable quantum theory


Bohm proposed a hidden variable theory of quantum physics (see Bohm inter-
pretation). According to Bohm, a key motivation for doing so was purely to
A
show the possibility of such theories. On this, Bohm (1980, p. 81) said "... it
should be kept in mind that before this proposal was made there had existed
the widespread impression that no conceptions of hidden variables at all, not
even if they were abstract, and hypothetical, could possibly be consistent with
the quantum theory". Bohm (1980, p. 110) also claimed that "the demon-
stration of the possibility of theories of hidden variables may serve in a more
general philosophical sense to remind us of the unreliability of conclusions
based on the assumption of the complete universality of certain features of a
DR
given theory, however general their domain of validity seems to be". Another
aspect of Bohm’s motivation was to point out a confusion he perceived to exist
in quantum theory. On the dominant approaches in quantum theory, he said:
"...we wish merely to point out that this whole line of approach re-establishes
at the abstract level of statistical potentialities the same kind of analysis into
separate and autonomous components in interaction that is denied at the more
concrete level of individual objects".

Quantum entanglement
Central to Bohm’s schema are correlations between observables of entities
which seem separated by great distances in the Explicate Order (such as a
particular electron here on earth and an alpha particle in one of the stars in the
Abell 1835 galaxy, the farthest galaxy from Earth known to humans), manifes-
tations of the Implicate Order. Within quantum theory there is entanglement
of such objects.

Implicate and Explicate Order according to David Bohm


200

This view of order necessarily departs from any notion which entails signalling,
and therefore causality. The correlation of observables does not imply a causal
influence, and in Bohm’s schema the latter represents ’relatively’ independent
events in space-time; and therefore Explicate Order.

FT
He also used the term unfoldment to characterise processes in which the Ex-
plicate Order becomes relevant (or "relevated"). Bohm likens unfoldment also
to the decoding of a television signal to produce a sensible image on a screen.
The signal, screen, and television electronics in this analogy represent the Im-
plicate Order whilst the image produced represents the Explicate Order. He
also uses an interesting example in which an ink droplet can be introduced
into a highly viscous substance (such as glycerine), and the substance rotat-
ed very slowly such that there is negligible diffusion of the substance. In this
example, the droplet becomes a thread which, in turn, eventually becomes in-
visible. However, by rotating the substance in the reverse direction, the droplet
can essentially reform. When it is invisible, according to Bohm, the order of
the ink droplet as a pattern can be said to be implicate within the substance.
In another analogy, Bohm asks us to consider a pattern produced by making
small cuts in a folded piece of paper and then, literally, unfolding it. Widely
A
separated elements of the pattern are, in actuality, produced by the same origi-
nal cut in the folded piece of paper. Here the cuts in the folded paper represent
the Implicate Order and the unfolded pattern represents the Explicate Order.

The hologram as analogy for the Implicate Order


Bohm employed the hologram as a means of characterising Implicate Order,
noting that each region of a photographic plate in which a hologram is ob-
DR
servable contains within it the whole three-dimensional image, which can be
viewed from a range of perspectives. That is, each region contains a whole
and undivided image. In Bohm’s words: "There is the germ of a new notion
of order here. This order is not to be understood solely in terms of a regular
arrangement of objects (eg., in rows) or as a regular arrangement of events
(e.g. in a series). Rather, a total order is contained, in some implicit sense, in
each region of space and time. Now, the word ’implicit’ is based on the verb ’to
implicate’. This means ’to fold inward’ ... so we may be led to explore the no-
tion that in some sense each region contains a total structure ’enfolded’ within
it". (Bohm, 1980, p. 149). Bohm noted that although the hologram conveys
undivided wholeness, it is nevertheless static.
In this view of order, laws represent invariant relationships between explicate
entities and structures, and thus Bohm maintained that in physics, the Expli-
cate Order generally reveals itself within well-constructed experimental con-
texts as, for example, in the sensibly observable results of instruments. With

Implicate and Explicate Order according to David Bohm


201

A
contains the whole image
FT
Figure 18 In a holographic reconstruction, each region of a photographic plate

respect to Implicate Order, however, Bohm (1980, p. 147) asked us to consider


the possibility instead "that physical law should refer primarily to an order of
undivided wholeness of the content of description similar to that indicated by
DR
the hologram rather than to an order of analysis of such content into separate
parts · · ·".

A common grounding for consciousness and matter


The Implicate Order represents the proposal of a general metaphysical concept
in terms of which it is claimed that matter and consciousness might both be
understood, in the sense that it is proposed that both matter and conscious-
ness: (i) enfold the structure of the whole within each region, and (ii) involve
continuous processes of enfoldment and unfoldment. For example, in the case
of matter, entities such as atoms may represent continuous enfoldment and un-
foldment which manifests as a relatively stable and autonomous entity that can
be observed to follow a relatively well-defined path in space-time. In the case
of consciousness, Bohm pointed toward evidence presented by Karl Pribram

Implicate and Explicate Order according to David Bohm


202

A FT
Figure 19 Karl Pribram and colleagues have presented evidence which indicates that memories
do not in general appear to be localized in specific regions of brains

that memories may be enfolded within every region of the brain rather than
being localized (for example in particular regions of the brain, cells, or atoms).
Bohm (1980, p. 205) went on to say: "As in our discussion of matter in gen-
eral, it is now necessary to go into the question of how in consciousness the
Explicate Order is what is manifest ... the manifest content of consciousness
is based essentially on memory, which is what allows such content to be held
DR
in a fairly constant form. Of course, to make possible such constancy it is also
necessary that this content be organized, not only through relatively fixed as-
sociation but also with the aid of the rules of logic, and of our basic categories
of space, time causality, universality, etc. ... there will be a strong background
of recurrent stable, and separable features, against which the transitory and
changing aspects of the unbroken flow of experience will be seen as fleeting
impressions that tend to be arranged and ordered mainly in terms of the vast
totality of the relatively static and fragmented content of [memories]". Bohm
also claimed that "as with consciousness, each moment has a certain Explicate
Order, and in addition it enfolds all the others, though in its own way. So the
relationship of each moment in the whole to all the others is implied by its total
content: the way in which it ’holds’ all the others enfolded within it". Bohm
characterises consciousness as a process in which at each moment, content that
was previously implicate is presently explicate, and content which was previ-
ously explicate has become implicate. He said: "One may indeed say that our

Implicate and Explicate Order according to David Bohm


203

memory is a special case of the process described above, for all that is recorded
is held enfolded within the brain cells and these are part of matter in general.
The recurrence and stability of our own memory as a relatively independent
sub-totality is thus brought about as part of the very same process that sus-

FT
tains the recurrence and stability in the manifest order of matter in general.
It follows, then, that the explicate and manifest order of consciousness is not
ultimately distinct from that of matter in general" (Bohm, 1980, p. 208).

Connections with other works


Many, along with Bohm himself, have seen strong connections between his
ideas and ideas from the East. There are particularly strong connections to
Buddhism, for which Einstein also shared sympathy. Some proponents of al-
ternative religions (such as shamanism) claim a connection with their belief
systems as well.
Bohm may have known that his idea is a striking analogy to "intensional and
extensional aboutness" to which R. A. Fairthorne (1969) insightfully referred
information scientists (although a Google search reveals that few paid atten-
tion to this suggestion). John Searle treated aboutness and network in his
A
Intentionality (1983), contemporarily with Bohm’s Wholeness (1983)! Sear-
le’s concept of aboutness is in sharp contrast to, and is as odd as Bohm’s idea
of wholeness. As the former is to the content, so the latter is to the context as
the ultimate determiner of meaning. The holistic view of context, hence an-
other striking analogy of wholeness, was first put forward in The Meaning of
Meaning by C. K. Ogden & I. A. Richards (1923), including the literary, psycho-
logical, and external. These are respectively analogous to Karl Popper’s world
3, 2, and 1 appearing in his Objective Knowledge (1972 and later ed.). Bohm’s
DR
worldview of "undivided wholeness" is contrasted with Popper’s three divided
worlds. The direct causality among these and other authorships may be ac-
tually evident in the Implicate Order, though apparently not in the Explicate
Order in spite of a great deal of reasonable doubt in terms of locality, ethnicity,
ideology, academic tendency, and so on. Bohm and Popper favored Einstein
above all.
Suppose that someone intends to convey a definite thought or story with the
following word string:
woman, street, crowd, traffic, noise, haste, thief, bag, loss, scream, police,
.....

which looks almost non-sensical as a whole. Then, what will happen to us lis-
teners? We have a dictionary, but we cannot simply sum up the meanings of
individual words. That "a whole is more than the sum of the parts" is too plain

Implicate and Explicate Order according to David Bohm


204

a saying. There seems to be no grammar to which the speaker might have con-
formed. He merely suggests rather than tells the story, which in other words is
implied or implicit in the word string. From this awkward symbology we can
guess the story with varying accuracies, if we are ready to take risks. In this

FT
case, the meaning of such symbology may be said to be connotative, implicit,
implicate or intensional, in contrast to denotative, explicit, explicate or exten-
sional. Consult a dictionary for these words. Note that the more context that
unfolds, the less uncertainty remains folded. Most importantly, note that inter-
pretation or making sense of Explicate in Implicate Order, that is, aboutness in
wholeness or in context is an outstanding analogy as well as the very principle
of subject indexing as a prerequisite of information retrieval that has now be-
come an everyday concern. This principle’s actual implication for and impact
on a number of other disciplines should be unfolded if any. Why not unfold
who on earth played an inspiring or leading role in shaping contextualism in
the spotlight?
A
DR

Figure 20 Cells stained for keratin and DNA: such parts of life exist
because of the whole, but also in order to sustain it

Bohm’s views also connect with those of Immanuel Kant in some key respects.
For example, Kant held that the parts of an organism, such as cells, simultane-
ously exist in order to sustain the whole, and depend upon the whole for their
own existence and functioning. Also, as noted by Bohm, Kant proposed that
the process of thought plays an active role in organizing knowledge, which
implies theoretical insights are instrumental to the process of acquiring factual

Implicate and Explicate Order according to David Bohm


205

knowledge. This perspective is also congruent with an analysis of the function


of measurement in physical science by Thomas Kuhn in 1961.

A FT
Figure 21 For Bohm, life is a continuous flowing process of enfoldment and unfoldment involving
relatively autonomous entities. DNA ’directs’ the environment to form a living thing. Life can be
said to be implicate in ensembles of atoms that ultimately form life.

There are also connections to views expressed by Stuart Kauffman, who not-
ed Kant’s perspective on organisms in his book At Home in the Universe in a
section given the evocative title An Unrepentant Holism. Kauffman’s concept
DR
of an autocatalytic set, as it was originally conceived in terms of molecules,
elaborates on Kant’s perspective in terms of modern scientific concepts. In his
later book Investigations, Kauffman attempts to define, or at least character-
ize, the notion of an autonomous agent. If viewed as "relatively autonomous",
this concept is potentially congruous with Bohm’s view. Bohm’s views are also
echoed in Kauffman’s (2000, p. 137) statement: "... our incapacity to prestate
the configuration space of the biosphere is not a failure to prestate the conse-
quences of the primitives, it appears to be a failure to prestate the primitives
themselves". Kauffman suggests that such a failure may stem from more gen-
erally applicable foundations applicable also within physics. Consistent with
Bohm, this potentially calls into question whether we should presuppose that
it is possible (even in principle) to formulate a final and complete theory of
everything.

Implicate and Explicate Order according to David Bohm


206

See also
• Brahman
• Buddhism

FT
• Holographic principle
• Immanuel Kant
• Laminar flow
• Mind’s eye
• Noumenon
• Parable of the cave
• Plato
• Samsara
• Taoism
• Unobservables

References
• Bohm, D. (1980). Wholeness and the Implicate Order. London: Routledge.
A ISBN 0-710-00971-2

• Kauffman, S. (1995). At Home in the Universe. New York: Oxford Universi-


ty Press. hardcover: ISBN 0-19-509599-5, paperback ISBN 0-19-511130-3

• Kauffman, S. (2000). Investigations. New York: Oxford University Press.

• Kuhn, T.S. (1961). The function of measurement in modern physical sci-


ence. ISIS, 52, 161-193.
DR
Further reading
• Talbot, M. (1991). The Holographic Universe. Harpercollins

External links
• Interview with David Bohm 99 – An interview with Bohm concerning this
particular subject matter conducted by F. David Peat.
• Excerpt from The Holographic Universe 100 – Parallels some of the experi-
ences of 18th century Swedish mystic, Emanuel Swedenborg, with David
Bohm’s ideas.

99 http://www.fdavidpeat.com/interviews/bohm.htm
100 http://www.soultravel.nu/2004/040907-swedenborg/index.asp

Implicate and Explicate Order according to David Bohm


207

• Mohamad Latiff’s Paradigm of Everything 101 - An Essay by a young Bohm-


enthusiast concerning the Interconnectedness of Everything using the Bohm
model using a practical analogy.

FT
Source: http://en.wikipedia.org/wiki/Implicate_and_Explicate_Order_according_to_David_Bohm

Principal Authors: Goethean, Holon, Floorsheim, Togo, Ciprianman

Incompleteness of quantum physics

Incompleteness of quantum physics is the assertion that the state of a phys-


ical system, as formulated by quantum mechanics, does not give a complete
description for the system. A complete description is one which uniquely de-
termines the values of all its measurable properties. The existence of inde-
terminacy for some measurements is a characteristic of quantum mechanics;
moreover, bounds for indeterminacy can be expressed in a quantitative form
Aby the Heisenberg uncertainty principle.
Incompleteness can be understood in two fundamentally different ways:

• QM is incomplete because it is not the "right" theory; the right theory would
provide descriptive categories to account for all observable behavior and
not leave "anything to chance".
• QM is incomplete, but it accurately reflects the way nature is.

Incompleteness understood as 1) is now considered highly controversial, since


DR
it contradicts the impossibility of a hidden variables theory which is shown by
Bell test experiments. There are many variants of 2) which is widely considered
to be the more orthodox view of quantum mechanics.

Einstein’s argument for the incompleteness of quan-


tum physics
Albert Einstein may have been the first person to carefully point out the radical
effect the new quantum physics would have on our notion of physical state.
For a historical background of Einstein’s thinking in regard to QM, see Jankiw
and Kleppner [2000], although his best known critique was formulated in the
EPR thought experiment. See Bell [1964].

101 http://mohamadlatiff.blogspot.com/2005/08/mohamad-latiffs-paradigm-of-everything.html

Incompleteness of quantum physics


208

According to Fuchs [2002], Einstein developed a very good argument for in-
completeness:
The best [argument of Einstein] was in essence this. Take two spatially sep-
arated systems A and B prepared in some entangled quantum state |ψ AB >.

FT
By performing the measurement of one or another of two observables on
system A alone, one can immediately write down a new state for system B.
Either the state will be drawn from one set of states

A
DR

Incompleteness of quantum physics


209

’s argument shows that quantum state is not a complete description of a physical system,
according to Fuchs [2002]:
Thus one must take it seriously that the new state (either a |φ i B > or |η i B >) represents
information about system B. In making a measurement on A, one learns something about B,

FT
but that is where the story ends. The state change cannot be construed to be something more
physical than that. More particularly, the final state itself for B cannot be viewed as more
than a reflection of some tricky combination of one’s initial information and the knowledge
gained through the measurement. Expressed in the language of Einstein, the quantum state
cannot be a “complete” description of the quantum system.

Reality of incompleteness
Although Einstein was one of the first to formulate the necessary incompleteness of quantum
physics, he never fully accepted it. In a 1926 letter to Max Born, he made a remark that is now
famous:
Quantum mechanics is certainly imposing. But an inner voice tells me it is not yet the real
thing. The theory says a lot, but does not really bring us any closer to the secret of the Old
One. I, at any rate, am convinced that He does not throw dice.

Einstein was mistaken according to Steven Hawking in Does God Play Dice 102,
Einstein’s view was what would now be called, a hidden variable theory. Hidden variable
theories might seem to be the most obvious way to incorporate the Uncertainty Principle into
Aphysics. They form the basis of the mental picture of the universe, held by many scientists,
and almost all philosophers of science. But these hidden variable theories are wrong. The
British physicist, John Bell, who died recently, devised an experimental test that would
distinguish hidden variable theories. When the experiment was carried out carefully, the
results were inconsistent with hidden variables. Thus it seems that even God is bound by the
Uncertainty Principle, and can not know both the position, and the speed, of a particle. So
God does play dice with the universe. All the evidence points to him being an inveterate
gambler, who throws the dice on every possible occasion.

Chris Fuchs [2002] summed up the reality of the necessary incompleteness of information in
quantum physics as follows, attributing this idea to Einstein "He [Einstein] was the first person to
DR
say in absolutely unambiguous terms why the quantum state should be viewed as information
(or, to say the same thing, as a representation of one’s beliefs and gambling commitments,
credible or otherwise).
Fuchs adds:
Incompleteness, it seems, is here to stay: The theory prescribes that no matter how much we
know about a quantum system—even when we have maximal information about it—there will
always be a statistical residue. There will always be questions that we can ask of a system for
which we cannot predict the outcomes. In quantum theory, maximal information is simply not
complete information [Caves and Fuchs 1996]. But neither can it be completed.

The kind of information about the physical world that is available to us according to Fuchs [2002]
is “the potential consequences of our experimental interventions into nature” which is the subject
matter of quantum physics.

Relational Quantum Physics


According to Relational Quantum Physics [Laudisa and Rovelli 2005], the way distinct physical
systems affect each other when they interact (and not of the way physical systems "are") exhausts
all that can be said about the physical world. The physical world is thus seen as a net of
interacting components, where there is no meaning to the state of an isolated system. A physical
system (or, more precisely, its contingent state) is described by the net of relations it entertains
Incompleteness
with the surrounding systems, and the physicalof quantum
structure of thephysics
world is identified as this net of
relationships. In other words, “Quantum physics is the theoretical formalization of the
experimental discovery that the descriptions that different observers give of the same events are
not universal.”
210

Source: http://en.wikipedia.org/wiki/Incompleteness_of_quantum_physics

FT
Interaction picture

In quantum mechanics, the Interaction picture (or Dirac picture) is an in-


termediate between the →Schrödinger picture and the →Heisenberg picture.
Whereas in the other two pictures either the state vector or the operators carry
time dependence, in the interaction picture both carry part of the time depen-
dence of observables.
Warning: Operator equations which hold in the interaction picture don’t nec-
essarily hold in the Schrödinger or the Heisenberg picture. This is because the
operators A I, A S and A H are not the same but are related by unitary transfor-
mations. Unfortunately, most textbooks and articles omit the subscript which
can often lead to confusion and mistakes when an unwary student applies an
equation for one picture to another picture.
A
Switching pictures
To switch into the interaction picture, we divide the Schrödinger picture Hamil-
tonian into two parts, HS = H0,S + H1,S . Then the state vector is defined as
|ψI (t)i = eiH0,S t/~ |ψS (t)i

Operators transform between the pictures as


DR
AI = eiH0 t/~ AS e−iH0 t/~ .

The →Schrödinger equation then becomes in this picture:


d
i~ dt |ψI (t)i = H1,I |ψI (t)i.

This equation is referred to as the Schwinger- Tomonaga equation.


The purpose of this picture is to shunt all the time dependence due to H 0 onto
the operators, leaving only H 1, I affecting the time-dependence of the state
vectors.
The interaction picture is convenient when considering the effect of a small
interaction term, H 1, S, being added to the Hamiltonian of a solved system,
H 0, S. By switching into the interaction picture, you can use time-dependent
perturbation theory to find the effect of H 1, I.

Interaction picture
211

References
• Townsend, John S. (2000). A Modern Approach to Quantum Mechanics,
2nd ed.. Sausalito, CA: University Science Books. ISBN 1891389130.

FT
See also
• →Bra-ket notation
• →Schrödinger equation
• Haag’s theorem

Source: http://en.wikipedia.org/wiki/Interaction_picture

Principal Authors: Laurascudder, Phys, Elroch, Charles Matthews, Pjacobi

Internal conversion
A This article is about the nuclear process. For the chemical process, see
Internal conversion (chemistry).

Internal conversion is a radioactive decay process where an excited nucleus


interacts with an electron in one of the lower electron shells causing the elec-
tron to be emitted. This is not to be confused with a photoelectric effect where
a photon emitted from the nucleus interacts with the electron.
DR
The internal conversion process is not actually the photoelectric ejection of an
atomic electron, as the nucleus does not actually emit a gamma ray in the first
place in this process. What happens is that the wavefunction of an inner shell
electron penetrates the nucleus (ie there is a finite probability of the electron
being found in the nucleus) and when this is the case the electron takes the
energy of the nuclear transition without an intermediary gamma ray being
produced. The energy of the emitted electron is equal to the transition energy
minus the binding energy of the electron. Most internal conversion electrons
come from the K shell as this electron has the highest probability of being found
inside the nucleus.
After the electron has been emitted, the atom is left with a vacancy in one of
the inner electron shells. This hole will be filled with an electron from one of
the higher shells and subsequently a characteristic x-ray or →Auger electron
will be emitted.

Internal conversion
212

Internal conversion is favoured when the energy gap between nuclear levels
is small, and is also the only mode of de-excitation for 0+ -> 0+ (i.e. E0)
transitions. It is the predominant mode of de-excitation whenever the initial
and final spin states are the same, but the multi-polarity rules for nonzero

FT
initial and final spin states do not necessarily forbid the emission of a gamma
ray in such a case.
The tendency towards internal conversion can be determined by the internal
conversion coefficient, which is empirically determined by the ratio of de-
excitations that go by the emission of electrons to those that go by gamma
emission.

References
Krane, Kenneth S. (1988). Introductory Nuclear Physics. J. Wiley & Sons. ISBN
0-471-80553-X.

External links
• HyperPhysics 107
ASee also
• Internal conversion coefficient

Source: http://en.wikipedia.org/wiki/Internal_conversion

Principal Authors: Sunborn, GangofOne, Jpau, AjAldous, Bensaccount


DR
Interpretation of quantum mechanics

An interpretation of quantum mechanics is an attempt to answer the ques-


tion, What exactly is quantum mechanics talking about? The question has its
historical roots in the nature of quantum mechanics itself which was consid-
ered as a radical departure from previous physical theories. However, quantum
mechanics has been described as "the most precisely tested and most successful
theory in the history of science" (c.f. Jackiw and Kleppner, 2000.)

107 http://hyperphysics.phy-astr.gsu.edu/hbase/nuclear/radact2.html#c5

Interpretation of quantum mechanics


213

Historical background
The operational meaning of the technical terms used by researchers in quan-
tum theory (such as wavefunctions and matrix mechanics) progressed through
various intermediate stages. For instance Schrödinger originally viewed the

FT
wavefunction associated to the electron as the charge density of an object
smeared out over an extended, possibly infinite, volume of space. Max Born
later proposed its interpretation as the probability distribution in the space of
the electron’s position. Other leading scientists, such as Albert Einstein, had
great difficulty in accepting some of the more radical consequences of the the-
ory, such as quantum indeterminacy. Even if these matters could be treated as
’teething troubles’, they have lent importance to the activity of interpretation.
It should not, however, be assumed that most physicists consider quantum
mechanics as requiring interpretation, other than very minimal instrumentalist
interpretations, which are discussed below. The Copenhagen interpretation, as
of 2006, appears to be the most popular one among scientists, followed by the
many worlds and consistent histories interpretations. But it is also true that
most physicists consider non-instrumental questions (in particular ontological
Aquestions) to be irrelevant to physics. They fall back on Paul Dirac’s point
of view, later expressed in the famous dictum: "Shut up and calculate" often
(perhaps erroneously) attributed to Richard Feynman (see 108).

Obstructions to direct interpretation


The perceived difficulties of interpretation reflect a number of points about the
orthodox description of quantum mechanics, including:
DR
• The abstract, mathematical nature of the description of quantum mechan-
ics.
• The existence of what appear to be non-deterministic and irreversible pro-
cesses in quantum mechanics.
• The phenomenon of entanglement, and in particular, the higher correla-
tions between remote events than would be expected in classical theory.
• The complementarity of possible descriptions of reality.

First, the accepted mathematical structure of quantum mechanics is based on


fairly abstract mathematics, such as →Hilbert spaces and operators on those
Hilbert spaces. In classical mechanics and electromagnetism, on the other
hand, properties of a point mass or properties of a field are described by re-
al numbers or functions defined on two or three dimensional sets. These have

108 http://www.physicstoday.org/vol-57/iss-5/p10.html

Interpretation of quantum mechanics


214

direct, spatial meaning, and in these theories there seems to be less need to
provide a special interpretation for those numbers or functions.
Further, the process of measurement plays an apparently essential role in the
theory. It relates the abstract elements of the theory, such as the wavefunction,

FT
to operationally definable values, such as probabilities. Measurement inter-
acts with the system state, in somewhat peculiar ways, as is illustrated by the
double-slit experiment.
The mathematical formalism used to describe the time evolution of a non-
relativistic system proposes two somewhat different kinds of transformations:

• Reversible transformations described by unitary operators on the state


space. These transformations are determined by solutions to the
→Schrödinger equation.

• Non-reversible and unpredictable transformations described by mathemat-


ically more complicated transformations (see quantum operations). Exam-
ples of these transformations are those that are undergone by a system as a
result of measurement.
A
A restricted version of the problem of interpretation in quantum mechanics
consists in providing some sort of plausible picture, just for the second kind
of transformation. This problem may be addressed by purely mathematical
reductions, for example by the many-worlds or the consistent histories inter-
pretations.
In addition to the unpredictable and irreversible character of measurement pro-
cesses, there are other elements of quantum physics that distinguish it sharply
DR
from classical physics and which cannot be represented by any classical pic-
ture. One of these is the phenomenon of entanglement, as illustrated in the
EPR paradox, which seemingly violates principles of local causality.
Another obstruction to direct interpretation is the phenomenon of complemen-
tarity, which seems to violate basic principles of propositional logic. Comple-
mentarity says there is no logical picture (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 "com-
plementary" sets A and B of propositions that can describe S, but not at the
same time. Examples of A and B are propositions involving a wave descrip-
tion 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.

Interpretation of quantum mechanics


215

Complementarity is not usually taken to mean that classical logic fails, al-
though Hilary Putnam did take that view in his paper Is logic empirical?. In-
stead complementarity means that composition of physical properties for S
(such as position and momentum both having values in certain ranges) using

FT
propositional connectives does not obey rules of classical propositional logic.
As is now well-known (Omnès, 1999) the "origin of complementarity lies in the
noncommutativity of operators" describing observables in quantum mechanics.

Problematic status of pictures and interpretations


The precise ontological status, of each one of the interpreting pictures, 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, seem to argue that
an interpretation is nothing more than a formal equivalence between sets of
A
rules for operating on experimental data. This would suggest that the whole
exercise of interpretation is unnecessary.

Instrumentalist interpretation
Any modern scientific theory requires at the very least an instrumentalist de-
scription which relates the mathematical formalism to experimental practice
and prediction. In the case of quantum mechanics, the most common in-
strumentalist description is an assertion of statistical regularity between state
DR
preparation processes and measurement processes. That is, if a measurement
of a real-valued quantity is performed many times, each time starting with the
same initial conditions, the outcome is a well-defined probability distribution
over the real numbers; moreover, quantum mechanics provides a computation-
al 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 matrices acting on H : these are referred to as observables.
Repeated measurement of an observable A for S prepared in state ψ yields a
distribution of values. The expectation value of this distribution is given by the
expression

Interpretation of quantum mechanics


216

hψ|A|ψi.

This mathematical machinery gives a simple, direct way to compute a statis-


tical property of the outcome of an experiment, once it is understood how to

FT
associate the initial state with a vector, and the measured quantity with an
observable (that is, a specific Hermitian matrix).
As an example of such a computation, the probability of finding the system
in a given state |φi is given by computing the expectation value of a (rank-1)
projection operator
Π = |φihφ|

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


P = hψ|Π|ψi = |hψ|φi|2 .

By abuse of language, the bare instrumentalist description can be referred to


as an interpretation, although this usage is somewhat misleading since instru-
mentalism explicitly avoids any explanatory role; that is, it does not attempt to
A
answer the question of what quantum mechanics is talking about.

Summary of common interpretations of QM


Properties of interpretations
An interpretation can be characterized by whether it satisfies certain properties,
such as:
DR
• 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 consists of the Hilbert space machinery of ket-


vectors, self-adjoint operators acting on the space of ket-vectors, unitary
time dependence of ket-vectors and measurement operations. In this con-
text a measurement operation can be regarded as a transformation which

Interpretation of quantum mechanics


217

carries a ket-vector into a probability distribution on ket-vectors. See also


quantum operations for a formalization of this concept.
• The interpreting structure includes states, transitions between states, mea-
surement operations and possibly information about spatial extension of

FT
these elements. A measurement operation here refers to an operation which
returns a value and results in a possible system state change. Spatial infor-
mation, for instance would be exhibited by states represented as functions
on configuration space. The transitions may be non-deterministic or proba-
bilistic or there may be infinitely many states. However, the critical assump-
tion of an interpretation is that the elements of I are regarded as physically
real.

In this sense, an interpretation can be regarded as a semantics for the mathe-


matical formalism.
In particular, the bare instrumentalist view of quantum mechanics outlined in
the previous section is not an interpretation at all since it makes no claims
about elements of physical reality.
The current use in physics of "completeness" and "realism" is often considered
A
to have originated in the paper (Einstein et al., 1935) which proposed the EPR
paradox. In that paper the authors proposed the concept "element of reality"
and "completeness" of a physical theory. Though they did not define "element of
reality", they did provide a sufficient characterization for it, namely a quantity
whose value can be predicted with certainty before measuring it or disturbing
it in any way. EPR define a "complete physical theory" as one in which every
element of physical reality is accounted for by the theory. In the semantic view
of interpretation, an interpretation of a theory is complete if every element
DR
of the interpreting structure is accounted for by the mathematical formalism.
Realism is a property of each one of the elements of the mathematical formal-
ism; any such element is real if it corresponds to something in the interpreting
structure. For instance, in some interpretations of quantum mechanics (such as
the many-worlds interpretation) the ket vector associated to the system state is
assumed to correspond to an element of physical reality, while in others it does
not.
Determinism is a property characterizing state changes due to the passage of
time, namely that the state at an instant of time in the future is a function of the
state at the present (see time evolution). It may not always be clear whether a
particular interpreting structure is deterministic or not, precisely because there
may not be a clear choice for a time parameter. Moreover, a given theory may
have two interpretations, one of which is deterministic, and the other not.
Local realism has two parts:

Interpretation of quantum mechanics


218

• The value returned by a measurement corresponds to the value of some


function on the state space. Stated in another way, this value is an element
of reality;
• The effects of measurement have a propagation speed not exceeding some

FT
universal bound (e.g., the speed of light). In order for this to make sense,
measurement operations must be spatially localized in the interpreting
structure.

A precise formulation of local realism in terms of a local hidden variable theory


was proposed by John Bell.
Bell’s theorem and its experimental verification restrict the kinds of properties
a quantum theory can have. For instance, Bell’s theorem implies quantum
mechanics cannot satisfy local realism.

Consistent histories
The consistent histories generalizes the conventional Copenhagen interpreta-
tion and attempts to provide a natural interpretation of quantum cosmology.
The theory is based on a consistency criterion that then allows the history of a
A
system to be described so that the probabilities for each history obey the addi-
tive rules of classical probability while being consistent with the →Schrödinger
equation.
According to this interpretation, the purpose of a quantum-mechanical theory
is to predict probabilities of various alternative histories.

Many worlds
The many-worlds interpretation (or MWI) is an interpretation of quantum me-
DR
chanics that rejects the non-deterministic and irreversible wavefunction col-
lapse associated with measurement in the Copenhagen interpretation in favor
of a description in terms of quantum entanglement and reversible time evolu-
tion of states. The phenomena associated with measurement are explained by
decoherence which occurs when states interact with the environment. As result
of the decoherence the world-lines of macroscopic objects repeatedly split into
mutally unobservable, branching histories – distinct universes within a greater
multiverse.

The Copenhagen Interpretation


The Copenhagen interpretation is an interpretation of quantum mechanics for-
mulated by Niels Bohr and Werner Heisenberg while collaborating in Copen-
hagen around 1927. Bohr and Heisenberg extended the probabilistic inter-
pretation of the wavefunction, proposed by Max Born. The Copenhagen in-
terpretation rejects questions like "where was the particle before I measured

Interpretation of quantum mechanics


219

its position" as meaningless. The act of measurement causes an instantaneous


"collapse of the wave function". This means that the measurement process ran-
domly picks out exactly one of the many possibilities allowed for by the state’s
wave function, and the wave function instantaneously changes to reflect that

FT
pick.

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

The Bohm interpretation


The Bohm interpretation of quantum mechanics is an interpretation postulated
by David Bohm in which the existence of a non-local universal wavefunction
A
allows distant particles to interact instantaneously. The interpretation gener-
alizes Louis de Broglie’s pilot wave theory from 1927, which posits that both
wave and particle are real. The wave function ’guides’ the motion of the par-
ticle, and evolves according to the Schrödinger equation. The interpretation
assumes a single, nonsplitting universe (unlike the Everett many-worlds inter-
pretation) and is deterministic (unlike the Copenhagen interpretation). It says
the state of the universe evolves smoothly through time, without the collapsing
of wavefunctions when a measurement occurs, as in the Copenhagen inter-
DR
pretation. However, it does this by assuming a number of hidden variables,
namely the positions of all the particles in the universe, which, like probability
amplitudes in other interpretations, can never be measured directly.

Transactional interpretation
The transactional interpretation of quantum mechanics (TIQM) by John
Cramer is an unusual interpretation of quantum mechanics that 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.

Consciousness causes collapse


Consciousness causes collapse is the speculative theory that observation by a
conscious observer is responsible for the wavefunction collapse. It is an attempt

Interpretation of quantum mechanics


220

to solve the Wigner’s friend paradox by simply stating that collapse occurs at
the first "conscious" observer. Supporters claim this is not a revival of sub-
stance dualism, since (in a ramification of this view) consciousness and objects
are entangled and cannot be considered as distinct. The consciousness causes

FT
collapse theory can be considered as a speculative appendage to almost any in-
terpretation of quantum mechanics and most physicists reject it as unverifiable
and introducing unnecessary elements into physics.

Relational Quantum Mechanics


According to the Stanford Encyclopedia of Philosophy, Relational Quantum Me-
chanics is an interpretation of Quantum mechanics which:
discards the notions of absolute state of a system, absolute value of its
physical quantities, or absolute event. The theory describes only the way
systems affect each other in the course of physical interactions. State and
physical quantities refer always to the interaction, or the relation, between
two systems. Nevertheless, the theory is assumed to be complete. The
physical content of quantum theory is understood as expressing the net of
relations connecting all different physical systems.
A(Stanford Encyclopedia of Philosophy, Federico Laudisa, Università degli
Studi di Milano-Bicocca)

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 sci-
DR
ence.” 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:

• The Copenhagen Variant


• Kochen-Dieks-Healey Interpretations
• Motivating Early Modal Interpretations, based on the work of R. Clifton, M.
Dickson and J. Bub.

Comparison
At the moment, there is no experimental evidence that would allow us to dis-
tinguish between the various interpretations listed below. To that extent, the
physical theory stands, and is consistent with, itself and with reality; troubles
come only when one attempts to "interpret" it. Nevertheless, there is active

Interpretation of quantum mechanics


221

research in attempting to come up with experimental tests which would allow


differences between the interpretations to be experimentally tested.
Some of the most common interpretations are summarized here (however, the
assignment of values in this table is not without controversy, for the precise
meanings of some of the concepts involved are unclear and, in fact, the subject

FT
of the very controversy itself):
Interpretation Deterministic? Waveform Real? Unique History? Avoids Avoids
Hidden Collapsing
Variables? Wavefunctions?
Copenhagen No No Yes Yes No
interpretation
(Waveform not
real)
Copenhagen No Yes Yes Yes No
interpretation
(Waveform
real)
Consistent Agnostic 1 Agnostic 1 No Yes Yes
histories
(Decoherent
approach)
A
Many-worlds
interpretation
(Decoherent
Yes Yes No Yes Yes

approach)
Bohm-de Yes Yes 2 Yes 3 No Yes
Broglie
interpretation
("Pilot-wave"
approach)
Transactional No Yes Yes Yes No
DR
interpretation
Consciousness No Yes Yes Yes No
causes collapse
1
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.
2
Both particle AND guiding wavefunction are real.
3
Unique particle history, but multiple wave histories.
Each interpretation has many variants. It is difficult to get a precise definition
of the Copenhagen Interpretation — in the table above, two variants are shown
— one that regards the waveform as being a tool for calculating probabilities
only, and the other regards the waveform as an "element of reality".

Interpretation of quantum mechanics


222

See also
• Afshar experiment
• Bell’s theorem

FT
• Bohm interpretation
• Bohr-Einstein debates
• Consistent Histories
• Copenhagen interpretation
• Many-minds interpretation
• Many-worlds interpretation
• Measurement problem
• →Penrose Interpretation
• Philosophical interpretation of classical physics
• Quantum computation
• →Quantum indeterminacy
• →Quantum mechanics
• Quantum metaphysics
• Transactional interpretation

AWavefunction collapse

Related lists

• List of physics topics


• Unsolved problems in physics

References
DR
• 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
• R. Carnap, The interpretation of physics, Foundations of Logic and Math-
ematics of the International Encyclopedia of Unified Science, University of
Chicago Press, 1939.
• D. Deutsch, The Fabric of Reality, Allen Lane, 1997. Though written for
general audiences, in this book Deutsch argues forcefully against instru-
mentalism.
• Dickson, M. 1994. Wavefunction tails in the modal interpretation, Proceed-
ings of the PSA 1994, Hull, D., Forbes, M., and Burian, R. (eds), Vol. 1, pp.
366-376. East Lansing, Michigan: Philosophy of Science Association.

Interpretation of quantum mechanics


223

• Dickson, M. and Clifton, R. 1998. Lorentz-invariance in modal interpre-


tations The Modal Interpretation of Quantum Mechanics, Dieks, D. and
Vermaas, P. (eds), pp. 9-48. Dordrecht: Kluwer Academic Publishers
• A. Einstein, B. Podolsky and N. Rosen, Can quantum-mechanical descrip-

FT
tion of physical reality be considered complete? Phys. Rev. 47 777, 1935.
• C. Fuchs and A. Peres, Quantum theory needs no ‘interpretation’ , Physics
Today, March 2000.
• Christopher Fuchs, Quantum Mechanics as Quantum Information (and only
a little more), arXiv:quant-ph/0205039 v1, (2002)
• N. Herbert. Quantum Reality: Beyond the New Physics, New York: Double-
day, ISBN 0385235690, LoC QC174.12.H47 1985.
• R. Jackiw and D. Kleppner, One Hundred Years of Quantum Physics, Sci-
ence, Vol. 289 Issue 5481, p893, August 2000.
• M. Jammer, The Conceptual Development of Quantum Mechanics. New
York: McGraw-Hill, 1966.
• M. Jammer, The Philosophy of Quantum Mechanics. New York: Wiley,
1974.
• W. M. de Muynck, Foundations of quantum mechanics, an empiricist ap-
A proach, Dordrecht: Kluwer Academic Publishers, 2002, ISBN 1-4020-0932-
1
• R. Omnès, Understanding Quantum Mechanics, Princeton, 1999.
• K. Popper, Conjectures and Refutations, Routledge and Kegan Paul, 1963.
The chapter "Three views Concerning Human Knowledge", addresses,
among other things, the instrumentalist view in the physical sciences.
• H. Reichenbach, Philosophic Foundations of Quantum Mechanics, Berkeley:
University of California Press, 1944.
DR
• M. Tegmark and J. A. Wheeler, 100 Years of Quantum Mysteries", Scientific
American 284, 68, 2001.
• van Fraassen, B. 1972. A formal approach to the philosophy of science, in
Paradigms and Paradoxes: The Philosophical Challenge of the Quantum Do-
main, Colodny, R. (ed.), pp. 303-366. Pittsburgh: University of Pittsburgh
Press.
• J. A. Wheeler and H. Z. Wojciech (eds), Quantum Theory and Mea-
surement, Princeton: Princeton University Press, ISBN 0691083169, LoC
QC174.125.Q38 1983.

External links

Interpretation of quantum mechanics


224

• Willem M. de Muynck 109 Broad overview, realist vs. empiricist interpreta-


tions, against oversimplified view of the measurement process
• Comparative interpretations 110
• Skeptical View of "New Age" Interpretations of QM 111

FT
• The many worlds of quantum mechanics 112
• Erich Joos’ Decoherence Website 113 Basic and indepth information on deco-
herence
• Quantum Mechanics for Philosophers 114 An argument for the superiority of
the Bohm interpretation.
• Many-Worlds Interpretation of Quantum Mechanics 115
• Numerous Many Worlds-related Topics and Articles 116
• Relational Quantum Mechanics 117

Source: http://en.wikipedia.org/wiki/Interpretation_of_quantum_mechanics

Principal Authors: CSTAR, Charles Matthews, Roadrunner, LC, Linas


AIntroduction to quantum mechanics

This article is intended as a general, non-technical introduction. For the


proper encyclopedia article, please see →Quantum mechanics.
DR

109 http://www.phys.tue.nl/ktn/Wim/muynck.htm#quantum
110 http://members.aol.com/jmtsgibbs/Interpretation.htm
111 http://www.csicop.org/si/9701/quantum-quackery.html
112 http://www.sankey.ws/qm.html
113 http://www.decoherence.de
114 http://home.sprynet.com/~owl1/qm.htm
115 http://plato.stanford.edu/entries/qm-manyworlds/
116 http://www.station1.net/DouglasJones/many.htm
117 http://plato.stanford.edu/entries/qm-relational/

Introduction to quantum mechanics


225

A FT
Quantum mechanics is a physical science dealing with the behaviour of matter
and waves on the scale of atoms and subatomic particles. It also forms the ba-
sis for the contemporary understanding of how large objects such as stars and
galaxies, and cosmological events such as the Big Bang, can be analyzed and
DR
explained. Its acceptance by the general physics community is due to its accu-
rate prediction of the physical behaviour of systems, including systems where
Newtonian mechanics fails. This difference between the success of classical
and quantum mechanics is most often observed in systems at the atomic scale
or smaller, or at very low or very high energies, or at extremely low tempera-
tures. Quantum mechanics is the basis of modern developments in chemistry,
molecular biology, and electronics, and the foundation for the technology that
has transformed the world in the last fifty years.

Background
Through a century of experimentation and applied science, quantum mechan-
ical theory has proven to be very successful and practical. The term "quantum
mechanics" was first coined by Max Born in 1924. Quantum mechanics is the

Introduction to quantum mechanics


226

foundation for other sciences including condensed matter physics, quantum


chemistry, and particle physics.
Despite the success of quantum mechanics, it does have some controversial el-
ements. For example, the behaviour of microscopic objects described in quan-

FT
tum mechanics is very different from our everyday experience, which may pro-
voke an incredulous reaction. Moreover, some of the consequences of quantum
mechanics appear to be inconsistent with the consequences of other successful
theories, such as Einstein’s Theory of Relativity, especially general relativity.
Some of the background of quantum mechanics dates back to the early 1800’s,
but the real beginnings of quantum mechanics date from the work of Max
Planck in 1900. Albert Einstein, Niels Bohr, and Louis de Broglie soon made
important contributions. However, it was not until the mid-1920’s that a
more complete picture emerged, and the true importance of quantum me-
chanics became clear. Some of the most prominent scientists to contribute
were Max Born, Paul Dirac, Werner Heisenberg, Wolfgang Pauli, and Erwin
Schrödinger [#endnote_Schrödinger1].
Later, the field was further expanded with work by Julian Schwinger, Mur-
A
ray Gell-Mann, and Richard Feynman, in particular, with the development of
Quantum Electrodynamics in 1947.
Early researchers differed in their explanations of the fundamental nature of
what we now call electromagnetic radiation. In 1690, Christian Huygens ex-
plained the laws of reflection and refraction on the basis of a wave theory.
Sir Isaac Newton believed that light consisted of particles which he designated
corpuscles. In 1827 Thomas Young and Augustin Fresnel made experiments
on interference that showed that a corpuscular theory of light was inadequate.
DR
Then in 1873 James Clerk Maxwell showed that by making an electrical circuit
oscillate it should be possible to produce electromagnetic waves. His theo-
ry made it possible to compute the speed of electromagnetic radiation purely
on the basis of electrical and magnetic measurements, and the computed val-
ue corresponded very closely to the empirically measured speed of light. In
1888, Heinrich Hertz made an electrical device that actually produced what
we would now call microwaves — essentially radiation at a lower frequency
than visible light. Everything up to that point suggested that Newton had been
entirely wrong to regard light as corpuscular. Then it was discovered that when
light strikes an electrical conductor it causes electrons to move away from their
original positions, and, furthermore, the phenomenon observed could only be
explained if the light delivered energy in definite packets. In a photoelectric
device such as the light meter in a camera, when light hits the metallic detector
electrons are caused to move. Greater intensities of light at one frequency can
cause more electrons to move, but they will not move any faster. In contrast,

Introduction to quantum mechanics


227

A FT
Figure 22 The interference that produces colored bands on bubbles cannot be explained
by a model that depicts light as a particle. It can be explained by a model that depicts it
as a wave. The drawing shows sine waves that resemble waves on the surface of water
being reflected from two surfaces of a film of varying width, but that depiction of the wave
nature of light is only a crude analogy.

higher frequencies of light can cause electrons to move faster. So intensity of


DR
light controls the amperes of current produced, but frequency of light controls
the voltage produced. This appeared to raise a contradiction when compared
to sound waves and ocean waves, where only intensity was needed to predict
the energy of the wave. In the case of light, frequency appeared to predict en-
ergy. Something was needed to explain this phenomenon and also to reconcile
experiments that had shown light to have a particle nature with experiments
that had shown it to have a wave nature.

Spectroscopy and Onward


It is fairly easy to see a spectrum produced by white light when it passes
through a prism, the beveled edge of a mirror or a special pane of glass, or
through drops of rain to form a rainbow. When samples of single elements
are caused to emit light they may emit light at several characteristic frequen-
cies. The frequency profile produced is characteristic of that element. Instead
of there being a wide band filled with colors from violet to red, there will be
isolated bands of single colors separated by darkness. Such a display is called

Introduction to quantum mechanics


228

a line spectrum. Some lines go beyond the visible frequencies and can only
be detected by special photographic film or other such devices. Scientists hy-
pothesized that an atom could radiate light the way the string on a fine violin
radiates sound – not only with a fundamental frequency (in which the entire
string moves the same way at once) but with several higher harmonics (formed

FT
when the string divides itself into halves and other divisions that vibrate in co-
ordination with each other as when one half of the string is going one way as
the other half of the string is going the opposite way). For a long time nobody
could find a mathematical way to relate the frequencies of the line spectrum of
any element.

Figure 23 NASA photo of the bright-line spectrum of hydrogen

In 1885, Johann Jakob Balmer (1825-1898) figured out how the frequencies
of atomic
 hydrogen are related to each other. The formula is a simple one:
1
A 1
L = R 4 − n2
1


where L is wavelength, R is Rydberg’s constant and n is an in-


teger (2, 3,· · ·n).

This formula can be generalized to apply to atoms that are more complicated
than hydrogen, but we will stay with hydrogen for this general exposition.
(That is the reason that the denominator in the first fraction is expressed as a
square.)
DR
The next development was the discovery of the →Zeeman effect, named after
Pieter Zeeman (1865-1943). The physical explanation of the Zeeman effect
was worked out by Hendrik Anton Lorentz (1853-1928). Lorentz hypothesized
that the light emitted by hydrogen was produced by vibrating electrons. It
was possible to get feedback on what goes on within the atom because moving
electrons create a magnetic field and so can be influenced by the imposition
of an external magnetic field in a manner analogous to the way that one iron
magnet will attract or repel another magnet.
The Zeeman effect could be interpreted to mean that light waves are originated
by electrons vibrating in their orbits, but classical physics could not explain why
electrons should not fall out of their orbits and into the nucleus of their atoms,
nor could classical physics explain why their orbits would be such as to produce
the series of frequencies derived by Balmer’s formula and displayed in the line
spectra. Why did the electrons not produce a continuous spectrum?

Introduction to quantum mechanics


229

Old quantum theory


Quantum mechanics developed from the study of electromagnetic waves
through spectroscopy which includes visible light seen in the colors of the rain-
bow, but also other waves including the more energetic waves like ultraviolet

FT
light, x-rays, and gamma rays plus the waves with longer wavelengths includ-
ing infrared waves, microwaves and radio waves. We are not, however, speak-
ing of sound waves, but only of those waves that travel at the speed of light.
Also, when the word "particle" is used below, it always refers to elementary or
subatomic particles.

Planck’s constant
Classical physics predicted that a black-body radiator would produce infinite
energy, but that result was not observed in the laboratory. If black-body radi-
ation was dispersed into a spectrum, then the amount of energy radiated at
various frequencies rose from zero at one end, peaked at a frequency related
to the temperature of the radiating object, and then fell back to zero. In 1900,
Max Planck developed an empirical equation that could account for the ob-
A
served energy curves, but he could not harmonize it with classical theory. He
concluded that the classical laws of physics do not apply on the atomic scale as
had been assumed.
In this theoretical account, Planck allowed all possible frequencies, all possible
wavelengths. However, he restricted the energy that is delivered. "In classical
physics,... the energy of a given oscillator depends merely on its amplitude,
and this amplitude is subject to no restriction." But, according to Planck’s the-
ory, the energy emitted by an oscillator is strictly proportional to its frequency.
DR
The higher the frequency, the greater the energy. To reach this theoretical
conclusion, he made an assumption about the inner structure of black-body
radiators:
He postulated that a radiating body consisted of an enormous number of el-
ementary oscillators, some vibrating at one frequency and some at another,
with all frequencies from zero to infinity being represented....The energy
E of any one oscillator was not permitted to take on any arbitrary value,
but was proportional to some integral multiple of the frequency f of the
oscillator. That is,
E = nhf,

where n = 1, 2, 3,... etc. The proportionality constant h is called Planck’s


constant.One of the most direct applications is finding the energy of photons.
If you know h (Planck’s constant), and you know the frequency of the photon,

Introduction to quantum mechanics


230

then you can calculate the energy of the photons. For instance, if a beam of
light illuminated a target, and the light frequency was 540 × 10 12 hertz, then
the energy of each photon would be h × 540 × 10 12 hertz. The value of h
itself is exceedingly small, about 6.6260693 x 10 -34 joule seconds in scientific

FT
notation). This means that the photons in the beam of light have a energy of
about 3.58 × 10−19 Joules which is approximately 2.23 eV.
When you describe the energy of a wave in this manner, it seems that the wave
is carrying its energy in a certain number of little packets per second. This
discovery then seemed to remake the wave into a particle. These packets of
energy carried along with the wave were called quanta by Planck. Quantum
mechanics began with the discovery that energy is delivered in packets whose
size is related to the frequencies of all electromagnetic waves (and to the col-
or of visible light since in that case frequency determines color). Be aware,
however, the descriptions in terms of wave and particle import macro world
concepts into the quantum world, where they have only provisional relevance
or appropriateness.
In early research on light, there were two competing ways to describe light, ei-
ther as a wave propagated through empty space, or as small particles traveling
A
in straight lines. Because Planck showed that the energy of the wave is made
up of packets, the particle analogy became favored to help understand how
light delivers energy in multiples of certain set values designated as quanta
of energy. Nevertheless, the wave analogy is also indispensable for helping to
understand other light phenomena. In 1905, Albert Einstein used Planck’s con-
stant to postulate that the energy in a beam of light occurs in concentrations
that he called photons. According to that account, a single photon of a given
frequency delivers an invariant amount of energy. In other words, individual
DR
photons can deliver more or less energy, but only depending on their frequen-
cies. Although the description that stemmed from Planck’s research sounds
like Newton’s corpuscular account, Einstein’s photon was still said to have a
frequency, and the energy of the photon was accounted proportional to that
frequency. The particle account had been compromised once again.
Both the idea of a wave and the idea of a particle are models derived from
our everyday experience. We cannot see photons. We can only investigate
their properties indirectly. We look at some phenomena, such as the rainbow
of colors that we see when a thin film of oil rests on the surface of a puddle of
water, and we can explain that phenomenon to ourselves by comparing light
to waves. We look at other phenomena, such as the way a photoelectric meter
in our camera works, and we explain it by analogy to particles colliding with
the detection screen in the meter. In both cases we take concepts from our
everyday experience and apply them to a world we have never seen.

Introduction to quantum mechanics


231

Neither form of explanation is entirely satisfactory. In general any model can


only approximate that which it models. A model is useful only within the range
of conditions where it is able to predict the real thing with accuracy. Newtonian
physics is still a good predictor of many of the phenomena in our everyday life.

FT
To remind us that both "wave" and "particle" are concepts imported from our
macro world to explain the world of atomic-scale phenomena, some physicists
such as George Gamow have used the term "wavicle" to refer to whatever it
is that is really there. In the following discussion, "wave" and "particle" may
both be used depending on which aspect of quantum mechanical phenomena
is under discussion.

Reduced Planck’s constant

A
DR
Figure 24 Relation between a cycle and a wave;
half of a circle describes half of the cycle of a wave

Planck’s constant originally represented the energy that a light wave carries as
a function of its frequency. A step in the development of this concept appeared
in Bohr’s work. Bohr was using a "planetary" or particle model of the electron,
and could not understand why a 2π factor was essential to his experimentally
derived formulae. Later, de Broglie postulated that electrons have frequencies,
just as do photons, and that the frequency of an electron must conform to the
conditions for a standing wave that can exist in a certain orbit. That is to say,
the beginning of one cycle of a wave at some point on the circumference of a
circle (since that is what an orbit is) must coincide with the end of some cycle.
There can be no gap, no length along the circumference that is not participating

Introduction to quantum mechanics


232

in the vibration, and there can be no overlap of cycles. So the circumference of


the orbit, C, must equal the wavelength, , of the electron multiplied by some
positive integer (n = 1, 2, 3...). Knowing the circumference one can calculate
wavelengths that fit that orbit, and knowing the radius, r, of the orbit one can

FT
calculate its circumference. To put all that in mathematical form, C = 2π*r =
n* and so = 2πr/n and the appearance of the 2π factor is seen to occur simply
because we need it to calculate possible wavelengths (and therefore possible
frequencies) when we already know the radius of an orbit.
Again in 1925 when Werner Heisenberg developed his full quantum theory,
calculations involving wave analysis called Fourier series were fundamental,
and so the "reduced" version of Planck’s constant (h/2π) became invaluable
because it includes a conversion factor to facilitate calculations involving wave
analysis. Finally, when this reduced Planck’s constant appeared naturally in
Dirac’s equation it was then given an alternate designation, "Dirac’s constant."
Therefore, it is appropriate to begin with an explanation of what this constant
is, even though we haven’t yet touched on the theories that made its use con-
venient.
As noted above, the energy of any wave is given by its frequency multiplied by
A
Planck’s constant. A wave is made up of crests and troughs. In a wave, a cycle
is defined by the return from a certain position to the same position such as
from the top of one crest to the next crest. A cycle actually is mathematically
related to a circle, and both have 360 degrees. A degree is a unit of measure
for the amount of turn needed to produce an arc of a certain length at a given
distance. A sine curve is generated by a point on the circumference of a circle
as that circle rotates. (See a demonstration at: Rotation Applet 118) There are
2 π radians per cycle in a wave, which is mathematically related to the way
DR
a circle has 360◦ (which are equal to two π radians). (A radian is simply the
angle you would get if you measured a distance along the circumference of the
circle equal to the radius of the circle, and then drew lines to the center of the
circle and looked at the angle thus formed.) Since one cycle is 2 π radians,
when h is divided by 2 π the two "2 π" factors will cancel out leaving just the
radian to contend with. So, dividing h by 2 π describes a constant that, when
multiplied by the frequency of a wave, gives the energy in joules per radian.
h
And h/2 π is h-bar or ~ = 2π .
The reduced Planck’s constant is written in mathematical formulas as ~, and
is read as "h-bar". The reduced Planck’s constant allows computation of the
energy of a wave in units per radian instead of in units per cycle. These two
constants h and h-bar are merely conversion factors between energy units and

118 http://www.math.utah.edu/~cherk/ccli/bob/Rotation/sin12.swf

Introduction to quantum mechanics


233

frequency units. The reduced Planck’s constant is used more often than h
(Planck’s constant) alone in quantum mechanical mathematical formulas for
many reasons, one of which is that angular velocity or angular frequency is
ordinarily measured in radians per second so using h-bar that works in radians

FT
too will save a computation to put radians into degrees or vice-versa. Also,
when equations relevant to those problems are written in terms of ~, the fre-
quently occurring 2 π factors in numerator and denominator can cancel out,
saving a computation. However, in other cases, as in the orbits of the Bohr
atom, h/π was obtained naturally for the angular momentum of the orbits.
Another expression for the relation between energy and wave length is giv-
en in electron volts for energy and angstroms for wavelength: E photon (eV)
= 12,400/(Å) – it appears not to involve h at all, but that is only because a
different system of units has been used and now, numerically, the appropriate
conversion factor is 12,400.

Bohr atom
A
DR

Figure 25 The Bohr model of the atom, show-


ing electron quantum jumping to ground state n=1

Introduction to quantum mechanics


234

In 1897 the particle called the electron was discovered. By means of the gold
foil experiment physicists discovered that matter is, volume for volume, largely
space. Once that was clear, it was hypothesized that negative charge entities
called electrons surround positively charged nuclei. So at first all scientists be-

FT
lieved that the atom must be like a miniature solar system. But that simple
analogy predicted that electrons would, within about one hundredth of a mi-
crosecond, crash into the nucleus of the atom. The great question of the early
20th century was, "Why do electrons normally maintain a stable orbit around
the nucleus?"
In 1913, Niels Bohr removed this substantial problem by applying the idea
of discrete (non-continuous) quanta to the orbits of electrons. This account
became known as the Bohr model of the atom. Bohr basically theorized that
electrons can only inhabit certain orbits around the atom. These orbits could
be derived by looking at the spectral lines produced by atoms.
Bohr explained the orbits that electrons can take by relating the angular mo-
mentum of electrons in each "permitted" orbit to the value of h, Planck’s con-
stant. He held that an electron in the lowest orbital has an angular momentum
equal to h/2π. Each orbit after the initial orbit must provide for an electron’s
A
angular momentum being an integer multiple of that lowest value. He depict-
ed electrons in atoms as being analogous to planets in a solar orbit. However,
he took Planck’s constant to be a fundamental quantity that introduces spe-
cial requirements at this subatomic level and that explains the spacing of those
"planetary" orbits.
Bohr’s analysis of electron orbits as circular
A little math on circular orbits. Bohr was very
familiar with the dynamics of simple circular
DR
orbits in an inverse square field as described in
classical mechanics.
Simply explained: To find the acceleration of a
circle, place it inside the shape of a square
where tangents meet, then find the linear speed
along one side of the square, then square the
speed of one side to complete the speed of the
entire square, then divide by the radius of the
circle placed in the square to get the speed
around the circle. Therefore, circular
(centripetal) acceleration is v squared over r
where v is speed and r is radius. The equation for
the centripetal acceleration is a = v 2 /r. That
is, acceleration is inversely proportional to the

Introduction to quantum mechanics


235

radius of the circle. If the radius is doubled,


then the acceleration is halved.
Also, Kepler’s Third Law is that the radius cubed
equals the circumference of the orbit squared.

FT
It immediately follows that the radius of any n
orbit is proportional to the orbit n squared, and
the speed in that orbit is proportional to 1/n. Speed
times radius gives angular momentum. That leaves
n-squared over n. It then follows that the angular
momentum for any orbit n is just proportional to
n. Bohr argued then that the angular momentum
in any orbit n was nKh, where h is Planck’s
constant and K is some multiplying factor, the
same for all the orbits, which was later
determined to be 1/2π.

Bohr considered one revolution in orbit to be equivalent to one cycle in an os-


cillator (as in Planck’s initial measurements to define the constant h) which is
A
in turn similar to one cycle in a wave. The number of revolutions per second is
(or defines) what we call the frequency of that electron or that orbital. Spec-
ifying that the frequency of each orbit must be an integer multiple of Planck’s
constant h would only permit certain orbits, and would also fix their size.
Bohr generalized Balmer’s formula for hydrogen by replacing denominator in
the term 1/4 with an explicit squared variable:
 
1 1 1
λ = RH m2 − n2 , m=1,2,3,4,5,..., and n > m
DR
where is the wavelength of the light and R H is the Rydberg constant for hy-
drogen. This generalization predicted many more line spectra than had been
previously detected, and experimental confirmation of this prediction followed.
It follows almost immediately that if λ is quantized as the formula above indi-
cates, then the momentum of any photon must be quantized. The frequency of
light, v, at a given wavelength λ is given by the relationship
c c
v= λ and :λ = v and multiplying by h/h = 1,

hc
λ= hv , and we know that

hc
E = hv so λ = E which we can rewrite as:

Introduction to quantum mechanics


236

h
λ= E/c
, and E/c = p (momentum) so

h h
λ= p or p = λ

FT
Beginning with line spectra, physicists were able to deduce empirically the
rules according to which the orbits of electrons are determined and to discover
something vital about the momentums involved–that they are quantized.
Bohr next realized how the angular momentum of an electron in its orbit, L,
is quantized, i.e., he determined that there is some constant value K such that
when it is multiplied by Planck’s constant, h, it will yield the angular momen-
tum that pertains to the lowest orbital. When it is multiplied by successive
integers it will then give the values of other possible orbitals. He later deter-
mined that K = 1/2π . (See the detailed argument at 119.)
Bohr’s theory represented electrons as orbiting the nucleus of an atom in a way
that was amazingly different from what we see in the world of our everyday
experience. He showed that when an electron changed orbits it did not move in
a continuous trajectory from one orbit around the nucleus to another. Instead,
Ait suddenly disappeared from its original orbit and reappeared in another or-
bit. Each distance at which an electron can orbit is a function of a quantized
amount of energy. The closer to the nucleus an electron orbits, the less en-
ergy it takes to remain in that orbital. Electrons that absorb a photon gain a
quantum of energy, so they jump to an orbit that is farther from the nucleus,
while electrons that emit a photon lose a quantum of energy and so jump to
an inner orbital. Electrons cannot gain or lose a fractional quantum of energy,
and so, it is argued, they cannot have a position that is at a fractional distance
DR
between allowed orbitals. Allowed orbitals were designated as whole numbers
using the letter n with the innermost orbital being designated n = 1, the next
out being n = 2, and so on. Any orbital with the same value of n is called an
electron shell.
Bohr’s model of the atom was essentially two-dimensional because it depicts
electrons as particles in circular orbits. In this context, two-dimensional means
something that can be described on the surface of a plane. One-dimensional
means something that can be described by a line. Because circles can be de-
scribed by their radius, which is a line segment, sometimes Bohr’s model of the
atom is described as one-dimensional.

119 http://galileo.phys.virginia.edu/classes/252/Bohr_Atom/Bohr_Atom.html

Introduction to quantum mechanics


237

Wave-particle duality

A
Figure 26
FT
Probability distribution of the Bohr atom

Niels Bohr determined that it is impossible to describe light adequately by the


sole use of either the wave analogy or of the particle analogy. Therefore he
DR
enunciated the principle of complementarity, which is a theory of pairs, such
as the pairing of wave and particle or the pairing of position and momentum.
Louis de Broglie worked out the mathematical consequences of these findings.
In quantum mechanics, it was found that electromagnetic waves could react in
certain experiments as though they were particles and in other experiments as
though they were waves. It was also discovered that subatomic particles could
sometimes be described as particles and sometimes as waves. This discovery
led to the theory of wave-particle duality by Louis-Victor de Broglie in 1924,
which states that subatomic entities have properties of both waves and particles
at the same time.
The Bohr atom model was enlarged upon with the discovery by de Broglie that
the electron has wave-like properties. In accord with de Broglie’s conclusions,
electrons can only appear under conditions that permit a standing wave. A
standing wave can be made if a string is fixed on both ends and made to vibrate

Introduction to quantum mechanics


238

(as it would in a stringed instrument). That illustration shows that the only
standing waves that can occur are those with zero amplitude at the two fixed
ends. The waves created by a stringed instrument appear to oscillate in place,
simply changing crest for trough in an up-and-down motion. A standing wave

FT
can only be formed when the wave’s length fits the available vibrating entity.
In other words, no partial fragments of wave crests or troughs are allowed. In
a round vibrating medium, the wave must be a continuous formation of crests
and troughs all around the circle. Each electron must be its own standing wave
in its own discrete orbital.

Development of modern quantum mechanics


Full quantum mechanical theory
Werner Heisenberg developed the full quantum mechanical theory in 1925 at
the young age of 23. Following his mentor, Niels Bohr, Werner Heisenberg
began to work out a theory for the quantum behavior of electron orbitals. Be-
cause electrons could not be observed in their orbits, Heisenberg went about
creating a mathematical description of quantum mechanics built on what could
A
be observed, that is, the light emitted from atoms in their characteristic atomic
spectra. Heisenberg studied the electron orbital on the model of a charged ball
on a spring, an oscillator, whose motion is anharmonic (not quite regular). For
a picture of the behavior of a charged ball on a spring see: Vibrating Charges 120.
Heisenberg first explained this kind of observed motion in terms of the laws of
classical mechanics known to apply in the macro world, and then applied quan-
tum restrictions, discrete (non-continuous) properties, to the picture. Doing so
causes gaps to appear between the orbitals so that the mathematical descrip-
DR
tion he formulated would then represent only the electron orbitals predicted
on the basis of the atomic spectra.
In 1925 Heisenberg published a paper (in Z. Phys. vol. 33, p. 879-893)
entitled "Quantum-mechanical re-interpretation of kinematic and mechanical
relations." So ended the old quantum theory and began the age of quantum
mechanics. Heisenberg’s paper gave few details that might aid readers in de-
termining how he actually contrived to get his results for the one-dimensional
models he used to form the hypothesis that proved so useful. In his paper,
Heisenberg proposed to "discard all hope of observing hitherto unobservable
quantities, such as the position and period of the electron," and restrict himself
strictly to actually observable quantities. He needed mathematical rules for
predicting the relations actually observed in nature, and the rules he produced
worked differently depending on the sequence in which they were applied. "It

120 http://www.colorado.edu/physics/2000/waves_particles/wavpart4.html

Introduction to quantum mechanics


239

quickly became clear that the non-commutativity (in general) of kinematical


quantities in quantum theory was the really essential new technical idea in
the paper." (Aitchison, p. 5) But it was unclear why this non-commutativity
was essential. Could it have a physical interpretation? At least the matter was

FT
made more palatable when Max Born discovered that the Heisenberg compu-
tational scheme could be put in a more familiar form present in elementary
mathematics.
The special type of multiplication that turned out to be required in his formula
was most elegantly described by using special arrays of numbers called matri-
ces. In ordinary situations it does not matter in which order the operations
involved in multiplication are performed, but matrix multiplication does not
commute. Essentially that means that it matters which order given operations
are performed in. Multiplying matrix A by matrix B is not the same as multiply-
ing matrix B by matrix A. In symbols, AxB is in general not equal to BxA. (The
important thing in quantum theory is that it turned out to matter whether one
measures velocity first and then measures position, or vice-versa.) The matrix
convention turned out to be a convenient way of organizing information and
making clear the exact sequence in which calculations must be made.
A
In matrix mathematics sets of numbers are given in rows and columns, and
there are conventions for the way multiplication of matrices is performed. If
everybody arranged their matrices entirely as they pleased, then understand-
ing every new matrix calculation would involve learning the personal plan of
the person who made the matrix, so certain conventions have evolved. Re-
verse the order of the multiplication of the matrices and the numerical results
will go wrong. In other words, Matrix multiplication is noncommutative (to be
precise, reversing the order will in general cause the multiplication to become
DR
undefined except for the special case where the matrices are square). Because
these complex operations are, by analogy, called "multiplication," it is tempting
to imagine that it ought to be irrelevant whether one multiplies matrix A by
matrix B, or one multiplies matrix B by matrix A. Because of the complications
involved in the rules of matrix multiplication, in almost every case the ordinary
mathematical expectation of commutation does not hold. (Sometimes matrices
even anticommute.) In Heisenberg’s matrix mechanics, the sets of numbers are
infinite, representing all possible positions of the electron, and those matrices
cannot be multiplied in reverse order and still produce the correct results. The
essential point is that Heisenberg first learned what ways he had to operate on
measured quantities to be able to account for the line spectra that had been
observed. The operations to be performed seemed complicated and arbitrary,
but when specialists realized that what he was doing could be represented in a

Introduction to quantum mechanics


240

coherent and structurally straightforward way by the use of matrix mathemat-


ics it became much easier to get through the calculations successfully and also
lessened the appearance of arbitrariness of the process.
Heisenberg approached quantum mechanics from the historical perspective

FT
that treated an electron as an oscillating charged particle. Bohr’s use of this
analogy had already allowed him to explain why the radii of the orbits of elec-
trons could only take on certain values. It followed from this interpretation
of the experimental results available and the quantum theory that Heisenberg
subsequently created that an electron could not be at any intermediate position
between two "permitted" orbits. Therefore electrons were described as "jump-
ing" from orbit to orbit. The idea that an electron might now be in one place
and an instant later be in some other place without having traveled between
the two points was one of the earliest indications of the "spookiness" of quan-
tum phenomena. Although the scale is smaller, the "jump" from orbit to orbit is
as strange and unexpected as would be a case in which someone stepped out
of a doorway in London onto the streets of Los Angeles.
Amplitudes of position and momentum that have a period of 2 π like a cycle in
a wave are called Fourier series variables. Heisenberg described the particle-
A
like properties of the electron in a wave as having position and momentum in
his matrix mechanics. When these amplitudes of position and momentum are
measured and multiplied together, they give intensity. However, he found that
when the position and momentum were multiplied together in that respective
order, and then the momentum and position were multiplied together in that
respective order, there was a difference or deviation in intensity between them
of h/2π. Heisenberg would not understand the reason for this deviation until
two more years had passed, but for the time being he satisfied himself with the
DR
idea that the math worked and provided an exact description of the quantum
behavior of the electron.
Matrix mechanics was the first complete definition of quantum mechanics, its
laws, and properties that described fully the behavior of the electron. It was
later extended to apply to all subatomic particles. Schrödinger subsequently
produced a quantum wave theory that was computationally easier and avoid-
ed some of the odd-sounding ideas like "quantum leaps" of an electron from
one orbit to another, and finally Dirac made the idea of non-commutativity
central to his own theory that proved the formulations of Heisenberg and of
Schrödinger to be special cases of his own.

Schrödinger wave equation


Because particles could be described as waves, later in 1925 Erwin Schrödinger
analyzed what an electron would look like as a wave around the nucleus of the

Introduction to quantum mechanics


241

atom. Using this model, he formulated his equation for particle waves. Rather
than explaining the atom by analogy to satellites in planetary orbits, he treated
everything as waves whereby each electron has its own unique wavefunction. A
wavefunction is described in Schrödinger’s equation by three properties (later

FT
Paul Dirac added a fourth). The three properties were (1) an "orbital" designa-
tion, indicating whether the particle wave is one that is closer to the nucleus
with less energy or one that is further from the nucleus with more energy, (2)
the shape of the orbital, i.e. an indication that orbitals were not just spherical
but other shapes, and (3) the magnetic moment of the orbital, which is a man-
ifestation of force exerted by the charge of the electron as it rotates around the
nucleus.
These three properties were called collectively the wavefunction of the elec-
tron and are said to describe the quantum state of the electron. "Quantum
state" means the collective properties of the electron describing what we can
say about its condition at a given time. For the electron, the quantum state is
described by its wavefunction which is designated in physics by the Greek let-
ter ψ (psi, pronounced "sigh"). The three properties of Schrödinger’s equation
that describe the wavefunction of the electron and therefore also describe the
Aquantum state of the electron as described in the previous paragraph are each
called quantum numbers. The first property which describes the orbital was
numbered according to Bohr’s model where n is the letter used to describe the
energy of each orbital. This is called the principal quantum number. The next
quantum number that describes the shape of the orbital is called the azimuthal
quantum number and it is represented by the letter l (lower case L). The shape
is caused by the angular momentum of the orbital. The rate of change of the
angular momentum of any system is equal to the resultant external torque
DR
acting on that system. In other words, angular momentum represents the re-
sistance of a spinning object to speed up or slow down under the influence of
external force. The azimuthal quantum number "l" represents the orbital angu-
lar momentum of an electron around its nucleus. However, the shape of each
orbital has its own letter as well. So for the letter "l" there are other letters to
describe the shapes of "l". The first shape is spherical and is described by the
letter s. The next shape is like a dumbbell and is described by the letter p. The
other shapes of orbitals become more complicated (see Atomic Orbitals 121) and
are described by the letters d, f, and g. To see the shape of a carbon atom,
see Carbon atom 122. The third quantum number of Schrödinger’s equation de-
scribes the magnetic moment of the electron and is designated by the letter m

121 http://orbitals.com/orb/
122 http://library.thinkquest.org/C0110925/carbon.htm

Introduction to quantum mechanics


242

and sometimes as the letter m with a subscript l because the magnetic moment
depends upon the second quantum number l.
In May 1926 Schrödinger published a proof that Heisenberg’s matrix mechanics
and Schrödinger’s wave mechanics gave equivalent results: mathematically

FT
they were the same theory. Both men claimed to have the superior theory.
Heisenberg insisted on the existence of discontinuous quantum jumps in his
particle-like examination of the oscillation of a charged electron giving more
precise definitions and Schrödinger insisted that a theory based on continuous
wave-like properties which he called "matter-waves" was better.

Uncertainty Principle
Main article: →Uncertainty principle

In 1927, Heisenberg made a new discovery on the basis of his quantum theory
that had further practical consequences of this new way of looking at matter
and energy on the atomic scale. In Heisenberg’s matrix mechanics formula,
Heisenberg had encountered an error or difference of h/2π between position
A
and momentum. This represented a deviation of one radian of a cycle when
the particle-like aspects of the wave were examined. Heisenberg analyzed this
difference of one radian of a cycle and divided the difference or deviation of
one radian equally between the measurement of position and momentum. This
had the consequence of being able to describe the electron as a point particle
in the center of one cycle of a wave so that its position would have a standard
deviation of plus or minus one-half of one radian of the cycle (1/2 of h-bar). A
standard deviation can be either plus or minus the measurement i.e. it can add
to the measurement or subtract from it. In three-dimensions a standard devia-
DR
tion is a displacement in any direction. What this means is that when a moving
particle is viewed as a wave it is less certain where the particle is. In fact, the
more certain the position of a particle is known, the less certain the momen-
tum is known. This conclusion came to be called "Heisenberg’s Indeterminacy
Principle," or Heisenberg’s Uncertainty Principle. To understand the real idea
behind the uncertainty principle imagine a wave with its undulations, its crests
and troughs, moving along. A wave is also a moving stream of particles, so you
have to superimpose a stream of particles moving in a straight line along the
middle of the wave. An oscillating ball of charge creates a wave larger than
its size depending upon the length of its oscillation. Therefore, the energy of a
moving particle is as large as the cycle of the wave, but the particle itself has a
location. Because the particle and the wave are the same thing, then the parti-
cle is really located somewhere in the width of the wave. Its position could be
anywhere from the crest to the trough. The math for the uncertainty principle

Introduction to quantum mechanics


243

says that the measurement of uncertainty as to the position of a moving parti-


cle is one-half the width from the crest to the trough or one-half of one radian
of a cycle in a wave.
For moving particles in quantum mechanics, there is simply a certain degree

FT
of exactness and precision that is missing. You can be precise when you take a
measurement of position and you can be precise when you take a measurement
of momentum, but there is an inverse imprecision when you try to measure
both at the same time as in the case of a moving particle like the electron. In
the most extreme case, absolute precision of one variable would entail absolute
imprecision regarding the other.
Heisenberg voice recording in an early lecture on the uncertainty principle pointing to a Bohr
model of the atom: "You can say, well, this orbit is really not a complete orbit. Actually at every
moment the electron has only an inactual position and an inactual velocity and between these
two inaccuracies there is an inverse correlation."

The consequences of the uncertainty principle were that the electron could no
longer be considered as in an exact location in its orbital. Rather the electron
had to be described by every point where the electron could possibly inhabit.
By creating points of probable location for the electron in its known orbital,
A
this created a cloud of points in a spherical shape for the orbital of a hydro-
gen atom which points gradually faded out nearer to the nucleus and farther
from the nucleus. This is called a probability distribution. Therefore, the Bohr
atom number n for each orbital became known as an n-sphere in the three
dimensional atom and was pictured as a probability cloud where the electron
surrounded the atom all at once.
This led to the further description by Heisenberg that if you were not making
measurements of the electron that it could not be described in one particular
DR
location but was everywhere in the electron cloud at once. In other words,
quantum mechanics cannot give exact results, but only the probabilities for the
occurrence of a variety of possible results. Heisenberg went further and said
that the path of a moving particle only comes into existence once we observe
it. However strange and counter-intuitive this assertion may seem, quantum
mechanics does still tell us the location of the electron’s orbital, its probability
cloud. Heisenberg was speaking of the particle itself, not its orbital which is in
a known probability distribution.
It is important to note that although Heisenberg used infinite sets of positions
for the electron in his matrices, this does not mean that the electron could be
anywhere in the universe. Rather there are several laws that show the electron
must be in one localized probability distribution. An electron is described by its
energy in Bohr’s atom which was carried over to matrix mechanics. Therefore,
an electron in a certain n-sphere had to be within a certain range from the
nucleus depending upon its energy. This restricts its location. Also, the number

Introduction to quantum mechanics


244

of places an electron can be is also called "the number of cells in its phase
space". The uncertainty principle set a lower limit to how finely one can chop
up classical phase space. Therefore, the number of places that an electron can
be in its orbital becomes finite due to the Uncertainty Principle. Therefore,

FT
an electron’s location in an atom is defined to be in its orbital and its orbital
although being a probability distribution does not extend out into the entire
universe, but stops at the nucleus and before the next n-sphere orbital begins
and the points of the distribution are finite due to the Uncertainty Principle
creating a lower limit.
Classical physics had shown since Newton that if you know the position of stars
and planets and details about their motions that you can predict where they
will be in the future. For subatomic particles, Heisenberg denied this notion
showing that due to the uncertainty principle one cannot know the precise
position and momentum of a particle at a given instant, so its future motion
cannot be determined, but only a range of possibilities for the future motion of
the particle can be described.
These notions arising from the uncertainty principle only arise at the subatomic
level and were a consequence of wave-particle duality. As counter-intuitive
A
as they may seem, quantum mechanical theory with its uncertainty principle
has been responsible for major improvements in the world’s technology from
computer components to fluorescent lights to brain scanning techniques.

Wavefunction collapse
Schrödinger’s wave equation with its unique wavefunction for a single elec-
tron is also spread out in a probability distribution like Heisenberg’s quantized
particle-like electron. This is because a wave is naturally a widespread dis-
DR
turbance and not a point particle. Therefore, Schrödinger’s wave equation
has the same predictions made by the uncertainty principle because uncertain-
ty of location is built into the definition of a widespread disturbance like a
wave. Uncertainty only needed to be defined from Heisenberg’s matrix me-
chanics because the treatment was from the particle-like aspects of the elec-
tron. Schrödinger’s wave equation shows that the electron is in the probabil-
ity cloud at all times in its probability distribution as a wave that is spread
out. Max Born discovered in 1928 that when you compute the square of
Schrödinger’s wavefunction (psi-squared), you get the electron’s location as
a probability distribution. Therefore, if a measurement of the position of an
electron is made as an exact location in space instead of as a probability dis-
tribution, it ceases to have wave-like properties. Without wave-like properties,
none of Schrödinger’s definitions of the electron being wave-like make sense

Introduction to quantum mechanics


245

anymore. The measurement of the position of the particle nullifies the wave-
like properties and Schrödinger’s equation then fails. Because the electron can
no longer be described by its wavefunction when measured due to it becoming
particle-like, this is called wavefunction collapse.

FT
Eigenstates and eigenvalues
The term eigenstate is derived from the German/Dutch word "eigen," which
means "inherent" or "characteristic." The word eigenstate is descriptive of the
measured state of some object that possesses quantifiable characteristics such
as position, momentum, etc. The state being measured and described must
be an "observable" (i.e. something that can be experimentally measured either
directly or indirectly like position or momentum), and must have a definite
value. In the everyday world, it is natural and intuitive to think of everything
being in its own eigenstate. Everything appears to have a definite position, a
definite momentum, a definite value of measure, and a definite time of occur-
rence. However, quantum mechanics affirms that it is impossible to pinpoint
exact values for the momentum of a certain particle like an electron in a given
location at a particular moment in time, or, alternatively, that it is impossible
A
to give an exact location for such an object when the momentum has been
measured. Due to the uncertainty principle, statements regarding both the po-
sition and momentum of particles can only be given in terms of a range of
probabilities, a "probability distribution". Eliminating uncertainty in one term
maximizes uncertainty in regard to the second parameter.
Therefore it became necessary to have a way to clearly formulate the difference
between the state of something that is uncertain in the way just described, such
as an electron in a probability cloud, and effectively contrast it to the state of
DR
something that is not uncertain, something that has a definite value. When
something is in the condition of being definitely "pinned-down" in some regard,
it is said to possess an eigenstate. For example, if the position of an electron
has been made definite, it is said to have an eigenstate of position.
A definite value, such as the position of an electron that has been successfully
located, is called the eigenvalue of the eigenstate of position. The German
word "eigen" was first used in this context by the mathematician David Hilbert
in 1904. Schrödinger’s wave equation gives wavefunction solutions, mean-
ing the possibilities where the electron might be, just as does Heisenberg’s
probability distribution. As stated above, when a wavefunction collapse occurs
because something has been done to locate the position of an electron, the
electron’s state becomes an eigenstate of position, meaning that the position
has a known value.

Introduction to quantum mechanics


246

The Pauli Exclusion Principle


There was a doublet, meaning a pair of lines, in the spectrum of a hydrogen
atom that was unaccounted for. This meant that there was more energy in the
electron orbital from magnetic moment than had previously been described.

FT
Wolfgang Pauli when studying alkali metals had introduced what he called a
"two-valued quantum degree of freedom" associated with the electron in the
outermost shell. "Degrees of freedom" simply means the number of possible in-
dependent ways a particle may move. This led to the Pauli Exclusion Principle
that predicted that no more than two electrons can inhabit the same orbital. It
also predicted that any neutron, electron, or proton (types of fermions) could
not exist in the same quantum state. We learned in Schrödinger’s equation that
there were three quantum states of the electron, but if two electrons could be
in the same orbital, there had to be another quantum number to distinguish
those two electrons from each other and to describe the extra magnetic mo-
ment shown in the atomic spectrum. In early 1925, the young physicist Ralph
Kronig had introduced a theory to Pauli that the electron rotates in space in
the same way that the earth rotates on its axis. This would account for the
missing magnetic moment and allow for two electrons in the same orbital to
A
be different if their spin was in opposite directions to each other thus satisfying
the Exclusion Principle.
The Pauli Exclusion Principle states that no electron (or other fermion) can be
in the same quantum state as another. This has an effect on the probability dis-
tribution of the electron further defining the number of cells in its phase space.
The minimum limit is the limit of the Uncertainty Principle and the Exclusion
Principle states that no two electrons can be within this same minimum space
defined by the Uncertainty Principle.
DR
Therefore, a single electron in its orbital when defined by its quantum state
which is its wavefunction which is defined by its four quantum numbers cannot
have the same four quantum numbers of another electron in that atom. Where
two electrons are in the same n-sphere and therefore share the same principal
quantum number, they must then have some other unique quantum number of
shape l, magnetic moment m or spin s. Even in the formation of degenerate
gases where the electrons are not in an orbital around the nucleus of an atom,
they must still follow the Pauli Exclusion Principle when in a confined space.

Dirac wave equation


In 1928, Paul Dirac worked out a variation of Schrödinger’s equation that ac-
counted for a fourth property of the electron in its orbital. Paul Dirac intro-
duced the fourth quantum number called the spin quantum number designated
by the letter s to the new Dirac equation of the wavefunction of the electron. In

Introduction to quantum mechanics


247

1930, Dirac combined Heisenberg’s matrix mechanics with Schrödinger’s wave


mechanics into a single quantum mechanical representation in his Principles of
Quantum Mechanics. The quantum picture of the electron was now complete.
All of the above development of quantum theory was based mainly on the

FT
atomic spectrum of the hydrogen atom. This is due to the fact that each atom of
each element produces a unique pattern of spectral lines when light from each
different kind of element is passed through a prism. Scientists could not study
the electron and nucleus of the atom itself because they cannot be seen. Even
today with High-resolution Scanning Tunneling Electron Microscopes we can
only get images of the atom as a blurry fuzzball. However, the spectral lines of
the atom reveal the orbits of electrons and the energy that can be expected. It
was basically this study of the spectroscopic analysis of first the hydrogen atom
and then the helium atom that led to quantum theory. Therefore, the mathe-
matical formula were made to fit the picture of the atomic spectrum. That is
why quantum mechanics is sometimes referred to as a form of mathematical
physics.

Quantum entanglement
A
Albert Einstein rejected Heisenberg’s Uncertainty Principle insofar as it seemed
to imply more than a necessary limitation on human ability to actually know
what occurs in the quantum realm. In a letter to Max Born in 1926, Ein-
stein claimed that God "does not play dice." Heisenberg’s quantum mechanics,
based on Bohr’s initial explanation, became known as the Copenhagen Inter-
pretation of quantum mechanics. Both Bohr and Einstein spent many years
defending and attacking this interpretation. After the 1930 Solvay conference,
Einstein never again challenged the Copenhagen interpretation on technical
DR
points, but did not cease a philosophical attack on the interpretation, on the
grounds of realism and locality. Einstein, in trying to show that it was not a
complete theory, recognized that the theory predicted that two or more parti-
cles which have interacted in the past exhibit surprisingly strong correlations
when various measurements are made on them. Einstein called this "spooky
action at a distance". In 1935, Schrödinger published a paper explaining the
argument which had been denoted the EPR paradox (Einstein-Podolsky-Rosen,
1935). Einstein showed that the Copenhagen Interpretation predicted quan-
tum entanglement which he was trying to prove was incorrect in that it would
defy the law of physics that nothing could travel faster than the speed of light.
Quantum entanglement means that when there is a change in one particle at a
distance from another particle then the other particle automatically changes to
counter-balance the system. In quantum entanglement, the act of measuring
one entangled particle defines its properties and seems to influence the proper-
ties of its partner or partners instantaneously, no matter how far apart they are.

Introduction to quantum mechanics


248

Because the two particles are in an entangled state, changes to the one cause
instantaneous effects on the other. Einstein had calculated that quantum theo-
ry would predict this, he saw it as a flaw and therefore challenged it. However,
instead of showing a weakness in quantum mechanics, this forced quantum

FT
mechanics to acknowledge that quantum entanglement did in fact exist and it
became another foundation theory of quantum mechanics. The 1935 paper is
currently Einstein’s most cited publication in physics journals.
Bohr’s original response to Einstein was that the particles were in a system.
However, Einstein’s challenge led to decades of substantial research into this
quantum mechanical phenomenon of quantum entanglement. This research
clarified by Yanhua Shih points out that the two entangled particles can be
viewed as somehow not separate, which removes the locality objection. This
means that no matter the distance between the entangled particles, they re-
main in the same quantum state so that one particle is not sending information
to another particle faster than the speed of light, but rather a change to one
particle is a change to the entire system or quantum state of the entangled
particles and therefore changes the state of the system without information
transference.
A
Interpretations
As a system becomes larger or more massive (action » h ) the classical picture
tends to emerge, with some exceptions, such as superfluidity. The emergence
of behaviour as we scale up that matches our classical intuition is called the
correspondence principle and is based on Ehrenfest’s theorem. This why we
can usually ignore quantum mechanics when dealing with everyday objects.
Even so, trying to make sense of quantum theory is an ongoing process which
DR
has spawned a number of interpretations of quantum theory, ranging from the
conventional Copenhagen Interpretation to hidden variables & many worlds.
There seems to be no end in sight to the philosophical musings on the sub-
ject; however the empirical or technical success of the theory is unrivalled; all
modern fundamental physical theories are quantum theories.

See also
• Atom
• →Quantum mechanics
• →Uncertainty principle
• Correspondence principle
• Quantum number
• Principal quantum number
• Azimuthal quantum number

Introduction to quantum mechanics


249

• Magnetic quantum number


• Spin quantum number
• Planck’s constant
• Standard Model

FT
• →Matrix mechanics
• Schrödinger’s equation
• →Quantum mechanics, philosophy and controversy
• →Interpretation of quantum mechanics

References
• Mara Beller, Quantum Dialogue: The Making of a Revolution. University of
Chicago Press, Chicago, 2001.
• Bohr, Niels (1958). Atomic Physics and Human Knowledge. John Wiley
and Sons. ISBN 000000000X.
• De Broglie, Louis. The Revolution in Physics, Noonday Press, 1953.
• Feynman, Richard P., QED: The Strange Theory of Light and Matter, Prince-
ton University Press, 1985. ISBN 0-691-08388-6

A Feigl, Herbert and May Brodbeck, Readings in the Philosophy of Science,
Appleton-Century-Crofts, 1953.
• Einstein, Albert. Essays in Science, Philosophical Library, 1934.
• Prof. Michael Fowler, The Bohr Atom, A series of lectures, 1999, University
of Virginia.
• Heisenberg, Werner. Physics and Philosophy, Harper and Brothers, 1958.
• S Lakshmibala, "Heisenberg, Matrix Mechanics and the Uncertainty Princi-
ple", Resonance, Journal of Science Education, Volume 9, Number 8, August
DR
2004.
• Richard L. Liboff, Introductory Quantum Mechanics, 2nd ed. 1992.
• Lindsay, Robert Bruce and Henry Margenau, Foundations of Physics, Dover,
1936.
• McEnvoy, J.P., and Zarate, Oscar. Introducing Quantum Theory, ISBN 1-
84046-577-9
• Carl Rod Nave, Hyperphysics-Quantum Physics, Department of Physics and
Astronomy, Georgia State University, CD 2005.
• F. David Peat, "From Certainty to Uncertainty: The Story of Science and
Ideas in the Twenty-First Century", Joseph Henry Press, 2002.
• Reichenbach, Hans, Philosophic Foundations of Quantum Mechanics, Uni-
versity of California Press, 1944.
• Schilpp, Paul Arthur, Albert Einstein: Philosopher-Scientist, Tudor Publish-
ing Commpany, 1949.
• Scientific American Reader, 1953.

Introduction to quantum mechanics


250

• Sears, Francis Weston, Optics, Addison-Wesley, 1949.


• Dr. Kenjiro Takada, Emeritus professor of Kyushu University, Microscop-
ic World-Introduction to Quantum Mechanics, Internet seminar, http:
//www2.kutl.kyushu-u.ac.jp/seminar/MicroWorld1_E/MicroWorld_1_E.

FT
html.
• "Uncertainty Prirnciple" Werner Heisenberg actual voice recording, http:
//www.thebigview.com/spacetime/index.html.
• J.H. Van Vleck, The Correspondence Principle in the Statistical Interpreta-
tion of Quantum Mechanics, Proc. Nat. Acad. Sci., Vol. 14, p.179, 1928.
• Carl Wieman and Katherine Perkins, "Transforming Physics Education",
Physics Today, November 2005,

Notes
↑ Events leading up to the December 1900 publication of Planck’s quantum
hypothesis are related by Werner Heisenberg in his Physics and Philosophy,
pp. 30f. Heisenberg believes that Planck was clearly aware that his ideas
would have very far-reaching consequences.
↑ It is remarkable that Einstein made significant contributions to quantum
A
mechanics at the same time he was revolutionizing physics with his relativity
theories. The far-reaching nature of his contributions to quantum mechanics is
noted by Richard Feynman in QED: The Strange Theory of Light and Matter, p.
112: "[The] phenomenon of ’stimulated emission’ was discovered by Einstein
when he launched the quantum theory proposing the photon model of light.
Lasers work on the basis of this phenomenon."
↑ Albert Einstein characterized Niels Bohr’s contributions to the quantum rev-
DR
olution by saying that history "will have to connect one of the most important
advances ever made in our knowledge of the nature of the atom with the name
of Niels Bohr." He added, "The boldly selected hypothetical basis of his specu-
lations soon became a mainstay for the physics of the atom....The theory of the
Röntgen spectra of the visible spectra, and the periodic system of the elements
is primarily based on the ideas of Bohr." See Einstein’s Essays in Science, p. 46f.
↑ Niels Bohr records the contributions of Louis de Broglie toward "a more
comprehensive quantum theory" that would take into account that "the wave-
corpuscle duality was not confined to the properties of radiation, but was
equally unavoidable in accounting for the behaviour of material particles." See
his Atomic Physics and Human Knowledge, p. 37 et passim.
↑ See Max Born, Atomic Physics, especially p. 90, where he says of quantum
mechanics that it is "in the nature of the case indeterministic, and therefore the
affair of statistics."

Introduction to quantum mechanics


251

↑ A two-page account of the highlights of Dirac’s work, including his specu-


lation that a positron would be found, appears in an article on "The Ultimate
Particles" by George W. Gray, in The Scientific American Reader, p. 100f. (Si-
mon and Schuster, 1953).

FT
↑ Werner Heisenberg is well known for his "indeterminancy principle" or "un-
certainty principle."
↑ Wolfgang Pauli’s name is most closely associated with what is known as the
"Pauli Exclusion Principle," according to which, it is impossible, in the words
of Louis de Broglie, "for two electrons to have rigorously identical quantized
states, i.e., defined by the same quantum numbers.... Translated into wave
mechanics, Pauli’s principle is expressed as follows: ’for electrons, the only
states realized in nature are the antisymmetric states.’" (See de Broglie’s The
Revolution in Physics, p. 267)
↑ Schrödinger’s cat was originally a character in an example intended to be
critical of an apparent difficulty in Heisenberg’s exposition of his principle of
indeterminancy. The story has been taken somewhat out of context and the cat
has assumed a minor literary life of its own. Schrödinger’s other contributions
A
to understanding quantum mechanics and to making the mathematics easier
to handle were, of course, much more important. The translation of his 1935
essay that includes the story is to be found at http://www.qedcorp.com/pcr
/pcr/qcat.html. Schrödinger describes a situation in which a cat will live or die
depending on whether a quantum mechanically probabalistic radioactive emis-
sion event occurs within the hour the cat is confined to a box. To Heisenberg’s
interpretation of quantum mechanics he objects: "If one has left this entire sys-
tem 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
DR
having in it the living and dead cat (pardon the expression) mixed or smeared
out in equal parts."
↑ Huygens’ principle is explained in Optics, by Francis Weston Sears, Addison-
Wesley, 1949, pp. 5f.
↑ Sears, Optics, p. 2.
↑ Sears, Optics, p. 2f.
↑ Lindsay and Margenau, Foundations of Physics, p. 388}}
↑ Sears, Mechanics, Wave Motion, and Heat, p. 537.
↑ See Sears, Optics, pp. 282-293.
↑ Planck is quoted by Louis de Broglie, The Revolution in Physics, p. 106. The
material in this paragraph summarizes de Broglie’s account given on pages 105
to 108. (Noonday Press, New York, 1953)

Introduction to quantum mechanics


252

↑ A. Einstein, Ann. d. Phys., 17, 132, (1905).


↑ J. P. McEvoy and Oscar Zarate, Introducing Quantum Theory, p. 114 and p.
118.
↑ A. P. French and Edwin F. Taylor, An Introduction to Quantum Physics,, p.

FT
18.
↑ For the length of time involved, see George Gamow’s One, Two,
Three...Infinity, p. 140.
↑ A. P. French and Edwin F. Taylor, An Introduction to Quantum Physics,, p.
23.
↑ A very clear explanation of interference in thin films may be found in Sears,
op. cit., p. 203ff.
↑ See Linus Pauling, The Nature of the Chemical Bond, p. 35f. He explains
that a circular electron orbit in the case of the first orbital of the hydrogen
atom would predict orbital angular momentum for the atom, which it does not
in fact have. He therefore depicts the electron as moving from the nucleus and
back in a straight line.
A
↑ The German and English forms of this quotation appear in slightly dif-
ferent versions from place to place, probably because Einstein repeated his
original remark several time. The earliest German version can be found at
http://www.goettingen.de/kultur/gott_wuerfelt_nicht.htm. In it, Einstein first
speaks of God and then says, "And this one does not (dice =) play dice."
↑ Hans Reichenback works out the mathematics in one sample case of quantum
entanglement. See his Philosophic Foundations of Quantum Mechanics, p. 170
ff.
DR
↑ Yanhua H. Shih (2001), "Quantum Entanglement and Quantum Teleporta-
tion" Annals of Physics 10 (2001) 1-2 pp.45-61 as referenced by Amir Aczel
(2003), Entanglement p.252 ISBN 0-452-28457-0

External links
• Development of Current Atomic Theory 123
• Quantum Mechanics 124
• Planck’s original paper on Planck’s constant 125

123 http://chemed.chem.purdue.edu/genchem/topicreview/bp/ch6/bohr.html
124 http://www.aip.org/history/heisenberg/p07.htm
125 http://dbhs.wvusd.k12.ca.us/webdocs/Chem-History/Planck-1901/Planck-1901.html

Introduction to quantum mechanics


253

• Everything you wanted to know about the quantum world 126 — Provided
by New Scientist.
• Quantum Articles 127

FT
Source: http://en.wikipedia.org/wiki/Introduction_to_quantum_mechanics

Principal Authors: Voyajer, Patrick0Moran, Paulc1001, Ancheta Wis, David R. Ingham

Josephson effect

The Josephson effect is a term given to the phenomenon of current flow


across two superconductors separated by a very thin insulating barrier.
This arrangement—two superconductors linked by a non-conducting oxide
barrier—is known as a Josephson junction; the current that crosses the barri-
er is the Josephson current. The terms are named eponymously after British
physicist Brian David Josephson, who predicted the existence of the effect in
A1962 128. It has important applications in quantum-mechanical circuits.

The effect
The basic equations 129 governing the dynamics of the Josephson effect are
~ ∂φ
U (t) = 2e ∂t , I(t) = Ic sin(φ(t))

where U (t) and I(t) are the voltage and current across the Josephson junction,
DR
φ(t) is the phase difference between the wave functions in the two supercon-
ductors comprising the junction, and Ic is a constant, the critical current of the
junction. The critical current is an important phenomenological parameter of
the device that can be affected by temperature as well as by an applied mag-
h
netic field. The physical constant, 2e is the magnetic flux quantum, the inverse
of which is the Josephson constant.
The three main effects predicted by Josephson follow from these relations:

126 http://www.newscientist.com/channel/fundamentals/quantum-world
127 http://www.thequantumsite.com/
128 B. D. Josephson. The discovery of tunnelling supercurrents(http://prola.aps.org/abstract/RMP/v46

/i2/p251_1). Rev. Mod. Phys. 1974; 46(2): 251-254.


129 Barone A, Paterno G. Physics and Applications of the Josephson Effect. New York: John Wiley & Sons;

1982.

Josephson effect
254

• The DC Josephson effect. This refers to the phenomenon of a direct cur-


rent crossing the insulator in the absence of any external electromagnetic
field, owing to tunneling. This DC Josephson current is proportional to
the sine of the phase difference across the insulator, and may take values

FT
between −Ic and Ic .
• The AC Josephson effect. With a fixed voltage UDC across the junctions,
the phase will vary linearly with time and the current will be an AC current
2e
with amplitude Ic and frequency h UDC . This means a Josephson junction
can act as a perfect voltage-to-frequency converter.
• The inverse AC Josephson effect. If the phase takes the form φ(t) =
φ0 + nωt + a sin(ωt), the voltage and current will be
ω(n + a cos(ωt)), I(t) = Ic ∞
~ P
• :U (t) = 2e m=−∞ Jn (a) sin(φ0 + (n + m)ωt)
The DC components will then be
~
• :UDC = n 2e ω, I(t) = Ic J−n (a) sin φ0 Hence, for distinct DC voltages,
the junction may carry a DC current and the junction acts like a perfect
frequency-to-voltage converter.

The device
A
The Josephson junction finds numerous important applications. Its proper-
ties are exploited in →SQUIDs used to measure magnetic flux at the quantum
level. This finds application in medicine for measurement of small currents
in the brain and the heart. A version using different superfluids can be used
as a quantum gyroscope. Josephson junctions are also used in Rapid Single
Flux Quantum integrated circuits, and some other of their properties can be
exploited to build photon or particle detectors. Josephson junctions are used
as microwave detectors in the giga- and terahertz range. When assembled in
DR
two dimensional arrays, "testboards" for the physical realization of mathemat-
ical model systems are created. When assembled in linear arrays (connected
in series) the inverse Josephson effect is used as a representation of the SI unit
volt. It is also speculated that Josephson junctions may allow the realisation of
qubits, the key elements of a future quantum computer.
There are two general types of Josephson junctions: overdamped and under-
damped. In overdamped junctions, the barrier is conducting (ie. it is a normal
metal or superconductor bridge). The effects of the junction’s internal electri-
cal resistance will be large compared to its small capacitance. An overdamped
junction will quickly reach a unique equilibrium state for any given set of con-
ditions.
The barrier of an underdamped junction is an insulator. The effects of the
junction’s internal resistance will be minimal. Underdamped junctions do not
have unique equilibrium states, but are hysteretic.

Josephson effect
255

A FT
Figure 27 The current - voltage curve of a Josephson junction at low temperature. The ver-
tical portions (zero voltage) of the curve represent Cooper pair tunnelling. There is a small
magnetic field applied, so that the maximum Josephson current is severely reduced. Hysteresis
is clearly visible around 100 microvolts. The portion of the curve between 100 and 300 micro-
volts is current independent, and is the regime where the device can be used as a detector.
DR
A Josephson junction can be transformed into the so-called Giaever tunneling
junction by the application of a small, well defined magnetic field. In such
a situation, the new device is called a superconducting tunneling junction
(STJ) 130 and is used as a very sensitive photon detector throughout a wide
range of the spectrum, from infrared to hard x-ray. Each photon breaks up
a number of Cooper pairs. This number depends on the ratio of the photon
energy to approximately twice the value of the gap parameter of the material of
the junction. The detector can be operated as a photon-counting spectrometer,
with a spectral resolution limited by the statistical fluctuations in the number of
released charges. The detector has to be cooled to extremely low temperature,
typically below 1 kelvin, to distinguish the signals generated by the detector
from the thermal noise. Small arrays of STJs have demonstrated their potential

130 European Space Agency. Payload and Advanced Concepts: Superconducting Tunnel Junction
(STJ)(http://sci.esa.int/science-e/www/object/index.cfm?fobjectid=33525). Last updated: Febru-
ary 17 2005.

Josephson effect
256

as spectro-photometers and could further be used in astronomy. They are also


used to perform energy dispersive X-ray spectroscopy and in principle they
could be used as elements in infrared imaging devices as well. 131

FT
See also
• Quantum gyroscope
• Quantum computer

References

Source: http://en.wikipedia.org/wiki/Josephson_effect

Principal Authors: Filou, Encephalon, Danko Georgiev MD, Rorro, Marudubshinki

Klein-Gordon equation
AThe Klein-Gordon equation (Klein-Fock-Gordon equation or sometimes
Klein-Gordon-Fock equation) is the relativistic version of the →Schrödinger
equation. It was named after Oskar Klein and Walter Gordon.

Details
The Schrödinger equation for a free particle is
p2 ∂
2m ψ = i~ ∂t ψ
DR
where
p = −i~∇ is the momentum operator (∇ being the del operator).

The Schrödinger equation suffers from not being relativistically covariant,


meaning it does not take into account Einstein’s special relativity.
It is natural to try to use the identity from special relativity
p
E = p2 c2 + m2 c4

for the energy; then, just inserting the quantum mechanical momentum oper-
ator, yields the equation

131 Enss C (ed). Cryogenic Particle Detection. Topics in Applied Physics Vol. 99. Springer; 2005. ISBN

3-540-20113-0

Klein-Gordon equation
257


p
(−i~∇)2 c2 + m2 c4 ψ = i~ ∂t ψ.

This, however, is a cumbersome expression to work with because of the square


root. In addition, this equation, as it stands, is nonlocal.

FT
Klein and Gordon instead worked with the more general square of this equa-
tion (the Klein-Gordon equation for a free particle), which in covariant notation
reads
(2 + µ2 )ψ = 0,

where
mc
µ= ~

and
1 ∂2
2 = c2 ∂t2
− ∇2 .

This operator is called the d’Alembert operator. Today this form is interpreted
A
as the relativistic field equation for a scalar (i.e. spin-0) particle.
The Klein-Gordon equation was allegedly first found by Schrödinger, before
he made the discovery of the equation that now bears his name. He reject-
ed it because he couldn’t make it include the spin of the electron. The way
Schrödinger found his equation was by making simplifications in the Klein-
Gordon equation.
In 1926, soon after the Schrödinger equation was introduced, Fock wrote an
article about its generalization for the case of magnetic fields, where forces
DR
were dependent on velocity, and independently derived this equation. Both
Klein and Fock used Kaluza and Klein’s method. Fock also determined the
gauge theory for the wave equation. The Klein-Gordon equation for a free
particle has a simple plane wave solution.

Relativistic free particle solution


The Klein-Gordon equation for a free particle can be written as
1 ∂2 m2 c2
∇2 ψ − c2 ∂t2
ψ = ~2
ψ

with the same solution as in the non-relativistic case:


ψ(r, t) = ei(k·r−ωt)

except with the constraint

Klein-Gordon equation
258

ω2 m2 c2
−k 2 + c2
= ~2
.

Just as with the non-relativistic particle, we have for energy and momentum:
hpi = hψ| − i~∇|ψi = ~k,

FT

hEi = hψ|i~ ∂t |ψi = ~ω.

Except that now when we solve for k and ω and substitute into the constraint
equation, we recover the relationship between energy and momentum for rel-
ativistic massive particles:
hEi2 = m2 c4 + hpi2 c2 .

For massless particles, we may set m = 0 in the above equations. We then


recover the relationship between energy and momentum for massless particles:
hEi = h|p|ic.
ASee also
• Dirac equation
• Oskar Klein
• →Quantum field theory

References
DR
• Sakurai, J. J. (1967). Advanced Quantum Mechanics. Addison Wesley.
ISBN 0201067102.

External links
• Linear Klein-Gordon Equation 132 at EqWorld: The World of Mathematical
Equations.

• Nonlinear Klein-Gordon Equation 133 at EqWorld: The World of Mathemati-


cal Equations.
• generalizing the Klein-Gordon equation 134 to include a generalized space

132 http://eqworld.ipmnet.ru/en/solutions/lpde/lpde203.pdf
133 http://eqworld.ipmnet.ru/en/solutions/npde/npde2107.pdf
134 http://www.public.asu.edu/~kevinlg/i256/grand_paper.pdf

Klein-Gordon equation
259

Source: http://en.wikipedia.org/wiki/Klein-Gordon_equation

Principal Authors: PAR, Charles Matthews, Kurtan, Light current, MarSch, Cyp, Phys

FT
Ladder operators

In linear algebra (and its application to quantum mechanics), a raising or


lowering operator (collectively known as ladder operators) is an operator
that increases or decreases the eigenvalue of another operator. In quantum
mechanics, the raising operator is sometimes called the creation operator, and
the lowering operator the annihilation operator. Well-known applications of
ladder operators in quantum mechanics are in the formalisms of the quantum
harmonic oscillator and angular momentum.
Suppose that two operators X and N have the commutation relation
[N, X] = cX
A
for some scalar c. Then the operator X will act in such a way as to shift the
eigenvalue of an eigenstate of N by c:
In other words, if |ni is an eigenstate of N with eigenvalue n then X|ni is an
eigenstate of N with eigenvalue n + c. A raising operator for N is an operator
X for which c is real and positive and a lowering operator is one for which c is
real and negative.
If N is a Hermitian operator then c must be real and the Hermitian adjoint of
DR
X obeys the commutation relation:
[N, X † ] = −cX † .

In particular, if X is a lowering operator for N then X † is a raising operator for


N (and vice-versa).

See also
• →Creation and annihilation operators

Source: http://en.wikipedia.org/wiki/Ladder_operators

Principal Authors: Fropuff, Aza, Yevgeny Kats, RedWordSmith

Ladder operators
260

Laplace-Runge-Lenz vector

In the classical mechanics of a particle moving under a central force, the

FT
Laplace-Runge-Lenz vector is defined as:
A = p × L − mk rr

where:

• r is the position vector of the particle of mass m,


• L is the angular momentum,
• k is a parameter that describes strength of the central force

The Laplace-Runge-Lenz vector is constant when only inverse-square central


forces such as gravity or electrostatics act on the particle. In quantum mechan-
ics, the Laplace-Runge-Lenz vector can be used to derive the spectrum of the
hydrogen atom (without using the →Schrödinger equation).
The conservation of both the angular momentum vector and the Laplace-
A
Runge-Lenz vector results from a special symmetry of the system when only
inverse-square central forces act on the particle. Other central forces reduce
the system to a simpler rotational symmetry that conserves angular momentum
but not the Laplace-Runge-Lenz vector.

Conservation of the Laplace-Runge-Lenz vector


By assumption, the force F acting on the particle is a central force
DR
dp
F= dt = f (r) rr

for some function f (r) of the radius r. The angular momentum L ≡ r × p is


d
conserved under central forces, i.e., dt L = 0. Therefore,
h   i
d dp r dr m dr 2 dr
dt (p × L) = dt × L = f (r) r × r × m dt = f (r) r r r · dt − r dt

where the momentum p ≡ m dr dt as usual and where we have simplified the


triple cross product using Lagrange’s formula
 
r × r × dr
dt = r r · dr 2 dr
dt − r dt

Exploiting the identity


d dr d
r2 = 2r dr

dt (r · r) = 2r · dt = dt dt

Laplace-Runge-Lenz vector
261

we arrive at
h i
d 1 dr r dr d r
(p × L) = −mf (r)r2 = −mf (r)r2 dt

dt r dt − r2 dt r

−k
In the special case of an inverse-square force f (r) = , this becomes

FT
r2
 
d d r d mkr

dt (p × L) = mk dt r = dt r

Therefore, A is conserved for inverse-square forces


 
d d d mkr
dt A ≡ dt (p × L) − dt r =0

Analogous conserved quantities can be defined for other central forces, but
none are as simple as the Laplace-Runge-Lenz vector A.

Properties of the Laplace-Runge-Lenz vector


By its definition, A is perpendicular to L, i.e., A · L = 0. (Recall that r · L = 0,
because L ≡ r × p).
The Laplace-Runge-Lenz vector can be used to derive the elliptical orbits of the
A
Kepler problem
A · r ≡ Ar cos θ = r · (p × L) − mkr

where θ is the angle between the position and Laplace-Runge-Lenz vectors.


Permuting the scalar triple product r·(p × L) = L·(r × p) = L2 and rearranging
yields the formula for an ellipse
 
1 mk A
r = L2 1 + mk cos θ
DR
The vector points toward the pericenter, from the geometric center of the or-
bit to the attracting, central body. The magnitude for a periodic orbit with
eccentricity e is given by:
|A| = mke

The seven quantities A, L and E are related by the equations A · L = 0 and


A2 = m2 k 2 + 2mEL2

giving five independent constants of motion. This is consistent with the six
degrees of freedom (the particle’s initial position and velocity vectors, each
with three components) that specify the orbit of the particle, after removing
the initial time (which is not determined by A, L and E).

Laplace-Runge-Lenz vector
262

All central forces conserve angular momentum, but the Laplace-Runge-Lenz


vector is conserved only for inverse-square central forces. As shown by
Bertrand’s theorem, inverse-square forces can produce closed, elliptical orbits;
the constancy of the Laplace-Runge-Lenz vector corresponds to the constancy

FT
of the axes of the ellipse. The introduction of even small deviations from the
inverse-square force causes the axes of the ellipse to precess, i.e., to vary with
time.

History of the Laplace-Runge-Lenz vector


Laplace calculated the components of the vector A explicitly and showed that it
was conserved in his Traite de Mecanique Celeste (1799). It was rediscovered
independently by Hamilton in 1845. A vector derivation was given in 1900
by Gibbs (the inventor of vectors). This derivation was repeated by Runge
in a popular German textbook on vectors, which was referenced by Lenz in a
paper on the perturbed hydrogen atom (1924). Thus, it is also known as the
Runge-Lenz vector.

Poisson brackets of the angular momentum and


A
Laplace-Runge-Lenz vector
The Poisson brackets of the three components of the angular momentum vector,
Li , i = 1, 2, 3 are
 
Li , Lj = ijk Lk

where ijk is the fully antisymmetric tensor (i.e., the Levi-Civita symbol). The
Poisson brackets are represented here with square brackets, both for consis-
DR
tency with the references below and because they will be interpreted as Lie
brackets in the next section.
Defining a reduced Laplace-Runge-Lenz vector D with the same units as angu-
lar momentum
A
D≡ √
2m|E|

the Poisson brackets of D with the angular momentum vector L can be written
in a similar form
 
Di , Lj = ijk Dk

The Poisson brackets of the reduced Laplace-Runge-Lenz vector D with itself


depend on whether the total energy E is negative (closed, elliptical orbits) or

Laplace-Runge-Lenz vector
263

positive (open, hyperbolic orbits). For negative energies, the Poisson brackets
are
 
Di , Dj = ijk Lk

FT
whereas, for positive energy, the Poisson brackets have the opposite sign
 
Di , Dj = −ijk Lk

The Casimir invariants for negative energies are defined


C1 ≡ D · D + L · L

C2 ≡ D · L

which have zero Poisson brackets with all components of D and L


[C1 , Li ] = [C1 , Di ] = [C2 , Li ] = [C2 , Di ] = 0

For the Kepler problem, C2 = 0, since the two vectors are perpendicular.
A
Quantum mechanics and the Laplace-Runge-Lenz
vector
A simple prescription for quantizing a classical system is to set the commutation
relations of the quantum mechanical operators equal to the Poisson bracket of
the corresponding classical variables, multiplied by i~.
By carrying out this quantization and calculating the eigenvalues of the C1
DR
Casimir operator for the Kepler problem, Wolfgang Pauli was able to derive the
spectrum of hydrogen-like atoms. Specifically, he was able to show that the
energy levels vary as 1/n2 where n is an integer. This elegant derivation was
obtained prior to the development of wave mechanics; see the Bohm reference
below for details.

Symmetry and the classical Laplace-Runge-Lenz


vector
Noether’s theorem states that every conserved quantity in a physical system
corresponds to a continuous symmetry of the system, and vice versa. For gen-
eral central forces, one of the conserved quantities is the angular momentum,
for which the corresponding Noether symmetry is the rotation group SO(3).
For the special case of inverse-square force laws, a greater symmetry is pos-
sible, which results in the conservation of both the angular momentum and

Laplace-Runge-Lenz vector
264

the Laplace-Runge-Lenz vector. For negative energies, the symmetry is that of


four-dimensional rotation group SO(4) which conserves the length of a four-
dimensional vector
|e|2 = e21 + e22 + e23 + e24

FT
By contrast, for positive energies, the symmetry is that of the group of Lorentz
transformations SO(3,1), which conserves the Minkowski length of a 4-vector
ds2 = e21 + e22 + e23 − e24

These symmetries may be identified as follows. The Poisson brackets above


can be interpreted as Lie brackets that define a Lie algebra that corresponds
to a symmetry group. For negative energies, that group is SO(4) whereas, for
positive energies, that group is SO(3,1).

See also
• Astrodynamics: Orbit, Eccentricity vector, Orbital elements

A→Quantum mechanics

References
• Goldstein H. (1980) Classical Mechanics, 2nd ed., Addison Wesley, pp. 102-
105, 421-422.

• Landau LD and Lifshitz EM (1976) Mechanics, 3rd. ed., Pergamon Press.


DR
ISBN 0080210228 (hardcover) and ISBN 0080291414 (softcover).

• Goldstein H. (1975) "Prehistory of the Runge-Lenz vector", Am. J. Phys.,


43, 735-738.

• Goldstein H. (1976) "More on the prehistory of the Runge-Lenz vector", Am.


J. Phys., 44, 1123-1124.

• Lenz W. (1924) "Über den Bewegungsverlauf und Quantenzustaende der


gestörten Keplerbewegung", Z. Phys., 24, 197-207.

• Pauli W., Jr. (1926) "Über das Wasserstoffspektrum vom Standpunkt der
neuen Quantenmechanik", Z. Phys., 36, 336–363.

• Bohm A. (1986) Quantum Mechanics: Foundations and Applications, 2nd


ed., Springer Verlag, pp. 208-222.

Laplace-Runge-Lenz vector
265

• John Baez, Mysteries of the gravitational 2-body problem 135

Source: http://en.wikipedia.org/wiki/Laplace-Runge-Lenz_vector

FT
Principal Authors: WillowW, 0.39, AugPi, Charles Matthews, Agentsoo

Large Area Neutron Detector

The Large Area Neutron Detector is also know as LAND. It is the name of a
detector for neutrons installed at GSI (Institute for Heavy Ion Research) situat-
ed in Arheilgen close to the city of Darmstadt, Germany.
The detector is built of 10 planes with 20 paddles each. The paddles have
a size of 10x10x200cm 3 and are composed of a converter (iron) and plastic
scintillator material. Within the paddle the 5mm thick converters serve as a
dense target for neutrons (uncharged particles) leading to processes that eject
charged particles through the process of hadronic showers. The interspersed
A5mm plastic scintillator stripes produce light for the passing charged particles.
The stripes of one paddle are viewed at both ends by photomultipliers.
The detector system was built in 1990.
A research group at GSI is named after this detector. They aim at studying the
nuclear structure of radioactive (short-lived) nuclei.

See also
DR
• T. Blaich et al., Nuclear Instruments and Methods in Physics Research Sec-
tion A: Accelerators, Spectrometers, Detectors and Associated Equipment,
A314(1992), p. 136-154, Elsevier NORTH-HOLLAND, ISSN: 0168-9002

• GSI 136

Source: http://en.wikipedia.org/wiki/Large_Area_Neutron_Detector

135 http://math.ucr.edu/home/baez/gravitational.html
136 http://www.gsi.de

Large Area Neutron Detector


266

Lindblad equation

In quantum mechanics, the Lindblad equation or master equation in the Lind-

FT
blad form is the most general type of master equation allowed by quantum
mechanics to describe non-unitary (dissipative) evolution of the density matrix
ρ (such as ensuring normalisation and hermiticity of ρ). It reads:

ρ̇ = − ~i [H, ρ] − ~1 n,m hn,m ρLm Ln + Lm Ln ρ − 2Ln ρLm + h.c.
P

where ρ is the density matrix, H is the Hamiltonian part, Lm are operators


defined by the system to model as are the constants hn,m . It is a quantum
analog of the Liouville equation in classical mechanics. A related equation
describes the time evolution of the expectation values of observables, it is given
by the →Ehrenfest theorem.
The most common Lindblad equation is that describing the damping of a quan-
tum harmonic oscillator, it has L0 = a, L1 = a† , h0,1 = −(γ/2)(n̄ + 1),
h1,0 = −(γ/2)n̄ with all others hn,m = 0. Here n̄ is the mean number of excita-
A
tions in the reservoir damping the oscillator and γ is the decay rate. Additional
Lindblad operators can be included to model various forms of dephasing and
vibrational relaxation. These methods have been incorporated into grid-based
density matrix propagation methods.

Source: http://en.wikipedia.org/wiki/Lindblad_equation

Principal Authors: Laussy, Charles Matthews, Hansbethe


DR
London moment

The →London moment is a quantum-mechanical phenomenon whereby a


spinning superconductive metal sphere generates a magnetic field whose axis
lines up exactly with the spin axis. The term may also refer to the magnetic
moment of any rotation of any superconductor, caused by the electrons lagging
behind the rotation of the object.

Etymology
Named for the physical scientist Fritz London, and moment as in magnetic
moment.

London moment
267

See also
• London force
• Proton conductor

FT
References

Source: http://en.wikipedia.org/wiki/London_moment

Many-body problem

This article is about the many-body problem in quantum mechanics. For


the n-body problem in classical mechanics, see n-body problem.

Definition
A
The many-body problem may be defined as the study of the effects of inter-
action between bodies on the behaviour of a many-body system, i.e. a closed
system which does not contains just a few bodies in action, such as the colli-
sions discussed in classical mechanics. Due to the amount of particles/bodies
contained in such a system, we cannot any longer describe the mechanics of the
system, by using a small system of equations, as we do in classical mechanics.
The sheer number and high-energy chaotic interaction of each body with an-
DR
other forces us to use result in such mathimatical techniques as the canonical
transformation technique. Another preferred method for solving the problem,
is simply to ignore any interactions present. The many-body problem is usu-
ally posed in quantum mechanics, as the question of solving for more complex
problems than the hydrogen atom — for example, the chemistry of all realistic
molecules, such as a molecule of plastic.

Quotes
"It would indeed be remarkable if Nature fortified herself against further
advances in knowledge behind the analytical difficulties of the many-body
problem."

- Max Born, 1960

Many-body problem
268

See also
• Hartree-Fock approximation

FT
References and further reading
• Jenkins, Stephen. "3. The Many Body Problem and Density Functional
Theory". Many Body Problem and Density Functional Theory 137
• D. J. Thouless, "The quantum mechanics of many-body systems", 2d ed.,
New York, Academic Press, 1972. ISBN 0126915601
• D. J. Thouless, "The quantum mechanics of many-body systems", New York,
Academic Press, 1961. LCCN 61012282 /L/r842

External links
• Evidence for Efimov quantum states in an ultracold gas of Cs atoms 138
A
Source: http://en.wikipedia.org/wiki/Many-body_problem

Principal Authors: The Anome, Hephaestos, Omegatron, Grendelkhan, Pt, Conscious, Oneboy, Sev-
entyThree

Mathematical formulation of quantum me-


chanics
DR
One of the remarkable characteristics of the mathematical formulation of
quantum mechanics, which distinguishes it from mathematical formulations
of theories developed prior to the early 1900s, is its use of abstract mathemat-
ical structures, such as →Hilbert spaces and operators on these spaces. Many
of these structures had not even been considered before the twentieth century.
In a general sense they are drawn from functional analysis, a subject within
pure mathematics that developed in parallel with, and was influenced by the
needs of quantum mechanics. In brief, physical observables such as energy and
momentum were no longer considered as functions on some phase space, but
as eigenvalues of operators which act on such functions.

137 http://newton.ex.ac.uk/research/qsystems/people/jenkins/mbody/mbody3.html
138 http://www2.uibk.ac.at/exphys/ultracold/

Mathematical formulation of quantum mechanics


269

This formulation of quantum mechanics, called canonical quantization, con-


tinues to be used today, and still forms the basis of ab-initio calculations in
atomic, molecular and solid-state physics. At the heart of the description is
an idea of quantum state which, for systems of atomic scale, is radically dif-

FT
ferent from the previous models of physical reality. While the mathematics
is a complete description and permits calculation of many quantities that can
be measured experimentally, there is a definite limit to access for an observer
with macroscopic instruments. This limitation was first elucidated by Heisen-
berg through a thought experiment, and is represented mathematically by the
non-commutativity of quantum observables.
Prior to the emergence of quantum mechanics as a separate theory, the math-
ematics used in physics consisted mainly of differential geometry and partial
differential equations; probability theory was used in statistical mechanics. Ge-
ometric intuition clearly played a strong role in the first two and, according-
ly, theories of relativity were formulated entirely in terms of geometric con-
cepts. The phenomenology of quantum physics arose roughly between 1895
and 1915, and for the 10 to 15 years before the emergence of quantum theo-
ry (around 1925) physicists continued to think of quantum theory within the
A
confines of what is now called classical physics, and in particular within the
same mathematical structures. The most sophisticated example of this is the
Sommerfeld-Wilson-Ishiwara quantization rule, which was formulated entirely
on the classical phase space.

History of the formalism


The "old quantum theory" and the need for new mathematics
DR
Main article: Old quantum theory

In the decade of 1890, Planck was able to derive the blackbody spectrum and
solve the classical ultraviolet catastrophe by making the unorthodox assump-
tion that, in the interaction of radiation with matter, energy could only be
exchanged in discrete units which he called quanta. Planck postulated a direct
proportionality between the frequency of radiation and the quantum of ener-
gy at that frequency. The proportionality constant, h is now called Planck’s
constant in his honour.
In 1905, Einstein explained certain features of the photoelectric effect by as-
suming that Planck’s light quanta were actual particles, which he called pho-
tons.

Mathematical formulation of quantum mechanics


270

A
Figure 28
FT
A sketch to justify spectroscopy observations for hydrogen atoms

In 1913, Bohr calculated the spectrum of the hydrogen atom with the help of
a new model of the atom in which the electron could orbit the proton only
on a discrete set of classical orbits, determined by the condition that angular
momentum was an integer multiple of Planck’s constant. Electrons could make
DR
quantum leaps from one orbit to another, emitting or absorbing single quanta
of light at the right frequency.
All of these developments were phenomenological and flew in the face of the
theoretical physics of the time. Bohr and Sommerfeld went on to modify clas-
sical mechanics in an attempt to deduce the Bohr model from first principles.
They proposed that, of all closed classical orbits traced by a mechanical system
in its phase space, only the ones that enclosed an area which was a multiple
of Planck’s constant were actually allowed. The most sophisticated version
of this formalism was the so-called Sommerfeld-Wilson-Ishiwara quantization.
Although the Bohr model of the hydrogen atom could be explained in this way,
the spectrum of the helium atom (classically an unsolvable 3-body problem)
could not be predicted. The mathematical status of quantum theory remained
uncertain for some time.

Mathematical formulation of quantum mechanics


271

In 1923 de Broglie proposed that wave-particle duality applied not only to


photons but to electrons and every other physical system.
The situation changed rapidly in the years 1925-1930, when working math-
ematical foundations were found through the groundbreaking work of Erwin

FT
Schrödinger and Werner Heisenberg and the foundational work of John von
Neumann, Hermann Weyl and Paul Dirac, and it became possible to unify sev-
eral different approaches in terms of a fresh set of ideas.

The "new quantum theory"


Erwin Schrödinger’s wave mechanics originally was the first successful at-
tempt at replicating the observed quantization of atomic spectra with the
help of a precise mathematical realization of de Broglie’s wave-particle duality.
Schrödinger proposed an equation (now bearing his name) for the wave asso-
ciated to an electron in an atom according to de Broglie, and explained energy
quantization by the well-known fact that differential operators of the kind ap-
pearing in his equation had a discrete spectrum. However, Schrödinger himself
initially did not understand the fundamental probabilistic nature of quantum
mechanics, as he thought that the (squared amplitude of the) wavefunction
A
of an electron must be interpreted as the charge density of an object smeared
out over an extended, possibly infinite, volume of space. It was Max Born who
introduced the probabilistic interpretation of the (squared amplitude of the)
wave function as the probability distribution of the position of a pointlike ob-
ject. With hindsight, Schrödinger’s wave function can be seen to be closely
related to the classical Hamilton-Jacobi equation.
Werner Heisenberg’s matrix mechanics formulation, introduced contempora-
neously to Schrödinger’s wave mechanics and based on algebras of infinite ma-
DR
trices, was certainly very radical in light of the mathematics of classical physics.
In fact, at the time linear algebra was not generally known to physicists in its
present form.
The reconciliation of the two approaches is generally associated to Paul Dirac,
who wrote a lucid account in his 1930 classic Principles of Quantum mechanics.
In it, he introduced the bra-ket notation, together with an abstract formulation
in terms of the →Hilbert space used in functional analysis, and showed that
Schödinger’s and Heisenberg’s approaches were two different representations
of the same theory. Dirac’s method is now called canonical quantization. The
first complete mathematical formulation of this approach is generally credited
to John von Neumann’s 1932 book Mathematical Foundations of Quantum Me-
chanics, although Hermann Weyl had already referred to Hilbert spaces (which
he called unitary spaces) in his 1927 classic book. It was developed in parallel

Mathematical formulation of quantum mechanics


272

with a new approach to the mathematical spectral theory based on linear op-
erators rather than the quadratic forms that were David Hilbert’s approach a
generation earlier.
Though theories of quantum mechanics continue to evolve to this day, there

FT
is a basic framework for the mathematical formulation of quantum mechanics
which underlies most approaches and can be traced back to the mathematical
work of John von Neumann. In other words, discussions about interpretation
of the theory, and extensions to it, are now mostly conducted on the basis of
shared assumptions about the mathematical foundations.

Later developments
The application of the new quantum theory to electromagnetism resulted in
quantum field theory, which was developed starting around 1930. Quantum
field theory has driven the development of more sophisticated formulations
of quantum mechanics, of which the one presented here is a simple special
case. In fact, the difficulties involved in implementing any of the following
formulations cannot be said yet to have been solved in a satisfactory fashion
except for ordinary quantum mechanics.


AFeynman path integrals
• axiomatic, algebraic and constructive quantum field theory
• geometric quantization
• quantum field theory in curved spacetime
• C* algebra formalism

On a different front, von Neumann originally dispatched quantum measure-


ment with his infamous postulate on the collapse of the wavefunction, raising
DR
a host of philosophical problems. Over the intervening 70 years, the problem
of measurement became an active research area and itself spawned some new
formulations of quantum mechanics.

• Relative state/Many-worlds interpretation of quantum mechanics


• Decoherence
• Consistent histories formulation of quantum mechanics
• Quantum logic formulation of quantum mechanics

A related topic is the relationship to classical mechanics. Any new physical


theory is supposed to reduce to successful old theories in some approximation.
For quantum mechanics, this translates into the need to study the so-called
classical limit of quantum mechanics. Also, as Bohr emphasized, human cog-
nitive abilities and language are inextricably linked to the classical realm, and
so classical descriptions are intuitively more accessible than quantum ones. In

Mathematical formulation of quantum mechanics


273

particular, quantization, namely the construction of a quantum theory whose


classical limit is a given and known classical theory, becomes an important area
of quantum physics in itself.
Finally, some of the originators of quantum theory (notably Einstein and

FT
Schrödinger) were unhappy with what they thought were the philosophical
implications of quantum mechanics. In particular, Einstein took the position
that quantum mechanics must be incomplete, which motivated research into
so-called hidden-variable theories. The issue of hidden variables has become
in part an experimental issue with the help of quantum optics.

• de Broglie- Bohm- Bell pilot wave formulation of quantum mechanics


• Bell’s inequalities
• Kochen-Specker theorem

Mathematical structure of quantum mechanics


A physical system is generally described by three basic ingredients: states; ob-
servables; and dynamics (or law of time evolution) or, more generally, a group
of physical symmetries. A classical description can be given in a fairly direct
A
way by a phase space model of mechanics: states are points in a symplec-
tic phase space, observables are real-valued functions on it, time evolution is
given by a one-parameter group of symplectic transformations of the phase
space, and physical symmetries are realized by symplectic transformations. A
quantum description consists of a →Hilbert space of states, observables are
self adjoint operators on the space of states, time evolution is given by a one-
parameter group of unitary transformations on the Hilbert space of states, and
physical symmetries are realized by unitary transformations.
DR
Postulates of quantum mechanics
The following summary of the mathematical framework of quantum mechanics
can be partly traced back to von Neumann’s postulates.

• Each physical system is associated with a (topologically) separable complex


→Hilbert space H with inner product hφ | ψi. Rays (one-dimensional sub-
spaces) in H are associated with states of the system. In other words, phys-
ical states can be identified with equivalence classes of vectors of length 1
in H, where two vectors represent the same state if they differ only by a
phase factor. Separability is a mathematically convenient hypothesis, with

Mathematical formulation of quantum mechanics


274

the physical interpretation that countably many observations are enough to


uniquely determine the state.
• The Hilbert space of a composite system is the Hilbert space tensor product
of the state spaces associated with the component systems. For a non-

FT
relativistic system consisting of a finite number of distinguishable particles,
the component systems are the individual particles.
• Physical symmetries act on the Hilbert space of quantum states unitarily or
antiunitarily (supersymmetry is another matter entirely).
• Physical observables are represented by densely-defined self-adjoint opera-
tors on H.

The expected value (in the sense of probability theory) of the observable A
for the system in state represented by the unit vector |ψi ∈ H is

hψ | A | ψi

By spectral theory, we can associate a probability measure to the values of


A in any state ψ. We can also show that the possible values of the observ-
A
able A in any state must belong to the spectrum of A. In the special case A
has only discrete spectrum, the possible outcomes of measuring A are its
eigenvalues.

More generally, a state can be represented by a so-called density operator,


which is a trace class, nonnegative self-adjoint operator ρ normalized to be
of trace 1. The expected value of A in the state ρ is
DR
tr(Aρ)

If ρψ is the orthogonal projector onto the one-dimensional subspace of H


spanned by |ψi, then

tr(Aρψ ) = hψ | A | ψi

Density operators are those that are in the closure of the convex hull of
the one-dimensional orthogonal projectors. Conversely, one-dimensional
orthogonal projectors are extreme points of the set of density operators.
Physicists also call one-dimensional orthogonal projectors pure states and
other density operators mixed states.

Mathematical formulation of quantum mechanics


275

One can in this formalism state Heisenberg’s uncertainty principle and prove
it as a theorem, although the exact historical sequence of events, concerning
who derived what and under which framework, is the subject of historical in-
vestigations outside the scope of this article.

FT
Superselection sectors. The correspondence between states and rays needs
to be refined somewhat to take into account so-called superselection sectors.
States in different superselection sectors cannot influence each other, and the
relative phases between them are unobservable.

Pictures of dynamics
In the so-called →Schrödinger picture of quantum mechanics, the dynamics is
given as follows:
The time evolution of the state is given by a differentiable function from the
real numbers R, representing instants of time, to the Hilbert space of system
states. This map is characterized by a differential equation as follows: If |ψ (t)i
denotes the state of the system at any one time t, the following →Schrödinger
equation holds:
A d
i~ dt |ψ(t)i = H |ψ(t)i

where H is a densely-defined self-adjoint operator, called the system Hamil-


tonian, i is the imaginary unit and ~ is the reduced Planck constant. As an
observable, H corresponds to the total energy of the system.
Alternatively, by Stone’s theorem one can state that there is a strongly contin-
uous one-parameter unitary group U (t): H → H such that
|ψ(t + s)i = U (t) |ψ(s)i
DR
for all times s, t. The existence of a self-adjoint Hamiltonian H such that
U (t) = e−(i/~)tH

is a consequence of Stone’s theorem on one-parameter unitary groups.


The →Heisenberg picture of quantum mechanics focuses on observables and
instead of considering states as varying in time, it regards the states as fixed
and the observables as changing. To go from the Schrödinger to the Heisen-
berg picture one needs to define time-independent states and time-dependent
operators thus:
|ψi = |ψ(0)i

Mathematical formulation of quantum mechanics


276

A(t) = U (−t)AU (t)

It is then easily checked that the expected values of all observables are the
same in both pictures

FT
hψ | A(t) | ψi = hψ(t) | A | ψ(t)i

and that the time-dependent Heisenberg operators satisfy


d
i~ dt A(t) = [A(t), H]

This assumes A is not time dependent in the Schrödinger picture. Notice the
commutator expression is purely formal when one of the operators is unbound-
ed. One would specify a representation for the expression to make sense of it.
The so-called Dirac picture or interaction picture has time-dependent states
and observables, evolving with respect to different Hamiltonians. This picture
is most useful when the evolution of the states can be solved exactly, confin-
ing any complications to the evolution of the operators. For this reason, the
Hamiltonian for states is called "free Hamiltonian" and the Hamiltonian for
A
observables is called "interaction Hamiltonian". In symbols:
d
i~ dt |ψ(t)i = H0 |ψ(t)i

d
i~ dt A(t) = [A(t), Hint ]

The interaction picture does not always exist, though. In interacting quantum
field theories, Haag’s theorem states that the interaction picture does not exist.
DR
This is because the Hamiltonian cannot be split into a free and an interacting
part within a superselection sector.
The Heisenberg picture is the closest to classical mechanics, but the
Schrödinger picture is considered easiest to understand by most people, to
judge from pedagogical accounts of quantum mechanics. The Dirac picture
is the one used in perturbation theory, and is specially associated to quantum
field theory.
Similar equations can be written for any one-parameter unitary group of sym-
metries of the physical system. Time would be replaced by a suitable coor-
dinate parameterizing the unitary group (for instance, a rotation angle, or a
translation distance) and the Hamiltonian would be replaced by the conserved
quantity associated to the symmetry (for instance, angular or linear momen-
tum).

Mathematical formulation of quantum mechanics


277

Representations
The original form of the →Schrödinger equation depends on choosing a par-
ticular representation of Heisenberg’s canonical commutation relations. The
Stone-von Neumann theorem states all irreducible representations of the finite-

FT
dimensional Heisenberg commutation relations are unitarily equivalent. This
is related to quantization and the correspondence between classical and quan-
tum mechanics, and is therefore not strictly part of the general mathematical
framework.
The quantum harmonic oscillator is an exactly-solvable system where the pos-
sibility of choosing among more than one representation can be seen in all its
glory. There, apart from the Schrödinger (position or momentum) represen-
tation one encounters the Fock (number) representation and the Bargmann-
Segal (phase space or coherent state) representation. All three are unitarily
equivalent.

Time as an operator
The framework presented so far singles out time as the parameter that every-
A
thing depends on. It is possible to formulate mechanics in such a way that time
becomes itself an observable associated to a self-adjoint operator. At the clas-
sical level, it is possible to arbitrarily parameterize the trajectories of particles
in terms of an unphysical parameter s, and in that case the time t becomes an
additional generalized coordinate of the physical system. At the quantum level,
translations in s would be generated by a "Hamiltonian" H -E, where E is the
energy operator and H is the "ordinary" Hamiltonian. However, since s is an
unphysical parameter, physical states must be left invariant by "s-evolution",
and so the physical state space is the kernel of H -E (this requires the use of a
DR
rigged Hilbert space and a renormalization of the norm).
This is related to quantization of constrained systems and quantization of
gauge theories. It is also possible to formulate a quantum theory of "events"
where time becomes an observable( see D. Edwards ).

The problem of measurement


The picture given in the preceding paragraphs is sufficient for description of a
completely isolated system. However, it fails to account for one of the main
differences between quantum mechanics and classical mechanics, that is the
effects of measurement. The von Neumann description of quantum measure-
ment of an observable A, when the system is prepared in a pure state ψ is the
following:

• Let A have spectral resolution

Mathematical formulation of quantum mechanics


278
R
A= λd EA (λ)
,

where E A is the resolution of the identity (also called projection-valued mea-

FT
sure) associated to A. Then the probability of the measurement outcome lying
in an interval B of R is |E A (B) ψ| 2. In other words, the probability is obtained
by integrating the characteristic function of B against the countably additive
measure
hψ | EA ψi
.

• If the measured valued is contained in B, then immediately after the mea-


surement, the system will be in the (generally non-normalized) state E A
ψ.

If the measured value does not lie in B, replace B by its complement for the
above state.
A
For example, suppose the state space is the n-dimensinal complex Hilbert space
C n and A is a Hermitian matrix with eigenvalues i , with corresponding eigen-
vectors ψ i . The projection-valued measure associated with A is E A is then
EA (B) = |ψi ihψi |
,

where B is a Borel set containing only the single eigenvalue i . If the system is
DR
prepared in state
|ψi

Then the probability of a measurement returning the value i can be calculated


by integrating the spectral measure
hψ | EA ψi

over B i . This gives trivially


hψ|ψi ihψi | ψi = |hψ | ψi i|2 .

The characteristic property of the von Neumann measurement scheme is that


repeating the same measurement will give the same results. This is also called
the projection postulate.

Mathematical formulation of quantum mechanics


279

A more general formulation replaces the projection-valued measure with a


positive-operator valued measure(POVM). To illustrate, take again the finite-
dimensional case. Here we would replace the rank-1 projections |ψi ihψi | by a
finite set of positive operators Fi Fi∗ , whose sum is still the identity operator

FT
as before. Just as a set of possible outcomes λ1 · · · λn is associated to a PVM,
the same can be said for a POVM. Suppose the measurement outcome is λi .
Instead of collapsing to the (unnormalized) state |ψi ihψi |ψi after the measure-
ment, the system now will be in the state Fi |ψi. Since the Fi Fi∗ ’s need not be
mutually orthogonal projections, the projection postulate of von Neumann no
longer holds.
Both of the above approaches apply to general mixed ensembles.
In von Neumann’s approach, the state transformation due to measurement is
distinct from that due to time evolution in several ways. For example, time evo-
lution is determinisic and unitary whereas measurement is non-deterministic
and non-unitary. However, since both types of state transformation take one
quantum state to another, this difference was viewed by many as unsatisfactory.
The POVM formalism views measurement as one among many other quantum
operations, which are described by completely positive maps which do not in-
A
crease the trace.

The relative state interpretation


An alternative interpretation of measurement is Everett’s relative state interpre-
tation, which was later dubbed the "many-worlds interpretation" of quantum
mechanics.

List of mathematical tools


DR
Part of the folklore of the subject concerns the mathematical physics textbook
Courant-Hilbert, put together by Richard Courant from David Hilbert’s Göttin-
gen University courses. The story is told (by mathematicians) that physicists
had dismissed the material as not interesting in the current research areas, un-
til the advent of Schrödinger’s equation. At that point it was realised that the
mathematics of the new quantum mechanics was already laid out in it. It is
also said that Heisenberg had consulted Hilbert about his matrix mechanics,
and Hilbert observed that his own experience with infinite-dimensional matri-
ces had derived from differential equations, advice which Heisenberg ignored,
missing the opportunity to unify the theory as Weyl and Dirac did a few years
later. Whatever the basis of the anecdotes, the mathematics of theory was
conventional at the time, where the physics was radically new.
The main tools include:

Mathematical formulation of quantum mechanics


280

• linear algebra: complex numbers, eigenvectors, eigenvalues


• functional analysis: →Hilbert spaces, linear operators, spectral theory
• differential equations: Partial differential equations, Separation of vari-
ables, Ordinary differential equations, Sturm-Liouville theory, eigenfunc-

FT
tions
• harmonic analysis: Fourier transforms

See also: list of mathematical topics in quantum theory.

References
• S. Auyang, How is Quantum Field Theory Possible?, Oxford University
Press, 1995.
• D. Edwards, The Mathematical Foundations of Quantum Mechanics, Syn-
these, 42 (1979),pp.1-70.
• G. Emch, Algebraic Methods in Statistical Mechanics and Quantum Vield
Theory, Wiley-Interscience, 1972.
• R. Jost, The General Theory of Quantized Fields, American Mathematical


ASociety, 1965.
A. Gleason, Measures on the Closed Subspaces of a Hilbert Space, Journal
of Mathematics and Mechanics, 1957.
• G. Mackey, Mathematical Foundations of Quantum Mechanics, W. A. Ben-
jamin, 1963 (paperback reprint by Dover 2004).
• J. von Neumann, Mathematical Foundations of Quantum Mechanics,
Princeton University Press, 1955. Reprinted in paperback form.
• R. F. Streater and A. S. Wightman, PCT, Spin and Statistics and All That,
Benjamin 1964 (Reprinted by Princeton University Press)
DR
• M. Reed and B. Simon, Methods of Mathematical Physics, vols 1-IV, Aca-
demic Press 1972.
• H. Weyl, The Theory of Groups and Quantum Mechanics, Dover Publica-
tions, 1950.

Source: http://en.wikipedia.org/wiki/Mathematical_formulation_of_quantum_mechanics

Principal Authors: Miguel, CSTAR, Charles Matthews, Phys, Mct mht

Mathematical formulation of quantum mechanics


281

Matrix mechanics
Matrix mechanics is a formulation of quantum mechanics created by Werner

FT
Heisenberg in 1925.
Matrix mechanics was the first complete definition of quantum mechanics, its
laws, and properties that described fully the behavior of subatomic particles
by associating their properties with matrices. It has been shown to be exactly
equivalent to the Schroedinger wave formulation of quantum mechanics and is
the basis of the bra-ket notation used to summarize quantum mechanical wave
functions.

Development of matrix mechanics


When it was introduced by Werner Heisenberg, Max Born and Pascual Jordan
in 1925, matrix mechanics was not immediately accepted and was a source of
great controversy.
Schrödinger’s later introduction of wave mechanics was favored because there
were no visual aids to fall back on in matrix mechanics and the mathematics
A
were unfamiliar to most physicists.
Matrix mechanics consists of an array of quantities which when appropriate-
ly manipulated gave the observed frequencies and intensities of spectral lines.
Heisenberg said himself that once and for all he had gotten rid of all electron
orbits that did not exist. However, in Heisenberg’s theory, the result of multipli-
cation changes depending on its order. This means that the physical quantities
in Heisenberg’s theory are not ordinary variables but mathematical matrices.
DR
Heisenberg developed matrix mechanics by interpreting the electron as a par-
ticle with quantum behavior. It is based on sophisticated matrix computations
which introduce discontinuities and quantum jumps.
In atomic physics, through spectroscopy, it was known that observational da-
ta related to atomic transitions arise from interactions of the atoms with light
quanta. Heisenberg was the first to say that the atomic spectrum which showed
spectral lines only in places where photons were being absorbed or emitted as
electrons changed orbitals were the only relevant objects to be defined. Heisen-
berg recognized that the matrix formulation was built on the premise that all
physical observables must be represented by matrices. The set of eigenval-
ues of the matrix representing an observable is the set of all possible values
that could arise as outcomes of experiments conducted on a system to mea-
sure the observable. Since the outcome of an experiment to measure a real
observable must be a real number, Hermitian matrices would represent such
observables as their eigenvalues are real. If the result of a measurement is a

Matrix mechanics
282

certain eigenvalue, the corresponding eigenvector represents the state of the


system immediately after the measurement.
Instead of using three dimensional orbitals, Heisenberg’s matrix mechanics de-
scribed the space in which the state of a quantum system inhabits as being

FT
one-dimensional as in the case of an anharmonic oscillator. To illustrate, con-
sider the simple example of a point particle like an electron that is free to move
on a line. An observable in this case could be the position of the particle, rep-
resented by the matrix X. Since the particle could be anywhere on the line,
the possible outcome of a measurement of its position could be any one of
an infinite set of eigenvalues of X, denoted by x. Thus X must be an infinite-
dimensional matrix, and hence so is the corresponding linear vector space.
Thus even one-dimensional motion could have an infinite-dimensional linear
vector space associated with it. This made operators, functions, and spaces
necessary to describe quantum mechanics.
The act of measurement in matrix mechanics is taken to ’collapse’ the state
of the system to that eigenvector (or eigenstate). Anyone familiar with
Schroedinger’s wave equation which came later in 1925 will be familiar with
this concept in the form of wavefunction collapse. If one were to make simul-
A
taneous measurements of two or more observables, the system will collapse to
a common eigenstate of these observables right after the measurement.
Further, from matrix theory we know that eigenvectors corresponding to dis-
tinct eigenvalues of a Hermitian matrix are orthogonal to each other which is
analogous to the x, y, z axes of the Cartesian coordinate system except with an
infinite number of distinct eigenvalues, and hence as many mutually orthogo-
nal eigenvectors directed along different independent directions in the linear
vector space.
DR
Prior to measurement, the system could have been in a linear superposition of
different eigenstates, with impossible to know coefficients. The precise state of
the quantum system before the measurement cannot be known. The Copen-
hagen interpretation is concerned only with outcomes of experiments.
Since a single measurement of any observable A yields one of the eigenvalues
of A as the outcome, and collapses the state of the system to the correspond-
ing eigenstate, subsequent measurements made immediately thereafter would
continue to yield the same eigenvalue. So the correct thing to do would be to
prepare a collection of a very large number of identical copies of the system
and conduct a single trial on each copy. The arithmetic mean of all the results
thus obtained is the average value we see, denoted by (A).
The Uncertainty Principle in matrix mechanics stems from the fact that, in gen-
eral, two matrices A and B do not obey the arithmetical law of commutation.

Matrix mechanics
283

The commutator A B - B A = [A, B] does not equal 0. The famous commuta-


tion relation that is the basis for Quantum Mechanics and the later Uncertainty
Principle
P is:
k p(n, k)q(k, n) − q(n, k)p(k, n) = h/2πi

FT
In 1925, Werner Heisenberg was not yet 24 years old.

Mathematical details
In quantum mechanics in the Heisenberg picture the state vector, |ψ> does
not change with time, and an observable A satisfies
 
d
A = (i~)−1 [A, H] + ∂A .
dt ∂t classical

In some sense, the Heisenberg picture is more natural and fundamental than
the →Schrödinger picture, especially for relativistic theories. Lorentz invari-
ance is manifest in the Heisenberg picture.
Moreover, the similarity to classical physics is easily seen: by replacing the
commutator above by the Poisson bracket, the Heisenberg equation becomes
an equation in Hamiltonian mechanics.
A
By the Stone-von Neumann theorem, the Heisenberg picture and the
→Schrödinger picture are unitarily equivalent.
See also →Schrödinger picture.

Deriving Heisenberg’s equation


Suppose we have an observable A (which is a Hermitian linear operator). The
expectation value of A for a given state |ψ(t)> is given by:
DR
hAit = hψ(t)|A|ψ(t)i

or if we write following the →Schrödinger equation


|ψ(t)i = e−iHt/~ |ψ(0)i

(where H is the Hamiltonian and h\hbar is Planck’s constant divided by 2*pi )


we get
hAit = hψ(0)|eiHt/~ Ae−iHt/~ |ψ(0)i

and so we define
A(t) := eiHt/~ Ae−iHt/~

Now,

Matrix mechanics
284
 
d
dt A(t) = ~i HeiHt/~ Ae−iHt/~ + ∂A
∂t classical + ~i eiHt/~ A · (−H)e−iHt/~

(differentiating according to the product rule),


 
= ~i eiHt/~ (HA − AH) e−iHt/~ + ∂A ∂t classical =
i
(HA(t) − A(t)H) +

FT
~
 
∂A
∂t classical

(the last passage is valid since exp(-iHt/hbar) commutes with H )


 
= ~i [H, A(t)] + ∂A∂t classical

(where [X,Y] is the commutator of two operators and defined as [X, Y ] :=


XY − Y X)
So we get
 
d
dt A(t) = ~i [H, A(t)] + ∂A
∂t classical

Though matrix mechanics does not include concepts such as the wave function
Aof Erwin Schrödinger’s wave equation, the two approaches were proven to be
mathematically equivalent by the mathematician John von Neumann.

See also
• →Interaction picture
• →Quantum mechanics
• →Schrödinger equation
• →Bra-ket notation
DR
• Basic quantum mechanics

External links
• An Overview of Matrix Mechanics 139
• Matrix Methods in Quantum Mechanics 140
• Heisenberg Quantum Mechanics 141

Source: http://en.wikipedia.org/wiki/Matrix_mechanics

Principal Authors: Voyajer, Zowie, Elroch, Snoyes, Neilc

139 http://www.cobalt.chem.ucalgary.ca/ziegler/educmat/chm386/rudiment/tourquan/matmech.htm
140 http://www.cobalt.chem.ucalgary.ca/ziegler/educmat/chm386/rudiment/quanmath/matrix.htm
141 http://www.aip.org/history/heisenberg/p08.htm

Matrix mechanics
285

Matrix model

The term matrix model may refer to one of several concepts:

FT
• In theoretical physics, a matrix model is a system (usually a quantum me-
chanical system) with matrix-valued physical quantities. See, for example,
Lax pair.

• The "old" matrix models are relevant for string theory in two spacetime
dimensions. The "new" matrix model is a synonym for Matrix theory.

• Matrix population models are used to model wildlife and human population
dynamics.

Source: http://en.wikipedia.org/wiki/Matrix_model

Principal Authors: QFT, Conscious, Charles Matthews, A2Kafir, Shenme


A
Maxwell-Boltzmann statistics

In statistical mechanics, Maxwell-Boltzmann statistics describes the statisti-


cal distribution of material particles over various energy states in thermal equi-
librium, when the temperature is high enough and density is low enough to
render quantum effects negligible. Maxwell-Boltzmann statistics are therefore
DR
applicable to almost any terrestrial phenomena for which the temperature is
above a few tens of kelvins.
The expected number of particles with energy i for Maxwell-Boltzmann statis-
tics is Ni where:
Ni gi gi e−i /kT
N = = Z
e(i −µ)/kT

where:

• Ni is the number of particles in state i


• i is the energy of the i -th state
• gi is the degeneracy of state i, the number of microstates with energy i
• µ is the chemical potential
• k is Boltzmann’s constant

Maxwell-Boltzmann statistics
286

• T is absolute temperature
• N is the total number of particles
P
N= i Ni

FT
• Z is the partition function

Z=
P −i /kT
i gi e

• e (...) is the exponential function

Equivalently, the distribution is sometimes expressed as


Ni 1 e−i /kT
N = = Z
e(i −µ)/kT

where the index i now specifies an individual microstate rather than the set of
all states with energy i
Fermi-Dirac and Bose-Einstein statistics apply when quantum effects have to
A
be taken into account and the particles are considered "indistinguishable". The
quantum effects appear if the concentration of particles (N/V) ≥ n q (where
n q is the quantum concentration). The quantum concentration is when the
interparticle distance is equal to the thermal de Broglie wavelength i.e. when
the wavefunctions of the particles are touching but not overlapping. As the
quantum concentration depends on temperature; high temperatures will put
most systems in the classical limit unless they have a very high density e.g.
a White dwarf. Fermi-Dirac statistics apply to fermions (particles that obey
DR
the Pauli exclusion principle), Bose-Einstein statistics apply to bosons. Both
Fermi-Dirac and Bose-Einstein become Maxwell-Boltzmann statistics at high
temperatures or low concentrations.
Maxwell-Boltzmann statistics are often described as the statistics of "distin-
guishable" classical particles. In other words the configuration of particle A in
state 1 and particle B in state 2 is different from the case where particle B is
in state 1 and particle A is in state 2. When this idea is carried out fully, it
yields the proper (Boltzmann) distribution of particles in the energy states, but
yields non-physical results for the entropy, as embodied in →Gibbs paradox.
These problems disappear when it is realized that all particles are in fact in-
distinguishable. Both of these distributions approach the Maxwell-Boltzmann
distribution in the limit of high temperature and low density, without the need

Maxwell-Boltzmann statistics
287

for any ad hoc assumptions. Maxwell-Boltzmann statistics are particularly use-


ful for studying gases F-D statistics are most often used for the study of elec-
trons in solids. As such, they form the basis of semiconductor device theory
and electronics.

FT
A derivation of the Maxwell-Boltzmann distribution
In this particular derivation, the Boltzmann distribution will be derived using
the assumption of distinguishable particles, even though the ad hoc correction
for Boltzmann counting is ignored, the results remain valid.
Suppose we have a number of energy levels, labelled by index i , each level
having energy i and containing a total of Ni particles. To begin with, lets
ignore the degeneracy problem. Assume that there is only one way to put Ni
particles into energy level i.
The number of different ways of performing an ordered selection of one object
from N objects is obviously N . The number of different ways of selecting
2 objects from N objects, in a particular order, is thus N (N − 1) and that
of selecting n objects in a particular order is seen to be N !/(N − n)!. The
A
number of ways of selecting 2 objects from N objects without regard to order
is N (N − 1) divided by the number of ways 2 objects can be ordered, which is
2!. It can be seen that the number of ways of selecting n objects from N objects
without regard to order is the binomial coefficient: N !/n!(N − n)!. If we have
a set of boxes numbered 1, 2, . . . , k, the number of ways of selecting N1 objects
from N objects and placing them in box 1, then selecting N2 objects from the
remaining N − N1 objects and placing them in box 2 etc. is
     
(N −N1 )! Nk !
W = N !(NN−N !
)! N !(N −N −N )!
. . . N k !0!
1 1 2 1 2
DR
Qk
= N! i=1 (1/Ni !)

where the extended product is over all boxes containing one or more objects.
If the i -th box has a "degeneracy" of gi , that is, it has gi sub-boxes, such that
any way of filling the i -th box where the number in the sub-boxes is changed is
a distinct way of filling the box, then the number of ways of filling the i -th box
must be increased by the number of ways of distributing the Ni objects in the
gi boxes. The number of ways of placing Ni distinguishable objects in gi boxes
is giNi . Thus the number of ways (W ) that N atoms can be arranged in energy
levels each level i having gi distinct states such that the i -th level has Ni atoms
is:
Q g Ni
W = N ! Ni i !

Maxwell-Boltzmann statistics
288

For example, suppose we have three particles, a, b, and c, and we have three
energy levels with degeneracies 1, 2, and 1 respectively. There are 6 ways to
arrange the three particles
. . . . . .

FT
c. .c b. .b a. .a
ab ab ac ac bc bc

The six ways are calculated from the formula:


Q g Ni  2 1 0
W = N ! Ni i ! = 3! 12! 2
1!
1
0! = 6

We wish to find the set of Ni for which W is maximised, subject to the con-
straint that there be a fixed number of particles, and a fixed energy. The maxi-
ma of W and ln(W ) are achieved by the same values of Ni and, since it is easier
to accomplish mathematically, we will maximise the latter function instead. We
constrain our solution using Lagrange multipliers forming the function:
P P
f (Ni ) = ln(W ) + α(N − Ni ) + β(E − Ni i )

Using Stirling’s approximation for the factorials and taking the derivative with
A
respect to Ni , and setting the result to zero and solving for Ni yields the
Maxwell-Boltzmann population numbers:
gi
Ni = eα+βi

It can be shown thermodynamically that β = 1/kT where k is Boltzmann’s


constant and T is the temperature, and that α = -µ/kT where µ is the chemical
potential, so that finally:
DR
gi
Ni =
e(i −µ)/kT

Note that the above formula is sometimes written:


gi
Ni =
ei /kT /z

where z = exp(µ/kT ) is the absolute activity.


Alternatively, we may use the fact that
P
i Ni = N

to obtain the population numbers as


−i /kT
Ni = N gi e Z

where Z is the partition function defined by:

Maxwell-Boltzmann statistics
289

Z=
P −i /kT
i gi e

Another derivation

FT
In the above discussion, the Boltzmann distribution function was obtained via
directly analysing the multiplicities of a system. Alternatively, one can make
use of the canonical ensemble. In a canonical ensemble, a system is in thermal
contact with a reservoir. While energy is free to flow between the system and
the reservoir, the reservoir is thought to have infinitely large heat capacity as
to maintain constant temperature, T, for the combined system.
In the present context, our system is assumed to be have energy levels i with
degeneracies gi . As before, we would like to calculate the probability that our
system has energy i .
If our system is in state s1 , then there would be a corresponding number of
microstates available to the reservoir. Call this number ΩR (s1 ). By assumption,
the combined system (of the system we are interested in and the reservoir) is
isolated, so all microstates are equally probable. Therefore, for instance, if
ΩR (s1 ) = 2 ΩR (s2 ), we can conclude that our system is twice as likely to be in
A
state s2 than s1 . In general, if P (si ) is the probability that our system is in
state si ,
P (s1 ) ΩR (s1 )
P (s2 )
= ΩR (s2 )
.

Since the entropy of the reservoir SR = k ln ΩR , the above becomes


P (s1 ) eSR (s1 ) /k
= = e(SR (s1 )−SR (s2 ))/k .
P (s2 ) eSR (s1 ) /k
DR
Next we recall the thermodynamic identity:
1
dSR = T (dUR + P dVR − µdNR ).

In a canonical ensemble, there is no exchange of particles, so the dNR term is


zero. Similarly, dVR = 0. This gives
1
(SR (s1 ) − SR (s2 ) = T (UR (s1 ) − UR (s2 )) = − T1 (E(s1 ) − E(s2 ))

, where UR (si ) and E(si ) denote the energies of the reservoir and the system
at si , respectively. For the second equality we have used the conservation of
energy. Substituting into the first equation relating P (s1 ), P (s2 ):
P (s1 ) e−E(s1 )/kT
=
P (s2 ) e−E(s2 )/kT

, which implies, for any state s of the system

Maxwell-Boltzmann statistics
290

1 −E(s)/kT
P (s) = Ze

, where Z is an appropriately chosen "constant" to make total probability 1. (Z


is constant provided that the temperature T is invariant.) It is obvious that

FT
Z = s e−E(s)/kT
P

where the index s run through all microstates of the system. (Z is sometimes
called the Boltzmann sum over states.) If we index the summation via the
energy eigenvalues instead of all possible states, degeneracy must be taken into
account. The probability of our system having energy i is simply the sum of
the probabilities of all corresponding microstates:
P (i ) = 1 −i /kT
Z gi e

where, with obvious modification, Z =


P −j /kT . This is the same result as
j gj e
before.

Comments


ANotice that in this formulation, the initial assumption "... suppose the sys-
tem has total N particles..." is dispensed with. Indeed, the number of par-
ticles possessed by the system plays no role in arriving at the distribution.
Rather, how many particles would occupy states with energy i follows as
an easy consequence.

• What has been presented above is essentially a derivation of the canoni-


cal partition function. As one can tell by comparing the definitions, the
DR
Boltzman sum over states is really no different from the canonical partition
function.

• Exactly the same approach can be used to derive Fermi-Dirac and Bose-
Einstein statistics. However, there one would replace the canonical ensem-
ble with the grand canonical ensemble, since there is exchange of particles
between the system and the reservoir. Also, the system one considers in
those cases is a single particle state, not a particle. (In the above discus-
sion, we could have assumed our system to be a single atom.)

Maxwell-Boltzmann statistics
291

Limits of applicability
The Bose-Einstein and Fermi-Dirac distributions may be written:
gi
Ni =
e(i −µ)/kT ±1

FT
Assuming the minimum value of i is small, it can be seen that the conditions
under which the Maxwell-Boltzmann distribution is valid is when
e−µ/kT  1

For an ideal gas, we can calculate the chemical potential using the development
in the Sackur-Tetrode article to show that:
   
∂E
µ = ∂N = −kT ln NVΛ3
S,V

where E is the total internal energy, S is the entropy, V is the volume, and Λ
is the thermal de Broglie wavelength. The condition for the applicability of the
Maxwell Boltzmann distribution for an ideal gas is again shown to be
V
 1.
A N Λ3

See also
• →Bose-Einstein statistics
• →Fermi-Dirac statistics
DR
Source: http://en.wikipedia.org/wiki/Maxwell-Boltzmann_statistics

Principal Authors: Mct mht, PAR, Jheald, LifeStar, Sam Korn

Measurement in quantum mechanics

The framework of quantum mechanics requires a careful definition of mea-


surement, and a thorough discussion of its practical and philosophical impli-
cations.

Measurement in quantum mechanics


292

Formalism of measurement
Measurable quantities ("observables") as operators
An observable quantity is represented mathematically by an Hermitian or self

FT
adjoint operator. The set of the operator’s eigenvalues represent the set of
possible outcomes of the measurement. For each eigenvalue there is a corre-
sponding eigenstate (or "eigenvector"), which will be the state of the system
after the measurement. Some properties of this representation are

• The eigenvalues of Hermitian matrices are real. The possible outcomes of a


measurement are precisely the eigenvalues of the given obervable.
• A Hermitian matrix can be unitarily diagonalized (See Spectral theorem),
generating an orthonormal basis of eigenvectors which spans the state
space of the system. In general, the state of a system can be represented
as a linear combination of eigenvectors of any Hermitian operator. Physi-
cally, this is to say that any state can be expressed as a superposition of the
eigenstates of an observable.
A
Important examples are:

• The Hamiltonian operator, representing the total energy of the system; with
2

the special case of the nonrelativistic Hamiltonian operator: Ĥ = 2m +
V (x̂).

• The momentum operator: p̂ = ~i ∂x (in the position basis).
• The position operator: x̂, where x̂ = −~ ∂
i ∂p (in the momentum basis).
DR
Operators can be noncommuting. In the finite dimensional case, two Hermi-
tian operators commute if they have the same set of {normalized} eigenvec-
tors. Noncommuting observables are said to be incompatible and can not be
measured simultaneously. This can be seen via the uncertainty principle.

Eigenstates and projection


Assume the system is prepared in state |ψi. Let Ô be a measurement operator,
an observable, with eigenstates |ni for n = 1, 2, 3, ... and corresponding eigen-
values O1 , O2 , O3 , .... If the measurement outcome is ON , the system will then
"collapse" to the state |N i after measurement.
The case of a continuous spectrum is more involved, since, physically speak-
ing, the basis has uncountably many eigenstates, but the general concept is
the same. In the position representation, for instance, the eigenstates can be
represented by the set of delta functions, indexed by all possible positions of

Measurement in quantum mechanics


293

the particle. In the experimental setting, the resolution of any given measure-
ment is finite, and therefore the continuous space may be divided into discrete
segments. Another solution is to approximate any lab experiments by a "box"
potential (which bounds the volume in which the particle can be found, and

FT
thus ensures a discrete spectrum).

Wavefunction collapse

Given any quantum state which is a superposition of eigenstates at


time t

|ψi = c1 e−iE1 t |1i + c2 e−iE2 t |2i + c3 e−iE3 t |3i + · · · ,

if we measure, for example, the energy of the system and receive E 2

(this result is chosen randomly according to probability given by


A |c |
Pr(En ) = P n
2

|c |2
k k
),

then the system’s quantum state after the measurement is

|ψi = e−iE2 t |2i

so any repeated measurement of energy will yield E 2.


DR
Figure 1. The process of wavefunction collapse illustrated.
The process in which a quantum state becomes one of the eigenstates of the op-
erator corresponding to the measured observable is called "collapse", or "wave-
function collapse". The final eigenstate appears randomly with a probability
equal to the square of its overlap with the original state. The process of col-
lapse has been studied in many experiments, most famously in the double-slit
experiment. The wavefunction collapse raises serious questions of determinism
and locality, as demonstrated in the EPR paradox and later in GHZ entangle-
ment.
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

Measurement in quantum mechanics


294

and provides an explanation for the absence of quantum coherence after mea-
surement. While this theory correctly predicts the form and probability distri-
bution of the final eigenstates, it does not explain the randomness inherent in
the choice of final state.

FT
There are two major approaches toward the "wavefunction collapse":

• Accept it as it is. This approach was supported by Niels Bohr and his
Copenhagen interpretation which accepts the collapse as one of the elemen-
tary properties of nature (at least, for small enough systems). According to
this, there is an inherent randomness embedded in nature, and physical
observables exist only after they are measured (for example: as long as a
particle’s speed isn’t measured it doesn’t have any defined speed).
• Reject it as a physical process and relate to it only as an illusion. This
approach says that there is no collapse at all, and we only think there is.
Those who support this approach usually offer another interpretation of
quantum mechanics, which avoids the wavefunction collapse.

von Neumann measurement scheme


A
The von Neumann measurement scheme, an ancestor of quantum decoherence
theory, describes measurements by taking into account the measuring appara-
tus which is also treated as a quantum object. Let the quantum state be in
P
the superposition |ψi = n cn |ψn i, where |ψn i are eigenstates of the operator
that needs to be measured. In order to make the measurement, the measured
system described by |ψi needs to interact with the measuring apparatus de-
scribed by the quantum state |φi, so that the total wave function before the
interaction is |ψi|φi. After the interaction, the total wave function exhibits
DR
P
the unitary evolution |ψi|φi → n cn |ψn i|φn i, where |φn i are orthonormal
states of the measuring apparatus. The unitary evolution above is referred to
as premeasurement. One can also introduce the interaction with the environ-
ment |ei, so that, after the interaction, the total wave function takes a form
P
n cn |ψn i|φn i|en i, which is related to the phenomenon of decoherence. The
above is completely described by the →Schrödinger equation and there are not
any interpretational problems with this. Now the problematic wavefunction
collapse does not need to be understood as a process |ψi → |ψn i on the level
of the measured system, but can also be understood as a process |φi → |φn i
on the level of the measuring apparatus, or as a process |ei → |en i on the level
of the environment. Studying these processes provides considerable insight in-
to the measurement problem by avoiding the arbitrary boundary between the
quantum and classical worlds, though it does not explain the presence of ran-
domness in the choice of final eigenstate. If the set of states {|ψn i}, {|φn i}, or
{|en i} represents a set of states that do not overlap in space, the appearance

Measurement in quantum mechanics


295

of collapse can be generated by either the Bohm interpretation or the Everett


interpretation which both deny the reality of wavefunction collapse; they both,
though, predict the same probabilities for collapses to various states as does
the conventional interpretation. The Bohm interpretation is held to be correct

FT
only by a small minority of physicists, since there are difficulties with the gen-
eralization 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.

Example
Suppose that we have a particle in a box. If the energy of the particle is
2 2 2
measured to be EN = N2mLπ ~
2 then the corresponding state of the system is
q  
2 N πx
R
|ψN i = |xihx|ψN idx where hx|ψN i = hx|N i = L sin L , which is
determined by solving the Time-Independent →Schrödinger equation for the
given potential.
Alternatively, if instead of knowing the energy of the particle the particle’s
A
position is determined to be a distance S from the left wall of the box, the
R
corresponding system state is |ψS i = |xihx|ψS idx where hx|ψS i = hx|Si =
δ(S − x).
These two state functions |ψN i and |ψS i are distinct functions (of the posi-
tion x after we left multiply by the bra state hx|), but they are in general not
orthogonal to each other:
RL q  
2 N πx
R
hψS |ψN i = hS|N i = hS|xihx|N idx = 0 δ(S − x) L sin L dx =
q  
DR
2 N πS
L sin L .
The two systems are therefore distinct; a position measurement is instanta-
neous whereas a definite value of energy EN is established only in the limit of
an infinitely long observation period.
Completeness of eigenvectors of Hermitian operators guarantees that either
system state, being the eigenvector to one measurement operator, can be ex-
pressed as a linear combination of eigenvectors of the other measurement op-
erator:
 
2 P nπx
sin nπS
P RP R 
|Si = n |ni hn|Si = n |xihx|ni hn|Si dx = |xi L n sin L L dx =
R R
|xiδ(S − x)dx, i.e. |Si = |xiδ(S − x)dx
and
q  
|si L2 sin NLπs ds.
R R
|N i = |si hs|N i ds =

Measurement in quantum mechanics


296

The time dependence of the system states is determined by the Time Dependent
→Schrödinger equation. In the preceding example, with energy eigenvalues
En , it follows that the time dependent solution is
|ψ(t)i = n |nihn|ψS i e−itEn /~ ,
P

FT
where t represents the time since the particle’s location in space was measured.
Consequently
q  
hn|ψ(t)i = hn|ψS i e−itEn /~ = L2 sin nπSL e−itEn /~ 6= 0
at least for several distinct energy eigenstates |ni, for all values t, and for all
0 < S < L.
The particle state |ψS i therefore can not have evolved (in the above technical
sense) into state |ψN i (which is orthogonal to all energy eigenstates, except
itself), for any duration t. While this conclusion may be characterized accord-
ingly instead as "the wave function of the particle having been projected, or
having collapsed into" the energy eigenstate |ψN i, it is perhaps worth empha-
sizing that any definite value of energy EN can be established only in the limit
of a long-lasting trial and never for any finite value of time.
A
Optimal quantum measurement
What is the optimal quantum measurement to distinguish mixed states from
a given ensemble? This is a natural question of which the solution is well
understood, and is given by a semidefinite programming.
More specifically, suppose a mixed state ρi is drawn from the ensemble
with probability pi , we wish to find a POVM measurement {Πi } so that
P
P i ρi ) is maximized. This is clearly
i pi tr(Π P a semidefinit programming:
DR
max i pi tr(Πi ρi ) s.t. Πi ≥ 0, i Πi = I.
Interestingly, the dual problem has a nice description:
min tr(X) s.t. X − pi ρi ≥ 0.
Let Π̂i and X̂ be the solutions of the primal and the dual, we have
Π̂i · (X̂ − pi ρi ) = 0.
From this one can conclude that if all ρi are pure states, then Π̂i must also be
of rank 1. Furthermore, if ρi ’s are in addition independent, then the optimal
measure is a von Neumann measurement.

Philosophical problems of quantum measurements


What physical interaction constitutes a measurement?
Until the advent of quantum decoherence theory in the late 20th century, a ma-
jor conceptual problem of quantum mechanics and especially the Copenhagen

Measurement in quantum mechanics


297

interpretation was the lack of a distinctive criterion for a given physical interac-
tion to qualify as "a measurement" and cause a wavefunction to collapse. This
best illustrated by the →Schrödinger’s cat paradox.
Major philosophical and metaphysical questions surround this issue:

FT
• The concept of weak measurements.
• Macroscopic systems (such as chairs or cats) do not exhibit counterintuitive
quantum properties, which can only be observed in microscopic particles
such as electrons or photons. This invites the question of when a system is
"big enough" to behave classically and not quantum mechanically?

Quantum decoherence theory has successfully addressed other questions that


previously haunted quantum measurement theory:

• Does a measurement depend on the existence of a self-aware observer?

Answer: No. Coupling an isolated quantum system to another quantum sys-


tem with many degrees of freedom generically transfers the coherence of the
first system into mutual coherence of the two systems. The initially isolated
A
quantum system then appears to "collapse." Interpreting the second system as
a measurement apparatus, as in the von Neumann scheme, shows that no con-
sciousness or self-awareness is necessary for collapse of the first system.

• What interactions are strong enough to constitute a measurement?

This question is quantitatively answered by decoherence theory, given a model


for the measurement apparatus. The scaling of the measurement effects with
the system/apparatus interaction strength usually only weakly depends on the
DR
choice of a model for the apparatus, so one can give a generic description of
the strength of a measurement induced by a given interaction.

Does measurement actually determine the state?


The question of whether a measurement actually determines the state, is
deeply related to the Wavefunction collapse.
Most versions of the Copenhagen interpretation answer this question with an
unqualified "yes".
See also:

• Philosophies: Copenhagen interpretation.


• People (actualist philosophers): Henri Poincaré, Niels Bohr.

Measurement in quantum mechanics


298

The quantum entanglement problem


See EPR paradox.

See also

FT
• Measurement related problems and paradoxes
• Afshar experiment
• Measurement problem
• Wavefunction collapse
• EPR paradox
• Renninger negative-result experiment
• Elitzur-Vaidman bomb-testing problem
• →Schrödinger’s cat
• Interpretations of quantum mechanics
• Transactional interpretation
• Copenhagen interpretation
• Many-worlds interpretation
• Quantum mechanics formalism
A • →Quantum mechanics
• →Mathematical formulation of quantum mechanics
• →Schrödinger equation
• →Bra-ket notation
• Generalized measurement (POVM, Positive operator valued measure)

External links
DR
• Analog: A Farewell to Copenhagen? 142
• " The Double Slit Experiment 143". (physicsweb.org)
• Shahriar S. Afshar, " Waving Copenhagen Good-bye: Were the founders of
Quantum Mechanics wrong? 144" This link presents a view that is not en-
dorsed by the majority of physicists.
• " Variation on the similar two-pin-hole "which-way" experiment 145". (report-
ed in New Scientist; July 24), Reprint at irims.org 146

142 http://www.analogsf.com/0410/altview2.shtml
143 http://physicsweb.org/article/world/15/9/1
144 http://my.harvard.edu/cgi-bin/webevent/webevent.cgi?cmd=showevent&ncmd=calmonth&cal=9719

&y=2004&m=3&d=23&id=10416384&token=G6409379:1&sb=0&cf=cal&lc=calmonth&swe=1&set=0
&sa=0&sort=e,m,t&ws=0&sib=0&de=0&tf=0
145 http://www.sciencefriday.com/images/shows/2004/073004/AfsharExperimentSmall.jpg
146 http://www.irims.org/quant-ph/030503/

Measurement in quantum mechanics


299

• " Measurement in Quantum Mechanics 147" Henry Krips in the Stanford En-
cyclopedia of Philosophy
• Decoherence, the measurement problem, and interpretations of quantum
mechanics 148

FT
• Measurements and Decoherence 149
• Yonina C. Eldar, Alexandre Megretski, and George C. Verghese. Designing
optimal quantum detectors via semidefinite programming. IEEE Transac-
tions on Information Theorey, Vol. 49, No. 4, 1007–1012, 2003.

Source: http://en.wikipedia.org/wiki/Measurement_in_quantum_mechanics

Principal Authors: MathKnight, Fwappler, MichaelCPrice, William M. Connolley, Dave Kielpinski

Molecular Hamiltonian

The non relativistic molecular Hamiltonian for a multi-electron molecule in


Aatomic units is:
H = TN + Hel
where
1
∇2a
P
TN = a − 2M a

is the kinetic energy operator corresponding to the molecular dynamics and


H el is the electronic molecular Hamiltonian. M a are the masses of the nuclei.
∇2a is the Laplacian with respect to the cartesian nuclear coordinates associated
to the molecular geometry.
DR
The relativistic molecular hamiltonian differs because it contains terms that
depend upon the electron spin and the nuclear spin. The electron spin appears
naturally in the solution of the Dirac equation, but the nuclear spins are added
in phenomenologically as if the nuclei were "heavy" electrons.
The electronic molecular Hamiltonian can be replaced within the Born-
Oppenheimer approximation by the potential energy surfaces. The
→Schrödinger equation becomes then an equation describing the "motion" of
the nuclei only. The "motion" of the electrons has been taken into account
during the diagonalization of the electronic molecular Hamiltonian (see com-
putational chemistry for more details).

147 http://plato.stanford.edu/entries/qt-measurement/
148 http://arxiv.org/abs/quant-ph/0312059
149 http://arxiv.org/abs/quant-ph/0505070

Molecular Hamiltonian
300

The discrete eigenvalues of the molecular Hamiltonian are called molecular


energy levels.

References

FT
• Richard Moss, Advanced Molecular Quantum Mechanics, ISBN 412-10490-
3.

Source: http://en.wikipedia.org/wiki/Molecular_Hamiltonian

Principal Authors: Vb, Iain.mcnab, Karol Langner

Multiplicative quantum number

In quantum field theory, multiplicative quantum numbers are conserved


quantum numbers of a special kind. A given quantum number q is said to
A
be additive if in a particle reaction the sum of the q-values of the interacting
particles is the same before and after the reaction. Most conserved quantum
numbers are additive in this sense; the electric charge is one example. A mul-
tiplicative quantum number q is one for which the corresponding product,
rather than the sum, is preserved.
Any conserved quantum number is a symmetry of the Hamiltonian of the sys-
tem (see Noether’s theorem). Symmetry groups which are examples of the
abstract group called Z 2 give rise to multiplicative quantum numbers. This
DR
group consists of an operation, P, whose square is the identity, P 2 = 1. Thus,
all symmetries which are mathematically similar to parity (physics) give rise to
multiplicative quantum numbers.
In principle, mutliplicative quantum numbers can be defined for any Abelian
group. An example would be to trade the electric charge, Q, (related to
the Abelian group U(1) of electromagnetism), for the new quantum number
exp(2i π Q ). Then this becomes a multiplicative quantum number by virtue of
the charge being an additive quantum number. However, this route is usual-
ly followed only for discrete subgroups of U(1), of which Z 2 finds the widest
possible use.

Multiplicative quantum number


301

See also
• Parity, C-symmetry, →T-symmetry and G-parity

FT
References
• Group theory and its applications to physical problems, by M. Hamermesh
(Dover publications, 1990) ISBN 0486661814

Source: http://en.wikipedia.org/wiki/Multiplicative_quantum_number

Principal Authors: Bambaiah, Michael Hardy, Phys, Shimgray, Agentsoo

Neutral particle oscillations

In particle physics, neutral particle oscillation is the transmutation of a


A
neutral particle with nonzero internal quantum numbers into its antiparticle.
These oscillations and the associated mixing of particles gives insight into
the realization of discrete parts of the Poincare group, ie, parity (P), charge
conjugation (C) and time reversal invariance (T).

The phenomenon
DR

Neutral particles such as the kaon, neutron, bottom quark mesons or neutrinos
have internal quantum numbers called flavour. This means that the particle
and antiparticle are different. If both particle and antiparticle can decay into
the same final state, then it is possible for the decay and its time reversed
process to contribute to oscillations—
A → F → B → F → A → ...

Neutral particle oscillations


302

where A is the particle, B is the antiparticle, and F is the common set of particles
into which both can decay. The example of the neutral kaon is pictured here.
Such a process is actually connected to the mass renormalization of the states
A and B in quantum field theory. However, under certain circumstances it can

FT
be tackled through a simpler quantum mechanics model which neglects these
intermediate multi-particle quantum states and concentrates only on the states
A and B.

Quantum mechanical model


Consider a state |ψ(t)> = a(t)|A> + b(t)|B>. Its time evolution is gov-
erned by the Hamiltonian, H, through the action of the evolution operator
U(t) = exp(iHt) on |ψ(0)>. The 2×2 matrix Hamiltonian can be written as
 
HAA HAB
H= = M − 2i Γ,
HBA HBB

where the Hamiltonian can be decomposed into a mass matrix M and a decay
width matrix Γ, both of which are 2×2 Hermitean matrices. We introduce the
A
notation M AB = |M AB| e iα and Γ AB = |Γ AB| e i(α+β ).
A and B are both flavour eigenstates. Oscillations mix these states, and the
mass eigenstates are the states which propagate without mixing, ie, the eigen-
vectors of H.

CPT symmetry
The action of the discrete spacetime symmetries are
C|A> = -|B>, P|A> = -|A> and T|A> = +|A>.
DR
If the Hamiltonian is CPT symmetric, then (CPT)H(CPT) -1=H. The transforma-
tion properties above imply that CPT|A>=|B>. Then <A|H|A> = <B|H|B>,
so a test of CPT symmetry is that the masses and the decay widths of the parti-
cle and the antiparticle are equal. This is a major class of experimental tests of
CPT symmetry.
Any 2×2 matrix can be written in the form E 0I+Eu.σ, where I is the identity
matrix, σ i are the Pauli matrices and u is an unit vector. With CPT symmetry,
the diagonal
 elements of H are equal, so
sin φ r


MAB −iΓ∗
AB /2 iα |MAB |+ 12 |ΓAB |e−i(β+π/2)
u = cos φ ,
  e = MAB −iΓAB /2 = e |MAB |+ 21 |ΓAB |ei(β+π/2)
,
0

Neutral particle oscillations


303

where φ is a complex angle. H is diagonalized by rotating u into an unit vector


in the z-direction. The eigenvectors and eigenvalues are
|1, 2i = √1 (|Ai ± eiφ |Bi),
2

FT

E1,2 = MAA − 12 ΓAA ± eiα |MAB | + 21 |ΓAB |ei(β+π/2) ,

where the plus signs are for the state |1> and minus, for |2>. A change in the
phase convention, |B> → e -iθ|B> changes the definition of the eigenstates,
but not the eigenvalues. By appropriate choice of this phase, the angle φ can
always be set equal to zero, so that the eigenstates are orthogonal.

Oscillations, regeneration and CP violation

CPT symmetry breaking

See also
• Kaons, B-Bbar oscillations and neutrino oscillations

A CP violation and CPT symmetry

Source: http://en.wikipedia.org/wiki/Neutral_particle_oscillations

Principal Authors: Bambaiah, Phys, Pearle, Agentsoo, DavidWBrooks

Normalisable wavefunction
DR
In quantum mechanics, wave functions which describe real particles must be
normalisable 1: the probability of the particle to occupy any place must equal
1. Mathematically, in one dimension this is expressed as
RB ∗
A ψ (x)ψ(x) dx = 1

in which the integration parameters A and B indicate the interval in which the
particle must exist.
All wavefunctions which represent real particles must be normalizable, that
is, they must have a total probability of one - they must describe the proba-
bility of the particle existing as 100%. This trait enables anyone who solves
the →Schrödinger equation for certain boundary conditions to discard solu-
tions which do not have a finite integral at a given interval. For example, this

Normalisable wavefunction
304

disqualifies periodic functions as wave function solutions for infinite intervals,


while those functions can be solutions for finite intervals.

Derivation of normalisation

FT
In general, ψ is a complex function. However,
ψ ∗ ψ =| ψ |2

is real, greater than zero, and is known as a probability density function.


This means that
R∞
p(−∞ ≤ x ≤ ∞) = −∞ | ψ |2 dx. (1)

where p(x) is the probability of finding the particle at x. Equation (1) is given
by the definition of a probability density function. Since the particle exists,
its probability of being anywhere in space must be equal to 1. Therefore we
integrate over all space:
R∞
p(−∞ ≤ x ≤ ∞) = −∞ | ψ |2 dx = 1. (2)
A
If the integral is finite, we can multiply the wavefunction, ψ, by a constant
such that the integral is equal to 1. Alternatively, if the wavefunction already
contains an appropriate arbitrary constant, we can solve equation (2) to find
the value of this constant which normalises the wavefunction.

Example of normalisation
A particle is restricted to a 1D region between x=0 and x=l; its wavefunction
DR
is: (
Aei(kx−ωt) , 0 ≤ x ≤ l
ψ(x, t) =
0, elsewhere.

To normalise the wavefunction we need to find the value of the arbitrary con-
stant A, i.e., solve
R∞ 2
−∞ | ψ | dx = 1

to find A.
Substituting ψ into | ψ |2 we get
| ψ |2 = A2 ei(kx−ωt) e−i(kx−ωt) = A2

so

Normalisable wavefunction
305
R0 Rl R∞
−∞ 0dx + 0 A2 dx + l 0dx = 1

therefore
 
1
A2 l = 1 ⇒ A = √ .

FT
l

The normalised
( wavefunction
 is:
1
√ ei(kx−ωt) , 0≤x≤l
ψ(x, t) = l
0, elsewhere.

Proof that wavefunction normalisation doesn’t change


associated properties
If normalisation of a wavefunction changed the properties associated with the
wavefunction, the process becomes pointless as we still cannot yield any infor-
mation about the properties of the particle associated with the un-normailied
wavefunction. It is therefore important to establish that the properties associ-
A
ated with the wavefunction are not altered by normalisation.
All properties of the particle such as probability distribution, momentum, en-
ergy, expectation value of position etc. are derived from the Schrödinger wave
equation. The properties are therefore unchanged if the Schrödinger wave
equation is invariant under normalisation.
The Schrödinger wave equation is
−~ d2 ψ
2m dx2 + V (x)ψ(x) = Eψ(x).
DR
If ψ is normalised and replaced with Aψ, then
d2 (Aψ)
  2
d(Aψ) dψ d(Aψ)
dx = A dx and dx2 = dx
d
dx = A ddxψ2 .

The Schrödinger wave equation therefore becomes:


−~ d2 ψ
2m A dx2 + V (x)Aψ(x) = EAψ(x)
 
−~ d2 ψ
⇒A 2m dx2 + V (x)ψ(x) = A (Eψ(x))

−~ d2 ψ
⇒ 2m dx2 + V (x)ψ(x) = Eψ(x)

Normalisable wavefunction
306

which is the original Schrödinger wave equation. That is to say, the


Schrödinger wave equation is invariant under normalisation, and consequently
associated properties are unchanged.

Note

FT
Note 1: The spelling normalisable is a British variant spelling of normalizable.

See also
• →Quantum mechanics
• →Schrödinger equation
• →Wavefunction

Source: http://en.wikipedia.org/wiki/Normalisable_wavefunction

Principal Authors: Germen, Charles Matthews, Oleg Alexandrov, Stevertigo, Conscious


A
Normal mode

Normal modes in an oscillating system are special solutions where all the
parts of the system are oscillating with the same frequency (called normal
frequencies or allowed frequencies). The concept of normal modes is of
vital importance in wave theory, optics and quantum mechanics.
Finding normal modes in harmonic oscillation uses the strength of linear alge-
DR
bra and linear sets of differential equations. One can present the problem as a
matrix-vector equation and then solve for its eigenvectors. After finding them,
the normal modes are the eigenvectors where the normal frequencies are the
eigenvalues.

Example - normal modes of coupled oscillators


Consider two bodies (not affected by gravity), each of mass M, attached to
three springs with stiffness K. They are attached in the following manner:

Normal mode
307

A
Figure 29
FT
Various normal modes in a 1D-lattice.

where the edge points are fixed and cannot move. We’ll use x 1(t) to denote
the displacement of the leftmost mass, and x 2(t) to denote the displacement
of the rightmost.
DR
If we denote the second derivative of x(t) with respect to time as x”, the equa-
tions of motion are:
M x001 = −K(x1 ) − K(x1 − x2 )

M x002 = −K(x2 ) − K(x2 − x1 )

Since we expect oscillatory motion, we try:


x1 (t) = A1 eiωt

x2 (t) = A2 eiωt

Substituting these into the equations of motion gives us:

Normal mode
308

−ω 2 M A1 eiωt = −2KA1 eiωt + KA2 eiωt

−ω 2 M A2 eiωt = KA1 eiωt − 2KA2 eiωt

FT
Since the exponential factor is common to all terms, we omit it and simplify:
(ω 2 M − 2K)A1 + KA2 = 0

KA1 + (ω 2 M − 2K)A2 = 0

And
 in matrix representation:  
ω 2 M − 2K K A1
=0
K ω 2 M − 2K A2

For this equation to have a non-trivial solution, the determinant of the matrix
on the left (the characteristic polynomial of the system) must be equal to 0, so:
(ω 2 M − 2K)2 − K 2 = 0
A
Solving for ω, we have:
q
K
ω1 = M
q
3K
ω2 = M

If we substitute ω1 into the matrix and solve for (A1 , A2 ), we get (1, 1). If
DR
we substitute ω2 , we get (1, -1). (These vectors are eigenvectors, and the
frequencies are eigenvalues.)
The
 firstnormalmode
 is:
x1 (t) 1
= c1 cos (ω1 t + φ1 )
x2 (t) 1

and
 the
second 
normal
 mode is:
x1 (t) 1
= c2 cos (ω2 t + φ2 )
x2 (t) −1

The general solution is a superposition of the normal modes where c 1, c 2, φ 1,


and φ 2, are determined by the initial conditions of the problem.

Normal mode
309

The process demonstrated here can be generalized and formulated using the
formalism of Lagrangian mechanics or Hamiltonian mechanics.

Standing waves

FT
A standing wave is a continuous form of normal mode. In a standing wave,
all the space elements (i.e (x,y,z) coordinates) are oscillating in the same fre-
quency and in phase (reaching the equilibrium point together), but each has a
different amplitude.

A
The general form of a standing wave is:
Ψ(t) = f (x, y, z)(A cos(ωt) + B sin(ωt))

where f (x, y, z) represents the dependence of amplitude on location and the


cosine\sine are the oscillations in time.
DR
Physically, standing waves are formed by the interference (superposition) of
waves and their reflections (although one may also say the opposite; that a
moving wave is a superposition of standing waves). The geometric shape of the
medium determines what would be the interference pattern, thus determines
the f (x, y, z) form of the standing wave. This space-dependence is called a
normal mode.
Usually, for problems with continuous dependence on (x,y,z) there is no single
or finite number of normal mode, but there are infinitely normal modes. If
the problem is bounded (i.e it is defined on a finite section of space) there are
countably many (a discrete infinity of ) normal modes (usually numbered n
= 1,2,3,...). If the problem is not bounded, there is a continuous spectrum of
normal modes.
The allowed frequencies are dependent on the normal modes, as well on phys-
ical constants of the problem (density, tension, pressure, etc.) which sets the

Normal mode
310

phase velocity of the wave. The range of all possible normal frequencies is
called the frequency spectrum. Usually, each frequency is modulated by the
amplitude at which it has arisen, creating a graph of the power spectrum of the
oscillations.

FT
When relating to music, normal modes of a vibrating instruments (strings, air
pipes, drums, etc.) are called "harmonics".

Normal modes in quantum mechanics


In quantum mechanics, a state |ψi of a system is described by a wavefunction
ψ(x, t)which solves the →Schrödinger equation. The square of the absolute
value of ψ ,i.e.
P (x, t) = |ψ(x, t)|2

is the probability (density) to measure the particle in place x at time t.


Usually, when involving some sort of potential, the wavefunction is decom-
posed into a superposition of energy eigenstates, each oscillating with frequen-
cy of ω = En /~. Thus, we may write
A|ψ(t)i = n |ni hn|ψ(t = 0)i e−iEn t/~
P

The eigenstates have a physical meaning further than an orthonormal basis.


When the energy of the system is measured, the wavefunction collapses into
one of its eigenstates and so the particle wavefunction is described by the pure
eigenstate corresponding to the measured energy.

See also
DR
• Physical applications:
• Waves
• Optics
• harmonic oscillator
• vibrational spectroscopy
• quantum theory
• →Schrödinger equation
• →Wavefunction
• →Measurement in quantum mechanics
• harmonic series (music)
• Mathematical tools:
• linear algebra
• eigenvectors

Normal mode
311

• differential equations
• Fourier analysis
• Sturm-Liouville theory
• Boundary value problem

FT
External links
• Java simulation of coupled oscillators 150.
• Java simulation of the normal modes of a string 151, drum 152, and bar 153.

Source: http://en.wikipedia.org/wiki/Normal_mode

Principal Authors: MathKnight, Michael Hardy, Charles Matthews, Pfalstad, Omegatron

Nuclear physics
ANuclear physics is the branch of physics concerned with the nucleus of the
atom. It has three main aspects: probing the fundamental particles (protons
and neutrons) and their interactions, classifying and interpreting the properties
of nuclei, and providing technological advances.
Nuclei do not lend themselves to exact theoretical understanding, because they
are composed of many particles (mesons as well as protons and neutrons),
but are not large enough to be accurately described as periodic, as done with
crystals. So "nuclear models" that, singly or in combination, account for most
DR
nuclear behavior are used. Three of the four types of fundamental interaction
play important roles in nuclei, the strong, electromagnetic and, on a longer
time scale, weak.
Nuclei are held together by strong interactions (mostly exchanging pions), but
electromagnetic repulsion of the positively charged protons tends to push them
apart, according to Coulomb’s law. The stable nuclei all have close to the low-
est energy ratio of protons to neutron for their atomic weight. Nuclei near
enough to this ratio to be bound but not close enough to be stable, give off
electrons or positrons (beta decay) or take in electrons (and also give off neu-
trinos), to move closer to that ratio. This is the main place where the weak

150 http://www.falstad.com/coupled/
151 http://www.falstad.com/loadedstring/
152 http://www.falstad.com/circosc/
153 http://www.falstad.com/barwaves/

Nuclear physics
312

interactions come in. Nuclei that are too massive to be stable are pulled apart
by the coulomb repulsion of their protons and either fission or give off alpha
particles.
Though the number of energy levels is not infinite, as it is for the electron wave

FT
functions of atoms, most stable or nearly stable nuclei have many bound levels.
These usually decay toward the ground state by emitting gamma ray photons.
Protons and neutrons are fermions, with different value of the isospin quantum
number, so two protons and two neutrons can share the same space wave
function. In the rare case of a hypernucleus, a third baryon called a hyperon,
with a different value of the strangeness quantum number can also share the
wave function.
The binding energies of the protons and neutrons are on the order of 1 %
of their relativistic rest masses, so non-relativistic quantum mechanics can be
used with errors usually smaller than those from other approximations.
Often, nuclear physicists will use Nuclear Units where h, c, and the mass of the
proton m p have been set to unity.
A
History
Once the chemists of the 18th century had elucidated the chemical elements,
the rules governing their combinations in matter, and their systematic classifi-
cation (Mendeleev’s periodic table of elements) and John Dalton had, in 1803,
applied Democritus’s idea of atom to them, it was natural that the next step
would be a study of the fundamental properties of individual atoms of the vari-
ous elements, an activity that we would today classify as atomic physics. These
studies led to the discovery in 1896 by Becquerel of the radioactivity of certain
DR
species of atoms and to the further identification of radioactive substances by
the Curies in 1898. Ernest Rutherford next took up the study of radiation and
its properties; once he had achieved an understanding of the nature of the ra-
dioactivity, he turned around and used radiated particles to probe the atoms
themselves. In the process he proposed in 1911 the existence of the atom-
ic nucleus, the confirmation of which (through the painstaking experiments
of Geiger and Marsden) provided a new branch of science, nuclear physics.
Investigations into the properties of the nucleus have continued from Ruther-
ford’s time to the present. In the 1940s and 1950s, it was discovered that there
was yet another level of structure even more fundamental than the nucleus,
which is itself composed of protons and neutrons. Thus nuclear physics can be
regarded as the descendant of chemistry and atomic physics and in turn the
progenitor of particle physics.

Nuclear physics
313

Experiments with nuclei continue to contribute to the understanding of basic


interactions. Investigation of nuclear properties and the laws governing the
structure of nuclei is an active and productive area of research, and practical
applications, such as nuclear power, smoke detectors, cardiac pacemakers, and

FT
medical imaging devices, have become common.

See also
• Important publications in nuclear physics
• Nuclear fission
• Nuclear fusion
• Nuclear reactions
• Nuclear Structure
• Radioactivity
• Radioactive decay
• Nuclear force

• Models of the nucleus


A • Interacting boson model
• Liquid drop model
• Shell model

Applications

• Mossbauer effect
• Nuclear technology
• Nuclear engineering
DR
• Nuclear magnetic resonance
• Nuclear medicine
• Nuclear power
• Nuclear weapons

External links
• SCK.CEN Belgian Nuclear Research Centre 154 Mol, Belgium
• A Free E-Learning Course on Nuclear Physics 155
• Nuclear Physics-Nuclear Medicine Information 156

154 http://www.sckcen.be/
155 http://www.VirtualClassroom.de.vu/
156 http://www.nucmedinfo.com/Pages/physic.html

Nuclear physics
314

References
• Kenneth S. Krane, "Introductory Nuclear Physics", Wiley & Sons (1988).
• Tien T. Tong, "Fifty Years of Seeing Atoms", Physics Today, March 2006, p.

FT
31

Source: http://en.wikipedia.org/wiki/Nuclear_physics

Principal Authors: David R. Ingham, Philipum, Heron, Securiger, Icairns, DV8 2XL, Yhr, Zoicon5,
Android 93, Stevegiacomelli

Observable

In physics, particularly in quantum physics, a system observable is a proper-


ty of the system state that can be determined by some sequence of physical
operations. These operations might involve submitting the system to various
A
electromagnetic fields and eventually reading a value off some gauge. In sys-
tems governed by classical mechanics, any experimentally observable value can
be shown to be given by a real-valued function on the set of all possible system
states. In quantum physics, on the other hand, the relation between system
state and the value of an observable is more subtle, requiring some basic linear
algebra to explain. In the mathematical formulation of quantum mechanics,
states are given by non-zero vectors in a →Hilbert space V (where two vectors
are considered to specify the same state if, and only if, they are scalar multi-
DR
ples of each other) and observables are given by self-adjoint operators on V.
However, as indicated below, not every self-adjoint operator corresponds to a
physically meaningful observable. For the case of a system of particles, the
space V consists of functions called wave functions.
In quantum mechanics, measurement of observables exhibits some seemingly
mysterious phenomena. This often leads to many misconceptions about the
nature of quantum mechanics itself. The facts of the matter, however, are far
more prosaic. Specifically, if a system is in a state described by a wave function,
the measurement process affects the state in a non-deterministic, but statisti-
cally predictable way. In particular, after a measurement is applied, the state
description by a single wave function may be destroyed, being replaced by a
statistical ensemble of wave functions. The irreversible nature of measurement
operations in quantum physics is sometimes referred to as the measurement
problem and is described mathematically by quantum operations. By the struc-
ture of quantum operations, this description is mathematically equivalent to

Observable
315

that offered by relative state interpretation where the original system is re-
garded as a subsystem of a larger system and the state of the original system is
given by the partial trace of the state of the larger system.
Physically meaningful observables must also satisfy transformation laws which

FT
relate observations performed by different observers in different frames of ref-
erence. These transformation laws are automorphisms of the state space, that
is bijective transformations which preserve some mathematical property. In
the case of quantum mechanics, the requisite automorphisms are unitary (or
antiunitary) linear transformations of the Hilbert space V. Under Galilean rela-
tivity or special relativity, the mathematics of frames of reference is particularly
simple, and in fact restricts considerably the set of physically meaningful ob-
servables.

References
• S. Auyang, How is Quantum Field Theory Possible, Oxford University Press,
1995.
• G. Mackey, Mathematical Foundations of Quantum Mechanics, W. A. Ben-


A jamin, 1963.
V. Varadarajan, The Geometry of Quantum Mechanics vols 1 and 2,
Springer-Verlag 1985.

Source: http://en.wikipedia.org/wiki/Observable

Principal Authors: CSTAR, Michael Hardy, Phys, ABCD, Mushin


DR
Oil-drop experiment

The purpose of Robert Millikan’s oil-drop experiment (1909) was to measure


the electric charge of the electron. He did this by carefully balancing the grav-
itational and electric forces on tiny charged droplets of oil suspended between
two metal electrodes. Knowing the electric field, the charge on the droplet
could be determined. Repeating the experiment for many droplets, it was
found that the values measured were always multiples of the same number.
This was taken to be the charge on a single electron: 1.602 × 10 -19 coulombs
(SI unit for electric charge).
In 1923, Millikan won the Nobel Prize for physics in part because of this ex-
periment. This experiment has since been repeated by generations of physics
students, although it is rather expensive and difficult to do properly.

Oil-drop experiment
316

A version of the oil drop experiment has subsequently been used to search
for free quarks (which, if they exist, would have a charge of 1/3 e), without
success. Current theories of quarks predict that they are tightly bound and
will not exist in a free form, however it is interesting to note that in Millikan’s

FT
original notebooks he observed and recorded the existence of an oil droplet
which had a +1/3 partial charge, which he at the time dismissed as an error.

Experimental procedure

A
The apparatus
The diagram shows a simplified version of Millikan’s set up. A uniform electric
field is provided by a pair of horizontal parallel plates with a high potential
difference between them. A charged drop of oil is allowed to drift in between
them. By varying the potential, the drop can be made to rise, descend or stay
steady. The plates are held together by a ring of insulating material (not shown
in the diagram). There are two holes cut into the ring. A bright light source is
shone through one of the holes, and focused on the region where the oil drops
DR
drift between the plates. A low-powered microscope is inserted through the
other hole. The oil drops reflect the light and look like bright points on a dark
field of view through the microscope. The microscope has a graduated scale in
the eyepiece which allows for the velocity of the drop to be measured by timing
how long it takes to travel from one division to another.
The oil used is the type that is usually used in vacuum apparatus. This is
because this type of oil has an extremely low vapour pressure. Ordinary oil
would evaporate away under the heat of the light source and so the mass of
the oil drop would not remain constant over the course of the experiment.
Some oil drops will pick up a charge through friction with the nozzle as they
are sprayed, but more can be charged by allowing an ionising radiation source
(such as an x ray tube) to ionise the air in the chamber.

Oil-drop experiment
317

Method
Initially the oil drops are allowed to fall between the plates with the electric
field turned off. They very quickly reach a terminal velocity because of fric-
tion with the air in the chamber. The field is then turned on and, if it is large

FT
enough, some of the drops (the charged ones) will start to rise. (This is be-
cause the upwards electric force F E is greater for them than the downwards
gravitational force W ). A likely looking drop is selected and kept in the middle
of the field of view by alternately switching off the voltage until all the other
drops have fallen. The experiment is then continued with this one drop.
The drop is allowed to fall and its terminal velocity v 1 in the absence of an
electric field is calculated. The drag force acting on the drop can then be
worked out using Stokes’ law:

F = 6πrηv1

where v 1 is the terminal velocity (i.e. velocity in the absence of an electric


field) of the falling drop, η is the viscosity of the air, and r is the radius of
Athe drop.

The weight W is the volume V multiplied by the density ρ and the acceleration
due to gravity g. However what is needed is the apparent weight. The apparent
weight in air is the true weight minus the upthrust (which equals the weight
of air displaced by the oil drop). For a perfectly spherical droplet the apparent
weight can be written as:
W = 43 πr3 g(ρ − ρair )
DR
Now at terminal velocity the oil drop is not accelerating. So the total force
acting on it must be zero. So the two forces F and W must cancel one another
out.
F = W implies:
9ηv1
r2 = 2g(ρ−ρair )

Once r is calculated, W can easily be worked out.


Now the field is turned back on.
FE = qE

where q is the charge on the oil drop and E is the electric field between the
plates. For parallel plates

Oil-drop experiment
318

V
E= d

where V is the potential difference and d is the distance between the plates.
One conceivable way to work out q would be to adjust V until the oil drop

FT
remained steady. Then we could equate F E with W. But in practice this is
extremely difficult to do precisely. A more practical approach is to turn V up
slightly so that the oil drop rises with a new terminal velocity v 2. Then
qE − W = 6πrηv2

W v2
= v1

Millikan’s experiment and cargo cult science


Richard Feynman said in a commencement lecture he gave at Caltech in 1974
We have learned a lot from experience about how to handle some of the ways
we fool ourselves. One example: Millikan measured the charge on an electron
by an experiment with falling oil drops, and got an answer which we now know
Anot to be quite right. It’s a little bit off because he had the incorrect value for
the viscosity of air. It’s interesting to look at the history of measurements of
the charge of an electron, after Millikan. If you plot them as a function of time,
you find that one is a little bit bigger than Millikan’s, and the next one’s a little
bit bigger than that, and the next one’s a little bit bigger than that, until finally
they settle down to a number which is higher.
Why didn’t they discover the new number was higher right away? It’s a thing
that scientists are ashamed of - this history - because it’s apparent that people
DR
did things like this: When they got a number that was too high above Millikan’s,
they thought something must be wrong - and they would look for and find a
reason why something might be wrong. When they got a number close to
Millikan’s value they didn’t look so hard. And so they eliminated the numbers
that were too far off, and did other things like that. We’ve learned those tricks
nowadays, and now we don’t have that kind of a disease.

External links and references


• Karlsson, Magnus, " Millikan’s oildrop experiment 157". (Simplified version)

157 http://www.edu.falkenberg.se/gymnasieskolan/fysik/elektron/millikaneng.html

Oil-drop experiment
319

• Graphical simulation of the experiment - examples of the difficulties


• Thomsen, Marshall, " Good to the Last Drop 158". Millikan Stories as
"Canned" Pedagogy. Eastern Michigan University.
• CSR/TSGC Team, " Quark search experiment 159". The University of Texas at

FT
Austin.
• The oil-drop experiment appears in a list of Science’s 10 Most Beautiful
Experiments 160 originally published in the New York Times.

More external links


• Delpierre, G.R. and B.T. Sewell, " Millikan’s Oil Drop Experiment 161". 25
April 2005
• Engeness, T.E., " The Millikan Oil Drop Experiment 162". 25 April 2005
• Millikan R. A. (1913). "On the elementary electrical charge and the Avo-
gadro constant" 163. The Physical Review, Series II 2: 109–143., Paper by
Millikan discussing modifications to his original experiment to improve its
accuracy
• Cargo cult science 164, text of the Feynman lecture.
A• Millikan Oil Drop Experiment in space 165. A variation of this experiment has
been suggested for the International Space Station.

Source: http://en.wikipedia.org/wiki/Oil-drop_experiment

Principal Authors: Michael Hardy, Mjmcb1, Reddi, Popefelix, Theresa knott, Raul654, Gene Nygaard,
Linas
DR

158 http://www.physics.emich.edu/mthomsen/sege.htm
159 http://www.tsgc.utexas.edu/floatn/1997/teams/UT-austin.html
160 http://physics.nad.ru/Physics/English/top10.htm
161 http://www.physchem.co.za/Static%20Electricity/Millikan.htm
162 http://people.ccmr.cornell.edu/~muchomas/8.04/Lecs/lec_Millikan/Mill.html
163 http://www.aip.org/history/gap/PDF/millikan.pdf
164 http://www.physics.brocku.ca/etc/cargo_cult_science.html
165 http://scitation.aip.org/getabs/servlet/GetabsServlet?prog=normal&id=APCPCS000504000001000715000001

&idtype=cvips&gifs=yes

Oil-drop experiment
320

Open quantum system

In physics, an open quantum system is a quantum system which is found

FT
to be in interaction with an external quantum system, the environment. The
open quantum system can be viewed as a distinguished part of a larger closed
quantum system, the other part being the environment.
Open quantum systems are an important concept in quantum optics, measure-
ment theory, quantum statistical mechanics, quantum cosmology and semiclas-
sical approximations.

Source: http://en.wikipedia.org/wiki/Open_quantum_system

Principal Authors: Daniel Arteaga, Sn0wflake, Conscious

Optical theorem
A
In physics, the optical theorem is a very general law of wave scattering theory,
which relates the forward scattering amplitude to the total cross section of the
scatterer. It is usually written in the form

σtot = k Im f (0),

where f(0) is the scattering amplitude with an angle of zero, that is, the ampli-
tude of the wave scattered to the center of a distant screen. Because the optical
DR
theorem is derived using only conservation of energy, or in quantum mechan-
ics from conservation of probability, the optical theorem is widely applicable
and, in quantum mechanics, σtot includes both elastic and inelastic scattering.
Note that the above form is for an incident plane wave; a more general form
discovered by Werner Heisenberg can be written
Im f (k̂0 , k̂) = 4π
k
f (k̂0 , k̂00 )f (k̂00 , k̂) dk̂00 .
R

Notice that as a natural consequence of the optical theorem, an object that


scatters any light at all ought to have a nonzero forward scattering amplitude,
and so a bright central spot.

Optical theorem
321

History
The optical theorem was originally discovered independently by Sellmeier and
Lord Rayleigh in 1871. Lord Rayleigh recognized the forward scattering am-
plitude in terms of the index of refraction as

FT
n = 1 + 2πN f (0)/k 2

which he used in a study of the color and polarization of the sky. The equation
was later extended to quantum scattering theory by several individuals, and
came to be known as the Bohr-Peierls-Placzek relation after a 1939 publication.
It was first referred to as the Optical Theorem in print in 1955 by Hans Bethe
and de Hoffman, after it had been known as a "well known theorem of optics"
for some time.

Derivation
The theorem can be derived rather directly from a treatment of a scalar wave. If
a plane wave is incident on an object, then the wave amplitude a great distance
away from the scatterer is given approximately by
A ikr
ψ(r) ≈ eikz + f (θ) e r .

All higher terms, when squared, vanish more quickly than 1/r2 , and so are
negligible a great distance away. Notice that for large values of z and small
angles the binomial theorem gives us
2
+y 2
r = x2 + y 2 + z 2 ≈ z + x 2z
p
.
DR
We would now like to use the fact that the intensity is proportional to the
square of the amplitude ψ. Approximating the r in the denominator as z, we
have
f (θ) ikz ik(x2 +y 2 )/2z 2
|ψ|2 = |eikz + r e e |

f (θ) ik(x2 +y 2 )/2z f ∗ (θ) −ik(x2 +y 2 )/2z |f (θ)|2


=1+ z e + z e + z2
.

If we drop the 1/z 2 term and use the fact that A + A∗ = 2 Re A we have
f (θ) ik(x2 +y 2 )/2z
|ψ|2 ≈ 1 + 2 Re z e .

Now suppose we integrate over a screen in the x-y plane, which is small enough
for the small angle approximations to be appropriate, but large enough that we
can integrate the intensity from −∞ to ∞ with negligible error. In optics, this

Optical theorem
322

is equivalent to including many fringes of the diffraction pattern. To further


simplify matters, let’s approximate f (θ) = f (0). We quickly obtain
f (0) R ∞ 2 R∞ 2
|ψ|2 da ≈ A + 2 Re z −∞ eikx /2z dx −∞ eiky /2z dy
R

FT
where A is the area of the surface integrated over. The exponentials can be
treated as Gaussians and so
f (0)
|ψ|2 da = A + 2 Re z 2ziπ
R
k


=A− k Im f (0),

which is just the amount of energy that would reach the screen if none was
scattered, lessened by an amount (4π/k) Im f (0). Therefore, because of con-
servation of energy, that must be the total amount of energy scattered, and
thus it is the effective scattering cross section of the scatterer.

References

AR. G. Newton (1976). "Optical Theorem and Beyond". Am. J. Phys 44:
639-642.

• John David Jackson (1999). Classical Electrodynamics. Hamilton Printing


Company. ISBN 047130932X.

Source: http://en.wikipedia.org/wiki/Optical_theorem
DR
Principal Authors: Hyandat, Michael Hardy, Pflatau, That Guy, From That Show!, Phys

Parity (physics)

In physics, a parity transformation (also called parity inversion) is the simul-


taneous
  flip in
the sign
 of all spatial coordinates:
x −x
P : y  7→ −y 
z −z

A 3×3 matrix representation of P would have determinant equal to –1, and


hence cannot reduce to a rotation. In a two-dimensional plane, parity is the
same as a rotation by 180 degrees.

Parity (physics)
323

Simple symmetry relations


Under rotations, classical geometrical objects can be classified into scalars, vec-
tors, and tensors of higher rank. In classical physics, physical configurations
need to transform under representations of every symmetry group.

FT
In a quantum theory states in a →Hilbert space do not need to transform under
representations of the group of rotations, but only under projective representa-
tions. The word projective refers to the fact that if one projects out the phase of
each state, where we recall that the overall phase of a quantum state is not an
observable, then a projective representation reduces to an ordinary representa-
tion. All representations are also projective representations, but the converse is
not true, therefore the projective representation condition on quantum states
is weaker than the representation condition on classical states.
The projective representations of any group are isomorphic to the ordinary rep-
resentations of a central extension of the group. For example, projective repre-
sentations of the 3-dimensional rotation group, which is the special orthogonal
group SO(3 ), are ordinary representations of the special unitary group SU(2 ).
Projective representations of the rotation group that are not representations
A
are called spinors, and so quantum states may transform not only as tensors
but also as spinors.
If one adds to this a classification by parity, these can be extended, for example,
into notions of

• scalars (P = 1) and pseudoscalars (P = –1) which are rotationally invariant


• vectors (P = –1) and axial vectors (also called pseudo-vectors) (P = 1)
which both transform as vectors under rotation.
DR
One can
  reflections
define  such as
x −x
Vx : y  7→  y  ,
z z

which also have negative determinant. Then, combining them with rotations
one can generate the parity transformation defined earlier. In any even number
of dimensions, the first definition of parity has positive determinant, and hence
can be obtained as some rotation. One then uses reflections to extend the
notion of scalars and vectors to pseudo-scalars and pseudo-vectors.
Parity forms the Abelian group Z 2 due to the relation P 2 = 1. All Abelian
groups have only one dimensional irreducible representations. For Z 2, there
are two irreducible representations: one is even under parity (P φ = φ), the

Parity (physics)
324

other is odd (P φ = –φ). These are useful in quantum mechanics. However,


as is elaborated below, in quantum mechanics states need not transform under
actual representations of parity but only under projective representations and
so in principle a parity transformation may rotate a state by any phase.

FT
Classical mechanics
Newton’s equation of motion F = ma (if mass is constant) equates two vectors,
and hence is invariant under parity. The law of gravity also involves only vec-
tors and is also, therefore, invariant under parity. However angular momentum
is an axial vector.
L=r×p,

P(L) = (–r) × (–p) = L.

In classical electrodynamics, charge density ρ is a scalar, the electric field, E,


and current j are vectors, but the magnetic field, H is an axial vector. However,
Maxwell’s equations are invariant under parity because the curl of an axial
A
vector is a vector.

Quantum mechanics
Possible eigenvalues
In quantum mechanics, spacetime transformations act on quantum states. The
parity transformation, P, is a unitary operator in quantum mechanics, acting
on a state ψ as follows: P ψ(r) = ψ(-r). One must have P 2 ψ(r) = e i φ ψ(r),
DR
since an overall phase is unobservable.
The operator P 2, which reverses the parity of a state twice, leaves the spacetime
invariant and so is an internal symmetry which rotates its eigenstates by phases
e i φ. If P 2 is an element e i Q of a continuous U(1) symmetry group of phase
rotations then e -i Q/2 is part of this U(1) and so is also a symmetry. In particular
we can define P’=Pe -i Q/2 which is also a symmetry and so we can choose to call
P’ our parity operator instead of P. Notice that P’ 2=1 and so P’ has eigenvalues
±1. However when no such symmetry group exists, it may be that all parity
transformations have some eigenvalues which are phases other than ±1.

Consequences of parity symmetry


When parity generates the Abelian group Z 2, one can always take linear com-
binations of quantum states such that they are either even or odd under parity

Parity (physics)
325

A FT
Figure 30 Two dimensional representations of parity
are given by a pair of quantum states which go into
each other under parity. However, this representation
can always be reduced to linear combinations of states,
each of which is either even or odd under parity. One
says that all irreducible representations of parity are
one-dimensional.
DR
(see the figure). Thus the parity of such states is ±1. The parity of a multipar-
ticle state is the product of the parities of each state; in other words parity is a
multiplicative quantum number.
In quantum mechanics, Hamiltonians are invariant (symmetric) under a parity
transformation if P commutes with the Hamiltonian. In non-relativistic quan-
tum mechanics, this happens for any potential which is scalar, ie, V = V(r),
hence the potential is spherically symmetric. The following facts can be easily
proven:

• If |A> and |B> have the same parity, then <A| X |B> = 0 where X is the
position operator.
• For a state |L,m> of orbital angular momentum L with z-axis projection m,
P |L,m> = (-1) L|L, m>.

Parity (physics)
326

• If [H,P] = 0, then no transitions occur between states of opposite parity.


• If [H,P] = 0, then a non-degenerate eigenstate of H is also an eigenstate
of the parity operator; i.e., a non-degenerate eigenfunction of H is either
invariant to P or is changed in sign by P.

FT
Some of the non-degenerate eigenfunctions of H are unaffected (invariant) by
parity P and the others will be merely be reversed in sign when the Hamiltonian
operator and the parity operator commute:
P Ψ = c Ψ,

where c is a constant, the eigenvalue of P,


P P Ψ = P c Ψ.

Quantum field theory


The intrinsic parity assignments in this section are true for relativistic quan-
tum mechanics as well as quantum field theory.
A
If we can show that the vacuum state is invariant under parity (P |0> = |0>),
the Hamiltonian is parity invariant ([H,P] = 0) and the quantization conditions
remain unchanged under parity, then it follows that every state has good parity,
and this parity is conserved in any reaction.
To show that quantum electrodynamics is invariant under parity, we have to
prove that the action is invariant and the quantization is also invariant. For
simplicity we will assume that canonical quantization is used; the vacuum state
DR
is then invariant under parity by construction. The invariance of the action
follows from the classical invariance of Maxwell’s equations. The invariance
of the canonical quantization procedure can be worked out, and turns out to
depend on the transformation of the annihilation operator:
P a(p,±) P + = -a(-p,±)

where p denotes the momentum of a photon and ± refers to its polarization


state. This is equivalent to the statement that the photon has odd intrinsic
parity. Similarly all vector bosons can be shown to have odd intrinsic parity,
and all axial-vectors to have even intrinsic parity.
There is a straightforward extension of these arguments to scalar field theories
which shows that scalars have even parity, since
P a(p) P + = a(-p).

Parity (physics)
327

This is true even for a complex scalar field. (Details of spinors are dealt with
in the article on the Dirac equation, where it is shown that fermions and an-
tifermions have opposite intrinsic parity.)
With fermions, there is a slight complication because there is more than one

FT
pin group. (See the article on pin groups for more details.)

Parity in the standard model


Fixing the global symmetries
In the standard model of fundamental interactions there are precisely three
global internal U(1) symmetry groups available, with charges equal to the bary-
on number B, the lepton number L and the electric charge Q. The product of the
parity operator with any combination of these rotations is another parity oper-
ator. It is conventional to choose one specific combination of these rotations to
define a standard parity operator, and other parity operators are related to the
standard one by internal rotations. One way to fix a standard parity operator is
to assign the parities of three particles with linearly independent charges B, L
Aand Q. In general one assigns the parity of the most common massive particles,
the proton, the neutron and the electron, to be +1.
Steven Weinberg has shown that if P 2=(-1) F, where F is the fermion number
operator, then, since the fermion number is the sum of the lepton number plus
the baryon number, F=B+L, for all particles in the standard model and since
lepton number and baryon number are charges Q of continuous symmetries
e i Q, it is possible to redefine the parity operator so that P 2=1. However, if
there exist Majorana neutrinos, which experimentalists today believe is quite
DR
likely, their fermion number is equal to one because they are neutrinos while
their baryon and lepton numbers are zero because they are Majorana, and so
(-1) F would not be embedded in a continuous symmetry group. Thus Majorana
neutrinos would have parity ±i.

Parity of the pion


In the 1954 paper Absorption of Negative Pions in Deuterium: Parity of the
Pion 166, William Chinowsky and Jack Steinberger demonstrated that the pion π
has negative parity. They studied the decay of an atom made from a deuterium
nucleus d and a negatively charged pion π - in a state with zero orbital angular
momentum L =0 into two neutrons n
d π − −→ n n.

166 http://prola.aps.org/abstract/PR/v95/i6/p1561_1

Parity (physics)
328

Neutrons are fermions and so obey Fermi statistics, which implies that the final
state is antisymmetric. Using the fact that the deuteron has spin one and the
pion spin zero together with the antisymmetry of the final state they concluded
that the two neutrons must have orbital angular momentum L =1. The total

FT
parity is the product of the intrinsic parities of the particles and the extrinsic
parity (-1) L. Since the orbital momentum changes from zero to one in this
process, if the process is to conserve the total parity then the products of the
intrinsic parities of the initial and final particles must have opposite sign. A
deuteron nucleus is made from a proton and a neutron, and so using the fore-
mentioned convention that protons and neutrons have intrinsic parities equal
to +1 they argued that the parity of the pion is equal to minus the product of
the parities of the two neutrons divided by that of the proton and neutron in
deuterium, (-1)(1) 2/(1) 2, which is equal to minus one. Thus they concluded
that the pion is a pseudoscalar particle.

Parity violation
Parity is not a symmetry of our universe. Although it is conserved in electro-
magnetism, strong interactions and gravity, it turns out to be violated in weak
A
interactions. The Standard Model incorporates parity violation by expressing
the weak interaction as a chiral gauge interaction. Only the left-handed com-
ponents of particles and right-handed components of antiparticles participate
in weak interactions in the Standard Model.
The history of the discovery of parity violation is interesting. It was suggested
several times and in different contexts that parity might not be conserved, but
in the absence of compelling evidence these were not taken seriously. A care-
ful review by theoretical physicists Tsung Dao Lee and Chen Ning Yang went
DR
further, showing that while parity conservation had been verified in decays by
the strong or electromagnetic interactions, it was untested in the weak inter-
action. They proposed several possible direct experimental tests. They were
almost ignored, but Lee was able to convince his Columbia colleague Chien-
Shiung Wu to try it. She needed special cryogenic facilities and expertise, so
the experiment was done at the National Bureau of Standards.
In 1956-1957 Wu, E. Ambler, R. W. Hayward, D. D. Hoppes, and R. P. Hudson
found a clear violation of parity conservation in the beta decay of cobalt-60. As
the experiment was winding down, with doublechecking in progress, Wu in-
formed her colleagues at Columbia of their positive results. Three of them, R.
L. Garwin, Leon Lederman, and R. Weinrich modified an existing cyclotron ex-
periment and immediately verified parity violation. They delayed publication
until after Wu’s group was ready; the two papers appeared back to back.

Parity (physics)
329

After the fact, it was noted that an obscure 1928 experiment had in effect
reported parity violation in weak decays, but as the appropriate concepts had
not been invented yet, it had no impact. The discovery of parity violation
immediately explained the outstanding τ -θ puzzle in the physics of kaons.

FT
Intrinsic parity of hadrons
To every particle one can assign an intrinsic parity as long as nature pre-
serves parity. Although weak interactions do not, one can still assign a parity
to any hadron by examining the strong interaction reaction that produces it, or
through decays not involving the weak interaction, such as
π 0 → γγ.

See also
• Charge conjugation, time reversal and CPT symmetry
• Standard model of particle physics, and the electroweak theory
• Vector
A
References and external links
• CP violation, by I.I. Bigi and A.I. Sanda 167 [ISBN 0521443490]]

• Weinberg, S. (1995). The Quantum Theory of Fields. Cambridge University


Press. ISBN 0-521-67053-5.
DR
Source: http://en.wikipedia.org/wiki/Parity_%28physics%29

Principal Authors: Xerxes314, JarahE, Bambaiah, Raghunathan, Patrick

167 http://www.amazon.com/exec/obidos/tg/detail/-/0521443490

Parity (physics)
330

Particle in a box

In physics, the particle in a box (also known as the infinite potential well

FT
or the infinite square well) is a very simple problem consisting of a single
particle bouncing around inside of an immovable box, from which it cannot
escape, and which loses no energy when it collides with the walls of the box.
In classical mechanics, the solution to the problem is trivial: The particle moves
in a straight line, always at the same speed, until it reflects from a wall. When
it reflects from a wall, it always reflects at an equal but opposite angle to its
angle of approach, and its speed does not change.
The problem becomes very interesting when one attempts a quantum-
mechanical solution, since many fundamental quantum mechanical concepts
need to be introduced in order to find the solution. Nevertheless, it remains a
very simple and solvable problem. This article will only be concerned with the
quantum mechanical solution.
The problem may be expressed in any number of dimensions, but the simplest
problem is one dimensional, while the most useful solution is the particle in the
A
three dimensional box. In one dimension this amounts to the particle existing
on a line segment, with the "walls" being the endpoints of the segment.
In physical terms, the particle in a box is defined as a single point particle,
enclosed in a box inside of which it experiences no force whatsoever, i.e. it is
at zero potential energy. At the walls of the box, the potential rises to infinity,
forming an impenetrable wall. Using this description in terms of potentials
allows the →Schrödinger equation to be used to determine the solution.
DR
As mentioned above, if we were studying this system under the rules of classical
mechanics we would apply Newton’s laws of motion to the initial conditions
and the result would seem reasonable and intuitive. In quantum mechanics,
when the →Schrödinger equation is applied to the proposed system, the results
are not intuitive. In the first place, the particle can only have certain specific
energy levels, and the zero energy level is not one of them. Secondly, the
chances of detecting the particle in the box at any specific energy level is not
uniform - there are certain locations in the box where the particle might be
found, but there are also places where it can never be found. Both of these
results differ from the usual way we perceive the world, yet rest on principles
that have been extensively experimentally verified.

Formal Introduction
The particle in a box (or the infinite potential well or infinite square well)
is a simple idealized system that can be completely solved within quantum

Particle in a box
331

mechanics. It is the situation of a particle confined within a finite region of


space (the box) by an infinite potential that exists at the walls of the box. The
particle experiences no forces while inside the box, but is constrained by the
walls to remain in the box. This is similar to the situation of a gas confined in a

FT
container. For simplicity we start with the 1-dimensional case, where all motion
is constrained to a single dimension. Later we will extend the discussion to the
2 and 3 dimensional cases. See also the →Particle in a spherically symmetric
potential where the case is treated of a particle in a spherical box, or the particle
in a ring which shows the case for a particle in a 1D ring. The statistical
mechanics of many particles in a box is developed in the gas in a box article.
As we shall see, the solution of the →Schrödinger equation for the particle in
a box problem reveals some decidedly quantum behavior of the particle that
agrees with observation but contrasts sharply with the predictions of classical
mechanics. This is a particularly useful illustration because this behaviour is
not "forced" on the system, it arises naturally from the initial conditions. It
neatly demonstrates that quantum behaviour is a natural outcome of any wave-
like system, contrary to the common concept of a "quantum leap" where the
behavior is almost magical.
A
The quantum behavior in the box includes:

• Energy quantization - It is not possible for the particle to have any arbitrary
definite energy. Instead only discrete definite energy levels are allowed (if
the state is not a steady state, however, any energy past zero-point energy
is allowed on average).
• Zero-point energy - The lowest possible energy level of the particle, called
the zero-point energy, is nonzero.
DR
• Nodes - In contrast to classical mechanics the Schrödinger equation predicts
that for some energy levels there are nodes, implying positions at which the
particle can never be found.

One can solve analytically the Schrödinger equation for such a simple potential.
However trivial, this case is both of great technical value for the insights it
allows, and of paramount physical importance. Depending on the boundary
conditions, one can use the solutions to describe two important systems. If one
considers real valued solutions (of which detailed derivation is given below),
one describes actual potentials of heterostructures called quantum wells which
trap spatially particles, typically electrons and holes. If one considers complex
valued solutions, one describes conveniently a particle propagating freely in a
constrained volume (like a solid).

Particle in a box
332

Solutions
The particle in a 1-dimensional box
For the 1-dimensional case in the x direction, the time-independent

FT
Schrödinger equation can be written as:
2 2
~ d ψ
− 2m dx2
+ V (x)ψ = Eψ (1)

where

h
~= 2π

h is Planck’s constant

m is the mass of the particle

ψ is the complex-valued stationery time-independent wavefunction that


A we want to find

V (x) is a function describing the potential at each point x and

E is the energy, a real number.

For the case of the particle in a 1-dimensional box of length L, the potential is
zero inside the box, but rises abruptly to infinity at x = 0 and x = L. Thus for
DR
the region inside the box V (x) = 0 and Equation 1 reduces to:
2 2
~ d ψ
− 2m dx2
= Eψ (2)

This is a well studied differential equation and eigenvalue problem with a gen-
eral solution of:
ψ = A sin(kx) + B cos(kx)

k 2 ~2
E= 2m (3)

Here, A and B can be any complex numbers, and k can be any real number (k
must be real because E is real).
Now in order find the specific solution for the problem at hand, we must spec-
ify the appropriate boundary conditions and find the values for A and B that

Particle in a box
333

A Figure 31
FT
The Potential is 0 inside the box, and infinite elsewhere

satisfy those conditions. One usually resorts to one of the following two choic-
es, describing two kinds of systems. The first case, with which we shall pursue
our derivation, demands that ψ equal zero at x = 0 and x = L. A handwaving
argument to motivate these boundary conditions is that the particle is unlikely
to be found at a location with a high potential (the potential repulses the par-
DR
ticle), thus the probability of finding the particle, |ψ| 2, must be small in these
regions and decreases with increasing potential. For the case of an infinite po-
tential, |ψ| 2 must infinitesimally small or 0, thus ψ must also be zero in this
region. In summary,
ψ(0) = ψ(L) = 0 (4)

The second case, to which solutions are given in section free propagation
at the end of this article, does not compel the wavefunction to vanish at the
boundary. This means that when the particle reaches one border of the well, it
instantaneously disappears from this side to reappear on the opposite side, as
if the well was some kind of torus. The value of the solutions are discussed in
the appropriate section. We now resume derivation with vanishing boundary
conditions.

Particle in a box
334

Substituting the general solution from Equation 3 into Equation 2 and evaluat-
ing at x = 0 (ψ = 0), we find that B = 0 (since sin(0) = 0 and cos(0) = 1). It
follows that the wavefunction must be of the form:
ψ = A sin(kx) (5)

FT
and at x = L we find:
ψ = A sin(kL) = 0 (6)

One solution for Equation 6 is A = 0, however, this "trivial solution" would


imply that ψ = 0 everywhere (I.e. the particle isn’t in the box.) and can be
thrown out. If A 6= 0 then sin(kL) = 0, which is only true when:
kL = nπ where n = 1, 2, . . .


or k = L (7)

(note that n = 0 is ruled out because then ψ = 0 everywhere, corresponding


A
to the case where the particle is not in the box. Negative values of n are also
neglected, since they merely change the sign of sin(nx)). Now in order to find
A we must undertake a process called normalising the wavefunction. We rec-
ognize that the particle must exist somewhere in space. |ψ|2 is the probability
of finding the particle at a particular point in space, so the integral of this value
over all x must be equal to 1:
R∞ RL
1 = −∞ |ψ|2 dx = |A|2 0 sin2 kx dx = |A|2 L2
DR
or
q
2
|A| = L (8)


Thus, A may be any complex number with absolute value (2/L); these dif-

ferent values of A yield the same physical state, so we choose A = (2/L) to
simplify.
Finally, substituting the results from Equations 7 and 8 into Equation 3, the
complete set of energy eigenfunctions for the 1-dimensional particle in a box
problem is:
q
ψn = L2 sin nπx

L (9)

Particle in a box
335

n2 ~2 π 2 n 2 h2
En = 2mL2
= 8mL2
(10)

with
n = 1, 2, 3, . . .

FT
Note, that as mentioned previously, only "quantized" energy levels are possi-
ble. Also, since n cannot be zero, the lowest energy from Equation 10 is also
non-zero. This zero-point energy, as it is called, can be explained in terms
of the uncertainty principle. Because the particle is constrained within a finite
region, the variance in its position is upper-bounded. Thus due to the uncer-
tainty principle the variance in the particle’s momentum cannot be zero, so the
particle must contain some amount of energy that increases as the length of
the box, L, decreases.
Also, since ψ consists of sine waves, for any value of n greater than one, there
are regions within the box for which ψ and thus ψ 2 both equal zero, indicating
that for these energy levels, nodes exist in the box where the probability of
finding the particle is zero.
A
The particle in a 2-dimensional or 3-dimensional rectangular
box
For the 2-dimensional case the particle is confined to a rectangular surface of
length L x in the x-direction and L y in the y-direction. Again the potential is zero
inside the "box" and infinite at the walls. For the region inside the box, where
the potential is zero, the two dimensional analogue of Equation 2 applies:
 2 
~2 ∂ ψ ∂2ψ
− 2m 2 + = Eψ (11)
DR
2
∂x ∂y

In this case ψ is a function of both x and y, so ψ=ψ(x,y). In order to solve


Equation 11, we use the method of separation of variables. First, we assume
that ψ can be expressed as the product of two independent functions, the first
depending only on x and the second depending only on y; i.e.:
ψ(x, y) = X(x)Y (y) (12)

Substituting Equation 12 into Equation 11 and evaluating the partial deriva-


tives gives:
 2 
~2 2
− 2m Y ∂∂xX2 + X ∂∂yY2 = EXY (13)

which upon dividing by XY and rewriting d 2X /dx 2 as X " and d 2Y /dy 2 as Y "
becomes:

Particle in a box
336
 
~ 2
X 00 Y 00
− 2m X + Y =E (14)

Now we note that since X "/X is independent of y, varying y can only change
the Y "/Y term. However, from Equation 14 we see that changing Y "/Y without

FT
varying X "/X, would also change E, but E is a constant, so Y "/Y must also be
a constant, independent of y. The same argument can be applied to show that
X "/X is independent of x. Since X "/X and Y "/Y are constants, we can write:
00
2
~ X ~2 Y 00
− 2m X = Ex and − 2m Y = Ey (15)

where E x + E y = E. Expanding X " and Y " in terms of the derivatives and


rearranging gives:
2 2
~ ∂ X
− 2m ∂x2
= Ex X (16)

2 2
~ ∂ Y
− 2m ∂y 2
= Ey Y (17)

each of which are of the same form as the 1-dimensional Schrödinger equation
(Equation 2) we solved in the previous section. Thus, adapting the results from
A
the previous section gives:
q  
Xnx = L2x sin nLx πxx
(18)
q  
2 ny πy
Yny = Ly sin Ly (19)

Finally, since ψ=XY and E = E x + E y, we obtain the solutions:


q    
ny πy
ψnx ,ny = Lx4Ly sin nLx πx
DR
x
sin L y
(20)
 2  2 
h2 nx ny
Enx ,ny = 8m Lx + Ly (21)

The same separation of variables technique can be applied to the three di-
mensional case to give the energy eigenfunctions:
q      
ny πy
ψnx ,ny ,nz = Lx L8y Lz sin nLx πx
x
sin L y
sin nz πz
L z
(22)
 2  2  2 
~2 π 2 nx ny nz
Enx ,ny ,nz = 2m Lx + Ly + Lz (23)

with
ni = 1, 2, 3, . . .

Particle in a box
337

An interesting feature of the above solutions is that when two or more of the
lengths are the same (e.g. L x = L y ), there are multiple wavefunctions corre-
sponding to the same total energy. For example the wavefunction with n x =
2, n y = 1 has the same energy as the wavefunction with n x = 1, n y = 2. This

FT
situation is called degeneracy and for the case where exactly two degener-
ate wavefunctions have the same energy that energy level is said to be doubly
degenerate. Degeneracy results from symmetry in the system. For the above
case two of the lengths are equal so the system is symmetric with respect to a
90◦ rotation.

Free propagation
If the potential is zero (or constant) everywhere, one describes a free particle.
This leads to some difficulties of normalization of the momentum or energy
eigenfunctions. One way around is to constrain the particle in a finite volume
V of arbitrary (large) extension, in which it is free to propagate. It is expected
that in the limit of V → ∞ we recover the free particle while allowing in the
intermediate calculations the use of properly normalized states. Also, when
describing for instance a particle propagating in a solid, one does not expect
A
spatially localized states but instead completely delocalized states (within the
solid), meaning that the particles propagates inside it (since it can be every-
where with the same probability, conversely to the sine solutions we encoun-
tered where the particle has favored locations). This understanding follows
from the solutions of the Schrödinger equation for zero potential following
from the so-called Von-Karman boundary conditions; i.e., the wavefunction as-
sumes same values on opposite sides of the box but it is not required to be zero
here. One can then check that the following solutions obey eq. 1:
DR
in 1D : ψk (x) = √1 eikx ; k = 2nπ
L L ; n∈Z

√1 eik·r ; kx 2nx π
in 3D : ψk (x) = L3
= L ; ky =
2ny π 2nz π
L ; kz = L ; nx , ny , nz ∈ Z

The energy remains ~2 k 2 /2m (cf. eq. 3) but interestingly, now the k are twice
as before (cf. eq. 7). This is because in the previous case, n was strictly positive
whereas now it can be negative or zero (the ground state). The solutions
where the sine does not superpose to itself after a translation of L can not be
recovered with exponentials, since in this propagating particle interpretation,
the derivative is discontinuous at the border, meaning that the particle acquires
infinite velocity here. This shows how the two interpretations bear intrinsically
differing behaviours.

Particle in a box
338

References
• Griffiths, David J. (2004). Introduction to Quantum Mechanics (2nd ed.).
Prentice Hall. ISBN 0131118927.

FT
See also
• finite potential well
• particle in a ring

External links
• Scienceworld 168 (Infinite Potential Well)

• Scienceworld 169 (Finite Potential Well)

• 1-D quantum mechanics java applet 170 simulates particle in a box, as well
as other 1-dimensional cases.
A• 2-D particle in a box applet 171

Source: http://en.wikipedia.org/wiki/Particle_in_a_box

Principal Authors: PAR, Dgrant, Pfalstad, Heron, AxelBoldt, Paul August, Laussy, JabberWok

Particle in a one-dimensional lattice (peri-


DR
odic potential)

In quantum mechanics, the particle in a one-dimensional lattice is problem


that occurs in the model of a periodic crystal lattice. The problem can be
simplified from the 3D infinite potential barrier (particle in a box) to a one-
dimensional case. The potential is caused by ions in the periodic structure of
the crystal creating an electromagnetic field so electrons are subject to a regular
potential inside the lattice. This is an extension of the free electron model that
assumes zero potential inside the lattice.

168 http://scienceworld.wolfram.com/physics/InfiniteSquarePotentialWell.html
169 http://scienceworld.wolfram.com/physics/FiniteSquarePotentialWell.html
170 http://www.falstad.com/qm1d/
171 http://www.falstad.com/qm2dbox/

Particle in a one-dimensional lattice (periodic potential)


339

Problem definition
When talking about solid materials, the discussion is mainly around
crystals - periodic lattices. Here we will discuss a 1-dimensional
lattice of positive ions. Assuming the spacing between two ions

FT
is a, the potential in the lattice will look something like this:

The
mathematical representation of the potential is a periodic function with a
A
period a. According to Bloch’s theorem, the wavefunction solution of the
→Schrödinger equation when the potential is periodic, can be written as:
ψ(x) = eikx u(x)

Where u(x) is a periodic function which satisfies:


u(x + a) = u(x)
DR
u0 (x + a) = u0 (x)

When nearing the edges of the lattice, there are problems with the boundary
condition. Therefore, we can represent the ion lattice as a ring. If L is the
length of the lattice so that L » a, then the number of ions in the lattice is so
big, that when considering one ion, its surrounding is almost linear, and the
wavefuntion of the electron is unchanged. So now, instead of two boundary
conditions we get one circular boundary condition:
ψ(0) = ψ(L)

If N is the number of Ions in the lattice, then we have the relation: aN = L.


Replacing in the boundary condition and applying Bloch’s theorem will result
in a quantization for k :

Particle in a one-dimensional lattice (periodic potential)


340

ψ(0) = eik·0 u(0) = eikL u(L) = ψ(L)

u(0) = eikL u(N a) → eikL = 1

FT
 

⇒ kL = 2πn → k = Ln n = 0, ±1, ±2, ..., ± N2 .

Kronig-Penney model
In order to simplify the problem the potential function is approximated by a

A
rectangular potential:
Using Bloch’s theorem, we only need to find a solution for a single period, make
sure it is continuous and smooth, and to make sure the function u(x) is also
continuous and smooth. Considering a single period of the potential:
DR

Particle in a one-dimensional lattice (periodic potential)


341

A0<x<a−b: −~2
2m ψxx

⇒ ψ = Aeiαx + A0 e−iαx

−b < x < 0 : −~2


2m ψxx
FT
have two regions here. We will solve for each independently:
= Eψ

α2 =

= (E + V0 )ψ
2mE
~2

We
DR
 
2m(E+V0 )
⇒ ψ = Beiβx + B 0 e−iβx β2 = ~2

In order to find u(x) in each region we need to manipulate the probability


function:
 
ψ(0 < x < a − b) = Aeiαx + A0 e−iαx = eikx · Aei(α−k)x + A0 e−i(α+k)x

⇒ u(0 < x < a − b) = Aei(α−k)x + A0 e−i(α+k)x

And in the same manner:


u(−b < x < 0) = Bei(β−k)x + B 0 e−i(β+k)x

To complete the solution we need to make sure the probability function is


continuous and smooth, i.e:

Particle in a one-dimensional lattice (periodic potential)


342

ψ(0− ) = ψ(0+ ) ψ 0 (0− ) = ψ 0 (0+ )

And that u(x) and u( x) are periodic


u(−b) = u(a − b) u0 (−b) = u0 (a − b).

FT
These

conditions yield the following matrix:  
1 1 −1 −1 A

 α −α −β β  A0 
  =
i(α−k)(a−b)
 e e−i(α+k)(a−b) −e−i(β−k)b −ei(β+k)b   B 
i(α−k)(a−b) i(β+k)b B0
  (α − k)e (α + k)e−i(α+k)(a−b) −(β − k)e−i(β−k)b (β + k)e
0
0
 
0
0

In order for us not to have the trivial solution, the determinant of the matrix
must be 0. This leads us to the following expression:
α2 +β 2
cos(ka) = cos(βb) cos[α(a − b)] − 2αβ sin(βb) sin[α(a − b)]
A
In order to further simplify the expression, we will perform the following ap-
proximations:
b → 0 ; V0 → ∞ ; V0 b = constant

⇒ βb → 0 ; β 2 b = constant ; α2 b → 0 ; sin(βb) → βb ; cos(βb) → 1

The expression will now be:


DR
 
sin(αa) β 2 ab
cos(ka) = cos(αa) − P αa P = 2

See also
• Ralph Kronig Kronig-Penney model
• Free electron model
• Nearly-free electron model
• Crystal structure
• Potential

Particle in a one-dimensional lattice (periodic potential)


343

External links
• 1-D periodic potential applet 172

FT
Source: http://en.wikipedia.org/wiki/Particle_in_a_one-dimensional_lattice_%28periodic_potential
%29

Principal Authors: Michael Hardy, Pfalstad, Rubber hound, Sverdrup, Salty-horse

Particle in a ring

In quantum mechanics, the case of a particle in a one-dimensional ring is


similar to the particle in a box. The →Schrödinger equation for a free particle
which is restricted to a ring (technically, whose configuration space is the circle
S 1 ) is
2
~
− 2m ∇2 ψ = Eψ
A
Using polar coordinates on the 1 dimensional ring, the wave function depends
only on the angular coordinate, and so
1 ∂2
∇2 = r2 ∂θ2

Requiring that the wave function be periodic in θ with a period 2 π (from the
demand that the wave functions be single-valued functions on the circle), and
DR
that they be normalized leads to the conditions
R 2π 2
0 |ψ(θ)| dθ = 1 ,

and
ψ(θ) = ψ(θ + 2 π)

Under these conditions, the solution to the Schrödinger equation is given by


r

ψ(θ) = √1 e±i ~ 2mEθ

The energy eigenvalues E are quantized because of the periodic boundary con-
ditions, and they are required to satisfy

172 http://www.falstad.com/qm1dcrystal/

Particle in a ring
344
r
√ r

e±i ~ 2mEθ = e±i ~ 2mE(θ+2π) , or

r

e±i2π ~ 2mE = 1 = ei2πn

FT
This leads to the energy eigenvalues
n2 ~2
E= 2mr2
where n = 0, 1, 2, 3, . . .

The full wave functions are, therefore


ψ(θ) = √1 e±inθ

Except for the case n = 0, there are two quantum states for every value of n
(corresponding to e±inθ ). Therefore there are 2n+1 states with energies less
than an energy indexed by the number n.
The case of a particle in a one-dimensional ring is an instructive example when
studying the quantization of angular momentum for, say, an electron orbiting
the nucleus. The azimuthal wave functions in that case are identical to the
A
energy eigenfunctions of the particle on a ring.
Interestingly, the statement that any wavefunction for the particle on a ring can
be written as a superposition of energy eigenfunctions is exactly identical to
Fourier’s theorem about the development of any periodic function in a Fourier
series.
This simple model can be used to find approximate energy levels of some ring
molecules, such as benzene.
DR
See also
• Angular momentum
• Harmonic analysis.
• One-dimensional periodic case

Source: http://en.wikipedia.org/wiki/Particle_in_a_ring

Principal Authors: AmarChandra, Charles Matthews, Creidieki, Michael Hardy, Idril

Particle in a ring
345

Particle in a spherically symmetric poten-


tial

FT
In quantum mechanics, the particle in a spherically symmetric potential de-
scribes the dynamics of a particle in a central force field, i.e. with potential
depending only on the distance of the particle to the center of force (radi-
al dependency), having no angular dependency. In its quantum mechanical
formulation, it amounts to solving the →Schrödinger equation with potentials
V( r) which depend only on r, the modulus of r.
Three special cases arise, of special importance:

• V(r) =0, or solving the vacuum in the basis of spherical harmonics, which
serves as the basis for other cases.
• V (r) = V0 for r < r0 and 0 (or ∞) elsewhere, or particle in the spherical
equivalent of the square well, useful to describe scattering and bound states
in a nucleus or quantum dot.
A
• V(r) 1/r to describe bound states of atoms, especially hydrogen.

We outline the solutions in these cases, which should be compared to their


counterparts in cartesian coordinates, cf. particle in a box. This article relies
heavily on Bessel functions.

General considerations
The time independent solution of 3D Schrödinger equation with hamiltonian
p2 /2m0 +V (r) where m0 is the particle’s mass, can be separated in the variables
DR
r, θ and φ so that the wavefunction ψ reads:
ψ(r) = R(r)Ylm (θ, φ)

Ylm are the usual Spherical harmonics, while R needs be solved with the so-
called radial equation:
h 2
i
d d l(l+1)
− 2m~ r2 dr (r2 dr ) + ~2 2m r2 + V (r) R(r) = ER(r)
0 0

It has the shape of the 1D Schrödinger equation for the variable u(r) ≡ rR(r),
with a centrifugal term ~2 l(l + 1)/2m0 r2 added to V, but r ranges from 0 to ∞
rather than over R.
For more information about how one derive Spherical harmonics from spheri-
cal symmetry, see Angular momentum, since the spherical harmonics are the
eigenstates of the operator L 2.

Particle in a spherically symmetric potential


346

Vacuum case
Let us now consider V(r) =0 (if V0 , replace everywhere E with E − V0 ). Intro-
ducing the dimensionless variable
q

FT
ρ ≡ kr, k ≡ 2m~20 E r


the equation becomes a Bessel equation for J defined by J(ρ) ≡ ρR(r)
(whence the notational choice of J ):
h  i
2
ρ2 ddρJ2 + ρ dJ + ρ 2− l+ 1 2 J =0
dρ 2

which regular solutions for positive energies are given by so-called Bessel func-
tions of the first kind Jl+1/2 (ρ) so that the solutions written for R are the
p
so-called Spherical Bessel function R(r) = jl (kr) ≡ π/(2kr)Jl+1/2 (kr).
The solutions of Schrödinger equation in polar coordinates for a particle of
mass m0 in vacuum are labelled by three quantum numbers: discrete indices l
and m, and k varying continuously in [0, ∞]:
ψ(r) = jl (kr)Ylm (θ, φ)
A √
where k ≡ 2m0 E/~, jl are the spherical Bessel function and Ylm are the
spherical harmonics.
These solutions represent states of definite angular momentum, rather than of
definite (linear) momentum, which are provided by plane waves exp(ik · r).

Spherical square well


DR
Let us now consider the potential V (r) = V0 for r < r0 , i.e., inside a sphere of
radius r0 and zero outside.
We first consider bound states, i.e., states which display the particle mostly
inside the box (confined states). Those have an energy E less than the potential
outside the sphere, i.e., they have negative energy, and we shall see that there
are a discrete number of such states, which we shall compare to positive energy
with a continuous spectrum, describing scattering on the sphere (of unbound
states). Also worth noticing is that unlike Coulomb potential, featuring an
infinite number of discrete bound states, the spherical square well has only a
finite (if any) number because of its finite range (if it has finite depth).
The resolution essentially follows that of the vacuum with normalisation of
the total wavefunction added, solving two Schrödinger equations—inside and
outside the sphere—of the previous kind, i.e., with constant potential. Also the
following constraints hold:

Particle in a spherically symmetric potential


347

• The wavefunction must be regular at the origin.


• The wavefunction and its derivative must be continuous at the potential
discontinuity.
• The wavefunction must converge at infinity.

FT
The first constraint comes from the fact that Neumann N and Hankel H func-
tions are nonsingular at the origin. The physical argument that ψ must be
defined everywhere selected Bessel function of the first kind J over the other
possibilities in the vacuum case. For the same reason, the solution will be of
this kind inside the sphere:
q 
2m0 (E−V0 )
R(r) = Ajl ~2
r , r < r0

with A a constant to be determined later. Note that for bound states, V0 < E <
0.
Bound states bring the novelty as compared to the vacuum case that E is now
negative (in the vacuum it was to be positive). This, along with third con-
straint, selects Hankel function of the first kind as the only converging solution
A
at infinity (the singularity at the origin of these functions does not matter since
we are now outside the sphere):
q 
(1)
R(r) = Bhl i −2m ~ 2
0E
r , r > r0

Second constraint on continuity of ψ at r = r0 along with normalization al-


lows the determination of constants A and B. Continuity of the derivative (or
logarithmic derivative for convenience) requires quantization of energy.
DR
Infinite spherical square well
In case where the potential is infinitely deep, so that we can take V0 = 0
inside the sphere and ∞ outside, the problem becomes that of matching the
wavefunction inside the sphere (the spherical Bessel functions) with identically
zero wavefunction outside the sphere. Allowed energies are those for which
the radial wavefunction vanishes at the boundary. Thus, we use the zeros of
the spherical Bessel functions to find the energy spectrum and wavefunctions.
Calling ul,k the k th zero of jl , we have:
u2l,k ~2
El = 2m0 r02

So that one is reduced to the computations of these zeros ul,k and to their
ordering them (as illustrated graphically below) (note that zeros of j are the
same as those of J ).

Particle in a spherically symmetric potential


348

Zeros of the first spherical Bessel equations


Calling s, p, d, f, g, h, etc., states with l =0, 1, 2, 3, 4, 5, etc., respectively, we
obtain the following spectrum:

FT
Spectrum of the infinitely deep spherical square well

Source: http://en.wikipedia.org/wiki/Particle_in_a_spherically_symmetric_potential

Principal Authors: Laussy, Oleg Alexandrov, Charles Matthews, Fibonacci, Starwed

Path integral formulation

This article is about a formulation of quantum mechanics. For integrals


along a path, also known as line or contour integrals, see Line integral.
A
The path integral formulation of quantum mechanics was developed in 1948
by Richard Feynman. Some preliminaries were worked out earlier, in the
course of his doctoral thesis work with John Archibald Wheeler. It is a de-
scription of quantum theory which generalizes the action principle of classical
mechanics. It replaces the classical notion of a single, unique history for a sys-
tem with a sum, or functional integral, over an infinity of possible histories to
compute a quantum amplitude.
DR
This formulation has proved crucial to the subsequent development of theoret-
ical physics, since it provided the basis for the grand synthesis of the 1970’s
called the renormalization group which unified quantum field theory with sta-
tistical mechanics. It is no surprise, therefore, that path integrals have also
been used in the study of Brownian motion and diffusion.

Formulating quantum mechanics


The path integral method is an alternative formulation of quantum mechan-
ics. The canonical approach, pioneered by Schrödinger, Heisenberg and Paul
Dirac paid great attention to wave-particle duality and the resulting uncertainty
principle by replacing Poisson brackets of classical mechanics by commutators
between operators in quantum mechanics. The →Hilbert space of quantum
states and the superposition law of quantum amplitudes follows. The path in-
tegral starts from the superposition law, and exploits wave-particle duality to
build a generating function for quantum amplitudes.

Path integral formulation


349

A FT
Figure 32 These are just three of the paths that contribute to the quantum amplitude for
a particle moving from point A at some time t 0 to point B at some other time t 1.

Quantum amplitudes
Feynman proposed the following postulates:
DR
1. The probability for any fundamental event is given by the absolute
square of a complex amplitude.

2. The amplitude for some event is given by adding together all the histo-
ries which include that event.

3. R The amplitude a certain history contributes is proportional to


i
e ~ [L(q,q̇,t)]dt , where [L(q, q̇, t)]dt is the action of that history, or time
R

integral of the Lagrangian.

In order to find the overall probability amplitude for a given process, then,
one adds up, or integrates, the amplitude of postulate 3 over the space of all
possible histories of the system in between the initial and final states, including
histories that are absurd by classical standards. In calculating the amplitude

Path integral formulation


350

for a single particle to go from one place to another in a given time, it would be
correct to include histories in which the particle describes elaborate curlicues,
histories in which the particle shoots off into outer space and flies back again,
and so forth. The path integral includes them all. Not only that, it assigns

FT
all of them, no matter how bizarre, amplitudes of equal magnitude; only the
phase, or argument of the complex number, varies. The contributions wildly
different from the classical history are suppressed only by the interference of
similar histories (see below).
Feynman showed that his formulation of quantum mechanics is equivalent to
the canonical approach to quantum mechanics. An amplitude computed ac-
cording to Feynman’s principles will also obey the →Schrödinger equation for
the Hamiltonian corresponding to the given action.
Feynman’s postulates are somewhat ambiguous in that they do not define what
an "event" is or the exact proportionality constant in postulate 3. The pro-
portionality problem can be solved by simply normalizing the path integral by
dividing the amplitude by the square root of the total probability for something
to happen (resulting in that the total probability given by all the normalized
amplitudes will be 1, as we would expect). Generally speaking one can simply
A
define the "events" in an operational sense for any given experiment.
The equal magnitude of all amplitudes in the path integral tends to make it
difficult to define it such that it converges and is mathematically tractable.
For purposes of actual evaluation of quantities using path-integral methods, it
is common to give the action an imaginary part in order to damp the wilder
contributions to the integral, then take the limit of a real action at the end of
the calculation. In quantum field theory this takes the form of Wick rotation.
DR
There is some difficulty in defining a measure over the space of paths. In
particular, the measure is concentrated on "fractal-like" distributional paths.

Recovering the action principle


Feynman was initially attempting to make sense of a brief remark by Paul Dirac
about the quantum equivalent of the action principle in classical mechanics. In
the limit of action that is large compared to Planck’s constant ~, the path inte-
gral is dominated by solutions which are stationary points of the action, since
there the amplitudes of similar histories will tend to constructively interfere
with one another. Conversely, for paths that are far from being stationary points
of the action, the complex phase of the amplitude calculated according to pos-
tulate 3 will vary rapidly for similar paths, and amplitudes will tend to cancel.
Therefore the important parts of the integral—the significant possibilities—in
the limit of large action simply consist of solutions of the Euler-Lagrange equa-
tion, and classical mechanics is correctly recovered.

Path integral formulation


351

Action principles can seem puzzling to the student of physics because of their
seemingly teleological quality: instead of predicting the future from initial con-
ditions, one starts with a combination of initial conditions and final conditions
and then finds the path in between, as if the system somehow knows where

FT
it’s going to go. The path integral is one way of understanding why this works.
The system doesn’t have to know in advance where it’s going; the path inte-
gral simply calculates the probability amplitude for a given process, and the
stationary points of the action mark neighborhoods of the space of histories for
which quantum-mechanical interference will yield large probabilities.

Time Slicing Definition


For a particle in a smooth potential, the path integral is approximated by Feyn-
man as the small-step limit over zig-zag paths, which in one dimension is a
product of ordinary integrals. For the motion of the particle from position x0 at
time 0 to xn at time t, the time interval can be divided up into little segments
of fixed duration ∆t. This process is called time slicing. The path integral can
be computed as proportional to
R +∞ R +∞ R +∞ R +∞ i
R
(H(x1 ,...,xj ,t)dt)
lim dx dx dx . . . dx e
A ∆t→0,n→∞,n∆t=t −∞ 1 −∞ 2 −∞ 3 −∞ n−1 ~

where H is the entire history in which the particle zigzags from its initial to its
final position linearly between all the values of
xj = x(j∆t).

In the limit of ∆t going to zero, this becomes a functional integral. This limit
DR
does not, however, exist for the most important quantum-mechanical systems,
the atoms, due to the singularity of the Coulomb potential e2 /r at the origin.
The problem was solved in 1979 by Duru and Kleinert (see here 173 and here 174)
by choosing ∆t proportional to r and going to new coordinates whose square
length is equal to r (→Duru-Kleinert transformation).

Particle in Curved Space


For a particle in curved space the kinetic term depends on the position and
the above time slicing cannot be applied, this being a manifestation of the
notorious operator ordering problem in Schrödinger quantum mechanics. One
may, however, solve this problem by transforming the time-sliced flat-space

173 http://www.physik.fu-berlin.de/~kleinert/kleiner_re65/65.pdf
174 http://www.physik.fu-berlin.de/~kleinert/kleiner_reb5/psfiles/pthic13.pdf

Path integral formulation


352

path integral to curved space using a multivalued coordinate transformation


(nonholonomic mapping explained here 175).

The path integral and the partition function

FT
The path integral is just the generalization of the integral above to all quantum
mechanical problems—
RT
Z = DxeiS[x]/~ where S[x] = 0 dtL[x(t)]
R

is the action of the classical problem in which one investigates the path start-
ing at time t=0 and ending at time t=T, and Dx denotes integration over all
paths. In the classical limit, ~ → 0, the path of minimum action dominates the
integral, because the phase of any path away from this fluctuates rapidly and
different contributions cancel.
The connection with statistical mechanics follows. Perform the Wick rotation
t→it, i.e., make time imaginary. Then the path integral resembles the partition
function of statistical mechanics defined in a canonical ensemble with temper-
ature 1/T ~.
AClearly, such a deep analogy between quantum mechanics and statistical me-
chanics cannot be dependent on the formulation. In the canonical formulation,
one sees that the unitary evolution operator of a state is given by
|α; ti = eiHt/~ |α; 0i

where the state α is evolved from time t=0. If one makes a Wick rotation
here, and finds the amplitude to go from any state, back to the same state in
(imaginary) time iT is given by
DR
Z = Tr[e−HT /~ ]

which is precisely the partition function of statistical mechanics for the same
system at temperature quoted earlier. One aspect of this equivalence was also
known to Schrödinger who remarked that the equation named after him looked
like the diffusion equation after Wick rotation.

Quantum field theory


Today, the most common use of the path-integral formulation is in quantum
field theory.

175 http://www.physik.fu-berlin.de/~kleinert/b5/psfiles/pthic10.pdf

Path integral formulation


353

The propagator
A common use of the path integral is to calculate hq1 , t1 |q0 , t0 i, a quantity (here
written in bra-ket notation) known as the propagator. As such it is very useful
in quantum field theory, where the propagator is an important component of

FT
Feynman diagrams. One way to do this, which Feynman used to explain photon
and electron/positron propagators in quantum electrodynamics, is to apply the
path integral to the motion of a single particle—one, however, that can roam
back and forth through time as well as space in the course of its wanderings.
(Such behavior can be reinterpreted as the contribution of the creation and
annihilation of virtual particle-antiparticle pairs, so in this sense the single-
particle restriction has already been loosened.)

Functionals of fields
However, the path-integral formulation is also extremely important in direct
application to quantum field theory, in which the "paths" or histories being con-
sidered are not the motions of a single particle, but the possible time evolutions
of a field over all space. The action is referred to technically as a functional of
the field: S[φ] where the field φ(xµ ) is itself a function of space and time, and
A
the square brackets are a reminder that the action depends on all the field’s
values everywhere, not just some particular value. In principle, one integrates
Feynman’s amplitude over the class of all possible combinations of values that
the field could have anywhere in space-time.
Much of the formal study of QFT is devoted to the properties of the resulting
functional integral, and much effort (not yet entirely successful) has been made
toward making these functional integrals mathematically precise.
DR
Such a functional integral is extremely similar to the partition function in sta-
tistical mechanics. Indeed, it is sometimes called a partition function, and the
two are essentially mathematically identical except for the factor of i in the
exponent in Feynman’s postulate 3. Analytically continuing the integral to an
imaginary time variable (called a Wick rotation) makes the functional inte-
gral even more like a statistical partition function, and also tames some of the
mathematical difficulties of working with these integrals.

Expectation values
In quantum field theory, if the action is given by the functional S of field con-
figurations (which only depends locally on the fields), then the time ordered
vacuum expectation value of polynomially bounded functional F, <F >, is giv-
en by
[φ]eiS[φ]
R
DφF
hF i = R
DφeiS[φ]

Path integral formulation


354

R
The symbol Dφ here is a concise way to represent the infinite-dimensional
integral over all possible field configurations on all of space-time. As stated
above, we put the unadorned path integral in the denominator to normalize

FT
everything properly.

Schwinger-Dyson equations
Since this formulation of quantum mechanics is analogous to classical action
principles, one might expect that identities concerning the action in classical
mechanics would have quantum counterparts derivable from a functional inte-
gral. This is often the case.
In the language of functional analysis, we can write the Euler-Lagrange equa-
δ
tions as δφ S[φ] = 0 (the left-hand side is a functional derivative; the equation
means that the action is stationary under small changes in the field configu-
ration). The quantum analogues of these equations are called the Schwinger-
Dyson equations.
If the functional measure Dφ turns out to be translationally invariant (we’ll
A
assume this for the rest of this article, although this does not hold for, let’s say
nonlinear sigma models) and if we assume that after a Wick rotation
eiS[φ] ,

which now becomes


e−H[φ]
DR
for some H, goes to zero faster than any reciprocal of any polynomial for large
values of φ, we can integrate by parts (after a Wick rotation, followed by a
Wick rotation back) to get the following Schwinger-Dyson equations:
D E D E
δ δ
δφ F [φ] = −i F [φ] δφ S[φ]

for any polynomially bounded functional F.





F,i = −i F S,i

in the deWitt notation.


These equations are the analog of the on shell EL equations.
If J (called the source field) is an element of the dual space of the field config-
urations (which has at least an affine structure because of the assumption of

Path integral formulation


355

the translational invariance for the functional measure), then, the generating
functional
R Z of the source fields is defined to be:
Z[J] = Dφei(S[φ]+hJ,φi)

FT
Note that
δn Z
δJ(x1 )···δJ(xn )
[J] = in Z[J] hφ(x1 ) · · · φ(xn )iJ

or
Z ,i1 ...in [J] = in Z[J] φi1 · · · φin J

where
[φ]ei(S[φ]+hJ,φi)
R
DφF
hF iJ = R
Dφei(S[φ]+hJ,φi)

Basically, if DφeiS[φ] is viewed as a functional distribution (this shouldn’t be


taken too literally as an interpretation of QFT, unlike its Wick rotated statistical
mechanics analogue, because we have time ordering complications here!), then
A
hφ(x1 ) · · · φ(xn )i are its moments and Z is its Fourier transform.
If F is a functional of φ, then for an operator K, F[K] is defined to be the operator
which substitutes K for φ. For example, if
∂ k1 ∂ kn
F [φ] = k φ(x1 ) · · · φ(xn )
∂x11 ∂xknn

and G is a functional of J, then


h i k1
δ ∂ kn
F −i δJ G[J] = (−i)n ∂ k1 δJ(x
δ
··· δ
G[J].
DR
) ∂x1 1 ∂xknn δJ(xn )

Then, from the properties of the functional integrals, we get the "master"
Schwinger-Dyson equation:
h i
δS δ
δφ(x)
−i δJ Z[J] + J(x)Z[J] = 0

or
S,i [−i∂]Z + Ji Z = 0

If the functional measure is not translationally invariant, it might be possible


to express it as the product M [φ] Dφ where M is a functional and Dφ is
a translationally invariant measure. This is true, for example, for nonlinear
sigma models where the target space is diffeomorphic to R n. However, if the

Path integral formulation


356

target manifold is some topologically nontrivial space, the concept of a trans-


lation does not even make any sense.
In that case, we would have to replace the S in this equation by another func-
tional Ŝ = S − i ln(M )

FT
If we expand this equation as a Taylor series about J=0, we get the entire set
of Schwinger-Dyson equations.

Functional identity
If we perform a Wick rotation inside the functional integral, professors J. Garcia
and Gerard T´Hooft showed using a functional differential equation that:
D[x]e−S[x]/~ = −A[x] ∞ n+1 δ n e−J/~
R P
n=0 (~)
where :S is the Wick-rotated classical action of the particle,J is the classical ac-
tion with an extra term "x" and delta here is the functional derivative operator
R
:A[x] = exp(1/~ dtX(t)

Ward-Takahashi identities
See main article Ward-Takahashi identity
A
Now how about the on shell Noether’s theorem for the classical case? Does it
have a quantum analog as well? Yes, but with a caveat. The functional mea-
sure would have to be invariant under the one parameter group of symmetry
transformation as well.
Let’s just assume for simplicity here that the symmetry in question is local
(not local in the sense of a gauge symmetry, but in the sense that the trans-
formed value of the field at any given point under an infinitesimal transfor-
DR
mation would only depend on the field configuration over an arbitrarily small
neighborhood of the point in question). Let’s also assume that the action is
local in the sense that it is the integral over spacetime of a Lagrangian, and
that Q[L(x)] = ∂µ f µ (x) for some function f where f only depends locally on
φ (and possibly the spacetime position).
If we don’t assume any special boundary conditions, this would not be a "true"
symmetry in the true sense of the term in general unless f=0 or something.
Here, Q is a derivation which generates the one parameter group in question.
We could have antiderivations as well, such as BRST and supersymmetry.
R
Let’s also assume DφQ[F ][φ] = 0 for any polynomially bounded functional
F. This property is called the invariance of the measure. And this does not hold
in general. See anomaly (physics) for more details.
Then,
R
DφQ[F eiS ][φ] = 0

Path integral formulation


357

, which implies

hQ[F ]i + i F ∂V f µ dsµ = 0

R

FT
where the integral is over the boundary. This is the quantum analog of
Noether’s theorem.
R d
Now, let’s assume even further that Q is a local integral Q = d xq(x)
where q(x)[φ(y)]=δ (d)(x-y)Q[φ(y)] so that q(x)[S] = ∂µ j µ (x) where

j µ (x) = f µ (x) − ∂(∂µ φ) L(x)Q[φ] (this is assuming the Lagrangian only de-
pends on φ and its first partial derivatives! More general Lagrangians would
require a modification to this definition!). Note that we’re NOT insisting that
q(x) is the generator of a symmetry (i.e. we’re NOT insisting upon the gauge
principle), but just that Q is. And let’s also assume the even stronger assump-
tion
R that the functional measure is locally invariant:
Dφq(x)[F ][φ] = 0
.
A
Then, we’d have
hq(x)[F ]i + i hF q(x)[S]i = hq(x)[F ]i + i hF ∂µ j µ (x)i = 0

Alternatively,
δ δ δ
q(x)[S][−i δJ ]Z[J] + J(x)Q[φ(x)][−i δJ ]Z[J] = ∂µ j µ (x)[−i δJ ]Z[J] +
δ
J(x)Q[φ(x)][−i δJ ]Z[J] = 0
DR
The above two equations are the Ward-Takahashi identities.
Now for the case where f=0, we can forget about all the boundary conditions
and locality assumptions. We’d simply have
hQ[F ]i = 0.

Alternatively,
δ
R d
d xJ(x)Q[φ(x)][−i δJ ]Z[J] = 0

The path integral in quantum-mechanical interpre-


tation
In one philosophical interpretation of quantum mechanics, the "sum over his-
tories" interpretation, the path integral is taken to be fundamental and reality

Path integral formulation


358

is viewed as a single indistinguishable "class" of paths which all share the same
events. For this interpretation, it is crucial to understand what exactly an event
is. The sum over histories method gives identical results to canonical quantum
mechanics, and Sinha and Sorkin (what is the reference?) claim the inter-

FT
pretation explains the Einstein-Podolsky-Rosen paradox without resorting to
nonlocality.
Some advocates of interpretations of quantum mechanics emphasizing deco-
herence have attempted to make more rigorous the notion of extracting a
classical-like "coarse-grained" history from the space of all possible histories.

Suggested reading
• Feynman, R. P., and Hibbs, A. R., Quantum Physics and Path Integrals, New
York: McGraw-Hill, 1965 [ISBN 0-070-20650-3]. The historical reference,
written by the Master himself and one of his students.
• Hagen Kleinert, Path Integrals in Quantum Mechanics, Statistics, Polymer
Physics, and Financial Markets, 4th edition, World Scientific (Singapore,
2004); Paperback ISBN 981-238-107-4 (also available online: PDF-files 176)
A• Zinn Justin, Jean ; Path Integrals in Quantum Mechanics, Oxford University
Press (2004), [ISBN 0-19-856674-3]. A highly readable introduction to the
subject.
• Schulman, Larry S. ; Techniques & Applications of Path Integration, Jonh
Wiley & Sons (New York-1981) [ISBN ]. The modern reference on the sub-
ject.
• Grosche, Christian & Steiner, Frank ; Handbook of Feynman Path Integrals,
Springer Tracts in Modern Physics 145, Springer-Verlag (1998) [ISBN 3-
DR
540-57135-3]
• Ryder, Lewis H. ; Quantum Field Theory (Cambridge University Press,
1985), [ISBN 0-521-33859-X] Highly readable textbook, certainly the best
introduction to relativistic Q.F.T. for particle physics.
• Rivers, R.J. ; Path Integrals Methods in Quantum Field Theory, Cambridge
University Press (1987) [ISBN 0-521-22979-7]
• Albeverio, S. & Hoegh-Krohn. R. ; Mathematical Theory of Feynman Path
Integral, Lecture Notes in Mathematics 523, Springer-Verlag (1976) [ISBN
].
• Glimm, James, and Jaffe, Arthur, Quantum Physics: A Functional Integral
Point of View, New York: Springer-Verlag, 1981. [ISBN 0-387-90562-6].

176 http://www.physik.fu-berlin.de/~kleinert/b5

Path integral formulation


359

• Gerald W. Johnson and Michel L. Lapidus ; The Feynman Integral and


Feynman’s Operational Calculus, Oxford Mathematical Monographs, Ox-
ford University Press (2002) [ISBN 0-19-851572-3].
• Etingof, Pavel ; Geometry and Quantum Field Theory 177, M.I.T. Open-

FT
CourseWare (2002). This course, designed for mathematicians, is a rigor-
ous introduction to perturbative quantum field theory, using the language
of functional integrals.

Papers on-line
• Grosche, Christian ; An Introduction into the Feynman Path Integral, lecture
given at the graduate college Quantenfeldtheorie und deren Anwendung
in der Elementarteilchen- und Festkörperphysik, Universität Leipzig, 16-26
November 1992. Full text available at : hep-th/9302097 178.
• MacKenzie, Richard ; Path Integral Methods and Applications, lectures giv-
en at Rencontres du Vietnam: VIth Vietnam School of Physics, Vung Tau,
Vietnam, 27 December 1999 - 8 January 2000. Full text available at :
quant-ph/0004090 179.
A• DeWitt-Morette, Cécile ; Feynman’s path integral - Definition without lim-
iting procedure, Communication in Mathematical Physics 28(1) (1972) pp.
47–67. Full text available at : Euclide Project 180.
• Cartier, Pierre & DeWitt-Morette, Cécile ; A new perspective on Functional
Integration, Journal of Mathematical Physics 36 (1995) pp. 2137-2340.
Full text available at : funct-an/9602005 181.
DR
Source: http://en.wikipedia.org/wiki/Path_integral_formulation

Principal Authors: Matt McIrvin, Karl-H, AcidFlask, Bambaiah, Charles Matthews

177 http://ocw.mit.edu/OcwWeb/Mathematics/18-238Fall2002/CourseHome/index.htm
178 http://arxiv.org/abs/hep-th/9302097
179 http://arxiv.org/abs/quant-ph/0004090
180 http://projecteuclid.org/Dienst/UI/1.0/Summarize/euclid.cmp/1103858329
181 http://fr.arxiv.org/abs/funct-an/9602005

Path integral formulation


360

Penrose Interpretation

A FT
Figure 33 Diagram illustrating a less complex version of the ex-
periment to test penrose Interpretation
DR
Penrose Interpretation is an interpretation of quantum mechanics formulated
by Sir Roger Penrose. This theory is a possible step toward Quantum gravity, as
it describes quantum mechanics in terms of General relativity. It states that a
system requires energy to allow it to exist in more than one location. A macro-
scopic system, such as a human, connot exist in more than one position because
its significant gravitational field requires it to have a large amount of energy to
sustain, and will settle into one position within a trilionth of a second. How-
ever a microscopic system or particle (such as an electron) has an insignificant
gravitational field, and therefore requires so little energy that it could exist in
more than one location almost indefinitly; This is called superposition.
In Einstein’s theory, any object that has mass causes a warp in the structure
of space and time round it. This warping produces the effect we experi-
ence as gravity. Penrose points out that tiny objects-dust specks, atoms,
electrons-produce space-time warps as well. Ignoring these warps is where

Penrose Interpretation
361

most physicists go awry, he believes. If a dust speck is in two locations at


the same time, each one should create its own distortions in space-time,
yielding two superposed gravitational fields. According to Penrose’s theory,
it takes energy to sustain these dual fields. The stability of a system depends

FT
on the amount of energy involved: The higher the energy required to sus-
tain a system, the less stable it is. Over time, an unstable system tends to
settle back to its simplest, lowest-energy state-in this case, one object in one
location producing one gravitational field. If Penrose is right, gravity yanks
objects back into a single location, without any need to invoke observers or
parallel universes.
Penrose believes that the transition between macroscopic and quantum begins
on the scale of dust particles, that could exist in more than one location for as
long as one second (a very long time compared to larger objects). An experi-
ment has been developed to test this theory, in which a X-ray laser in space is
directed toward a tiny miror , and fissioned by a beam spliter from thousands
of miles away, in which the photons are directed toward to other mirrors and
reflected back. According to modern physics one photon will stike the tiny
mirror moving en route to another mirror and move the tiny mirror back as it
Areturns, so the tiny mirror exists in two location at one time. If gravity effects
the mirror, it will be unable to exist in two locations at once because gravity
holds it in place.

See also
• Interpretations of Quantum Mechanics
DR
References
• Folger, Tim. "If an Electron Can Be in 2 Places at Once, Why Can’t I?"
Discover. Vol. 25 No. 6 (June 2005). 32.

External links
• Molecules - Quantum Interpretations 182
• QM- the Penrose Interpretation 183
• Roger Penrose discusses his experiment on the BBC (25 minutes in) 184

182 http://universe-review.ca/F12-molecule.htm#interpretations
183 http://sci4um.com/about2884.html
184 http://www.bbc.co.uk/radio4/history/inourtime/inourtime_20020502.shtml

Penrose Interpretation
362

Source: http://en.wikipedia.org/wiki/Penrose_Interpretation

Peres-Horodecki criterion

FT
The Peres-Horodecki criterion is a necessary condition, for the joint density
matrix ρ of two systems A and B, to be separable. It is also called the PPT
criterion, for Positive Partial Transpose. In the 2x2 and 2x3 dimensional cases
the condition is also sufficient.
The criterion reads:
If ρ is separable, then the partial transpose
σm µ n ν := ρn µ m ν

of ρ, taken in some basis |miA ⊗ |µiB , has non negative eigenvalues.


In matrix notation, if we write a N X M mixed state ρ as block matrix:
 
ρ11 ··· ρ1n
A
 ..
ρ= .
ρn1
..
···
.
.. 
. 
ρnn

,where each ρi:j is M X M and n runs from 1 to N. The partial transpose of ρ is


then given by
ρT11 ρT1n
 
···
 .. .. 
ρP T = . ..
.
DR
. 
ρTn1 ··· ρTnn

So ρ is PPT if (I ⊗ T )(ρ) is positive, where T is the transposition map on


matrices.
That necessity of PPT for separability follows immediately from the fact that
if ρ is separable, then (I ⊗ Φ)(ρ) must be positive for all positive map Φ. The
transposition map is clearly a positive map.
Showing that being PPT is also sufficient for in the 2 X 2 and 2 X 3 (therefore
3 X 2) cases is more involved. It was shown by the Horodecki’s that for every
entangled state there exists an entanglement witness. This is a result of ge-
ometric nature and invokes the Hahn-Banach theorem (see reference below).
From the existence of entanglement witnesses, one can show that (I ⊗ Φ)(ρ)
being positive for all positive map Φ is not only necessary but also sufficient for

Peres-Horodecki criterion
363

separability of ρ. Furthermore, every positive map from the C*-algebra of 2 × 2


matrices to 2 × 2 or 3 × 3 matrices can be decomposed into a sum of completely
positive and completely copositive maps. In other words, every such map can
be written as

FT
Λ = Λ1 + Λ 2 ◦ T

,where Λ1 and Λ2 are completely positive and T is the transposition map. Com-
bining the above two facts, we can conclude PPT is also sufficient for separa-
bility in the 2 X 2 and 2 X 3 cases.
Due to the existence of non-decomposable positive maps in higher dimensions,
PPT is no longer sufficient in higher dimensions. In higher dimension, there
are entangled states which are PPT. Such states have some interesting propeties
including the fact that thay are bound entangled, i.e. they can not be distilled
for quantum communication purposes.

References


A Asher Peres, Separability Criterion for Density Matrices, Phys. Rev. Lett.
77, 1413–1415 (1996)
M. Horodecki, P. Horodecki, R. Horodecki, Separability of Mixed States:
Necessary and Sufficient Conditions, Physics Letters A 210, 1996.

Source: http://en.wikipedia.org/wiki/Peres-Horodecki_criterion

Principal Authors: Mct mht, Tinissimo, Stevey7788, Matthew Mattic, Charles Matthews
DR
Perturbation theory (quantum mechanics)

In quantum mechanics, perturbation theory is a set of approximation schemes


directly related to mathematical perturbation for describing a complicated
quantum system in terms of a simpler one. The idea is to start with a simple
system and gradually turn on an additional "perturbing" Hamiltonian repre-
senting a weak disturbance to the system. If the disturbance is not too large,
the various physical quantities associated with the perturbed system (e.g. its
energy levels and eigenstates) will be continuously generated from those of the
simple system. We can therefore study the former based on our knowledge of
the latter.

Perturbation theory (quantum mechanics)


364

Applications of perturbation theory


Perturbation theory is an extremely important tool for describing real quan-
tum systems, as it turns out to be very difficult to find exact solutions to the
→Schrödinger equation for Hamiltonians of even moderate complexity. The

FT
Hamiltonians to which we know exact solutions, such as the hydrogen atom,
the quantum harmonic oscillator and the particle in a box, are too idealized to
adequately describe most systems. Using perturbation theory, we can use the
known solutions of these simple Hamiltonians to generate solutions for a range
of more complicated systems. For example, by adding a perturbative electric
potential to the quantum mechanical model of the hydrogen atom, we can cal-
culate the tiny shifts in the spectral lines of hydrogen caused by the presence
of an electric field (the →Stark effect). This is only approximate because the
sum of a Coulomb potential with a linear potential is unstable although the
tunneling time (decay rate) is very long. This shows up as a broadening of
the energy spectrum lines, something which perturbation theory fails to notice
entirely.
The expressions produced by perturbation theory are not exact, but they can
A
lead to accurate results as long as the expansion parameter, say α, is very
small. Typically, the results are expressed in terms of finite power series in
α that seem to converge to the exact values when summed to higher order.
After a certain order n ∼ 1/α, however, the results become increasingly worse
since the series are usually divergent, being asymptotic series). There exist
ways to convert them into convergent series, which can be evalauted for large-
expansion parameters, most efficiently by variational perturbation theory.
In the theory of quantum electrodynamics (QED), in which the electron-photon
DR
interaction is treated perturbatively, the calculation of the electron’s magnetic
moment has been found to agree with experiment to eleven decimal places. In
QED and other quantum field theories, special calculation techniques known
as Feynman diagrams are used to systematically sum the power series terms.
Under some circumstances, perturbation theory is an invalid approach to take.
This happens when the system we wish to describe cannot be described by
a small perturbation imposed on some simple system. In quantum chromo-
dynamics, for instance, the interaction of quarks with the gluon field cannot
be treated perturbatively at low energies because the coupling constant (the
expansion parameter) becomes too large. Perturbation theory also fails to de-
scribe states that are not generated adiabatically from the "free model", includ-
ing bound states and various collective phenomena such as solitons. Imagine,
for example, that we have a system of free (i.e. non-interacting) particles, to
which an attractive interaction is introduced. Depending on the form of the
interaction, this may create an entirely new set of eigenstates corresponding to

Perturbation theory (quantum mechanics)


365

groups of particles bound to one another. An example of this phenomenon may


be found in conventional superconductivity, in which the phonon-mediated at-
traction between conduction electrons leads to the formation of correlated elec-
tron pairs known as Cooper pairs. When faced with such systems, one usually

FT
turns to other approximation schemes, such as the variational method and the
WKB approximation. This is because there is no analogue of a bound particle
in the unperturbed model and the energy of a soliton typically goes as the in-
verse of the expansion parameter. However, if we "integrate" over the solitonic
phenomena, the nonperturbative corrections in this case will be tiny; of the
2
order of e−1/g or e−1/g in the perturbation parameter g. Perturbation theory
can only detect solutions "close" to the unperturbed solution, even if there are
other solutions (which typically blow up as the expansion parameter goes to
zero).
The problem of non-perturbative systems has been somewhat alleviated by
the advent of modern computers. It has become practical to obtain numerical
non-perturbative solutions for certain problems, using methods such as density
functional theory. These advances have been of particular benefit to the field
of quantum chemistry. Computers have also been used to carry out perturba-
A
tion theory calculations to extraordinarily high levels of precision, which has
proven important in particle physics for generating theoretical results that can
be compared with experiment.

Time-independent perturbation theory


There are two categories of perturbation theory: time-independent and time-
dependent. In this section, we discuss time-independent perturbation theo-
ry, in which the perturbation Hamiltonian is static (i.e., possesses no time
DR
dependence.) Time-independent perturbation theory was invented by Erwin
Schrödinger in 1926, shortly after he produced his theories in wave mechan-
ics.
We begin with an unperturbed Hamiltonian H 0, which is also assumed to have
no time dependence. It has known energy levels and eigenstates, arising from
the time-independent Schrödinger equation:
(0)
H0 |n(0) i = En |n(0) i , n = 1, 2, 3, · · ·

For simplicity, we have assumed that the energies are discrete. The (0) super-
scripts denote that these quantities are associated with the unperturbed system.
We now introduce a perturbation to the Hamiltonian. Let V be a Hamiltonian
representing a weak physical disturbance, such as a potential energy produced
by an external field. (Thus, V is formally a Hermitian operator.) Let λ be a

Perturbation theory (quantum mechanics)


366

dimensionless parameter that can take on values ranging continuously from 0


(no perturbation) to 1 (the full perturbation). The perturbed Hamiltonian is
H = H0 + λV

FT
The energy levels and eigenstates of the perturbed Hamiltonian are again given
by the Schrödinger equation:
(H0 + λV ) |ni = En |ni

Our goal is to express E n and |n> in terms of the energy levels and eigenstates
of the old Hamiltonian. If the perturbation is sufficiently weak, we can write
them as power series in :
(0) (1) (2)
En = En + λEn + λ2 En + · · ·

|ni = |n(0) i + λ|n(1) i + λ2 |n(2) i + · · ·

When = 0, these reduce to the unperturbed values, which are the first term in
A
each series. Since the perturbation is weak, the energy levels and eigenstates
should not deviate too much from their unperturbed values, and the terms
should rapidly become smaller as we go to higher order.
Plugging the power series into the Schrödinger equation, we obtain
(0) (1)

(H0 +
 λV ) |n i + λ|n i + · · · 
(0) (1) (2)
= En + λEn + λ2 En + · · · |n(0) i + λ|n(1) i + · · ·

DR
Expanding this equation and comparing coefficients of each power of results
in an infinite series of simultaneous equations. The zeroth-order equation is
simply the Schrödinger equation for the unperturbed system. The first-order
equation is
(0) (1)
H0 |n(1) i + V |n(0) i = En |n(1) i + En |n(0) i

Multiply through by <n (0)|. The first term on the left-hand side cancels with
the first term on the right-hand side. (Recall, the unperturbed Hamiltonian is
hermitian). This leads to the first-order energy shift:
(1)
En = hn(0) |V |n(0) i

This is simply the expected value of the perturbation Hamiltonian while the
system is in the unperturbed state. This result can be interpreted in the fol-
lowing way: suppose the perturbation is applied, but we keep the system in

Perturbation theory (quantum mechanics)


367

the quantum state |n (0)>, which is a valid quantum state though no longer
an energy eigenstate. The perturbation causes the average energy of this state
to increase by <n (0)|V |n (0)>. However, the true energy shift is slightly differ-
ent, because the perturbed eigenstate is not exactly the same as |n (0)>. These

FT
further shifts are given by the second and higher order deviations.
To obtain the first-order deviation in the energy eigenstate, we insert our ex-
pression for the first-order energy shift back into the above equation between
the first-order coefficients of . We then make use of the resolution of the iden-
tity,
P 
V |n(0) i = (0) (0)
k |k ihk | V |n i
(0)

The result is
   
(0)
En − H0 |n(1) i = k6=n hk (0) |V |n(0) i |k (0) i
P

For the moment, suppose that this energy level is not degenerate, i.e. there is
no other eigenstate with the same energy. The operator on the left hand side
therefore has a well-defined inverse, and we get
A hk (0) |V |n(0) i
|n(1) i = k6=n (0) (0) |k (0) i
P
En −Ek

The first-order change in the n-th energy eigenket has a contribution from each
of the energy eigenstates k 6= n. Each term is proportional to the matrix ele-
ment <k (0)|V |n (0)>, which is a measure of how much the perturbation mixes
eigenstate n with eigenstate k ; it is also inversely proportional to the ener-
gy difference between eigenstates k and n, which means that the perturbation
DR
deforms the eigenstate to a greater extent if there are more eigenstates at near-
by energies. We see also that the expression is singular if any of these states
have the same energy as state n, which is why we assumed that there is no
degeneracy.
We can find the higher-order deviations by a similar procedure, though the
calculations become quite tedious with our current formulation. For example,
the second-order energy shift is
(2) P |hk (0) |V |n(0) i|2
En = k6=n (0) (0)
En −Ek

Effects of degeneracy
Suppose that two or more energy eigenstates are degenerate. Our above cal-
culation for the first-order energy shift is unaffected, but the calculation of the
change in the eigenstate is problematic because the operator

Perturbation theory (quantum mechanics)


368

(0)
En − H 0

does not have a well-defined inverse.


This is actually a conceptual, rather than mathematical, problem. Imagine

FT
that we have two or more perturbed eigenstates with different energies, which
are continuously generated from an equal number of unperturbed eigenstates
that are degenerate. Let D denote the subspace spanned by these degenerate
eigenstates. The problem lies in the fact that there is no unique way to choose a
basis of energy eigenstates for the unperturbed system. In particular, we could
construct a different basis for D by choosing different linear combinations of
the spanning eigenstates. In such a basis, the unperturbed eigenstates would
not continuously generate the perturbed eigenstates.
We thus see that, in the presence of degeneracy, perturbation theory does not
work with an arbitrary choice of energy basis. We must instead choose a basis
so that the perturbation Hamiltonian is diagonal in the degenerate subspace D.
In other words,
V |k (0) i = k |k (0) i + (terms not in D) ∀ |k (0) i ∈ D
A
In that case, our equation for the first-order deviation in the energy eigenstate
reduces to
   
(0)
En − H0 |n(1) i = k6∈D hk (0) |V |n(0) i |k (0) i
P

The operator on the left hand side is not singular when applied to eigenstates
outside D, so we can write
DR
hk (0) |V |n(0) i
|n(1) i = k6∈D (0) (0) |k (0) i
P
En −Ek

Time-dependent perturbation theory


Time-dependent perturbation theory, developed by Paul Dirac, studies the ef-
fect of a time-dependent perturbation V(t) applied to a time-independent
Hamiltonian H 0. Since the perturbed Hamiltonian is time-dependent, so are its
energy levels and eigenstates. Therefore, the goals of time-dependent pertur-
bation theory are slightly different from time-independent perturbation theory.
We are interested in the following quantities:

• The time-dependent expected value of some observable A, for a given initial


state.
• The time-dependent amplitudes of those quantum states that are energy
eigenkets (eigenvectors) in the unperturbed system.

Perturbation theory (quantum mechanics)


369

The first quantity is important because it gives rise to the classical result of an A
measurement performed on a macroscopic number of copies of the perturbed
system. For example, we could take A to be the displacement in the x-direction
of the electron in a hydrogen atom, in which case the expected value, when

FT
multiplied by an appropriate coefficient, gives the time-dependent electrical
polarization of a hydrogen gas. With an appropriate choice of perturbation (i.e.
an oscillating electric potential), this allows us to calculate the AC permittivity
of the gas.
The second quantity looks at the time-dependent probability of occupation for
each eigenstate. This is particularly useful in laser physics, where one is in-
terested in the populations of different atomic states in a gas when a time-
dependent electric field is applied. These probabilities are also useful for cal-
culating the "quantum broadening" of spectral lines (see line broadening).
We will briefly examine the ideas behind Dirac’s formulation of time-dependent
perturbation theory. Choose an energy basis {|n>} for the unperturbed system.
(We will drop the (0) superscripts for the eigenstates, because it is not mean-
ingful to speak of energy levels and eigenstates for the perturbed system.)
A
If the unperturbed system is in eigenstate |j > at time t = 0, its state at subse-
quent times varies only by a phase (we are following the →Schrödinger picture,
where state vectors evolve in time and operators are constant):
|j(t)i = e−iEj t/~ |ji

We now introduce a time-dependent perturbing Hamiltonian V(t). The Hamil-


tonian of the perturbed system is
H = H0 + V (t)
DR
Let |ψ(t) > denote the quantum state of the perturbed system at time t. It obeys
the time-dependent Schrödinger equation,

H|ψ(t)i = i~ ∂t |ψ(t)i

The quantum state at each instant can be expressed as a linear combination of


the basis {|n>}. We can write the linear combination as
|ψ(t)i = n cn (t)e−iEn t/~ |ni
P

where the c n(t) s are undetermined complex functions of t which we will re-
fer to as amplitudes (strictly speaking, they are the amplitudes in the Dirac
picture.) We have explicitly extracted the exponential phase factors exp(-
iE nt/<strike>h</strike>) on the right hand side. This is only a matter of

Perturbation theory (quantum mechanics)


370

convention, and may be done without loss of generality. The reason we go to


this trouble is that when the system starts in the state |j > and no perturbation
is present, the amplitudes have the convenient property that, for all t, c j(t) =
1 and c n(t) = 0 if n6=j.

FT
The absolute square of the amplitude c n(t) is the probability that the system is
in state n at time t, since
|cn (t)|2 = |hn|ψ(t)i|2

Plugging into the Schrödinger equation and using the fact that ∂/∂t acts by a
chain rule, we obtain
P  ∂cn 
−iEn t/~ |ni = 0
n i~ ∂t − cn (t)V (t) e

By resolving the identity in front of V, this can be reduced to a set of partial


differential equations for the amplitudes:
∂cn −i P −i(Ek −En )t/~
∂t = ~ k hn|V (t)|ki ck (t) e
A
The matrix elements of V play a similar role as in time-independent pertur-
bation theory, being proportional to the rate at which amplitudes are shifted
between states. Note, however, that the direction of the shift is modified by the
exponential phase factor. Over times much longer than the energy difference
E k-E n, the phase winds many times. If the time-dependence of V is sufficient-
ly slow, this may cause the state amplitudes to oscillate. Such oscillations are
useful for managing radiative transitions in a laser.
Up to this point, we have made no approximations, so this set of differential
DR
equations is exact. By supplying appropriate initial values c n(0), we could in
principle find an exact (i.e. non-perturbative) solution. This is easily done
when there are only two energy levels (n = 1, 2), and the solution is useful for
modelling systems like the ammonia molecule. However, exact solutions are
difficult to find when there are many energy levels, and one instead looks for
perturbative solutions, which may be obtained by putting the equations in an
integral form:
cn (t) = cn (0) + −i
P Rt 0 0 0 −i(Ek −En )t0 /~
~ k 0 dt hn|V (t )|ki ck (t ) e

By repeatedly substituting this expression for c n back into right hand side, we
get an iterative solution
(0) (1) (2)
cn (t) = cn + cn + cn + · · ·

where, for example, the first-order term is

Perturbation theory (quantum mechanics)


371

(1) −i P Rt 0
cn (t) = ~ k 0 dt0 hn|V (t0 )|ki ck (0) e−i(Ek −En )t /~

Many further results may be obtained, such as →Fermi’s golden rule, which
relates the rate of transitions between quantum states to the density of states

FT
at particular energies, and the Dyson series, obtained by applying the iterative
method to the time evolution operator, which is one of the starting points for
the method of Feynman diagrams.

Source: http://en.wikipedia.org/wiki/Perturbation_theory_%28quantum_mechanics%29

Principal Authors: CYD, Phys, A. Wilson, Karol Langner, Sigfpe

Photoelectric effect

The photoelectric effect is the emission of electrons from matter upon the
absorption of electromagnetic radiation, such as ultraviolet radiation or x-rays.
A An older term for the photoelectric effect was the Hertz effect, though this
phrase has fallen out of current use. 185

Introduction
Upon exposing a metallic surface to electromagnetic radiation that is above the
threshold frequency (which is specific to the type of surface and material), the
photons are absorbed and current is produced. No electrons are emitted for
radiation with a frequency below that of the threshold, as the electrons are
DR
unable to gain sufficient energy to overcome the electrostatic barrier presented
by the termination of the crystalline surface (the material’s work function). In
1905 it was known that the energy of the photoelectrons increased with in-
creasing frequency of incident light, but the manner of the increase was not
experimentally determined to be linear until 1915 when Robert Andrews Mil-
likan showed that Einstein was correct [3].
By conservation of energy, the energy of the photon is absorbed by the electron
and, if sufficient, the electron can escape from the material with a finite kinetic
energy. A single photon can only eject a single electron, as the energy of one
photon may only be absorbed by one electron. The electrons that are emitted
are often termed photoelectrons.
The photoelectric effect helped further wave-particle duality, whereby physical
systems (such as photons, in this case) display both wave-like and particle-like

185
http://scienceworld.wolfram.com/physics/HertzEffect.html

Photoelectric effect
372

FT
Figure 34 The photoelectric effect. Incoming EM radiation on the left ejects electrons, depicted
as flying off to the right, from a substance.
A
properties and behaviours, a concept that was used by the creators of quantum
mechanics. The photoelectric effect was explained mathematically by Albert
Einstein, who extended the work on quanta developed by Max Planck.

Explanation
The photons of the light beam have a characteristic energy given by the wave-
length of the light. In the photoemission process, if an electron absorbs the
DR
energy of one photon and has more energy than the work function, it is ejected
from the material. If the photon energy is too low, however, the electron is
unable to escape the surface of the material. Increasing the intensity of the
light beam does not change the energy of the constituent photons, only their
number, and thus the energy of the emitted electrons does not depend on the
intensity of the incoming light.
Electrons can absorb energy from photons when irradiated, but they follow an
"all or nothing" principle. All of the energy from one photon must be absorbed
and used to liberate one electron from atomic binding, or the energy is re-
emitted. If the photon is absorbed, some of the energy is used to liberate
it from the atom, and the rest contributes to the electron’s kinetic (moving)
energy as a free particle.

Photoelectric effect
373

Equations
In analysing the photoelectric effect quantitatively using Einstein’s method, the
following equivalent equations are used:
Energy of photon = Energy needed to remove an electron + Kinetic energy of

FT
the emitted electron
Algebraically:
hf = φ + Ekmax

where

• h is Planck’s constant,
• f is the frequency of the incident photon,
• φ = hf0 is the work function, or minimum energy required to remove an
electron from atomic binding,
• Ekmax = 12 mvm 2 is the maximum kinetic energy of ejected electrons,

• f 0 is the threshold frequency for the photoelectric effect to occur,


• m is the rest mass of the ejected electron, and
A
• vm is the velocity of the ejected electron.

Note: If the photon’s energy (hf ) is not greater than the work function (φ), no
electron will be emitted. The work function is sometimes denoted W .

History
Early observations
DR
In 1839, Alexandre Edmond Becquerel observed the photoelectric effect via an
electrode in a conductive solution exposed to light. In 1873, Willoughby Smith
found that selenium is photoconductive.

Hertz’s spark gaps


Heinrich Hertz, in 1887, made observations of the photoelectric effect and of
the production and reception of electromagnetic (EM) waves, published in the
journal Annalen der Physik. His receiver consisted of a coil with a spark gap,
whereupon a spark would be seen upon detection of EM waves. He placed
the apparatus in a darkened box in order to see the spark better; he observed,
however, that the maximum spark length was reduced when in the box. A glass
panel placed between the source of EM waves and the receiver absorbed ultra-
violet radiation that assisted the electrons in jumping across the gap. When
removed, the spark length would increase. He observed no decrease in spark

Photoelectric effect
374

length when he substituted quartz for glass, as quartz does not absorb UV ra-
diation.
Hertz concluded his months of investigation and reported the results obtained.
He did not further pursue investigation of this effect, nor did he make any

FT
attempt at explaining how the observed phenomenon was brought about.

JJ Thomson: electrons
In 1899, Joseph John Thomson investigated ultraviolet light in Crookes tubes.
Influenced by the work of James Clerk Maxwell, Thomson deduced that cath-
ode rays consisted of negatively charged particles, later called electrons, which
he called "corpuscles". In the research, Thomson enclosed a metal plate (a
cathode) in a vacuum tube, and exposed it to high frequency radiation. It was
thought that the oscillating electromagnetic fields caused the atoms’ field to
resonate and, after reaching a certain amplitude, caused a subatomic "corpus-
cle" to be emitted, and current to be detected. The amount of this current
varied with the intensity and color of the radiation. Larger radiation intensity
or frequency would produce more current.
ATesla’s radiant energy
On November 5 1901, Nikola Tesla received the U.S. Patent 685957 186 (Appa-
ratus for the Utilization of Radiant Energy) that describes radiation charging
and discharging conductors by "radiant energy". Tesla used this effect to charge
a capacitor with energy by means of a conductive plate. The patent specified
that the radiation included many different forms.

Von Lenard’s observations


DR
In 1902, Philipp von Lenard observed 187 the variation in electron energy with
light frequency. He used a powerful electric arc lamp which enabled him to
investigate large changes in intensity, and had sufficient power to enable him
to investigate the variation of potential with light frequency. His experiment
directly measured potentials, not electron kinetic energy: he found the elec-
tron energy by relating it to the maximum stopping potential (voltage) in a
phototube. He found that the calculated maximum electron kinetic energy is
determined by the frequency of the light. For example, an increase in frequency
results in an increase in the maximum kinetic energy calculated for an electron
upon liberation - ultraviolet radiation would require a higher applied stopping
potential to stop current in a phototube than blue light. However Lenard’s
results were qualitative rather than quantitative because of the difficulty in

186 http://patft.uspto.gov/netacgi/nph-Parser?patentnumber=685957
187 http://www.phys.virginia.edu/classes/252/photoelectric_effect.html

Photoelectric effect
375

performing the experiments: the experiments needed to be done on freshly cut


metal so that the pure metal was observed, but it oxidised in a matter of min-
utes even in the partial vacuums he used. The current emitted by the surface
was determined by the light’s intensity, or brightness: doubling the intensity

FT
of the light doubled the number of electrons emitted from the surface. Lenard
did not know of photons.

Einstein: light quanta


Albert Einstein’s mathematical description in 1905 of how it was caused by
absorption of what were later called photons, or quanta of light, in the inter-
action of light with the electrons in the substance, was contained in the paper
named "On a Heuristic Viewpoint Concerning the Production and Transforma-
tion of Light". This paper proposed the simple description of "light quanta"
(later called "photons") and showed how they could be used to explain such
phenomena as the photoelectric effect. The simple explanation by Einstein
in terms of absorption of single quanta of light explained the features of the
phenomenon and helped explain the characteristic frequency. Einstein’s expla-
nation of the photoelectric effect won him the Nobel Prize of 1921.
A
The idea of light quanta was motivated by Max Planck’s published law of black-
body radiation ("On the Law of Distribution of Energy in the Normal Spec-
trum". Annalen der Physik 4 (1901)) by assuming that Hertzian oscillators
could only exist at energies E proportional to the frequency f of the oscilla-
tor by E = hf, where h is Planck’s constant. Einstein, by assuming that light
actually consisted of discrete energy packets, wrote an equation for the photo-
electric effect that fit experiments. This was an enormous theoretical leap and
the reality of the light quanta was strongly resisted. The idea of light quan-
DR
ta contradicted the wave theory of light that followed naturally from James
Clerk Maxwell’s equations for electromagnetic behavior and, more generally,
the assumption of infinite divisibility of energy in physical systems. Even after
experiments showed that Einstein’s equations for the photoelectric effect were
accurate there was resistance to the idea of photons, since it appeared to con-
tradict Maxwell’s equations, which were believed to be well understood and
well verified.
Einstein’s work predicted that the energy of the ejected electrons would in-
crease linearly with the frequency of the light. Perhaps surprisingly, that had
not yet been tested. In 1905 it was known that the energy of the photoelec-
trons increased with increasing frequency of incident light, but the manner of
the increase was not experimentally determined to be linear until 1915 when
Robert Andrews Millikan showed that Einstein was correct 188.

Photoelectric effect
376

Effect on wave-particle question


The photoelectric effect helped propel the then-emerging concept of the dual
nature of light, that light exhibits characteristics of waves and particles at dif-
ferent times. The effect was impossible to understand in terms of the classical

FT
wave description of light, as the energy of the emitted electrons did not depend
on the intensity of the incident radiation. Classical theory predicted that the
electrons could ’gather up’ energy over a period of time, and then be emitted.
For such a classical theory to work a pre-loaded state would need to persist
in matter. The idea of the pre-loaded state was discussed in Millikan’s book
Electrons (+ & -) and in Compton and Allison’s book X-Rays in Theory and
Experiment. These ideas were abandoned.

Uses and effects


Solar cells (used in solar power) and light-sensitive diodes use the photoelec-
tric effect. They absorb photons from light and put the energy into electrons,
creating electric current.

Electroscopes
AElectroscopes are fork-shaped, hinged metallic leaves placed in a vacuum jar,
partially exposed to the outside environment. When an electroscope is charged
positively or negatively, the two leaves separate, as charge distributes evenly
along the leaves causing repulsion between two like poles. When ultraviolet
radiation (or any radiation above threshold frequency) shines onto the metallic
outside of the electroscope, a negatively charged scope will discharge and the
leaves will collapse, while nothing will happen to a positively charged scope
DR
(besides charge decay). The reason is that electrons will be liberated from the
negatively charged one, gradually making it neutral, while liberating electrons
from the positively charged one will make it even more positive, keeping the
leaves apart.

Photoelectron spectroscopy
Since the energy of the photoelectrons emitted is exactly the energy of the in-
cident photon minus the material’s work function or binding energy, the work
function of a sample can be determined by bombarding it with a monochro-
matic X-ray source or UV source (typically a helium discharge lamp), and mea-
suring the kinetic energy distribution of the electrons emitted.
This must be done in a high vacuum environment, since the electrons would
be scattered by air.

188 http://spiff.rit.edu/classes/phys314/lectures/photoe/photoe.html

Photoelectric effect
377

A typical electron energy analyzer is a concentric hemispherical analyser


(CHA), which uses an electric field to divert electrons different amounts de-
pending on their kinetic energies. For every element and core atomic orbital
there will be a different binding energy. The many electrons created from each

FT
will then show up as spikes in the analyzer, and can be used to determine the
elemental composition of the sample. 189

Spacecraft
The photoelectric effect will cause spacecraft exposed to sunlight to develop
a positive charge. This can get up to the tens of volts. This can be a major
problem, as other parts of the spacecraft in shadow develop a negative charge
(up to several kilovolts) from nearby plasma, and the imbalance can discharge
through delicate electrical components. The static charge created by the pho-
toelectric effect is self-limiting, though, because a more highly-charged object
gives up its electrons less easily. 190

Moon dust
Light from the sun hitting lunar dust causes it to become charged through the
Aphotoelectric effect. The charged dust then repels itself and lifts off the surface
of the Moon by electrostatic levitation. This manifests itself almost like an
"atmosphere of dust", visible as a thin haze and blurring of distant features,
and visible as a dim glow after the sun has set. This was first photographed
by the Surveyor program probes in the 1960s. It is thought that the smallest
particles are repelled up to kilometers high, and that the particles move in
"fountains" as they charge and discharge. 191 192
DR

189 Photoelectron spectroscopy(http://www.chem.qmw.ac.uk/surfaces/scc/scat5_3.htm)


190 Spacecraft charging(http://www.eas.asu.edu/~holbert/eee460/spc-chrg.html)
191 - Moon fountains(http://www.firstscience.com/site/articles/moonfountains.asp)
192 - Dust gets a charge in a vacuum(http://www.spacer.com/news/dust-00a.html)

Photoelectric effect
378

See also
Electronics: People:

• Photocurrent • Aleksandr Grigorievich Stoletov


• Photomultiplier • Albert Einstein

FT
• Solar cell • Heinrich Hertz
• Solar power • Ernest Lawrence
• Transducer • Robert Millikan
• Max Planck
Physics: • Joseph John Thomson

• Atom Lists:
• Corona discharge
• →Double-slit experiment • List of electronics topics
• Electron • List of optical topics
• Gamma ray • List of physics topics
• Nobel Prize in Physics • Timeline of solar cells
• Optical phenomenon • Scientific method list
• →Planck’s law of black body radiation • Timeline of mechanics and physics
• Photon
• →Quantum mechanics
• Radiant energy
• →Wave-particle duality
AExternal links and references
General

• Nave, R., " Wave-Particle Duality 193". HyperPhysics.


• Jpaul’s " Photovoltaics: Theory and Practice 194". Photoelectric effect 195.
• " Photoelectric effect 196". Physics 2000. University of Colorado, Boulder,
Colorado.
DR
• ACEPT W3 Group, " The Photoelectric Effect 197". Department of Physics and
Astronomy, Arizona State University, Tempe, AZ.
• Haberkern, Thomas, and N Deepak " Grains of Mystique: Quantum Physics
for the Layman 198". Einstein Demystifies Photoelectric Effect 199, Chapter 3.
• Department of Physics, " The Photoelectric effect 200". Physics 320 Laborato-
ry, Davidson College, Davidson.

193 http://hyperphysics.phy-astr.gsu.edu/hbase/mod1.html
194 http://www.students.uiuc.edu/~jpaul/theory.htm
195 http://www.students.uiuc.edu/~jpaul/photoelectric.htm
196 http://www.colorado.edu/physics/2000/quantumzone/photoelectric.html
197 http://acept.la.asu.edu/PiN/rdg/photoelectric/photoelectric.shtml
198 http://www.faqs.org/docs/qp/
199 http://www.faqs.org/docs/qp/chap03.html
200 http://www.phy.davidson.edu/ModernPhysicsLabs/hovere.html

Photoelectric effect
379

• Fowler, Michael, " The Photoelectric Effect 201". Physics 252, University of
Virginia.
• Brandl, Michael, " MISN-0-213 The Photoelectric Effect 202" (PDF file), Project
PHYSNET 203.

FT
• Quantum Chemistry I Lecute 204

Applets

• Curull, Xavi Espinal, " Photoelectric effect Applet 205". (Java)


• Fendt, Walter, and Taha Mzoughi, " The Photoelectric Effect 206". (Java)
• " Applet: Photo Effect 207". Open Source Distributed Learning Content Man-
agement and Assessment System. (Java)

Notes

Source: http://en.wikipedia.org/wiki/Photoelectric_effect

Principal Authors: Reddi, Enochlau, William M. Connolley, Heron, Omegatron


APlanck particle

A Planck particle is a hypothetical subatomic particle, defined as a tiny black


hole whose →Compton wavelength is the same as its Schwarzschild radius. Its
mass is thus (by definition) equal to the Planck mass, and its Compton wave-
length and Schwarzschild radius are equal (also by definition) to the Planck
length.
DR
See also
• Micro black hole

Source: http://en.wikipedia.org/wiki/Planck_particle

Principal Authors: Jaraalbe, Hidaspal

201 http://www.phys.virginia.edu/classes/252/photoelectric_effect.html
202 http://35.9.69.219/home/modules/pdf_modules/m213.pdf
203 http://www.physnet.org
204 http://cinarz.zdo.com/moodle/mod/resource/view.php?id=15
205 http://www.ifae.es/xec/phot2.html
206 http://www.walter-fendt.de/ph14e/photoeffect.htm
207 http://lectureonline.cl.msu.edu/~mmp/kap28/PhotoEffect/photo.htm

Planck particle
380

Planck postulate

The Planck Postulate (or Planck’s Postulate) was used by Max Planck in his

FT
derivation of his law of black body radiation. It is the postulate that the energy
of oscillators in a black body is quantised by:
E = nhν ,

where n = 1, 2, 3, ..., h is Planck’s constant, and ν is the frequency.

External links and sources


• Planck Postulate 208 — from Eric Weisstein’s World of Physics

Source: http://en.wikipedia.org/wiki/Planck_postulate
APlanck’s law of black body radiation

In physics, the spectral intensity of electromagnetic radiation from a black body


at temperature T is given by Planck’s law of black body radiation:
2hν 3 1
I(ν, T ) = c2 e kT

−1
DR
where the following table provides the definition and SI units of measure for
each symbol:

Symbol Meaning SI units of measure


I spectral radiance, J·s -1·m -2·sr -1·Hz -1
energy per unit time per unit surface area per unit solid angle per
unit frequency
ν frequency hertz
T temperature of the black body kelvin
h Planck’s constant joule per hertz
c speed of light meter per second
e base of the natural logarithm, 2.718282... dimensionless
k Boltzmann’s constant joule per kelvin

208 http://scienceworld.wolfram.com/physics/PlanckPostulate.html

Planck’s law of black body radiation


381

A
Figure 35
FT
Black body spectrum as a function of wavelength

The wavelength is related to the frequency by


λ = νc .
DR
The law is sometimes written in terms of the spectral energy density
4π 8πhν 3 1
u(ν, T ) = c I(ν, T ) = c3 hν
e kT −1

which has units of energy per unit volume per unit frequency (joule per cubic
meter per hertz).
The spectral energy density can also be expressed as a function of wavelength:
8πhc 1
u(λ, T ) = λ5 ehc/λkT −1

as shown in the derivation below.


Max Planck originally produced this law in 1900 (published in 1901) in an
attempt to improve upon an expression proposed by Wilhelm Wien which fit
the experimental data at short wavelengths but deviated from it at long wave-
lengths. He found that the above function, Planck’s function, fit the data for

Planck’s law of black body radiation


382

all wavelengths remarkably well. In constructing a derivation of this law, he


considered the possible ways of distributing electromagnetic energy over the
different modes of charged oscillators in matter. Planck’s law emerged when
he assumed that the energy of these oscillators was limited to a set of discrete,

FT
integer multiples of a fundamental unit of energy, E, proportional to the oscil-
lation frequency ν:
E = hν .

Planck made this quantization assumption five years before Albert Einstein hy-
pothesized the existence of photons as a means of explaining the photoelectric
effect. At the time, Planck believed that the quantization applied only to the
tiny oscillators that were thought to exist in the walls of the cavity (what we
now know to be atoms), and made no assumption that light itself propagates
in discrete bundles or packets of energy. Moreover, Planck did not attribute any
physical significance to this assumption, but rather believed that it was merely
a mathematical device that enabled him to derive a single expression for the
black body spectrum that matched the empirical data at all wavelengths.
A
Ultimately, Planck’s assumption of energy quantization and Einstein’s pho-
ton hypothesis became the fundamental basis for the later development of
Quantum Mechanics. Both scientists would eventually receive (separate) No-
bel prizes in recognition of these major contributions to the advancement of
physics.

Derivation (Statistical Mechanics)


(See also the gas in a box article for a general derivation.)
DR
Consider a cube of side L with conducting walls filled with electromagnetic
radiation. At the walls of the cube, the parallel component of the electric field
and the orthogonal component of the magnetic field must vanish. Analogous
to the wave function of a particle in a box, one finds that the fields are su-
perpositions of periodic functions. The wavelength λi in the three directions
i = 1 . . . 3 orthogonal to the walls can be:
2L
λi = ni

where the ni are integers. For each set of integers ni there are two linear inde-
pendent solutions (modes). According to quantum theory, the energy levels of
a mode are given by:
 hc q 2
En1 ,n2 ,n3 (r) = r + 12 2L n1 + n22 + n23 (1)

Planck’s law of black body radiation


383

The quantum number r can be interpreted as the number of photons in the


mode. The two modes for each set of ni correspond to the two polarization
states of the photon which has a spin of 1. Note that for r = 0 the energy of
the mode is not zero. This vacuum energy of the electromagnetic field is re-

FT
sponsible for the Casimir effect. In the following we will calculate the internal
energy of the box at temperature T relative to the vacuum energy.
According to statistical mechanics, the probability distribution over the energy
levels of a particular mode is given by:
exp(−βE(r))
Pr = Z(β)

Here
β ≡ 1/ (kT ).

The denominator Z (β), is the partition function of a single mode and makes
Pr properly normalized:
Z (β) = ∞ 1
P
r=0 exp [−βE (r)] = 1−exp[−βε]
A
Here we have defined
hc
q
ε ≡ 2L n21 + n22 + n23

which is the energy of a single photon. As explained here, the average energy
in a mode can be expressed in terms of the partition function:
d log(Z) ε
hEi = − dβ = exp(βε)−1
DR
This formula is a special case of the general formula for particles obeying
→Bose-Einstein statistics. Since there is no restriction on the total number
of photons, the chemical potential is zero.
The total energy in the box now follows by summing hEi over all allowed single
photon states. This can be done exactly in the thermodynamic limit L → ∞.
In this limit, ε becomes continuous and we can then integrate hEi over this
parameter. To calculate the energy in the box in this way, we need to evaluate
how many photon states there are in a given energy range. If we write the total
number of single photon states with energies between ε and ε + dε as g (ε) d,
where g (ε) is the density of states which we’ll evaluate in a moment, then we
can write:
R∞ ε
U = 0 exp(βε)−1 g (ε) dε (2)

Planck’s law of black body radiation


384

To calculate the density of states we rewrite equation (1) as follows:


hc
ε≡ 2L n

where n is the norm of the vector ~n = (n1 , n2 , n3 ):

FT
q
n = n21 + n22 + n23

For every vector n with integer components larger or equal than zero there are
two photon states. This means that the number of photon states in a certain
region of n-space is twice the volume of that region. An energy range of dε cor-
responds to shell of thickness dn = 2L hc dε in n-space. Because the components
of ~n have to be positive, this shell spans an octant of a sphere. The number of
photon states g (ε) d in an energy range dε is thus given by:
8πL3 2
g (ε) d = 2 81 4πn2 dn = h3 c3
ε dε

Inserting this in Eq. (2) gives:


R∞ ε3
U = L3 h8π3 3 0
c exp(βε)−1
dε (3)
A
From this equation one easily derives the spectral energy density as a function
of frequency u(ν, T ) and as a function of wavelength u(λ, T ):
U
R∞
L3
= 0 u(ν, T )dν

where:
8πhν 3 1
u(ν, T ) = c3 ehν/kT −1
DR
u(ν, T ) is known as the black body spectrum. It is a spectral energy density
function with units of energy per unit frequency per unit volume.
And:
U
R∞
L3
= 0 u(λ, T )dλ

where
8πhc 1
u(λ, T ) = λ5 ehc/λkT −1

This is also a spectral energy density function with units of energy per unit
wavelength per unit volume. Integrals of this type for Bose and Fermi gases can
be expressed in terms of polylogarithms. In this case, however, it is possible to

Planck’s law of black body radiation


385

calculate the integral in closed form. Let’s first make the integration variable
in Eq. (3) dimensionless by substituting ε = kT x:
8π(kT )4
u(T ) = (hc)3
J

FT
Here J is given by:
R∞ x3 π4
J = 0 exp(x)−1 dx = 15

We prove this result in the Appendix below. The total electromagnetic energy
inside the box is thus given by:
U 8π 5 (kT )4
V = 15(hc)3

where V = L3 is the volume of the box. (Note - This is not the Stefan-
Boltzmann law, which is the total energy radiated by a black body. See
that article for an explanation.) Since the radiation is the same in all direc-
tions, and propagates at the speed of light (c), the spectral intensity (ener-
gy/time/area/solid angle/frequency) is
AI(ν, T ) =
u(ν,T ) c

which yields
2hν 3 1
I(ν, T ) = c2 ehν/kT −1

Derivation (Thermodynamics)
The fact that the energy density of the box containing radiation is proportional
DR
to T 4 was derived by Ludwig Boltzmann in 1884 using thermodynamics. It
follows from classical electrodynamics that the radiation pressure P is related
to the internal energy density:
u
P = 3
The total internal energy of the box containing radiation can thus be written
as:
U = 3P V
Inserting this in the fundamental law of thermodynamics
dU = T dS − P dV
yields the equation:
dS = 4 PT dV + 3 VT dP
We can now use this equation to derive a Maxwell relation. We read off that:

Planck’s law of black body radiation


386

∂S P
∂V = 4 T
P
And

∂S V
∂P = 3 T
V

FT
The symmetry of second derivatives of S w.r.t. P and V then implies:

∂( P ) ∂( V )
4 ∂PT = 3 ∂VT
V P
Because the pressure is proportional to the internal energy density it depends
only on the temperature and not on the volume. In the derivative on the r.h.s.
the temperature is thus a constant. Evaluating the derivatives gives the differ-
ential equation:
1 dP 4
P dT = T
This implies that u = 3P ∝ T 4

History
Many popular science accounts of quantum theory, as well as some physics
textbooks, contain some serious errors in their discussions of the history of
A
Planck’s Law. Although these errors were pointed out over forty years ago by
historians of physics, they have proved to be difficult to eradicate. The article
by Helge Kragh cited below gives a lucid account of what actually happened.
Contrary to popular opinion, Planck did not quantize light. This is evident from
his original 1901 paper and the references therein to his earlier work. It is also
plainly explained in his book "Theory of Heat Radiation," where he explains
that his constant refers to Hertzian oscillators. The idea of quantization was
developed by others into what we now know as quantum mechanics. The
DR
next step along this road was made by Albert Einstein, who, by studying the
photoelectric effect, proposed a model and equation whereby light was not
only emitted but also absorbed in packets or photons. Then, in 1924, Satyendra
Nath Bose developed the theory of the statistical mechanics of photons, which
allowed a theoretical derivation of Planck’s law.
Contrary to another myth, Planck did not derive his law in an attempt to resolve
the "ultraviolet catastrophe", the name given to the paradoxical result that the
total energy in the cavity tends to infinity when the equipartition theorem of
classical statistical mechanics is applied to black body radiation. Planck did not
consider the equipartion theorem to be universally valid, so he never noticed
any sort of "catastrophe" — it was only discovered some five years later by
Einstein, Lord Rayleigh, and Sir James Jeans.

Planck’s law of black body radiation


387

Appendix
A simple way to calculate the integral
R∞ x3
J = 0 exp(x)−1 dx

FT
is as follows. After multiplying the numerator and denominator of the inte-
grand we can expand the integrand in powers of exp(−x).
R ∞ x3 exp(−x) R∞ 3 π4
J = 0 1−exp(−x) dx = ∞
P P∞ 1
n=1 0 x exp (−nx) dx = 6 n=1 n4 = 15

P∞ 1
Here we have used that n=1 n4 is the Riemann zeta function evaluated for
π4
the argument 4, which is given by 90 . This fact can be proven by considering
the contour integral
H π cot(πz)
CR z4

Where CR is a contour of radius R around the origin. In the limit R → ∞


the integral approaches zero. Using the residue theorem the integral can also
be written as a sum of residues at the poles of the integrand. The poles are
Aat zero, the positive and negative integers. The sum of the residues yields
precisely twice the desired summation plus the residue at zero. This means that
P∞ 1 π4 3
n=1 n4 equals minus 2 times the coefficient of x of the series expansion of
the series expansion of the cotangent function.

External link and references


• Planck’s original 1901 paper 209
DR
• Planck, Max, "On the Law of Distribution of Energy in the Normal Spec-
trum". Annalen der Physik, vol. 4, p. 553 ff (1901).
• Radiation of a Blackbody 210 - interactive simulation to play with Planck’s
law
• Scienceworld entry on the Planck Law 211
• Kragh, Helge Max Planck: The reluctant revolutionary 212 Physics World,
December 2000

Source: http://en.wikipedia.org/wiki/Planck%27s_law_of_black_body_radiation

209 http://dbhs.wvusd.k12.ca.us/webdocs/Chem-History/Planck-1901/Planck-1901.html
210 http://www.vias.org/simulations/simusoft_blackbody.html
211 http://scienceworld.wolfram.com/physics/PlanckLaw.html
212 http://www.physicsweb.org/articles/world/13/12/8/1

Planck’s law of black body radiation


388

Principal Authors: Metacomet, PAR, Unc.hbar, Diegueins, Rparson

Plum pudding model

A FT
DR
Figure 36 A schematic representation of the plum pudding model of the atom.

In physics, the Plum pudding model of the atom was proposed by J. J. Thom-
son, the discoverer of the electron in 1897. The plum pudding model was
proposed in March, 1904 before the discovery of the atomic nucleus. In this
model, the atom is composed of electrons surrounded by a soup of positive
charge to balance the electron’s charge, like plums surrounded by pudding.
The electrons were thought to be positioned throughout the atom, but with
many electron structures possible, particularly rotating rings (see below). In-
stead of a soup, the atom was also sometimes said to have had a cloud of
positive charge.
The model was disproved by the 1909 gold foil experiment, which was inter-
preted by Ernest Rutherford in 1911 to imply a very small nucleus of the atom
containing its full positive charge, thus leading implicitly to the →Rutherford

Plum pudding model


389

model of the atom, and eventually, by 1913, to the solar-system-like (but


quantum-limited) →Bohr model of the atom.
Thomson’s model was compared (though not by Thomson) to a British treat
called plum pudding, hence the name. It has also been called the chocolate

FT
chip cookie model, but only by those who have not read Thomson’s original
paper (On the Structure of the Atom: an Investigation of the Stability and
Periods of Oscillation of a number of Corpuscles arranged at equal intervals
around the Circumference of a Circle; with Application of the Results to the
Theory of Atomic Structure), published in the Philosophical Magazine (the
leading British science journal of the day). For an excerpt see 213.
A little-known (or now forgotten) fact about the original Thomson "plum pud-
ding" model is that it was dynamic, not static. The electrons were free to rotate
within the blob or cloud of positive substance. These orbits were stabilized
in the model by the fact that when an electron moved farther from the center
of the positive cloud, it felt a larger net positive inward force, because there
was more material of opposite charge, inside its orbit (A particle like a small
black hole would feel the same restorative force if it penetrated the body of
the Earth; such a particle would feel only the gravity of the Earth inside its
A
radius). In Thomson’s model, electrons were free to rotate in rings which were
further stabilized by interactions between the electrons, and spectra were to be
accounted for by energy differences of different ring orbits. Thomson attempt-
ed to make his model account for some of the major spectral lines known for
some elements, but was not notably successful at this. Still, Thomson’s model
(along with a similar Saturnian ring model for atomic electrons, put forward
by Nagaoka after the Maxwell model of Saturn’s rings), were earlier harbingers
of the later and more successful solar-system like →Bohr model of the atom.
DR
External links
• "On the Structure of the Atom: an Investigation of the Stability and Periods
of Oscillation of a number of Corpuscles arranged at equal intervals around
the Circumference of a Circle; with Application of the Results to the Theory
of Atomic Structure" 214 — J.J. Thomson’s 1904 paper proposing the plum
pudding model.

Source: http://en.wikipedia.org/wiki/Plum_pudding_model

Principal Authors: Linas, Salsb, Sbharris, Fastfission, Ragesoss

213 http://dbhs.wvusd.k12.ca.us/webdocs/Chem-History/Thomson-Structure-Atom.html
214 http://dbhs.wvusd.k12.ca.us/webdocs/Chem-History/Thomson-Structure-Atom.html

Plum pudding model


390

Position operator

Definition

FT
In quantum mechanics, the position operator corresponds to the position ob-
servable of a particle. Consider, for example, the case of a spinless particle
moving on a line. The state space for such a particle is L 2(R), the Hilbert
space of complex-valued and square-integrable (with respect to the Lebesgue
measure) functions on the real line. The position operator, Q, is then defined
by
Q(ψ)(x) = x · ψ(x),

with domain D(Q) = {ψ ∈ L2 |Qψ ∈ L2 }.


Since all continuous functions with compact support lie in D(Q), Q is densely
defined. Q, being simply multiplication by x, is a self adjoint operator, thus
satisfying the requirement of a quantum mechanical observable. Immediately
from the definition we can deduce that the spectrum consists of the entire
A
real line and that Q has purely continuous spectrum, therefore no eigenvalues.
The three dimensional case is defined analogously. We shall keep the one-
dimensional assumption in the following discussion.

Measurement
As with any observable, In order to discuss measurement, we need to calculate
the spectral resolution of Q :
R
Q = λdΩQ (λ).
DR
Since Q is just multiplication by x, its spectral resolution is simple. For a Borel
subset B of the real line, let C B denote the indicator function of B. We see that
the projection-valued measure ΩQ are given by
ΩQ (B)ψ = C B · ψ

,i.e. ΩQ (B) is multiplication by the indicator fuction of B. Therefore, if the


system is prepared in state Ψ, then the probability of the measured position of
the particle being in a Borel set B is
|ΩQ (B)ψ|2 = |C B · ψ|2 = B |ψ|2 dµ
R

,µ being the Lebesgue measure. After the measurement, the wave function
Ω (B)ψ
collapses to ||ΩQ (B)ψ|| , where || · || is the Hilbert space norm.
Q

Position operator
391

Unitary equivalence with momentum operator


For a particle on a line, the momentum operator P is defined by

P ψ = −i~ ∂x ψ

FT
,with appropriate domain. P and Q are unitarily equivalent, with the unitary
operator being given explicitly by the Fourier transform. Thus they have the
same spectrum. In physical language, P acting on momentum space wave
functions is the same as Q acting on position space wave functions (under the
image of Fourier transform).

Source: http://en.wikipedia.org/wiki/Position_operator

Potential energy surface

A potential energy surface is generally used within the adiabatic or Born-


Oppenheimer approximation in quantum mechanics and statistical mechanics
A
to model chemical reactions and interactions in simple chemical and physical
systems. There is a natural correspondence between potential energy surfaces
as they exist (as polynomial surfaces) and their application in potential theory,
which associates and studies harmonic functions in relation to these surfaces.
For example, the Morse potential and the simple harmonic potential well are
common one-dimensional potential energy surfaces (potential energy curves)
in applications of quantum chemistry and physics.
DR
Source: http://en.wikipedia.org/wiki/Potential_energy_surface

Principal Authors: V8rik, Cypa, Charles Matthews

Potential well

A potential well is the region surrounding a local minimum of potential energy.


Energy captured in a potential well is unable to convert to another type of
energy (kinetic energy in the case of a gravitational potential well) because it
is captured in the local minimum of a potential well. Therefore, a body may
not proceed to the global minimum of potential energy, as it would naturally
tend to due to entropy.

Potential well
392

Overview
Energy may be released from a potential well if sufficient energy is added to
the system such that the local minimum is surmounted. In quantum physics,
potential energy may escape a potential well without added energy due to the

FT
probabilistic characteristics of quantum particles; in these cases a particle may
be imagined to tunnel through the walls of a potential well.
The graph of a 2D potential energy function is a potential energy surface that
can be imagined as the Earth’s surface in a landscape of hills and valleys. Then
a potential well would be a valley surrounded on all sides with higher terrain,
which thus could be filled with water (i.e., be a lake) without any water flowing
away toward another, lower minimum (i.e. sea level).
In the case of gravity, the region around a mass is a gravitional potential well,
unless the density of the mass is so low that tidal forces from other masses are
greater than the gravity of the body itself.
A potential hill is the opposite of a potential well, the region surrounding a
local maximum.
A
Quantum confinement
Quantum confinement is when electrons and holes in a semiconductor are con-
fined by a potential well in 1D (quantum well), 2D (quantum wire), or 3D
(quantum dot). That is, quantum confinement occurs when one or more of the
dimensions of a nanocrystal is made very small so that it approaches the size
of an exciton in bulk crystal, called the Bohr exciton radius. A quantum well
is a structure where the height is approximately the Bohr exciton radius while
the length and breadth can be large. A quantum wire is a structure where the
DR
height and breadth is made small while the length can be long. A quantum dot
is a structure where all dimensions are near the Bohr exciton radius, typically
a small sphere.

See also
• a Graphical representation of a potential well

References
• W. E. Buhro and V. L. Colvin, Semiconductor nanocrystals: Shape
matters 215, Nat. Mater., 2003, 2, 138 139.

215 http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=pubmed&dopt=Abstract&list_uids=12612665

Potential well
393

Source: http://en.wikipedia.org/wiki/Potential_well

Principal Authors: Patrick, Laurascudder, Linas, Fasten, Bantman

FT
POVM
In functional analysis and quantum measurement theory, a POVM (Positive Op-
erator Value Measure) is a measure whose values are non-negative self-adjoint
operators on a →Hilbert space. It is the most general formulation of a mea-
surement in the theory of quantum physics. The need for the POVM formalism
arises from the fact that projective measurements on a larger system will act
on a subsystem in ways that cannot be described by projective measurement
on the subsystem alone. They are used in the field of Quantum information.
In rough analogy, a POVM is to a projective measurement what a density matrix
is to a pure state. Density matrices can describe part of a larger system that is
in a pure state (see purification of quantum state); analogously, POVMs on a
A
physical system can describe the effect of a projective measurement performed
on a larger system.

Definition
In the simplest case, a POVM is a set of Hermitian positive semidefinite opera-
P {Fi } on a Hilbert space H that sum to unity,
tors
n
i=1 Fi = IH .
DR
This formula is similar to the decomposition of a →Hilbert space into a set of
orthogonal projectors,
PN
i=1 Ei = IH ,

and if i 6= j,
Ei Ej = 0.

An important difference is that the elements of a POVM are not necessarily


orthogonal, with the consequence that the number of elements in the POVM,
n, can be larger than the dimension, N, of the →Hilbert space they act in.
In general, POVMs can be defined in situations where outcomes can occur
in a non-discrete space. The relevant fact is that measurements determine
probability measures on the outcome space:

POVM
394

Definition. Let (X, M ) be measurable space; that is M is a σ-algebra of subsets


of X. A POVM is a function F defined on M whose values are bounded non-
negative self-adjoint operators on a Hilbert space H such that F(X ) = I H and
for every ξ ∈ H,

FT
E 7→ hF (E)ξ | ξi

is a non-negative countably additive measure on the σ-algebra M.

POVMs and measurement


As in the theory of projective measurement, the probability the outcome asso-
ciated with measurement of operator Fi occurs is,
P (i) = T r(Fi ρ),

where ρ is the density matrix describing the state of the measured system.
An element of a POVM can always be written as,
Fi = Mi† Mi ,
A
for some operator Mi , known as a Kraus Operator. The state of the system after
the measurement ρ0 is transformed according to,
Mi ρMi†
ρ0 = .
tr(Mi ρMi† )

Neumark’s dilation theorem


DR
An alternate spelling of this is Naimark’s Theorem

Neumark’s dilation theorem states that measuring a POVM consisting of a set


of n>N operators acting on a N-dimensional →Hilbert space can always be
achieved by performing a projective measurement on a Hilbert space of dimen-
sion n then consider the reduced state.
In practice, however, obtaining a suitable projection-valued measure from a
given POVM is usually done by coupling to the original system an ancilla. Con-
sider a Hilbert space HA that is extended by HB . The state of total system
is ρAB and ρA = T rA (ρAB ). The probability the projective measurement π̂i
succeeds is,
P (i) = T rA (T rB (π̂i ρAB )).

POVM
395

An implication of Neumark’s theorem is that the associated POVM in subspace


A, Fi , must have the same probability of success.
P (i) = T rA (Fi ρA )).

FT
An example: Unambiguous quantum state discrim-
ination
The task of unambiguous quantum state discrimination (UQSD) is to discern
conclusively which state, of given set of pure states, a quantum system (which
we call the input) is in. The impossibility of perfectly discriminating between
a set of non-orthogonal states is the basis for quantum information protocols
such as quantum cryptography, quantum coin-flipping, and quantum money.
This example will show that a POVM has a higher success probability for per-
forming UQSD than any possible projective measurement.
A
DR

Figure 37 The projective measurement strategy for unambiguously discriminating between


nonorthogonal states.

First let us consider a trivial case. Take a set that consists of two orthogonal
states |ψi and |ψ T i. A projective measurement of the form,

POVM
396

 = a|ψ T ihψ T | + b|ψihψ|,

will result in eigenvalue a only when the system is in |ψ T i and eigenvalue


b only when the system is in |ψi. In addition, the measurement always dis-

FT
criminates between the two states (i.e. with 100% probability). This latter
ability is unnecessary for UQSD and, in fact, is impossible for anything but or-
thogonal states. Now consider a set that consists of two states |ψi and |φi in
two-dimensional Hilbert space that are not orthogonal. i.e.,
|hφ|ψi| = cos(θ),

for θi0. These states could a system, such as the spin of spin-1/2 particle (e.g.
an electron), or the polarization of a photon. Assuming that the system has
an equal likelihood of being in each of these two states, the best strategy for
UQSD using only projective measurement is to perform each of the following
measurements,
π̂ψT = |ψ T ihψ T |,
A
π̂φT = |φT ihφT |,

50% of the time. If π̂φT is measured and results in an eigenvalue of 1, than


it is certain that the state must have been in |ψi. However, an eigenvalue of
zero is now an inconclusive result since this can come about from the system
could being in either of the two states in the set. Similarly, a result of 1 for
π̂ψT indicates conclusively that the system is in |φi and 0 is inconclusive. The
probability that this strategy returns a conclusive result is,
DR
1−|hφ|ψi|2
Pproj = 2 .

In contrast, a strategy based on POVMs has a greater probability of success


given by,
PP OV M = 1 − |hφ|ψi|.

This is the minimum allowed by the rules of quantum indeterminacy and the
uncertainty principle. This strategy is based on a POVM consisting of,
1−|φihφ|
F̂ψ = 1+|hφ|ψi|

1−|ψihψ|
F̂φ = 1+|hφ|ψi|

POVM
397

F̂inconcl. = 1 − F̂ψ − F̂φ ,

where the result associated with F̂i indicates the system is in state i with cer-
tainty.

A FT
Figure 38 The POVM strategy for unambiguous-
ly discriminating between nonorthogonal states.
DR
These POVMs can be created by extending the two-dimensional Hilbert space.
This can be visualized as follows: The two states fall in the x-y plane with an
angle of θ between them and the space is extended in the z-direction. (The
total space is the direct sum of spaces defined by the z-direction and the x-y
plane.) The measurement first unitarily rotates the states towards the z-axis
so that |ψi has no component along the y-direction and |φi has no component
along the x-direction. At this point, the three elements of the POVM correspond
to projective measurements along x-direction, y-direction and z-direction, re-
spectively.
For a specific example, take a stream of photons, each of which are polarized
along either along the horizontal direction or at 45 degrees. On average there
are equal numbers of horizontal and 45 degree photons. The projective strate-
gy corresponds to passing the photons through a polarizer in either the vertical

POVM
398

direction or -45 degree direction. If the photon passes through the vertical po-
larizer it must have been at 45 degrees and vice versa. The success probability
is (1 − 1/2)/2 = 25%. The POVM strategy for this example is more complicat-
ed and requires another optical mode (known as an ancilla). It has a success

FT
probability of 1 − 1/ 2 = 29.3%.

See also
• Quantum measurement
• →Mathematical formulation of quantum mechanics
• Quantum logic
• Density matrix
• Quantum operation
• Projection-valued measure

References
• POVMs
A
• J.Preskill, Lecture Note for Physics: Quantum Information and Compu-
tation, http://theory.caltech.edu/people/preskill
• K.Kraus, States, Effects, and Operations, Lecture Notes in Physics 190,
Springer (1983)
• E.B.Davies, Quantum Theory of Open Systems, Academic Press (1976).
• Neumark’s theorem
• A. Peres. Neumark’s theorem and quantum inseparability. Foundations
of Physics, 12:1441–1453, 1990.
DR
• A. Peres. Quantum Theory: Concepts and Methods. Kluwer Academic
Publishers, 1993.
• I. M. Gelfand and M. A. Neumark, On the imbedding of normed rings
into the ring of operators in Hilbert space, Rec. Math. [Mat. Sbornik]
N.S. 12(54) (1943), 197–213.
• Unambiguous quantum state-discrimination
• I. D. Ivanovic, Phys. Lett. A 123 257 (1987).
• D. Dieks, Phys. Lett. A 126 303 (1988).
• A. Peres, Phys. Lett. A 128 19 (1988).

Source: http://en.wikipedia.org/wiki/POVM

POVM
399

Probability amplitude

In quantum mechanics, a probability amplitude is a complex-valued function

FT
that describes an uncertain or unknown quantity. For example, each parti-
cle has a probability amplitude describing its position. This amplitude is then
called wave function. This is a complex-valued function of the position coordi-
nates.
For a probability amplitude ψ, the associated probability density function is
ψ*ψ, which is equal to |ψ| 2. This is sometimes called just probability density 1,
and may be found used without normalization.
If |ψ| 2 has a finite integral over the whole of three-dimensional space, then
it is possible to choose a normalising constant, c, so that by replacing ψ by
cψ the integral becomes 1. Then the probability that a particle is within a
particular region V is the integral over V of |ψ| 2. Which means, according to
the Copenhagen interpretation of quantum mechanics, that, if some observer
tries to measure the quantity associated with this probability amplitude, the
A
result of the measurement will lie within  with a probability P() given by
P () =  |ψ(x)|2 dx
R

Probability amplitudes which are not square integrable are usually interpreted
as the limit of a series of functions which are square integrable. For instance
the probability amplitude corresponding to a plane wave corresponds to the
’non physical’ limit of a monochromatic source of particles. Another example:
The Siegert wave functions describing a resonance are the limit for t → ∞ of
DR
a time-dependent wave packet scattered at an energy close to a resonance. In
these cases, the definition of P() given above is still valid.
The change over time of this probability (in our example, this corresponds to
a description of how the particle moves) is expressed in terms of ψ itself, not
just the probability function |ψ| 2. See →Schrödinger equation.
In order to describe the change over time of the probability density it is ac-
ceptable to define the probability flux (also called probability current). The
probability flux j is defined as:
j= ~
m · 1
2i (ψ ∗ ∇ψ − ψ∇ψ ∗ ) = ~ ∗
m Im (ψ ∇ψ)

and measured in units of (probability)/(area*time) = r -2t -1.


The probability flux satisfies a quantum continuity equation, i.e.:

Probability amplitude
400


∇·j+ ∂t P (x, t) =0

where P(x,t) is the probability density and measured in units of (probabili-


ty)/(volume) = r -3. This equation is the mathematical equivalent of probabil-

FT
ity conservation law.
It is easy to show that for a plane wave function,
|ψi = A exp (ikx − iωt)

the probability flux is given by


j(x, t) = |A|2 k~
m

The bi-linear form of the axiom has interesting consequences as well.

Notes

• Note 1: Max Born was awarded part of the 1954 Nobel Prize in Physics for
this work.
A
Source: http://en.wikipedia.org/wiki/Probability_amplitude

Principal Authors: Charles Matthews, P3d0, That Guy, From That Show!, Onebyone, Michael Hardy,
RJFJR, NymphadoraTonks, Conscious, Paul A

Probability current
DR
In quantum mechanics, the probability current (sometimes called probability
flux) is a useful concept which describes the flow of probability density. In
particular, if one pictures the probability density as an inhomogeneous fluid,
then the probability current is the rate of flow of this fluid (the density times
the velocity).

Definition
The probability current, ~j, is defined as
 
~j = ~ Ψ∗ ∇Ψ~ − Ψ∇Ψ ~ ∗
2mi

in the position basis and satisfies the quantum mechanical continuity equation

Probability current
401
∂ρ ~ · ~j = 0
∂t +∇

with the probability density ρ defined as


ρ = |Ψ|2 .

FT
The divergence theorem implies the continuity equation is equivalent to the
integral equation
∂ ~ =0
|Ψ|2 dV + ~j · dA
R R
∂t V S

where the V is any volume and S is the boundary of V . This is the conservation
law for probability density in quantum mechanics.
In particular, if Ψ is a wavefunction describing a single particle, the integral
in the first term of the preceding equation (without the time derivative) is the
probability of obtaining a value within V when the position of the particle is
measured. The second term is then the rate at which probability is flowing out
of the volume V . Altogether the equation states that the time derivative of the
change of the probability of the particle being measured in V is equal to the
A
rate at which probability flows into V .

Examples
Plane wave
The probability current associated with the (three dimensional) plane wave
~
Ψ = eik·~r
DR
is
 
~ ~ i~k·~r ~ ~ −i~k·~r ~~k
~j = ~
2mi e−ik·~r ∇e − eik·~r ∇e = m.

This is just the particle’s momentum


p~ = ~~k

divided by its mass, i.e. its "velocity" (insofar as a quantum mechanical particle
has one). Note that the probability current is nonzero despite the fact that
plane waves are stationary states and hence
d|Ψ|2
dt =0

Probability current
402

everywhere. This demonstrates that a particle may be in "motion" even if its


spacial probability density has no explicit time dependence.

Particle in a box

FT
The energy eigenstates of a particle in a box of one spatial dimension and of
length L are
q
Ψn = L2 sin nπ

L x

The associated probability currents are


∂Ψ∗n
 
~
jn = 2mi Ψ∗n ∂Ψ
∂x − Ψn ∂x
n
=0

since Ψn = Ψ∗n .

Derivation of continuity equation


In this section the continuity equation is derived from the definition of proba-
bility current and the basic principles of quantum mechanics.
Suppose Ψ is the wavefunction for a single particle in the position basis (i.e.
A
Ψ is a function of x, y, and z). Then
P = V |Ψ|2 dV
R

is the probability that a measurement of the particle’s position will yield a value
within V. The time derivative of this is
dP ∂
R 2
R  ∂Ψ ∗ ∂Ψ∗

dt = ∂t V |Ψ| dV = V ∂t Ψ + Ψ ∂t dV
DR
where the last equality follows from the product rule and the fact that the
shape of V is presumed to be independent of time (i.e. the time derivative can
be moved through the integral). In order to simplify this further consider the
time dependent →Schrödinger equation
−~2 2
i~ ∂Ψ
∂t = 2m ∇ Ψ + V Ψ

and use it to solve for the time derivative of Ψ :


∂Ψ i~ 2 i
∂t = 2m ∇ Ψ − ~ V Ψ

dP
When substituted back into the preceding equation for dt this gives
dP ∗ 2 2 ∗
R ~ 
dt = − V 2mi Ψ ∇ Ψ − Ψ∇ Ψ dV .

Now from the product rule for the divergence operator

Probability current
403
 
~ − Ψ∇Ψ
∇ · Ψ∗ ∇Ψ ~ ∗ = ∇Ψ
~ ∗ · ∇Ψ
~ + Ψ∗ ∇2 Ψ − ∇Ψ
~ · ∇Ψ
~ ∗ − Ψ∇
~ 2 Ψ∗

and since the first and third terms cancel:


 
dP ~ · ~ Ψ∗ ∇Ψ ~ − Ψ∇Ψ ~ ∗ dV
R
=− ∇

FT
dt V 2mi

If we now recall the expression for P and note that the argument of the diver-
gence operator is just ~j this becomes
R  ∂|Ψ|2 
+ ~ · ~j dV = 0

V ∂t

which is the integral form of the continuity equation. The differential form
follows from the fact that the preceding equation holds for all V , and hence
the integrand must vanish everywhere:
∂|Ψ|2 ~ · ~j = 0.
∂t +∇

Source: http://en.wikipedia.org/wiki/Probability_current
A
Projective Hilbert space

In mathematics and the foundations of quantum mechanics, the projective


Hilbert space P (H ) of a complex →Hilbert space is the set of equivalence
classes of vectors v in H, with v 6= 0, for the relation given by
DR
v w when v = w

with a scalar, that is, a complex number (which must therefore be non-zero).
Here the equivalence classes for are also called projective rays.
This is the usual construction of projective space, applied to a Hilbert space.
The physical significance of the projective Hilbert space is that in quantum
theory, the wave functions ψ and ψ represent the same physical state, for any
6= 0. There is not a unique normalized wavefunction in a given ray, since we
can multiply by with absolute value 1. This freedom means that projective
representations enter quantum theory.
The same construction can be applied also to real Hilbert spaces.
In the case H is finite-dimensional, that is, H = Hn , the set of projective rays
may be treated just as any other projective space; it is a homogeneous space for

Projective Hilbert space


404

a unitary group or orthogonal group, in the complex and real cases respectively.
For the finite-dimensional complex Hilbert space, one writes
P (Hn ) = CP n−1

FT
so that, for example, the two-dimensional projective Hilbert space (the space
describing one qubit) is the complex projective line CP 1 . This is known as the
Bloch sphere, which treats the subject in greater detail.
Complex projective Hilbert space may be given a natural metric, the Fubini-
Study metric. The product of two projective Hilbert space is given by the Segre
mapping.

Source: http://en.wikipedia.org/wiki/Projective_Hilbert_space

Principal Authors: Charles Matthews, Asbestos, Linas

Pure gauge
A
In physics, pure gauge is the set of field configurations obtained by a gauge
transform on the null field configuration. So it is a particular "gauge orbit" in
the field configuration’s space. In the abelian case, where Aµ (x) → A0µ (x) =
Aµ (x)+∂µ f (x), the pure gauge is the set of field configurations A0µ (x) = ∂µ f (x)
for all f (x).

Source: http://en.wikipedia.org/wiki/Pure_gauge
DR
Principal Authors: LeeHunter, Michael Hardy, Sn0wflake, Oleg Alexandrov, Jag123

Quantum

In physics, a quantum refers to an indivisible and perhaps elementary entity.


For instance, a "light quantum", being a unit of light (that is, a photon). In
combinations like "quantum mechanics", "quantum optics", etc., it distinguishes
a more specialized field of study.
The word comes from the Latin "quantus", for "how much".
Behind this, one finds the fundamental notion that a physical property may be
"quantized", referred to as "quantization". This means that the magnitude can
take on only certain numerical values, rather than any value, at least within a

Quantum
405

range. For example, the energy of an electron bound to an atom (at rest) is
quantized. This accounts for the stability of atoms, and matter in general.
An entirely new conceptual framework was developed around this idea, dur-
ing the first half of the 1900s. Usually referred to as quantum "mechanics",

FT
it is regarded by virtually every professional physicist as the most fundamen-
tal framework we have for understanding and describing nature, for the very
practical reason that it works. It is "in the nature of things", not a more or less
arbitrary human preference.

Discovery of quantum theory


Quantum theory, the branch of physics based on quantization, began in 1900
when Max Planck published his theory explaining the emission spectrum of
black bodies. In that paper Planck used the Natural system of units invented
by him the previous year.
The consequences of the differences between classical and quantum mechan-
ics quickly became obvious. But it was not until 1926, by the work of Werner
Heisenberg, Erwin Schrödinger, and others, that quantum mechanics became
A
correctly formulated and understood mathematically. Despite tremendous ex-
perimental success, the philosophical interpretations of quantum theory are
still widely debated.
Planck was reluctant to accept the new idea of quantization, as were many
others. But, with no acceptable alternative, he continued to work with the
idea, and found his efforts were well received. Eighteen years later, he called
it, "a few weeks of the most strenuous work" of his life, when he accepted the
Nobel Prize in Physics for his contributions. During those few weeks, he even
DR
had to discard much of his own theoretical work from the preceding years.
Quantization turned out to be the only way to describe the new, and detailed
experiments which were just then being performed. He did this practically
overnight, openly reporting his change of mind to his scientific colleagues, in
the October, November, and December meetings of the German Physical Soci-
ety, in Berlin, where the black body work was being intensely discussed. In this
way, careful experimentalists (including F. Paschen, O.R. Lummer, Ernst Pring-
sheim, Heinrich Rubens, and F. Kurlbaum), and a reluctant theorist, ushered
in the greatest revolution science has ever seen.

The quantum black-body radiation formula


When a body is heated, it emits radiant heat, a form of electromagnetic radia-
tion in the infrared region of the EM spectrum. All of this was well understood
at the time, and of considerable practical importance. When the body becomes

Quantum
406

red-hot, the red wavelength parts start to become visible. This had been stud-
ied over the previous years, as the instruments were being developed. Howev-
er, most of the heat radiation remains infrared, until the body becomes as hot
as the surface of the Sun (about 6000 ◦ C, where most of the light is green in

FT
color). This was not achievable in the laboratory at that time. What is more,
measuring specific infrared wavelengths was only then becoming feasible, due
to newly developed experimental techniques. Until then, most of the electro-
magnetic spectrum was not measurable, and therefore blackbody emission had
not been mapped out in detail.
The quantum black-body radiation formula, being the very first piece of quan-
tum mechanics, appeared Sunday evening October 7, 1900, in a so-called back-
of-the-envelope calculation by Planck. It was based on a report by Rubens
(visiting with his wife) of the very latest experimental findings in the infrared.
Later that evening, Planck sent the formula on a postcard, which Rubens had
the following morning. A couple of days later, he could tell Planck that it
worked perfectly. As it does to this day.
At first, it was just a fit to the data. Only weeks later did it turn out to enforce
quantization.
A
That the latter became possible involved a certain amount of luck (or skill,
even though Planck himself called it "a fortuitous guess at an interpolation for-
mula"). It only had that drastic "side effect" because the formula happened to
become fundamentally correct, in regard to the as yet non-existent quantum
theory. And normally, that much is not at all expected. The skill lay in simpli-
fying the mathematics, so that this could happen. And here Planck used hard
won experience from the previous years. Briefly stated, he had two mathemat-
ical expressions:
DR
• (i) from the previous work on the red parts of the spectrum, he had x;
• (ii) now, from the new infrared data, he got x 2.

Combining these as x(a+x), he still has x, approximately, when x is much


smaller than a ( the red end of the spectrum). But now also x 2, again ap-
proximately, when x is much larger than a (in the infrared). The luck part is
that, this procedure turned out to actually give something completely right, far
beyond what could reasonably be expected. The formula for the energy E, in a
single mode of radiation at frequency f, and temperature T, can be written
hf
E= hf
e kT −1

This is (essentially) what is being compared with the experimental measure-


ments. There are two parameters to determine from the data, written in the

Quantum
407

present form by the symbols used today: h is the new Planck’s constant, and
k is Boltzmann’s constant. Both have now become fundamental in physics,
but that was by no means the case at the time. The "elementary quantum of
energy" is hf. But such a unit does not normally exist, and is not required for

FT
quantization.

The birthday of quantum mechanics


From the experiments, Planck deduced the numerical values of h and k. Thus
he could report, in the German Physical Society meeting on December 14,
1900, where quantization (of energy) was revealed for the first time, values of
the Avogadro-Loschmidt number, the number of real molecules in a mole, and
the unit of electrical charge, which were more accurate than those known until
then. This event has been referred to as "the birthday of quantum mechanics".

Quantization in antiquity
In a sense, it can be said that the quantization idea is very old. A string under
tension, and fixed at both ends, will vibrate at certain quantized frequencies,
corresponding to various standing waves. This, of course, is the basis of music.
A
The basic idea was regarded as essential by the Pythagoreans, who are reported
to have held numbers in high esteem.
It is a curious fact that the famous formula, named after Pythagoras, for the side
lengths of a right triangle, today serves as a cornerstone of quantum mechanics
as well.
The very existence of atoms and molecules can be ascribed to various forms of
quantization contrary to notions of matter as some form of continuous medium.
DR
This was also understood already in antiquity, particularly by Leucippus and
Democritus, although not generally appreciated, even by physicists, until the
late 19th- and early 20th- centuries, shortly before the invention of quantum
mechanics.
It should be mentioned, though, that later works within the Epicurean school
of thought played a significant role in forming the physics and chemistry of
the Renaissance period in Europe. In particular the famous tutorial poem "De
rerum natura" by the Roman author Titus Lucretius Carus.
Ancient India had a very highly developed atomic doctrine in the school of
Vaisheshika associated with the sage Kanada.

Quantum
408

References
• J. Mehra and H. Rechenberg, The Historical Development of Quantum The-
ory, Vol.1, Part 1, Springer-Verlag New York Inc., New York 1982.

FT
• 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.
• 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.

See also
• →Quantum mechanics
• →Quantum state
• Quantum number
• Quantum cryptography


AQuantum electronics
Quantum computer
• Quantum immortality
• Magnetic flux quantum
• Quantization
• →Subatomic particle
• →Elementary particle
DR
Source: http://en.wikipedia.org/wiki/Quantum

Principal Authors: Dennis Estenson II, Finn-Zoltan, Deathphoenix, Bensaccount, Stevertigo,


Freakofnurture, Scorpionman, Truthflux, Salvatore Ingala, Ashmoo

Quantum
409

Quantum 1/f noise

Quantum 1/f noise is an intrinsic part of quantum mechanics. It comes from

FT
scattering of different particles of one another in solid state physics. Quantum
1/f noise is a source of Chaos in such systems.

Other noise data sets


1/f noise has also recently been discovered in higher ordered self constructing
functions, as well as complex systems, both biological, chemical, and physical.

The theory
The basic derivation of Quantum 1/f was made by Peter Handel from the Uni-
versity of Missouri - St. Louis, and published in Physical Review Letters A, in
August 1980.
For more on Quantum 1/f, see: P.H. Handel: "1/f Macroscopic Quantum Fluc-
tuations of Electric Currents Due to Bremsstrahlung with Infrared Radiative
A
Corrections", Zeitschrift fuer Naturforschung 30a, p.1201 (1975)

See also
• shot noise
• 1/f noise
• white noise
• Johnson-Nyquist noise
• signal-to-noise ratio
DR
• noise level
• noise power
• noise-equivalent power
• phase noise
• thermal noise
• list of noise topics
• audio system measurements
• Colors of noise

Source: http://en.wikipedia.org/wiki/Quantum_1/f_noise

Principal Authors: Gordon Stangler, Charles Matthews, Wendell, Ardric47, Conscious

Quantum 1/f noise


410

Quantum acoustics

In physics, quantum acoustics is the study of sound under conditions such

FT
that quantum mechanical effects are germane. For most applications, classical
mechanics are sufficient to accurately describe the physics of sound. However
very high frequency sounds, or sounds made at very low temperatures may be
subject to quantum effects.
216
A symposium on quantum acoustics is held in Poland each year

See also
• Superfluid

References
• Quantum acoustics by Humphrey J. Maris in the McGraw-Hill Encyclopedia
A of Science & Technology Online 217
• Handbook of Acoustics by Malcolm Crocker has a chapter on quantum
acoustics.

Source: http://en.wikipedia.org/wiki/Quantum_acoustics

Quantum biology
DR
Quantum biology is the science of studying biological processes in terms of
quantum mechanics. In exploring quantum mechanical explanations for bio-
logical phenomena, the nascent science of quantum biology represents one of
the first efforts to apply quantum theory to systems much more macroscopic
than the atomic or subatomic realms generally described by quantum theory.
The following biological phenomena have been described in terms of quantum
processes (although, to the extent that quantum theory is correct, all macro-
scopic phenenoma would be the result of quantum processes):

216 http://gnom.matfiz.polsl.gliwice.pl/afik/2005/index.html
217 http://www.accessscience.com/Encyclopedia/5/56/Est_562350_frameset.html?doi

Quantum biology
411

• the absorbance of frequency-specific radiation (i.e., photosynthesis and vi-


sion);
• the conversion of chemical energy into motion;
• magnetoreception in animals.

FT
Quantum biological research is extremely limited by computer processing pow-
er; the analytical power required to model quantum effects increases exponen-
tially with the number of particles involved.

References
• nanoword.net 218
• Theoretical and Computational Biophysics Group, University of Illinois at
Urbana-Champaign 219

Source: http://en.wikipedia.org/wiki/Quantum_biology
A
Quantum chaos

Quantum chaos is an interdisciplinary branch of physics, arising from so-called


semi-classical models.
Classical mechanics has historically been one of the fundamental theories of
physics, and is complete in the sense that all its axioms are mutually consistent
and not in need of further incremental refinement. However, many of the most
DR
difficult unsolved problems in contemporary physics and applied mathematics
in fact originate in classical mechanics, particularly in the field of deterministic
chaos. Laws of classical mechanics govern the macroscopic world of everyday
experience.
An important question of quantum mechanics is how to obtain the laws of
classical mechanics as limiting cases of the more fundamental laws governing
the microscopic constituents of matter. The correspondence principle is an
expression of this goal, which strongly influenced the early development of
quantum mechanical theories and their applications. However, the classical
limit of a quantum description may lead to a mechanical system with chaotic
dynamics.

218 http://www.nanoword.net/library/weekly/aa062500a.htm
219 http://www.ks.uiuc.edu/Research/quantum_biology/

Quantum chaos
412

During the first half of the twentieth century, chaotic behavior in mechanics
was recognized (in celestial mechanics), but not well-understood. The founda-
tions of modern quantum mechanics were laid in that period, essentially leav-
ing aside the issue of the quantum-classical correspondence in systems whose

FT
classical limit exhibits chaos.
This question defines the field of quantum chaos, which has emerged in the
second half of the twentieth century, aided to a large extent by renewed interest
in classical nonlinear dynamics (chaos theory), and by quantum experiments
bordering on the macroscopic size regime where laws of classical mechanics
are expected to emerge. This transition regime between classical and quantum
systems is also called semiclassical physics.
Similar questions arise in many different branches of physics, ranging from
nuclear to atomic, molecular and solid-state physics, and even to acoustics, mi-
crowaves and optics. This is what makes quantum chaos an interdisciplinary
field, unified by wave phenomena that can be interpreted as fingerprints of
classical chaos. Such phenomena can be identified in spectroscopy by analyz-
ing the statistical distribution of spectral lines. Other phenomena show up in
the time evolution of a quantum system, or in its response to various types
Aof external forces. In some contexts, such as acoustics or microwaves, wave
patterns are directly observable and exhibit irregular amplitude distributions.
Important observations often associated with classically chaotic quantum sys-
tems are level repulsion in the spectrum, dynamical localization in the time
evolution (e.g. ionization rates of atoms), and enhanced stationary wave in-
tensities in regions of space where classical dynamics exhibits only unstable
trajectories (wave function scarring).
DR
An alternative name for quantum chaos, proposed by Sir Michael Berry, is quan-
tum chaology.

History
Important methods applied in the theoretical study of quantum chaos in-
clude random-matrix theory (significant contributions by Oriol Bohigas, see
also American Scientist 220) and periodic-orbit theory (pioneered by Martin
Gutzwiller).

220 http://www.americanscientist.org/template/AssetDetail/assetid/21879/page/1;jsessionid=aaa-ZYP5NrRxh8

Quantum chaos
413

References
• A. Einstein (1917). "Zum Quantensatz von Sommerfeld und Epstein". Ver-
handlungen der Deutschen Physikalischen Gesellschaft 19: 82-92. Reprint-

FT
ed in The Collected Papers of Albert Einstein, A. Engel translator, (1997)
Princeton University Press, Princeton. 6 p.434. (Provides an elegant re-
formulation of the Bohr-Sommerfeld quantization conditions, as well as an
important insight into the quantization of non-integrable (chaotic) dynam-
ical systems.)
• Joeseph B. Keller (1958). "". Annals of Physics (NY) 4: 180. (An indepen-
dent rediscovery of the A. Einstein quantization conditions.)
• Joeseph B. Keller (1960). "". Annals of Physics (NY) 9: 24.
• Martin C. Gutzwiller (1971). "". Journal of Mathematical Physics 12: 343.
• Martin C. Gutzwiller, Chaos in Classical and Quantum Mechanics, (1990)
Springer-Verlag, New York ISBN=0-387-97173-4.

Source: http://en.wikipedia.org/wiki/Quantum_chaos
A Principal Authors: Linas, ChicXulub, Rmrfstar, Daniel tzvi, Neilc

Quantum Critical Point

The Quantum critical point is the lowest temperature point (1-3 degrees
Kelvin) at which a change of state-of-existence occurs. Experimentally, this
DR
has been demonstrated with the Han purple pigment. 221 When exposed to
both super-low temperatures and very high magnetic fields (above 23 Tesla),
the Han Purple pigment actually loses a dimension, transforming from 3D to
2D.

References

Source: http://en.wikipedia.org/wiki/Quantum_Critical_Point

221 National high magnetic field laboratory, press release May 31, 2006(http://www.magnet.fsu.edu
/news/pressreleases/2006may31.html) (accessed June 15, 2005)

Quantum Critical Point


414

Quantum entanglement

Quantum entanglement is a quantum mechanical phenomenon in which the

FT
quantum states of two or more objects have to be described with reference to
each other, even though the individual objects may be spatially separated. This
leads to correlations between observable physical properties of the systems.
For example, it is possible to prepare two particles in a single quantum state
such that when one is observed to be spin-up, the other one will always be
observed to be spin-down and vice versa, this despite the fact that it is impos-
sible to predict, according to quantum mechanics, which set of measurements
will be observed. As a result, measurements performed on one system seem
to be instantaneously influencing other systems entangled with it. Quantum
entanglement does not enable the transmission of classical information faster
than the speed of light (see discussion in next section below).
Quantum entanglement is closely concerned with the emerging technologies of
quantum computing and quantum cryptography, and has been used to experi-
mentally realize quantum teleportation. At the same time, it prompts some of
A
the more philosophically oriented discussions concerning quantum theory. The
correlations predicted by quantum mechanics, and observed in experiment, re-
ject the principle of local realism, which is that information about the state of a
system should only be mediated by interactions in its immediate surroundings.
Different views of what is actually occurring in the process of quantum entan-
glement can be related to different interpretations of quantum mechanics.

Background
DR
Entanglement is one of the properties of quantum mechanics which caused
Einstein and others to dislike the theory. In 1935, Einstein, Podolsky, and Rosen
formulated the EPR paradox, a quantum-mechanical thought experiment with
a highly counterintuitive and apparently nonlocal outcome. Einstein famously
derided entanglement as "spooky action at a distance."
On the other hand, quantum mechanics has been highly successful in produc-
ing correct experimental predictions, and the strong correlations associated
with the phenomenon of quantum entanglement have in fact been observed.
One apparent way to explain quantum entanglement is an approach known
as "hidden variable theory", in which unknown deterministic microscopic pa-
rameters would cause the correlations. However, in 1964 Bell derived an up-
per limit, known as Bell’s inequality, on the strength of correlations for any

Quantum entanglement
415

theory obeying "local realism" (see principle of locality). Quantum entangle-


ment can lead to stronger correlations that violate this limit, so that quan-
tum entanglement is experimentally distinguishable from a broad class of local
hidden-variable theories. Results of subsequent experiments have overwhelm-

FT
ingly supported quantum mechanics. It is known that there are a number of
loopholes in these experiments. High efficiency and high visibility experiments
are now in progress which should accept or reject those loopholes. For more
information, see the article on Bell test experiments.
Observations on entangled states naively appear to conflict with the property
of Einsteinian relativity that information cannot be transferred faster than the
speed of light. Although two entangled systems appear to interact across large
spatial separations, no useful information can be transmitted in this way, so
causality cannot be violated through entanglement. This occurs for two sub-
tle reasons: (i) quantum mechanical measurements yield probabilistic results,
and (ii) the no cloning theorem forbids the statistical inspection of entangled
quantum states.
Although no information can be transmitted through entanglement alone, it is
possible to transmit information using a set of entangled states used in con-
A
junction with a classical information channel. This process is known as quan-
tum teleportation. Despite its name, quantum teleportation cannot be used to
transmit information faster than light, because a classical information channel
is involved.

Pure States
The following discussion builds on the theoretical framework developed in the
articles bra-ket notation and mathematical formulation of quantum mechanics.
DR
Consider two noninteracting systems A and B, with respective →Hilbert spaces
HA and HB . The Hilbert space of the composite system is the tensor product
HA ⊗ HB

If the first system is in state |ψiA and the second in state |φiB , the state of the
composite system is
|ψiA ⊗ |φiB ,

which is often also written as


|ψiA |φiB .

States of the composite system which can be represented in this form are called
separable states, or product states.

Quantum entanglement
416

Pick observables (and corresponding Hermitian operators) ΩA acting on HA ,


and ΩB acting on HB . According to the spectral theorem, we can find a basis
{|iiA } for HA composed of eigenvectors of ΩA , and a basis {|jiB } for HB
composed of eigenvectors of ΩB . We can then write the above pure state as

FT
P P 
( i ai |iiA ) j bj |jiB ,

for some choice of complex coefficients ai and bj . This is not the most general
state of HA ⊗ HB , which has the form
P
i,j cij |iiA ⊗ |jiB .

If such a state is not separable, it is known as an entangled state.


For example, given two basis vectors {|0iA , |1iA } of HA and two basis vectors
{|0iB , |1iB } of HB , the following is an entangled state:
 
√1 |0iA ⊗ |1iB − |1iA ⊗ |0iB .
2

If the composite system is in this state, it is impossible to attribute to either


A
system A or system B a definite pure state. Instead, their states are superposed
with one another. In this sense, the systems are "entangled".
Now suppose Alice is an observer for system A, and Bob is an observer for
system B. If Alice performs the measurement ΩA , there are two possible out-
comes, occurring with equal probability:

• Alice measures 0, and the state of the system collapses to |0iA |1iB
• Alice measures 1, and the state of the system collapses to |1iA |0iB .
DR
If the former occurs, any subsequent measurement of ΩB performed by Bob
always returns 1. If the latter occurs, Bob’s measurement always returns 0.
Thus, system B has been altered by Alice performing her measurement on
system A., even if the systems A and B are spatially separated. This is the
foundation of the EPR paradox.
The outcome of Alice’s measurement is random. Alice cannot decide which
state to collapse the composite system into, and therefore cannot transmit in-
formation to Bob by acting on her system. (There is a possible loophole: if
Bob could make multiple duplicate copies of the state he receives, he could
obtain information by collecting statistics. This loophole is closed by the no
cloning theorem, which forbids the creation of duplicate states.) Causality is
thus preserved, as claimed above.
In more formal mathematical settings, it is noted that the correct setting for
pure states in quantum mechanics is projective Hilbert space endowed with

Quantum entanglement
417

the Fubini-Study metric. The product of two pure states is then given by the
Segre embedding.

Ensembles

FT
As mentioned above, a state of a quantum system is given by a unit vector in a
Hilbert space. More generally, if one has a large number of copies of the same
system, then the state of this ensemble is described by a density matrix, which
is a positive matrix (or trace class, when the state space is infinite dimensional)
and has trace 1. Again, by the spectral theorem, such a matrix takes the general
form:
P
ρ = i wi |αi ihαi |,

where the wi ’s sum up to 1 (in the infinite dimensional case, we would take the
closure of such states in the trace norm). We can interpret ρ as representing
an ensemble where wi is the proportion of the ensemble whose states are |αi i.
When a mixed state has rank 1, it therefore describes a pure ensemble. When
there is less than total information about the state of a quantum system we
A
need density matrices to represent the state (see experiment discussed below).
Following the definition in previous section, for a bipartite composite system,
mixed states are just density matrices on HA ⊗ HB . Extending the definition of
separability from the pure case, we say that a mixed state is separable if it can
be written as
ρ = i p i ρA B
P
i ⊗ ρi

,where ρA B
i ’s and ρi ’s are they themselves states on the subsystems A and B
DR
respectively. In other words, a state is separable if it is probability distribution
over uncorrelated states, or product states. We can assume without loss of gen-
erality that ρA B
i and ρi are pure ensembles. A state is then said to be entangled
if it is not separable. In general, finding out whether or not a mixed state is
entangled is considered difficult. Formally, it has been shown to be NP-hard.
For the 2 × 2 and 2 × 3 cases, a necessary and sufficient criterion for separability
is given by the famous PPT (Positive Partial Transpose) condition.
Experimentally, a mixed ensemble might be realized as follows. Consider a
"black-box" apparatus that spits electrons towards an observer. The electrons’
Hilbert spaces are identical. The apparatus might produce electrons that are all
in the same state; in this case, the electrons received by the observer are then
a pure ensemble. However, the apparatus could produce electrons in different
states. For example, it could produce two populations of electrons: one with
state |z+i (spins aligned in the positive z direction), and the other with state

Quantum entanglement
418

|y−i (spins aligned in the negative y direction.) Generally, there can be any
number of populations, each corresponding to a different state. Therefore we
now have a mixed ensemble.

Reduced Density Matrices

FT
Consider as above systems A and B each with a Hilbert space HA , HB . Let the
state of the composite system be
|Ψi ∈ HA ⊗ HB .

As indicated above, in general there is no way to associate a pure state to the


component system A. However, it still is possible to associate a density matrix.
Let
ρT = |Ψi hΨ|.

which is the projection operator onto this state. The state of A is the partial
trace of ρT over the basis of system B:
P
ρA ≡ j hj|B (|ΨihΨ|) |jiB = TrB ρT .
A
ρA is sometimes called the reduced density matrix of ρ on subsystem A. Collo-
quially, we "trace out" system B to obtain the reduced density matrix on A.
For example, the density matrix of A for the entangled state discussed above is
 
ρA = (1/2) |0iA h0|A + |1iA h1|A

This demonstrates that, as expected, the reduced density matrix for an entan-
DR
gled pure ensemble is a mixed ensemble. Also not surprisingly, the density
matrix of A for the pure product state |ψiA ⊗ |φiB discussed above is
ρA = |ψiA hψ|A .

Entropy
In this section we briefly discuss entropy of a mixed state and how it can be
viewed as a measure of entanglement.
In classical information theory, to a probability distribution p1 , · · · , pn , one can
associate the Shannon entropy:
P
H(p1 , · · · , pn ) = − i pi log pi ,

where the logarithm is taken in base 2.

Quantum entanglement
419

Since one can think of a mixed state ρ as a probability distribution over an


ensemble, this leads naturally to the definition of the von Neumann entropy:
S(ρ) = −Tr (ρ log ρ) ,

FT
where the logarithm is again taken in base 2. In general, to calculate log ρ,
one would use the Borel functional calculus. If ρ acts on a finite dimensional
Hilbert space and has eigenvalues λ1 , · · · , λn , then we recover the Shannon
entropy:
P
S(ρ) = −Tr (ρ log ρ) = i λi log λi .

Since an event of probability 0 should not contribute to the entropy, we adopt


the convention that 0 log 0 = 0. This extends to the infinite dimensional case
R
as well: if ρ has spectral resolution ρ = λdPλ , then we assume the same
convention when calculating
R
ρ log ρ = λ log λdPλ .

Entropy provides one tool which can be used to quantify entanglement (al-
A
though other entanglement measures exist). As in statistical mechanics, one
can say that the more uncertainty (number of microstates) possessed by the
system, the larger the entropy. For example, the entropy of any pure state is
zero, which is unsurprising since there is no uncertainty about a system in a
pure state. The entropy of any of the two subsystems of the entangled state
discussed above is ln 2 (which can be shown to be the maximum entropy for
2 × 2 mixed states). If the overall system is pure, the entropy of one subsystem
can be used to measure its degree of entanglement with the other subsystems.
DR
It turns out that, for pure states, the von Neumann entropy of reduced states
is the unique measure of entanglement. On the other hand, uniqueness does
not hold for mixed states. Physically speaking, this is because the uncertainty
in the mixed state gives us entropy in itself, irrespective of whether or not the
state is entangled.
As an aside, the information-theoretic definition is closely related to entropy in
the sense of statistical mechanics (comparing the two definitions, we note that,
in the present context, it is customary to set the Boltzmann constant k = 1).
For example, by properties of the Borel functional calculus, we see that for any
unitary operator U,
S(ρ) = S(U ρU ∗ ).

Quantum entanglement
420

Indeed, without the above property, the von Neumann entropy would not be
well-defined. In particular, U could be the time evolution operator of the sys-
tem, i.e.
 
U (t) = exp −iHt

FT
~

where H is the Hamiltonian of the system. This associates the reversibility of


a process with its resulting entropy change, i.e. a process is reversible if and
only if it leaves the entropy of the system invariant. This provides a connection
between quantum information theory and thermodynamics.

Applications of entanglement
Entanglement has many applications in quantum information theory. Mixed
state entanglement can be viewed as a resource for quantum communication.
With the aid of entanglement, otherwise impossible tasks may be achieved.
Among the most well known such applications of entanglement are super-
dense coding and quantum state teleportation. Efforts to quantify this resource
are often termed entanglement theory. See for example Entanglement Theory
A
Tutorials 222.
The Reeh-Schlieder theorem of quantum field theory is sometimes seen as the
QFT analogue of quantum entanglement.

See also
• →Entanglement witness
• Fubini-Study metric
DR
• →Separable states

• →Squashed entanglement

References
• M. Horodecki, P. Horodecki, R. Horodecki, "Separability of Mixed States:
Necessary and Sufficient Conditions", Physics Letters A 210, 1996.

• L. Gurvits, "Classical deterministic complexity of Edmonds’ Problem and


quantum entanglement", Proceedings of the thirty-fifth annual ACM sym-
posium on Theory of computing, 2003.

222 http://www.imperial.ac.uk/quantuminformation

Quantum entanglement
421

External links
• An Interview With Brian Clegg, Author of "The God Effect : Quantum En-
tanglement, Science’s Strangest Phenomenon" 223 California Literary Review

FT
• Multiple entanglement and quantum repeating 224
• How to entangle photons experimentally 225

Source: http://en.wikipedia.org/wiki/Quantum_entanglement

Principal Authors: Mct mht, CYD, Roadrunner, CSTAR, Caroline Thompson, Cortonin, Charles
Matthews

Quantum field theory

Quantum field theory (QFT) is the application of quantum mechanics to


fields. It provides a theoretical framework, widely used in particle physics
Aand condensed matter physics, in which to formulate consistent quantum the-
ories of many-particle systems, especially in situations where particles may be
created and destroyed. Non-relativistic quantum field theories are needed in
condensed matter physics— for example in the BCS theory of superconduc-
tivity. Relativistic quantum field theories are indispensable in particle physics
(see the standard model), although they are known to arise as effective field
theories in condensed matter physics.
DR
Origin
Quantum field theory originated in the problem of computing the energy radi-
ated by an atom when it dropped from one quantum state to another of lower
energy. This problem was first examined by Max Born and Pascual Jordan in
1925. In 1926, Max Born, Werner Heisenberg and Pascual Jordan wrote down
the quantum theory of the electromagnetic field neglecting polarization and
sources to obtain what would today be called a free field theory. In order to
quantize this theory, they used the canonical quantization procedure. In 1927,
Paul Dirac gave the first consistent treatment of this problem. Quantum field
theory followed unavoidably from a quantum treatment of the only known
classical field, ie, electromagnetism. The theory was required by the need to

223 http://calitreview.com/Interviews/clegg_8029.htm
224 http://www.physorg.com/news63037231.html
225 http://physicsweb.org/articles/world/11/3/9/1/world%2D11%2D3%2D9%2D3

Quantum field theory


422

treat a situation where the number of particles changes. Here, one atom in the
initial state becomes an atom and a photon in the final state.
It was obvious from the beginning that the quantum treatment of the electro-
magnetic field required a proper treatment of relativity. Jordan and Wolfgang

FT
Pauli showed in 1928 that commutators of the field were actually Lorentz in-
variant. By 1933, Niels Bohr and Leon Rosenfeld had related these commu-
tation relations to a limitation on the ability to measure fields at space-like
separation. The development of the Dirac equation and the hole theory drove
quantum field theory to explain these using the ideas of causality in relativi-
ty, work that was completed by Wendell Furry and Robert Oppenheimer using
methods developed for this purpose by Vladimir Fock. This need to put togeth-
er relativity and quantum mechanics was a second motivation which drove the
development of quantum field theory. This thread was crucial to the eventual
development of particle physics and the modern (partially) unified theory of
forces called the standard model.
In 1927 Jordan tried to extend the canonical quantization of fields to the wave
function which appeared in the quantum mechanics of particles, giving rise to
the equivalent name second quantization for this procedure. In 1928 Jordan
A
and Eugene Wigner found that the Pauli exclusion principle demanded that
the electron field be expanded using anti-commuting creation and annihilation
operators. This was the third thread in the development of quantum field
theory— the need to handle the statistics of multi-particle systems consistently
and with ease. This thread of development was incorporated into many-body
theory, and strongly influenced condensed matter physics and nuclear physics.

What QFT is
DR
Just as quantum mechanics deals with operators acting upon a (separable)
→Hilbert space, QFT also deals with operators acting upon a →Hilbert space.
However, in the case of QFT, the operators are generated by what is known as
operator-valued fields, that is, operators which are parametrized by a space-
time point. Intuitively, this means that operators can be localized. This defi-
nition applies even to the cases of theories which aren’t quantizations, and as
such, is pretty general.
This is sometimes stated as "position is an operator in QM but is a parameter
in QFT" but this statement, while accurate, can be very misleading. QM deals
with particles and one of the properties of a particle is its position as a func-
tion of time and in QM, this becomes the position operator as a function of
time (it’s constant in the Schrödinger picture and varying in the Heisenberg
picture). QFT, on the other hand, deals with fields on a fundamental level and
particles only emerge as localized excitations (aka quanta aka quasiparticles)

Quantum field theory


423

of the ground state (aka the vacuum) and it’s precisely these quantum fields
which correspond to the operator valued functions. Put more simply, instead
of looking at the operators generated by
x̂(t) and p̂(t),

FT
we now look at operators generated by
φ̂(x, t)

And just as in QM, we may work in the →Schrödinger picture, the


→Heisenberg picture or the interaction picture (in the context of perturbation
theory). Only the Heisenberg picture is manifestly Lorentz covariant.
The energy is given by the Hamiltonian operator, which can be generated from
the quantum fields, and corresponds to the generator of infinitesimal time
translations. (the condition that the generator of infinitesimal time translations
can be generated by the quantum fields rules out many unphysical theories,
which is a good thing) We further assume that this Hamiltonian is bounded
from below and has a lowest energy eigenstate (this rules out theories which
A
are unstable and have no stable solutions, which is also a good thing), which
may or may not be degenerate. (although there are physical QFTs which have
a lower bound to the Hamiltonian but don’t have a lowest energy eigenstate,
like N=1 super QCD theories with too few quarks...) This lowest energy eigen-
state is called the vacuum in particle physics and the ground state in condensed
matter physics. (QFT appears in the continuum limit of condensed matter sys-
tems)
This simple explanation of what QFT really is, is often obscured in treatments
DR
which jump straight to the path integral approach, which is a good computa-
tional technique but often obscures the underlying ideas.
QFT most definitely isn’t the same thing as classical field theory or classical
field theory with some "minor" quantum corrections, which is a mistake many
high energy physicists are prone to making at times, especially when working
in the semiclassical approximation.

Technical statement
Quantum field theory corrects several limitations of ordinary quantum mechan-
ics, which we will briefly discuss now. The →Schrödinger equation, in its most
commonly encountered form, is
h 2 i
|p| ∂
2m + V (r) |ψ(t)i = i~ ∂t |ψ(t)i

Quantum field theory


424

where |ψi denotes the quantum state (notation) of a particle with mass m, in
the presence of a potential V .
The first problem occurs when we seek to extend the equation to large num-
bers of particles. As described in the article on identical particles, quantum

FT
mechanical particles of the same species are indistinguishable, in the sense
that the state of the entire system must be symmetric (bosons) or antisymmet-
ric (fermions) when the coordinates of its constituent particles are exchanged.
These multi-particle states are extremely complicated to write. For example,
the general quantum state of a system of N bosons is written as
rQ
j
Nj ! P
|φ1 · · · φN i = N! p∈Sn |φp(1) i · · · |φp(N ) i

where |φi i are the single-particle states, Nj is the number of particles occupying
state j, and the sum is taken over all possible permutations p acting on N
elements. In general, this is a sum of N ! (N factorial) distinct terms, which
quickly becomes unmanageable as N increases. Large numbers of particles are
needed in condensed matter physics where typically the number of particles is
on the order of Avogadro’s number, approximately 10 23.
A
The second problem arises when trying to reconcile the Schrödinger equation
with special relativity. It is possible to modify the Schrödinger equation to in-
clude the rest energy of a particle, resulting in the →Klein-Gordon equation or
the Dirac equation. However, these equations have many unsatisfactory quali-
ties; for instance, they possess energy eigenvalues which extend to –∞, so that
there seems to be no easy definition of a ground state. Such inconsistencies
occur because these equations neglect the possibility of dynamically creating
or destroying particles, which is a crucial aspect of relativity. Einstein’s famous
DR
mass-energy relation predicts that sufficiently massive particles can decay in-
to 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. Such processes must be accounted for in a truly
relativistic quantum theory. This problem brings to the fore the notion that a
consistent relativistic quantum theory, even of a single particle, must be a many
particle theory.

Quantizing a classical field theory


Canonical quantization
Quantum field theory solves these problems by consistently quantizing a field.
By interpreting the physical observables of the field appropriately, one can cre-
ate a (rather successful) theory of many particles. Here is how it is:

Quantum field theory


425

1. Each normal mode oscillation of the field is interpreted as a particle with


frequency f.
2. The quantum number n of each normal mode (which can be thought of as a
harmonic oscillator) is interpreted as the number of particles.

FT
The energy associated with the mode of excitation is therefore = (n + 1/2)~ω
which directly follows from the energy eigenvalues of a one dimensional har-
monic oscillator in quantum mechanics. With some thought, one may similarly
associate momenta and position of particles with observables of the field.
Having cleared up the correspondence between fields and particles (which is
different from non-relativistic QM), we can proceed to define how a quantum
field behaves.
Two caveats should be made before proceeding further:

• Each of these "particles" obeys the usual uncertainty principle of quantum


mechanics. The "field" is an operator defined at each point of spacetime.
• Quantum field theory is not a wildly new theory. Classical field theory is the
same as classical mechanics of an infinite number of dynamical quantities
A (say, tiny elements of rubber on a rubber sheet). Quantum field theory is
the quantum mechanics of this infinite system.

The first method used to quantize field theory was the method now called
canonical quantization (earlier known as second quantization). This method
uses a Hamiltonian formulation of the classical problem. The later technique
of Feynman path integrals uses a Lagrangian formulation. Many more methods
are now in use; for an overview see the article on quantization.
DR
Canonical quantization for bosons
Suppose we have a system of N bosons which can occupy mutually orthogonal
single-particle states |φ1 i, |φ2 i, |φ3 i, and so on. The usual method of writing
a multi-particle state is to assign a state to each particle and then impose ex-
change symmetry. As we have seen, the resulting wavefunction is an unwieldy
sum of N ! terms. In contrast, in the second quantized approach we will simply
list the number of particles in each of the single-particle states, with the un-
derstanding that the multi-particle wavefunction is symmetric. To be specific,
suppose that N = 3, with one particle in state |φ1 i and two in state|φ2 i. The
normal way of writing the wavefunction is
√1 [|φ1 i|φ2 i|φ2 i + |φ2 i|φ1 i|φ2 i + |φ2 i|φ2 i|φ1 i]
3

In second quantized form, we write this as

Quantum field theory


426

|1, 2, 0, 0, 0, · · ·i

which means "one particle in state 1, two particles in state 2, and zero particles
in all the other states."

FT
Though the difference is entirely notational, the latter form makes it easy for us
to define creation and annihilation operators, which add and subtract particles
from multi-particle states. These creation and annihilation operators are very
similar to those defined for the quantum harmonic oscillator, which added and
subtracted energy quanta. However, these operators literally create and anni-
hilate particles with a given quantum state. The bosonic annihilation operator
a2 and creation operator a†2 have the following effects:

a2 |N1 , N2 , N3 , · · ·i = N2 | N1 , (N2 − 1), N3 , · · ·i


a†2 |N1 , N2 , N3 , · · ·i = N2 + 1 | N1 , (N2 + 1), N3 , · · ·i

We may well ask whether these are operators in the usual quantum mechanical
sense, i.e. linear operators acting on an abstract →Hilbert space. In fact, the
A
answer is yes: they are operators acting on a kind of expanded Hilbert space,
known as a →Fock space, composed of the space of a system with no particles
(the so-called vacuum state), plus the space of a 1-particle system, plus the
space of a 2-particle system, and so forth. Furthermore, the creation and anni-
hilation operators are indeed Hermitian conjugates, which justifies the way we
have written them.
The bosonic creation and annihilation operators obey the commutation relation
h i h i
a†i , a†j = 0 , ai , a†j = δij
 
DR
ai , aj = 0 ,

where δ 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 quan-
tum of energy in a harmonic oscillator.
The final step toward obtaining a quantum field theory is to re-write our orig-
inal N -particle Hamiltonian in terms of creation and annihilation operators
acting on a Fock space. For instance, the Hamiltonian of a field of free (non-
interacting) bosons is
H = k Ek a†k ak
P

where Ek is the energy of the k-th single-particle energy eigenstate. Note that

Quantum field theory


427

a†k ak | · · · , Nk , · · ·i = Nk | · · · , Nk , · · ·i.

Canonical quantization for fermions

FT
It turns out that the creation and annihilation operators for fermions must
be defined differently, in order to satisfy the Pauli exclusion principle. For
fermions, the occupation numbers Ni can only take on the value 0 or 1, since
particles cannot share quantum states. We then define the fermionic annihila-
tion operators c and creation operators c† by
cj |N1 , N2 , · · · , Nj = 0, · · ·i = 0

cj |N1 , N2 , · · · , Nj = 1, · · ·i = (−1)(N1 +···+Nj−1 ) |N1 , N2 , · · · , Nj = 0, · · ·i

c†j |N1 , N2 , · · · , Nj = 0, · · ·i = (−1)(N1 +···+Nj−1 ) |N1 , N2 , · · · , Nj = 1, · · ·i

c†j |N1 , N2 , · · · , Nj = 1, · · ·i = 0
A
The fermionic creation and annihilation operators obey an anticommutation
relation,
n o n o
c†i , c†j = 0 , ci , c†j = δij

ci , cj = 0 ,

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.
DR
Significance of creation and annihilation operators
When we re-write a Hamiltonian using a Fock space and creation and annihi-
lation operators, as in the previous example, the symbol N , which stands for
the total number of particles, drops out. This means that the Hamiltonian is
applicable to systems with any number of particles. Of course, in many com-
mon situations N is a physically important and perfectly well-defined quantity.
For instance, if we are describing a gas of atoms sealed in a box, the number
of atoms had better remain a constant at all times. This is certainly true for
the above Hamiltonian. Viewing the Hamiltonian as the generator of time evo-
lution, we see that whenever an annihilation operator ak destroys a particle
during an infinitesimal time step, the creation operator a†k to the left of it in-
stantly puts it back. Therefore, if we start with a state of N non-interacting
particles then we will always have N particles at a later time.

Quantum field theory


428

On the other hand, it is often useful to consider quantum states where the
particle number is ill-defined, i.e. linear superpositions of vectors from the
Fock space that possess different values of N . For instance, it may happen that
our bosonic particles can be created or destroyed by interactions with a field
of fermions. Denoting the fermionic creation and annihilation operators by c†k

FT
and ck , we could add a "potential energy" term to our Hamiltonian such as:
V = k,q Vq (aq + a†−q )c†k+q ck
P

This describes processes in which a fermion in state k either absorbs or emits a


boson, thereby being kicked into a different eigenstate k + q. In fact, this is the
expression for the interaction between phonons and conduction electrons in a
solid. The interaction between photons and electrons is treated in a similar
way; it 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 generally end up with a superposition of states with
different numbers of bosons at later times. On the other hand, the number of
fermions is conserved in this case.
A
In condensed matter physics, states with ill-defined particle numbers are also
very 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.

Field operators
We can now define field operators that create or destroy a particle at a particu-
lar point in space. In particle physics, these are often more convenient to work
DR
with than the creation and annihilation operators, 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 φ(r) is
φ(r) ≡ i eiki ·r ai
P

The bosonic field operators obey the commutation relation


[φ(r), φ(r0 )] = 0 ,
 †
φ (r), φ† (r0 ) = 0 , φ(r), φ† (r0 ) = δ 3 (r − r0 )
  

where δ(x) stands for the Dirac delta function. As before, the fermionic rela-
tions are the same, with the commutators replaced by anticommutators.

Quantum field theory


429

It should be emphasized that 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 just a scalar field. However, they are closely related, and are indeed
commonly denoted with the same symbol. If we have a Hamiltonian with a

FT
space representation, say
~2 P 2 P
H = − 2m i ∇i + i<j U (|ri − rj |)

where the indices i and j run over all particles, then the field theory Hamilto-
nian is
~2
d r φ(r)† ∇2 φ(r) + d3r d3r0 φ(r)† φ(r0 )† U (|r − r0 |)φ(r0 )φ(r)
R 3 R R
H = − 2m

This looks remarkably like an expression for the expectation value of the ener-
gy, with φ 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.

Quantization of classical fields


A
So far, we have shown how one goes from an ordinary quantum theory to a
quantum field theory. There are certain systems for which no ordinary quan-
tum theory exists. These are the "classical" fields, such as the electromagnetic
field. There is no such thing as a wavefunction for a single photon in classical
electromagnetisim, so a quantum field theory must be formulated right from
the start.
The essential difference between an ordinary system of particles and the elec-
tromagnetic field is the number of dynamical degrees of freedom. For a system
DR
of N particles, there are 3N coordinate variables corresponding to the posi-
tion of each particle, and 3N conjugate momentum variables. One formulates
a classical Hamiltonian using these variables, and obtains a quantum theory
by turning the coordinate and position variables into quantum operators, and
postulating commutation relations between them such as
 
qi , pj = δij

For an electromagnetic field, the analogue of the coordinate variables are the
values of the electrical potential φ(x) and the vector potential A(x) at every
point x. This is an uncountable set of variables, because x is continuous. This
prevents us from postulating the same commutation relation as before. The
way out is to replace the Kronecker delta with a Dirac delta function. This ends
up giving us a commutation relation exactly like the one for field operators!
We therefore end up treating "fields" and "particles" in the same way, using

Quantum field theory


430

the apparatus of quantum field theory. Only by accident electrons were not
regarded as de Broglie waves and photons governed by geometrical optics were
not the dominant theory when QFT was developed.

Path integral methods

FT
The axiomatic approach
There have been many attempts to put quantum field theory on a firm mathe-
matical footing by formulating a set of axioms for it. These attempts fall into
two broad classes.
The first class of axioms (most notably the Wightman, Osterwalder-Schrader,
and Haag-Kastler systems) tried to formalize the physicists’ notion of an
"operator-valued field" within the context of functional analysis. These ax-
ioms enjoyed limited success. It was possible to prove that any QFT satisfying
these axioms satisfied certain general theorems, such as the spin-statistics theo-
rem and the PCT theorems. Unfortunately, it proved extraordinarily difficult to
show that any realistic field theory (e.g. quantum chromodynamics) satisfied
these axioms. Most of the theories which could be treated with these analyt-
A
ic axioms were physically trivial: restricted to low-dimensions and lacking in
interesting dynamics. Constructive quantum field theory is the construction of
theories which satisfy one of these sets of axioms. Important work was done in
this area in the 1970s by Segal, Glimm, Jaffe and others.
In the 1980s, a second wave of axioms were proposed. These axioms (associat-
ed most closely with Atiyah and Segal, and notably expanded upon by Witten,
Borcherds, and Kontsevich) are more geometric in nature, and more closely
resemble the path integrals of physics. They have not been exceptionally use-
DR
ful to physicists, as it is still extraordinarily difficult to show that any realistic
QFTs satisfy these axioms, but have found many applications in mathematics,
particularly in representation theory, algebraic topology, and geometry.
Finding the proper axioms for quantum field theory is still an open and diffi-
cult problem in mathematics. In fact, one of the Clay Millennium Prizes offers
$1,000,000 to anyone who proves the existence of a mass gap in Yang-Mills
theory. It seems likely that we have not yet understood the underlying struc-
tures which permit the Feynman path integrals to exist.

Renormalization
Some of the problems and phenomena eventually addressed by renormaliza-
tion actually appeared earlier in the classical electrodynamics of point particles
in the 19th and early 20th century. The basic problem is that the observable
properties of an interacting particle cannot be entirely separated from the field

Quantum field theory


431

A FT
Figure 39 Edward Witten

that mediates the interaction. The standard classical example is the energy of a
charged particle. To cram a finite amount of charge into a single point requires
an infinite amount of energy; this manifests itself as the infinite energy of the
particle’s electric field. The energy density grows to infinity as one gets close
DR
to the charge.
A single particle state in quantum field theory incorporates within it multipar-
ticle states. This is most simply demonstrated by examining the evolution of a
single particle state in the interaction picture—
|ψ(t)i = eiHI t |ψ(0)i = 1 + iHI t − 12 HI2 t2 − 3!i HI3 t3 + 4!1 HI4 t4 + · · · |ψ(0)i.
 

Taking the overlap with the initial state, one retains the even powers of H I.
These terms are responsible for changing the number of particles during prop-
agation, and are therefore quintessentially a product of quantum field theory.
Corrections such as these are incorporated into wave-function renormalization

Quantum field theory


432

and mass renormalization. Similar corrections to the interaction Hamiltoni-


an, H I, include vertex renormalization, or, in modern language, effective field
theory.

FT
Gauge theories
A gauge theory is a theory which 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, so 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 electrodynam-
ics, the theory is also invariant under a local change of phase, that is - one may
shift the phase of all wave functions so that in every point in space-time the
shift is different. 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 effect the derivative. In quantum electrodynamics this
gauge field is the electromagnetic field. The change of local change of variables
is termed gauge transformation.
A
In quantum field theory the excitations of fields represent particles. The parti-
cle 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
DR
at all, and they therefore have no physical meaning. Such fluctuations are usu-
ally 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 anoma-
ly. 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 intercation, making the theory non unitary and again
inconsistent (see optical theorem).

Quantum field theory


433

In general, the gauge transformations of a theory consist several different trans-


formations, which may not be commutative. These transformations are togeth-
er described by a mathematical object known as a gauge group. Infinitesimal
gauge transformations are the gauge group generators. Therefore the num-

FT
ber of gauge bosons is the group rank (i.e. number of generators forming an
orthogonal basis).
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 U (1)×SU (2) (a direct prod-
uct of U(1) and SU(2)).
• Gravity, whose classical theory is general relativity, admits the equivalence
principle which is a form of gauge symmetry.
A
Supersymmetry
Supersymmetry assumes that every fundamental fermion has a superpartner
which 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
DR
consistency of string theory.
The way supersymmetry protects the hierarchies is the following: since for ev-
ery 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
ruinning 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.

Quantum field theory


434

Beyond local field theory

History

FT
More details can be found in the article on the history of quantum field theory.
Quantum field theory was created by Dirac when he attempted to quantize
the electromagnetic field in the late 1920s. The early development of the field
involved Fock, Jordan, Pauli, Heisenberg, Bethe, Tomonaga, Schwinger, Feyn-
man, and Dyson. This phase of development culminated with the construction
of the theory of quantum electrodynamics in the 1950s.
Gauge theory was formulated and quantized, leading to the unification of
forces embodied in the standard model of particle physics. This effort started
in the 1950s with the work of Yang and Mills, was carried on by Martinus
Veltman and a host of others during the 1960s and completed during the 1970s
by the work of Gerard ’t Hooft, Frank Wilczek, David Gross and David Politzer.
Parallel developments in the understanding of phase transitions in condensed
matter physics led to the study of the renormalization group. This in turn led
to the grand synthesis of theoretical physics which unified theories of particle
A
and condensed matter physics through quantum field theory. This involved the
work of Michael Fisher and Leo Kadanoff in the 1970s which led to the seminal
reformulation of quantum field theory by Kenneth Wilson.
The study of quantum field theory is alive and flourishing, as are applications of
this method to many physical problems. It remains one of the most vital areas
of theoretical physics today, providing a common language to many branches
of physics.
DR
See also
• List of quantum field theories
• Feynman path integral
• Quantum chromodynamics
• Quantum electrodynamics
• Schwinger-Dyson equation
• Relationship between string theory and quantum field theory
• Abraham-Lorentz force

Quantum field theory


435

Suggested reading
• Wilczek, Frank ; Quantum Field Theory, Review of Modern Physics 71
(1999) S85-S95. Review article written by a master of Q.C.D., Nobel laure-

FT
ate 2003 226. Full text available at : hep-th/9803075 227

• Ryder, Lewis H. ; Quantum Field Theory (Cambridge University Press,


1985), [ISBN 0-521-33859-X] Highly readable textbook, certainly the best
introduction to relativistic Q.F.T. for particle physics.

• Zee, Anthony ; Quantum Field Theory in a Nutshell, Princeton University


Press (2003) [ISBN 0-691-01019-6].
• Peskin, M and Schroeder, D. ;An Introduction to Quantum Field Theory
(Westview Press, 1995) [ISBN 0201503972]

• Weinberg, Steven ; The Quantum Theory of Fields (3 volumes) Cambridge


University Press (1995). A monumental treatise on Q.F.T. written by a
leading expert, Nobel laureate 1979 228.
A• Loudon, Rodney ; The Quantum Theory of Light (Oxford University Press,
1983), [ISBN 0198511558]

• D.A. Bromley (2000). Gauge Theory of Weak Interactions. Springer. ISBN


3540676724.

• Gordon L. Kane (1987). Modern Elementary Particle Physics. Perseus


Books. ISBN 0201117495.
DR
External links
• Siegel, Warren ; Fields 229 (also available from arXiv:hep-th/9912205)
• ’t Hooft, Gerard ; The Conceptual Basis of Quantum Field Theory, Hand-
book of the Philosophy of Science, Elsevier (to be published). Review article
written by a master of gauge theories, [http://nobelprize.org/physics/lau-
reates/1999/thooft-autobio.html’’Nobel laureate 1999]. Full text available
in 230.

226 http://nobelprize.org/physics/laureates/2004/wilczek-autobio.html
227 http://fr.arxiv.org/abs/hep-th/9803075
228 http://nobelprize.org/physics/laureates/1979/weinberg-lecture.html
229 http://insti.physics.sunysb.edu/%7Esiegel/errata.html
230 http://www.phys.uu.nl/~thooft/lectures/basisqft.pdf’’pdf’’

Quantum field theory


436

• Srednicki, Mark ; Quantum Field Theory 231


• Kuhlmann, Meinard ; Quantum Field Theory 232, Stanford Encylopedia
of Philosophy

FT
Source: http://en.wikipedia.org/wiki/Quantum_field_theory

Principal Authors: CYD, Bambaiah, Stupidmoron, Phys, Odddmonster, Lethe, Arnero, Charles
Matthews, Itinerant1, AmarChandra

Quantum fluctuation

In quantum physics, a quantum fluctuation is the temporary change in the


amount of energy in a point in space, arising from Werner Heisenberg’s uncer-
tainty principle.
According to one formulation of the principle, energy and time can be related
by the relation
A ∆E∆t ≈ h

That means that conservation of energy can appear to be violated, but only
for small times. This allows the creation of particle-antiparticle pairs of virtual
particles. The effects of these particles are measurable, for example, in the
effective charge of the electron, different from its "naked" charge.
In the modern view, energy is always conserved, but the eigenstates of the
DR
Hamiltonian (energy observable) aren’t the same as (don’t commute with) the
particle number operators.
Quantum fluctuations may have been very important in the origin of the struc-
ture of the universe: according to the model of inflation the ones that existed
when inflation began were amplified and formed the seed of all current ob-
served structure.

Quantum fluctuations of a field


A reasonably clear distinction can be made between quantum fluctuations and
thermal fluctuations of a quantum field (at least for a free field; for interacting

231 http://gabriel.physics.ucsb.edu/~mark/qft.html
232 http://plato.stanford.edu/entries/quantum-field-theory/

Quantum fluctuation
437

fields, renormalization complicates matters a lot). For the quantized Klein-


Gordon field, we can calculate the probability density that we would observe a
configuration ϕt (x) at a time t in terms of its fourier transform ϕ̃t (k) to be
h R d3 k ∗ p i
ρ0 [ϕt ] = exp − ~1 (2π) 3 ϕ̃t (k) |k|2 + m2 ϕ̃t (k) .

FT
In contrast, for the classical Klein-Gordon field at non-zero temperature, the
Gibbs probability density that we would observe a configuration ϕt (x) at a
time t is
h R d3 k ∗ i
1 2 2
ρE [ϕt ] = exp [−H[ϕt ]/kT ] = exp − kT (2π)3 ϕ̃t (k) 2 (|k| + m ) ϕ̃t (k) .
1

The amplitude of quantum fluctuations is controlled by the amplitude of


Planck’s constant ~, just as the amplitude of thermal fluctuations is controlled
by kT . Note that the following three points are closely related:
(1) Planck’s constant has units
p of action instead of units of energy,
(2) the quantum kernel is |k|2 + m2 instead of 21 (|k|2 + m2 ) (the quantum
kernel is nonlocal from a classical heat kernel viewpoint, but it is local in the
sense that it does not allow signals to be transmitted),
A
(3) the quantum vacuum state is Lorentz invariant (although not manifestly in
the above), whereas the classical thermal state is not (the classical dynamics is
Lorentz invariant, but the Gibbs probability density is not a Lorentz invariant
initial condition).
We can construct a classical continuous random field that has the same prob-
ability density as the quantum vacuum state, so that the principal difference
from quantum field theory is the measurement theory (measurement in quan-
tum theory is different from measurement for a classical continuous random
DR
field, in that classical measurements are always mutually compatible — in
quantum mechanical terms they always commute). Quantum effects that are
consequences only of quantum fluctuations, not of subtleties of measurement
incompatibility, can alternatively be modelled by classical continuous random
fields.

Quantum annealing: A novel utilization of quan-


tum fluctuations
Quantum fluctuations are recently being used to anneal glassy systems (physi-
cal glass, or equivalently, hard combinatorial optimization problem with ragged
energy/cost landscape) to their minimal/ground states. Thus it provides a gen-
eral algorithmic scheme for classical/quantum computers.

Quantum fluctuation
438

See also
• Casimir effect
• Virtual particle

FT
• Quantum annealing

Source: http://en.wikipedia.org/wiki/Quantum_fluctuation

Principal Authors: Arnab das, AstroNomer, Eequor, Michael Hardy, RoboDick, Phys, Nowhither, Andre
Engels

Quantum foam

Quantum foam, also referred to as spacetime foam, is a concept in quantum


mechanics, devised by John Wheeler in 1955. It is sometimes likened to the
old concept of the ether/aether.
A
The foam is a qualitative description of the turbulence that the phenomenon
creates at extremely small distances of the order of the Planck length. At such
small scales of time and space the uncertainty principle allows particles and
energy to briefly come into existence, and then annihilate, without violating
conservation laws. As the scale of time and space being discussed shrinks, the
energy of the virtual particles increases. At sufficiently small scale space is not
smooth as would be expected from observations at larger scales.
DR
Foaming through the universe
Quantum foam is theorized to create masses of virtual particles. They are
particle-antiparticle pairs, and prior to their annihilation, exist for a short pe-
riod of time, on the order of the Planck time. They are created randomly from
photons; the higher the energy of the photon from which they are created, the
longer the time they will exist prior to annihilation.
These virtual particles make their existence known by the Casimir effect. It is
thought that there are constant quantum fluctuations in "empty" space, even at
the energetic homogeneity referred to as absolute zero. Due to this, quantum
fluctuations are often described using the term "zero-point energy".
The "foamy" spacetime would look like a complex turbulent storm-tossed sea.
Some physicists theorize the formation of wormholes therein; speculation aris-
ing from this includes the possibility of hyperspatial links to other universes.
As far as realistic phenomena are concerned, it’s thought that the hyperspatial

Quantum foam
439

nature of the quantum foam may account for such diverse physical principles
as inertia, propagation of light, and time flow.
Reginald Cahill has developed a theory called Process physics, which describes
space as a quantum foam system in which gravity is an inhomogeneous flow

FT
of the quantum foam into matter. According to this theory, the so-called spiral
galaxy rotation-velocity anomaly may be explained without the need for dark
matter.
Various scientists have theorized that quantum foam is an incredibly powerful
source of zero-point energy. It has been estimated that one cubic centime-
ter of space contains enough zero point energy to boil all the world’s oceans.
However, estimates of this energy vary widely due to the huge disparity in
the calculations of the quantum foam density, which vary more than 1:10 100.
Physicist Michio Kaku thinks that this enormous uncertainty in the estimation
of quantum-foam density would represent the largest disparity for any quantity
in all of physics.

See also
A•

Dirac sea
Hawking radiation
• Hyperspace theory
• Planck time
• "Rolling ball" topology
• Vacuum energy
• Wormhole
DR
References
• John Archibald Wheeler with Kenneth Ford. Geons, Black Holes, and Quan-
tum Foam. 1995.
• Reginald T. Cahill. Gravity as Quantum Foam In-Flow. June 2003. 233
• Process Physics 234 Resource Index

Source: http://en.wikipedia.org/wiki/Quantum_foam

Principal Authors: GregorB, Peak, ErkDemon, Stevertigo, Platypus222

233 http://www.scieng.flinders.edu.au/cpes/people/cahill_r/HPS15.pdf
234 http://www.mountainman.com.au/process_physics/

Quantum foam
440

Quantum Hall effect

The quantum Hall effect is a quantum-mechanical version of the Hall effect,

FT
observed in two-dimensional electron systems subjected to low temperatures
and strong magnetic fields, in which the Hall conductance σ takes on the quan-
tized values
e2
σ=ν h,

where e is the elementary charge and h is Planck’s constant. In the "ordinary"


quantum Hall effect, known as the integer quantum Hall effect, ν takes on
integer values (ν = 1, 2, 3, etc.). There is another type of quantum Hall effect,
known as the fractional quantum Hall effect, in which ν can occur as a vulgar
fraction (ν = 2/7, 1/3, 2/5, 3/5, 5/2 etc.)
The quantization of the Hall conductance has the important property of being
incredibly precise. Actual measurements of the Hall conductance have been
found to be integer or fractional multiples of e 2/h to nearly one part in a bil-
Alion. This phenomenon, referred to as "exact quantization", has been shown
to be a subtle manifestation of the principle of gauge invariance. It has al-
lowed for the definition of a new practical standard for electrical resistance:
the resistance unit h/e 2, roughly equal to 25 812.8 ohms, is referred to as the
von Klitzing constant R K 235 (after Klaus von Klitzing, the discoverer of exact
quantization) and since 1990, a fixed conventional value R K-90 236 is used in
resistance calibrations worldwide. The quantum Hall effect also provides an
extremely precise independent determination of the fine structure constant, a
DR
quantity of fundamental importance in quantum electrodynamics.
The integer quantization of the Hall conductance was originally predicted by
Ando, Matsumoto, and Uemura in 1975, on the basis of an approximate calcu-
lation. Several workers subsequently observed the effect in experiments carried
out on the inversion layer of MOSFETs. It was only in 1980 that von Klitzing,
working with samples developed by Michael Pepper and Gerhard Dorda, made
the totally unexpected discovery that the Hall conductivity was exactly quan-
tized. For this finding, von Klitzing was awarded the 1985 Nobel Prize in
Physics. The link between exact quantization and gauge invariance was subse-
quently found by Robert Laughlin.
The fractional effect is due to completely different physics, and was experi-
mentally discovered in 1982 by Daniel Tsui and Horst Störmer, in experiments

235 http://physics.nist.gov/cgi-bin/cuu/Value?rk|search_for=RK
236 http://physics.nist.gov/cgi-bin/cuu/Value?rk90|search_for=RK

Quantum Hall effect


441

performed on gallium arsenide heterostructures developed by Arthur Gossard.


The effect was explained by Robert B. Laughlin in 1983, using a novel quan-
tum liquid phase that accounts for the effects of interactions between electrons.
Tsui, Störmer, and Laughlin were awarded the 1998 Nobel Prize for their work.

FT
Although it was generally assumed that the discrete resistivity jumps found in
the Tsui experiment were due to the presence of fractional charges, it was not
until 1997 that R. de-Picciotto, et. al., indirectly observed fractional charges
through measurements of quantum shot noise. The fractional quantum hall
effect continues to be influential in theories about topological order.

References
• T. Ando, Y. Matsumoto, and Y. Uemura, J. Phys. Soc. Jpn. 39, 279 (1975)
DOI: 10.1143/JPSJ.39.279 237
• K. von Klitzing, G. Dorda, and M. Pepper, Phys. Rev. Lett. 45, 494 (1980)
DOI: 10.1103/PhysRevLett.45.494 238
• R.B. Laughlin, Phys. Rev. B. 23, 5632 (1981) DOI:
10.1103/PhysRevB.23.5632 239
A• D.C. Tsui, H.L. Stormer, and A.C. Gossard, Phys. Rev. Lett. 48, 1559 (1982)
DOI: 10.1103/PhysRevLett.48.1559 240
• R.B. Laughlin, Phys. Rev. Lett. 50, 1395 (1983) DOI:
10.1103/PhysRevLett.50.1395 241
• R. de-Picciotto, M. Reznikov, M. Heiblum, V. Umansky, G. Bunin and D.
Mahalu, Nature 389, 162-164 (1997)
DR
Source: http://en.wikipedia.org/wiki/Quantum_Hall_effect

Principal Authors: CYD, Michael Hardy, Jaraalbe, Shaddack, Glenn

237 http://dx.doi.org/10.1143/JPSJ.39.279
238 http://dx.doi.org/10.1103/PhysRevLett.45.494
239 http://dx.doi.org/10.1103/PhysRevB.23.5632
240 http://dx.doi.org/10.1103/PhysRevLett.48.1559
241 http://dx.doi.org/10.1103/PhysRevLett.50.1395

Quantum Hall effect


442

Quantum harmonic oscillator

The quantum harmonic oscillator is the quantum mechanical analogue of

FT
the classical harmonic oscillator. It is one of the most important model systems
in quantum mechanics because, as in classical mechanics, a wide variety of
physical situations can be reduced to it either exactly or approximately. In
particular, a system near an equilibrium configuration can often be described
in terms of one or more harmonic oscillators. Furthermore, it is one of the few
quantum mechanical systems for which a simple exact solution is known.
The following discussion of the quantum harmonic oscillator relies on the arti-
cle mathematical formulation of quantum mechanics.

One-dimensional harmonic oscillator

Diatomic molecules
242
In diatomic molecules, the natural frequency can be found by:
A q
ω = mkr

where
ω = 2πf is the angular frequency,

k is the bond force constant, and


DR
mr is the reduced mass.

Hamiltonian and energy eigenstates


In the one-dimensional harmonic oscillator problem, a particle of mass m is
subject to a potential V (x) = (1/2)mω 2 x 2. The Hamiltonian of the particle is:
p2
H= 2m + 21 mω 2 x2

where x is the position operator, and p is the momentum operator (p =



−i~ ∂x ). The first term represents the kinetic energy of the particle, and the
second term represents the potential energy in which it resides. In order to

242 http://hyperphysics.phy-astr.gsu.edu/hbase/quantum/hosc.html

Quantum harmonic oscillator


443

A FT
Figure 40 Wavefunction representa-
tions for the first six bound eigenstates,
n = 0 to 5. The horizontal axis shows
the position x. The graphs are not nor-
malised
DR
find the energy levels and the corresponding energy eigenstates, we must solve
the time-independent →Schrödinger equation,
H |ψi = E |ψi.

We can solve the differential equation in the coordinate basis, using a power
series method. It turns out that there is a family of solutions,
1/4  
mωx2
hx|ψn i = √ 1n · mω
p mω 
π~ · exp − 2~ · Hn ~ x
2 n!

n = 0, 1, 2, . . .

The first six solutions (n = 0 to 5) are shown on the right. The functions Hn
are the Hermite polynomials:

Quantum harmonic oscillator


444

FT
Figure 41 Probability densities |ψ n(x)| 2 for the bound
eigenstates, beginning with the ground state (n = 0) at
the bottom and increasing in energy toward the top. The
horizontal axis shows the position x, and brighter colors
represent higher probability densities.

2 dn −x2
Hn (x) = (−1)n ex dxn e
A
They should not be confused with the Hamiltonian, which is also denoted by
H. The corresponding energy levels are
En = ~ω n + 21 .


This energy spectrum is noteworthy for two reasons. Firstly, the energies are
"quantized", and may only take the discrete values of ~ω times 1/2, 3/2, 5/2,
and so forth. This is a feature of many quantum mechanical systems. In the
DR
following section on ladder operators, we will engage in a more detailed ex-
amination of this phenomenon. Secondly, the lowest achievable energy is not
zero, but ~ω/2, which is called the "ground state energy" or zero-point energy.
It is not obvious that this is significant, because normally the zero of energy is
not a physically meaningful quantity, only differences in energies. Neverthe-
less, the ground state energy has many implications, particularly in quantum
gravity.
Note that the ground state probability density is concentrated at the origin.
This means the particle spends most of its time at the bottom of the potential
well, as we would expect for a state with little energy. As the energy increases,
the probability density becomes concentrated at the "classical turning points",
where the state’s energy coincides with the potential energy. This is consistent
with the classical harmonic oscillator, in which the particle spends most of its
time (and is therefore most likely to be found) at the turning points, where it
is the slowest. The correspondence principle is thus satisfied.

Quantum harmonic oscillator


445

Ladder operator method


The power series solution, though straightforward, is rather tedious. The "lad-
der operator" method, due to Paul Dirac, allows us to extract the energy eigen-
values without directly solving the differential equation. Furthermore, it is

FT
readily generalizable to more complicated problems, notably in quantum field
theory. Following this approach, we define the operators a and its adjoint a †
p mω i

a = 2~ x + mω p
a†
p mω i
= 2~ x − mω p

The operator a is not Hermitian since it and its adjoint a † are not equal.
In deriving the form of a †, we have used the fact that the operators x and p,
which represent observables, are Hermitian. These observable operators can
be expressed as a linear combination of the ladder operators as
q
~
a† + a

x =
q2mω
~mω
a† − a

p = i 2
A
The x and p operators obey the following identity, known as the canonical
commutation relation:
[x, p] = i~.

The square brackets in this equation are a commonly-used notational device,


known as the commutator, defined as
DR
[A, B] ≡ AB − BA.

Using the above, we can prove the identities


H = ~ω a† a + 1/2


a, a† = 1.
 

Now, let |ψE i denote an energy eigenstate with energy E. The inner product of
any ket with itself must be non-negative, so
(a |ψE i , a |ψE i) = hψE | a† a |ψE i ≥ 0.

Expressing a †a in terms of the Hamiltonian:

Quantum harmonic oscillator


446
 
H 1 E 1
hψE | ~ω − 2 |ψE i = ~ω − 2 ≥ 0,

so that E ≥ ~ω/2. Note that when (a |ψE i) is the zero ket (i.e. a ket with length
zero), the inequality is saturated, so that E = ~ω/2. It is straightforward to

FT
check that there exists a state satisfying this condition; it is the ground (n = 0)
state given in the preceding section.
Using the above identities, we can now show that the commutation relations
of a and a † with H are:
[H, a]†  = −~ωa
H, a = ~ωa†
.

Thus, provided (a |ψE i) is not the zero ket,


H(a |ψE i) = ([H, a] + aH) |ψE i
= (−~ωa + aE) |ψE i
= (E − ~ω)(a |ψE i)
.
A
Similarly, we can show that
H(a† |ψE i) = (E + ~ω)(a† |ψE i).

In other words, a acts on an eigenstate of energy E to produce, up to a mul-


tiplicative constant, another eigenstate of energy E − ~ω, and a † acts on an
eigenstate of energy E to produce an eigenstate of energy E + ~ω. For this
DR
reason, a is called a "lowering operator", and a † a "raising operator". The two
operators together are called "ladder operators". In quantum field theory, a and
a † are alternatively called "annihilation" and "creation" operators because they
destroy and create particles, which correspond to our quanta of energy.
Given any energy eigenstate, we can act on it with the lowering operator, a, to
produce another eigenstate with ~ω, less energy. By repeated application of the
lowering operator, it seems that we can produce energy eigenstates down to E
= -∞. However, this would contradict our earlier requirement that E ≥ ~ω/2.
Therefore, there must be a ground-state energy eigenstate, which we label |0i
(not to be confused with the zero ket), such that
a |0i = 0(zero ket).

Quantum harmonic oscillator


447

In this case, subsequent applications of the lowering operator will just pro-
duce zero kets, instead of additional energy eigenstate. Furthermore, we have
shown above that
H |0i = (~ω/2) |0i

FT
Finally, by acting on |0i with the raising operator and multiplying by suitable
normalization factors, we can produce an infinite set of energy eigenstates
{|0i , |1i , |2i , ..., |ni , ...}, such that
H |ni = ~ω(n + 1/2) |ni

which matches the energy spectrum which we gave in the preceding section.

Natural length and energy scales


The quantum harmonic oscillator possesses natural scales for length and ener-
gy, which can be used to simplify the problem. These can be found by nondi-
mensionalization. The result is that if we measure energy in units of ~ω and
distance in units of (~/ (mω))1/2 , then the Schrödinger equation becomes:
AH = − 21 ∂u
∂ 2
1 2
2 + 2u ,

and the energy eigenfunctions and eigenvalues become


hx|ψn i = √ 1 π −1/4 exp(−u2 /2)Hn (u)
2n n!

En = n + 21 .
DR
To avoid confusion, we will not adopt these natural units in this article. How-
ever, they frequently come in handy when performing calculations.

N -dimensional harmonic oscillator


The one-dimensional harmonic oscillator is readily generalizable to N dimen-
sions, where N = 1, 2, 3, ... . In one dimension, the position of the particle
was specified by a single coordinate, x. In N dimensions, this is replaced by
N position coordinates, which we label x 1, ..., x N . Corresponding to each po-
sition coordinate is a momentum; we label these p 1, ..., p N . The canonical
commutation relations between these operators are
[xi , pj ] = i~δi,j
[xi , xj ] = 0
[pi , pj ] = 0

Quantum harmonic oscillator


448

The Hamiltonian for this system is


P  p2i 
H= N 1 2 2
i=1 2m + 2 mω xi .

FT
As the form of this Hamiltonian makes clear, the N -dimensional harmonic os-
cillator is exactly analogous to N independent one-dimensional harmonic oscil-
lators with the same mass and spring constant. In this case, the quantities x 1,
..., x N would refer to the positions of each of the N particles. This is a happy
property of the r 2 potential, which allows the potential energy to be separated
into terms depending on one coordinate each.
This observation makes the solution straightforward. For a particular set of
quantum numbers {n} the energy eigenfunctions for the N -dimensional oscil-
lator are expressed in terms of the 1-dimensional eigenfunctions as:
hx|ψ{n} i = N
Q
i=1 hxi |ψni i

ai
a†i
A
In the ladder operator method,
=
=
p mω
2~ xi +
p mω
 we define N sets of ladder operators,
i
mω pi 
i
2~ xi − mω pi
.

By a procedure analogous to the one-dimensional case, we can then show that


each of the a i and a † i operators lower and raise the energy by ω respectively.
The energy levels of the system are
DR
h i
E = ~ω (n1 + · · · + nN ) + N2 .

ni = 0, 1, 2, . . .

As in the one-dimensional case, the energy is quantized. The ground state


energy is N times the one-dimensional energy, as we would expect using the
analogy to N independent one-dimensional oscillators. There is one further
difference: in the one-dimensional case, each energy level corresponds to a
unique quantum state. In N -dimensions, except for the ground state, the en-
ergy levels are degenerate, meaning there are several states with the same
energy.
The degeneracy can be calculated relatively easily, as an example, consider the
3-dimensional case: Define n = n 1 + n 2 + n 3. All states with the same n

Quantum harmonic oscillator


449

will have the same energy. For a given n, we choose a particular n 1. Then
n 2 + n 3 = n - n 1. There are n - n 1 + 1 possible groups {n 2, n 3}. n 2 can take
on the values 0 to n - n 1, and for each n 2 the value of n 3 is fixed. The degree
of degeneracy therefore is:

FT
n(n+1)
gn = nn1 =0 n−n1 +1 = nn1 =0 n+1− nn1 =0 n1 = (n+1)(n+1)− 2 =
P P P
(n+1)(n+2)
2

Related problems
The quantum harmonic oscillator can be extended in many interesting ways.
We will briefly discuss two of the more important extensions, the anharmonic
oscillator and coupled harmonic oscillators.

Anharmonic oscillator
As mentioned in the introduction, a system residing "near" the minimum of
some potential may be treated as a harmonic oscillator. In this approximation,
we Taylor-expand the potential energy around the minimum and discard terms
A
of third or higher order, resulting in an approximate quadratic potential. Once
we have studied the system in this approximation, we may wish to investigate
the corrections due to the discarded higher-order terms, particularly the third-
order term.
The anharmonic oscillator Hamiltonian is the harmonic oscillator Hamiltonian
with an additional x 3 potential:
p2
H= 2m + 12 mω 2 x2 + λx3
DR
If the harmonic approximation is valid, the coefficient is small compared to
the quadratic term. We may therefore use perturbation theory to determine
the corrections to the states and energy levels imposed by the anharmonic
term. This task may be simplified by using the ladder operators to rewrite the
anharmonic term as
 3
2
~
λ 2mω (a + a† )3 .

It turns out that the correction to the energies vanish to first-order in . The
second-order corrections are given by the usual formula in perturbation theory:
1
∆E (2) = λ2 hψE | x3 E−H 0
x3 |ψE i .

Quantum harmonic oscillator


450

This is straightforward, though tedious, to evaluate. One failing of this method,


however, is that it does not take into account the possibility of the particle
tunnelling out, since it is no longer bound on both sides.

Coupled harmonic oscillators

FT
In this problem, we consider N equal masses which are connected to their
neighbors by springs, in the limit of large N. The masses form a linear chain in
one dimension, or a regular lattice in two or three dimensions.
As in the previous section, we denote the positions of the masses by x 1, x 2, ...,
as measured from their equilibrium positions (i.e. x k = 0 if particle k is at its
equilibrium position.) In two or more dimensions, the xs are vector quantities.
The Hamiltonian of the total system is
p2i
H= N 1 2P 2
P
i=1 2m + 2 mω {ij}(nn) (xi − xj )

The potential energy is summed over "nearest-neighbor" pairs, so there is one


term for each spring.
Remarkably, there exists a coordinate transformation to turn this problem in-
A
to a set of independent harmonic oscillators, each of which corresponds to a
particular collective distortion of the lattice. These distortions display some
particle-like properties, and are called phonons. Phonons occur in the ionic
lattices of many solids, and are extremely important for understanding many
of the phenomena studied in solid state physics.

See also
DR
• Gas in a harmonic trap
• →Creation and annihilation operators
• →Coherent state
• Morse potential

References
• Griffiths, David J. (2004). Introduction to Quantum Mechanics (2nd ed.).
Prentice Hall. ISBN 013805326X.
• Liboff, Richard L. (2002). Introductory Quantum Mechanics. Addison-
Wesley. ISBN 0805387145.

Quantum harmonic oscillator


451

External links
• Quantum Harmonic Oscillator 243

FT
Source: http://en.wikipedia.org/wiki/Quantum_harmonic_oscillator

Principal Authors: CYD, Michael Hardy, HappyCamper, PAR, Dmn

Quantum indeterminacy

Quantum indeterminacy is the apparent necessary incompleteness in the de-


scription of a physical system, that has become one of the characteristics of
the standard description of quantum physics. Prior to quantum physics, it was
thought that (a) a physical system had a determinate state which uniquely de-
termined all the values of its measurable properties, and conversely (b) the
values of its measurable properties uniquely determined the state. Albert Ein-
Astein may have been the first person to carefully point out the radical effect the
new quantum physics would have on our notion of physical state.
Quantum indeterminacy can be quantitatively characterized by a probability
distribution on the set of outcomes of measurements of an observable. The
distribution is uniquely determined by the system state, and moreover quantum
mechanics provides a recipe for calculating this probability distribution.
Indeterminacy in measurement was not an innovation of quantum mechan-
ics, since it had established early on by experimentalists that errors in mea-
DR
surement may lead to indeterminate outcomes. However, by the latter half of
the eighteenth century, measurement errors were well understood and it was
known that they could either be reduced by better equipment or accounted
for by statistical error models. In quantum mechanics, however, indeterminacy
is of a much more fundamental nature, having nothing to do with errors or
disturbance.

Measurement
An adequate account of quantum indeterminacy requires a theory of measure-
ment. Many theories have been proposed since the beginning of quantum
mechanics and quantum measurement continues to be an active research area

243 http://hyperphysics.phy-astr.gsu.edu/hbase/quantum/hosc.html

Quantum indeterminacy
452

in both theoretical and experimental physics (Braginski and Khalili 1992.) Pos-
sibly the first systematic attempt at a mathematical theory was developed by
John von Neumann. The kind of measurements he investigated in (von Neu-
mann, 1955) are now called projective measurements. That theory was based

FT
in turn on the theory of projection-valued measures for self-adjoint operators
which had been recently developed (by von Neumann and independently by
Marshall Stone) and the Hilbert space formulation of quantum mechanics (at-
tributed by von Neumann to Paul Dirac).
In this formulation, the state of a physical system corresponds to a vector of
length 1 in a →Hilbert space H over the complex numbers. An observable is
represented by a self-adjoint operator A on H. If H is finite dimensional, by the
spectral theorem, A has an orthonormal basis of eigenvectors. If the system
is in state ψ, then immediately after measurement the system will occupy a
state which is an eigenvector e of A and the observed value will be the corre-
sponding eigenvalue of the equation A e = e. It is immediate from this that
measurement in general will be non-deterministic. Quantum mechanics, more-
over, gives a recipe for computing a probability distribution Pr on the possible
outcomes given the initial system state is ψ. The probability is
A
Pr(λ) = hE(λ)ψ | ψi

where E() is the projection onto the space of eigenvectors of A with eigenvalue
.
DR

Quantum indeterminacy
453

Example

A FT
Bloch sphere showing eigenvectors for Pauli Spin matrices. The Bloch sphere is
a two-dimensional surface the points of which correspond to the state space of
a spin 1/2 particle. At the state ψ the values of σ 1 are +1 whereas the values
DR
of σ 2 and σ 3 take the values +1, -1 with probability 1/2. In this example,
we consider a single spin 1/2 particle (such as an electron) in which we only
consider the spin degree of freedom. The corresponding Hilbert space is the
two-dimensional Hilbert space C 2, with each quantum state corresponding to
a unit vector in C 2 (unique up to phase). In this case, the state space can be
geometrically represented as the surface of a sphere, as shown in the figure on
the right.
The Pauli
 spinmatrices    
0 1 0 −i 1 0
σ1 = , σ2 = , σ3 =
1 0 i 0 0 −1

are self-adjoint and correspond to spin-measurements along the 3 coordinate


axes.

Quantum indeterminacy
454

The Pauli matrices all have the eigenvalues +1, -1.

• For σ 1, these eigenvalues correspond to the eigenvectors

√1 (1, 1), √1 (1, −1)

FT
2 2

• For σ 3, they correspond to the eigenvectors

(1, 0), (0, 1)

Thus in the state


ψ= √1 (1, 1),
2

σ 1 has the determinate value +1, while measurement of σ 3 can produce either
+1, -1 each with probability 1/2. In fact, there is no state in which measure-
ment of both σ 1 and σ 3 have determinate values.
There are various questions that can be asked about the above indeterminacy


A
assertion.

Can the indeterminacy be understood as similar to an error in measurement


explainable by an error parameter? More precisely, is there a hidden pa-
rameter that could account for the statistical indeterminacy in a completely
classical way?
• Can the indeterminacy be understood as a disturbance of the system being
measured?
DR
Von Neumann formulated the question 1) and provided an argument why the
answer had to be no, if one accepted the formalism he was proposing, al-
though his argument contained a flaw. The definitive negative answer to 1)
has been established by experiment that Bell’s inequalities are violated (see
Bell test experiments.) The answer to 2) depends on how disturbance is un-
derstood (particularly since measurement is disturbance), but in the most nat-
ural interpretation the answer is also no. To see this, consider two sequences
of measurements: (A) which measures exclusively σ 1 and (B) which measures
only σ 3 of a spin system in the state ψ. The measurement outcomes of (A) are
all +1, while the statistical distribution of the measurements (B) is still divided
between +1, -1 with probability 1/2.

Other examples of indeterminacy


Quantum indeterminacy can also be illustrated in terms of a particle with a
definitely measured momentum for which there must be a fundamental limit

Quantum indeterminacy
455

to how precisely its location can be specified. This quantum uncertainty prin-
ciple can be expressed in terms of other variables, for example, a particle with
a definitely measured energy has a fundamental limit to how precisely one can
specify how long it will have that energy. The units involved in quantum un-

FT
certainty are on the order of Planck’s constant (found experimentally to be 6.6
x 10 -34 J·s).

Indeterminacy and incompleteness


Quantum indeterminacy is the assertion that the state of a system does not
determine a unique collection of values for all its measurable properties. In-
deed in the quantum mechanical formalism, for a given quantum state, each
one of these measurable values will be obtained non-deterministically in ac-
cordance with a probability distribution which is uniquely determined by the
system state. Note that the state is destroyed by measurement, so when we
refer to a collection of values, each measured value in this collection must be
obtained using a freshly prepared state.
This indeterminacy might be regarded as a kind of essential incompleteness in
our description of a physical system. Notice however, that the indeterminacy
A
as stated above only applies to values of measurements not to the quantum
state. For example, in the spin 1/2 example discussed above, the system can
be prepared in the state ψ by using measurement of σ 1 as a filter which retains
only those particles such that σ 1 yields +1. By the von Neumann (so-called)
postulates, immediately after the measurement the system is assuredly in the
state ψ.
However, Einstein did believe that quantum state cannot be a complete de-
DR
scription of a physical system and, it is commonly thought, never came to terms
with quantum mechanics. In fact, Einstein, Boris Podolsky and Nathan Rosen
did show that if quantum mechanics is correct, then the classical view of how
the real world works (at least after special relativity) is no longer tenable. This
view included the following two ideas:

• A measurable property of a physical system whose value can be predicted


with certainty is actually an element of reality (this was the terminology
used by EPR).
• Effects of local actions have a finite propagation speed.

This failure of the classical view was one of the conclusions of the EPR thought
experiment in which two remotely located observers, now commonly referred
to 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 was a conclusion of EPR, using the formal apparatus of quantum theory,

Quantum indeterminacy
456

that once Alice measured spin in the x direction, Bob’s measurement in the x
direction was determined with certainty, whereas immediately before Alice’s
measurement Bob’s outcome was only statistically determined. From this it
follows that either value of spin in the x direction is not an element of reality

FT
or that the effect of Alice’s measurement has infinite speed of propagation.

Indeterminacy for mixed states


We have described indeterminacy for a quantum system which is in a pure
state. Mixed states are a more general kind of state obtained by a statistical
mixture of pure states. For mixed states the "quantum recipe" for determining
the probability distribution of a measurement is determined as follows:
Let A be an observable of a quantum mechanical system. A is given by a dense-
ly defined self-adjoint operator on H. The spectral measure of A is a projection-
valued measure
R defined by the condition
EA (U ) = U
λd E(λ),

for every Borel subset U of R. Given a mixed state S, we introduce the distri-
A
bution of A under S as follows:
DA (U ) = Tr(EA (U )S).

This is a probability measure defined on the Borel subsets of R which is the


probability distribution obtained by measuring A in S.

See also
DR
• Quantum mind
• just about any of the quantum mechanics articles, including
• →Quantum entanglement
• →Complementarity (physics)
• Interpretations of quantum mechanics
• Quantum measurement
• Counterfactual definiteness
• EPR paradox

References

Quantum indeterminacy
457

• A. Aspect, Bell’s inequality test: more ideal than ever, Nature 398 189
(1999). 244
• V. Braginski and F. Khalili, Quantum Measurements, Cambridge Universi-
ty Press, 1992.

FT
• G. Bergmann, The Logic of Quanta, American Journal of Physics, 1947.
Reprinted in Readings in the Philosophy of Science, Ed. H. Feigl and M.
Brodbeck, Appleton-Century-Crofts, 1953. Discusses measurement, accura-
cy and determinism.
• J.S. Bell, On the Einstein-Poldolsky-Rosen paradox, Physics 1 195 (1964).
• A. Einstein, B. Podolsky, and N. Rosen, Can quantum-mechanical de-
scription of physical reality be considered complete? 245 Phys. Rev. 47 777
(1935). 246
• G. Mackey, Mathematical Foundations of Quantum Mechanics, W. A. Ben-
jamin, 1963 (paperback reprint by Dover 2004).
• J. von Neumann, Mathematical Foundations of Quantum Mechanics,
Princeton University Press, 1955. Reprinted in paperback form. Originally
published in German in 1932.
• R. Omnès, Understanding Quantum Mechanics, Princeton University Press,
A 1999.

External links
• Common Misconceptions Regarding Quantum Mechanics 247 See especially
part III "Misconceptions regarding measurement".
DR
Source: http://en.wikipedia.org/wiki/Quantum_indeterminacy

Principal Authors: CarlHewitt, CSTAR, JWSchmidt, DV8 2XL, William M. Connolley

244 http://www-ece.rice.edu/~kono/ELEC565/Aspect_Nature.pdf
245 http://www.drchinese.com/David/EPR.pdf
246 http://prola.aps.org/abstract/PR/v47/i10/p777_1
247 http://www.oberlin.edu/physics/dstyer/TeachQM/misconnzz.pdf

Quantum indeterminacy
458

Quantum leap

In physics, a quantum leap or quantum jump is a change of an electron with-

FT
in an atom from one energy state to another. This is a discontinuous change
in which the electron goes from one energy level to another without passing
through any intermediate levels. This phenomenon contradicted expectations
set by classical theories, that the electron’s energy should be able to vary con-
tinuously. Quantum leaps of electrons cause the emission of electromagnetic
radiation in quantized units called photons. All emission of light occurs as a
result of quantum leaps.
More generally, a quantum leap is the smallest possible change, as when one’s
bank account balance goes from $500.00 (five hundred dollars) to $500.01
(five hundred dollars and one cent). There are no possible amounts interme-
diate between those.

Vernacular usage
In the vernacular, the term quantum leap has come to mean an abrupt change,
A
especially an advance or augmentation. The term dates back to early-to-mid-
20th century. The vernacular usage is not always in accord with the original
meaning, in that a large abrupt change is often implied. A quantum leap in
quantum mechanics is by definition the smallest change possible. The usages
agree, however, in that both describe an advance that happens all at once,
rather than gradually over time. A ’quantum leap in technology’ is thus a revo-
lutionary advance, rather than an evolutionary one.
DR
See also
• Fluorescence
• Phosphorescence
• Stimulated emission
• Sinclair QL

External links
• Are there quantum jumps? 248

Source: http://en.wikipedia.org/wiki/Quantum_leap

248 http://www.mikomma.de/schroe/quantumjumps.htm

Quantum leap
459

Principal Authors: Michael Hardy, Srleffler, Maveric149, Hinakana, Djinn112

Quantum level

FT
Quantum levels are fixed levels with a logarithmic, descending quantum pat-
tern in the visible spectrum of light that can be observed through a spectrome-
ter while looking at intense flows of electricity through the various halides on
the periodic table in a vacuum tube. They also have some use in chemistry
when dealing with the movement of electrons to different orbital levels around
the atom and the energy levels involved in such actions.

See also
• →Quantum mechanics
• Absorption spectrum
• Emission spectrum
A
Source: http://en.wikipedia.org/wiki/Quantum_level

Principal Authors: Enochlau, JYOuyang, Michael Hardy, Borofkin, Charles Matthews

Quantum mechanics
DR
For a non-technical introduction to the topic, please see Introduction to
Quantum mechanics.

Quantum mechanics is a fundamental branch of theoretical physics that re-


places classical mechanics and classical electromagnetism at the atomic and
subatomic levels. It is the underlying mathematical framework of many
fields of physics and chemistry, including condensed matter physics, atomic
physics, molecular physics, computational chemistry, quantum chemistry, par-
ticle physics, and nuclear physics. Along with general relativity, quantum me-
chanics is one of the pillars of modern physics.

Quantum mechanics
460

A FT
Figure 42 Fig. 1: The wavefunctions of an electron in a hydrogen atom possessing
definite energy (increasing downward: n=1,2,3,...) and angular momentum (increasing
across: s, p, d,...). Brighter areas correspond to higher probability density for a position
measurement. Wavefunctions like these are directly comparable to Chladni’s figures of
acoustic modes of vibration in classical physics and are indeed modes of oscillation as
well: they possess a sharp energy and thus a sharp frequency. The angular momentum
and energy are quantized, and only take on discrete values like those shown (as is the
DR
case for resonant frequencies in acoustics).

Introduction
The term quantum (Latin, "how much ") refers to discrete units that the theory
assigns to certain physical quantities, such as the energy of an atom at rest (see
Figure 1, at right). The discovery that waves could be measured in particle-like
small packets of energy called quanta led to the branch of physics that deals
with atomic and subatomic systems which we today call Quantum Mechanics.
The foundations of quantum mechanics were established during the first half
of the 20th century by Max Planck, Albert Einstein, Niels Bohr, Werner Heisen-
berg, Erwin Schrödinger, Max Born, John von Neumann, Paul Dirac, Wolfgang
Pauli and others. Some fundamental aspects of the theory are still actively
studied.

Quantum mechanics
461

Quantum mechanics is a more fundamental theory than Newtonian mechan-


ics and classical electromagnetism, in the sense that it provides accurate and
precise descriptions for many phenomena that these "classical" theories simply
cannot explain on the atomic and subatomic level. It is necessary to use quan-

FT
tum mechanics to understand the behavior of systems at atomic length scales
and smaller. For example, if Newtonian mechanics governed the workings of
an atom, electrons would rapidly travel towards and collide with the nucleus.
However, in the natural world the electron normally remains in a stable orbit
around a nucleus – seemingly defying classical electromagnetism.
Quantum mechanics was initially developed to explain the atom, especially
the spectra of light emitted by different atomic species. The quantum theory of
the atom developed as an explanation for the electron’s staying in its orbital,
which could not be explained by Newton’s laws of motion and by classical
electromagnetism.
In the formalism of quantum mechanics, the state of a system at a given time
is described by a complex number wave functions (sometimes referred to as
orbitals in the case of atomic electrons), and more generally, elements of a
complex vector space. This abstract mathematical object allows for the calcu-
A
lation of probabilities of outcomes of concrete experiments. For example, it
allows to compute the probability of finding an electron in a particular region
around the nucleus at a particular time. Contrary to classical mechanics, one
cannot in general make predictions of arbitrary accuracy. For instance electrons
cannot in general be pictured as localized particles in space but rather should
be thought of as "clouds" of negative charge spread out over the entire orbit.
These clouds represent the regions around the nucleus where the probability
of "finding" an electron is the largest. The Heisenberg’s Uncertainty Principle
DR
quantifies the inability to precisely locate the particle.
The other exemplar that led to quantum mechanics was the study of electro-
magnetic 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 quan-
ta, Albert Einstein exploited this idea to show that an electromagnetic wave
such as light could be described by a particle called the photon with a dis-
crete energy dependent on its frequency. This led to a theory of unity between
subatomic particles and electromagnetic waves called wave-particle duality in
which particles and waves were neither one nor the other, but had certain
properties of both. While quantum mechanics describes the world of the very
small, it also is needed to explain certain "macroscopic quantum systems" such
as superconductors and superfluids.
Broadly speaking, quantum mechanics incorporates four classes of phenomena
that classical physics cannot account for: (i) the quantization (discretization)

Quantum mechanics
462

of certain physical quantities, (ii) wave-particle duality, (iii) the uncertainty


principle, and (iv) quantum entanglement. Each of these phenomena will be
described in greater detail in subsequent sections.
Since the early days of quantum theory, physicists have made many attempts

FT
to combine it with the other highly successful theory of the twentieth century,
Albert Einstein’s General Theory of Relativity. While quantum mechanics is en-
tirely consistent with special relativity, serious problems emerge when one tries
to join the quantum laws with general relativity, the more elaborate descrip-
tion of nature which includes gravity. Resolving these inconsistencies has been
a major goal of twentieth- and twenty-first-century physics. Despite the propos-
al of many novel ideas, the unification of quantum mechanics—which reigns in
the domain of the very small—and general relativity—a superb description of
the very large—remains a tantalizing future possibility. (See quantum gravity,
string theory.)
Because everything is composed of quantum-mechanical particles, the laws
of classical physics must approximate the laws of quantum mechanics in the
appropriate limit. This is often expressed by saying that in case of large quan-
tum numbers quantum mechanics "reduces" to classical mechanics and classical
A
electromagnetism . This requirement is called the correspondence, or classical
limit.
Quantum mechanics can be formulated in either a relativistic or non-relativistic
manner. Relativistic quantum mechanics (quantum field theory) provides the
framework for some of the most accurate physical theories known. Still, non-
relativistic quantum mechanics is also used due to its simplicity and when rel-
ativistic effects are negligible. We will use the terms quantum mechanics,
quantum physics, and quantum theory synonymously, to refer to both rela-
DR
tivistic and non-relativistic quantum mechanics. It should be noted, however,
that certain authors refer to "quantum mechanics" in the more restricted sense
of non-relativistic quantum mechanics. Also, in quantum mechanics, the use of
the term particle typically refers to an elementary or subatomic particle.

Description of the theory


There are a number of mathematically equivalent formulations of quantum
mechanics. One of the oldest and most commonly used formulations is the
transformation theory invented by Cambridge theoretical physicist Paul Dirac,
which unifies and generalizes the two earliest formulations of quantum me-
chanics, matrix mechanics (invented by Werner Heisenberg) and wave me-
chanics (invented by Erwin Schrödinger).

Quantum mechanics
463

In this formulation, the instantaneous state of a quantum system encodes the


probabilities of its measurable properties, or "observables". Examples of ob-
servables include energy, position, momentum, and angular momentum. Ob-
servables can be either continuous (e.g., the position of a particle) or discrete

FT
(e.g., the energy of an electron bound to a hydrogen atom).
Generally, quantum mechanics does not assign definite values to observables.
Instead, it makes predictions about probability distributions; that is, the proba-
bility of obtaining each of the possible outcomes from measuring an observable.
Naturally, these probabilities will depend on the quantum state at the instant
of the measurement. There are, however, certain states that are associated
with a definite value of a particular observable. These are known as "eigen-
states" of the observable ("eigen" meaning "own" in German). In the everyday
world, it is natural and intuitive to think of everything being in an eigenstate
of every observable. Everything appears to have a definite position, a definite
momentum, and a definite time of occurrence. However, Quantum Mechanics
does not pinpoint the exact values for the position or momentum of a certain
particle in a given space in a finite time, but, rather, it only provides a range of
probabilities of where that particle might be. Therefore, it became necessary to
A
use different words for a) the state of something having an uncertainty relation
and b) a state that has a definite value. The latter is called the "eigenstate" of
the property being measured.
A concrete example will be useful here. Let us consider a free particle. In quan-
tum mechanics, there is wave-particle duality so the properties of the particle
can be described as a wave. Therefore, its quantum state can be represented
as a wave, of arbitrary shape and extending over all of space, called a wave-
function. The position and momentum of the particle are observables. The
DR
Uncertainty Principle of quantum mechanics states that both the position and
the momentum cannot simultaneously be known with infinite precision at the
same time. However, we can measure just the position alone of a moving free
particle creating an eigenstate of position with a wavefunction that is very large
at a particular position x, and zero everywhere else. If we perform a position
measurement on such a wavefunction, we will obtain the result x with 100%
probability. In other words, we will know the position of the free particle. This
is called an eigenstate of position. If the particle is in an eigenstate of position
then its momentum is completely unknown. An eigenstate of momentum, on
the other hand, has the form of a plane wave. It can be shown that the wave-
length is equal to h/p, where h is Planck’s constant and p is the momentum of
the eigenstate. If the particle is in an eigenstate of momentum then its position
is completely blurred out.

Quantum mechanics
464

Usually, a system will not be in an eigenstate of whatever observable we are


interested in. However, if we measure the observable, the wavefunction will
immediately become an eigenstate of that observable. This process is known
as wavefunction collapse. If we know the wavefunction at the instant before

FT
the measurement, we will be able to compute the probability of collapsing into
each of the possible eigenstates. For example, the free particle in our previ-
ous example will usually have a wavefunction that is a wave packet centered
around some mean position x 0, neither an eigenstate of position nor of momen-
tum. When we measure the position of the particle, it is impossible for us to
predict with certainty the result that we will obtain. It is probable, but not cer-
tain, that it will be near x 0, where the amplitude of the wavefunction is large.
After we perform the measurement, obtaining some result x, the wavefunction
collapses into a position eigenstate centered at x.
Wave functions can change as time progresses. An equation known as the
→Schrödinger equation describes how wave functions change in time, a role
similar to Newton’s second law in classical mechanics. The Schrödinger equa-
tion, applied to our free particle, predicts that the center of a wave packet
will move through space at a constant velocity, like a classical particle with no
A
forces acting on it. However, the wave packet will also spread out as time pro-
gresses, 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.
Some wave functions produce probability distributions that are constant in
time. Many systems that are treated dynamically in classical mechanics are
described by such "static" wave functions. For example, a single electron in
DR
an unexcited atom is pictured classically as a particle moving in a circular tra-
jectory around the atomic nucleus, whereas in quantum mechanics it is de-
scribed by a static, spherically symmetric wavefunction surrounding the nucle-

Quantum mechanics
465

A
us (
FT ). (Note
that only the lowest angular momentum states, labeled s, are spherically sym-
metric).
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 wave-
function will be at any later time. During a measurement, the change of the
DR
wavefunction into another one is not deterministic, but rather unpredictable,
i.e., random.
The probabilistic nature of quantum mechanics thus stems from the act of mea-
surement. 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 exten-
sively studied. Interpretations of quantum mechanics have been formulated
to 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 indepen-
dent entity. For details, see the article on measurement in quantum mechanics.

Quantum mechanics
466

Quantum mechanical effects


As mentioned in the introduction, there are several classes of phenomena that
appear under quantum mechanics which have no analogue in classical physics.
These are sometimes referred to as "quantum effects".

FT
The first type of quantum effect is the quantization of certain physical quanti-
ties. Quantization first arose in the mathematical formulae of Max Planck in
1900 as discussed in the introduction. Max Planck was analyzing how the ra-
diation emitted from a body was related to its temperature, in other words, he
was analyzing the energy of a wave. The energy of a wave could not be infi-
nite, so Planck used the property of the wave we designate as the frequency to
define energy. Max Planck discovered a constant that when multiplied by the
frequency of any wave gives the energy of the wave. This constant is referred
to by the letter h in mathematical formulae. It is a cornerstone of physics. By
measuring the energy in a discrete non-continuous portion of the wave, the
wave took on the appearance of chunks or packets of energy. These chunks of
energy resembled particles. So energy is said to be quantized because it only
comes in discrete chunks instead of a continuous range of energies.
A
In the example we have given, of a free particle in empty space, both the po-
sition and the momentum are continuous observables. However, if we restrict
the particle to a region of space (the so-called "particle in a box" problem),
the momentum observable will become discrete; it will only take on the val-
h
ues n 2L , where L is the length of the box, h is Planck’s constant, and n is an
arbitrary nonnegative integer number. Such observables are said to be quan-
tized, and they play an important role in many physical systems. Examples of
quantized observables include angular momentum, the total energy of a bound
DR
system, and the energy contained in an electromagnetic wave of a given fre-
quency.
Another quantum effect is the uncertainty principle, which is the phenomenon
that consecutive measurements of two or more observables may possess a fun-
damental limitation on accuracy. In our free particle example, it turns out that
it is impossible to find a wavefunction that is an eigenstate of both position and
momentum. This implies that position and momentum can never be simulta-
neously measured with arbitrary precision, even in principle: as the precision
of the position measurement improves, the maximum precision of the momen-
tum measurement decreases, and vice versa. Those variables for which it holds
(e.g., momentum and position, or energy and time) are canonically conjugate
variables in classical physics.

Quantum mechanics
467

Another quantum effect is the wave-particle duality. It has been shown that,
under certain experimental conditions, microscopic objects like atoms or elec-
trons exhibit particle-like behavior, such as scattering. ("Particle-like" in the
sense of an object that can be localized to a particular region of space.) Under

FT
other conditions, the same type of objects exhibit wave-like behavior, such as
interference. We can observe only one type of property at a time, never both
at the same time.
Another quantum effect is quantum entanglement. In some cases, the wave
function of a system composed of many particles cannot be separated into in-
dependent wave functions, one for each particle. In that case, the particles are
said to be "entangled". If quantum mechanics is correct, entangled particles
can display remarkable and counter-intuitive properties. For example, a mea-
surement made on one particle can produce, through the collapse of the total
wavefunction, an instantaneous effect on other particles with which it is entan-
gled, even if they are far apart. (This does not conflict with special relativity
because information cannot be transmitted in this way.)

Mathematical formulation
A
Main article: →Mathematical formulation of quantum mechanics. See also the
discussion in Quantum logic.
In the mathematically rigorous formulation of quantum mechanics, developed
by Paul Dirac and John von Neumann, the possible states of a quantum me-
chanical system are represented by unit vectors (called "state vectors") residing
in a complex separable →Hilbert space (variously called the "state space" or the
"associated Hilbert space" of the system) well defined upto a complex number
of norm 1 (the phase factor). In other words, the possible states are points
DR
in the projectivization of a Hilbert 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 densely defined Hermitian (or
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
operator’s spectrum is discrete, the observable can only attain those discrete
eigenvalues.
The time evolution of a quantum state is described by the →Schrödinger equa-
tion, in which the Hamiltonian, the operator corresponding to the total energy
of the system, generates time evolution.

Quantum mechanics
468

The inner product between two state vectors is a complex number known as
a probability amplitude. During a measurement, the probability that a system
collapses from a given initial state to a particular eigenstate is given by the
square of the absolute value of the probability amplitudes between the initial

FT
and final states. The possible results of a measurement are the eigenvalues
of the operator - 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 Schrödinger equation acts on the entire probability amplitude, not merely
its absolute value. Whereas the absolute value of the probability amplitude en-
codes information about probabilities, its phase encodes information about the
interference between quantum states. This gives rise to the wave-like behavior
of quantum states.
It turns out that analytic solutions of Schrödinger’s equation are only available
for a small number of model Hamiltonians, of which the quantum harmonic
A
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, for example,
the addition of a weak potential energy. Another method is the "semi-classical
DR
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.
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 fundamental rules of quantum mechanics are very broad. They state that
the state space of a system is a Hilbert space and the observables are Hermitian
operators acting on that space, but do not tell us which Hilbert space or which
operators. These must be chosen appropriately in order to obtain a quantitative

Quantum mechanics
469

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 becomes large.
This "large system" limit is known as the classical or correspondence limit. One

FT
can therefore start from an established classical model of a particular system,
and attempt to guess the underlying quantum model that gives rise to the
classical model in the correspondence limit.
Unsolved problems in physics: In the correspondence limit of quantum me-
chanics: Is there a preferred interpretation of quantum mechanics? How does
the quantum description of reality, which includes elements such as the super-
position of states and wavefunction collapse, give rise to the reality we per-
ceive?
When quantum mechanics was originally formulated, it was applied to mod-
els 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.
A
Early attempts to merge quantum mechanics with special relativity involved
the replacement of the Schrödinger 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 unsatis-
factory qualities stemming from their neglect of the relativistic creation and
annihilation of particles. A fully relativistic quantum theory required the de-
velopment 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
DR
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 1/r 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

Quantum mechanics
470

called quantum chromodynamics, and describes the interactions of the subnu-


clear particles: quarks and gluons. The weak nuclear force and the electromag-
netic force were unified, in their quantized forms, into a single quantum field
theory known as electroweak theory.

FT
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 com-
plete theory of quantum gravity is hindered by apparent incompatibilities be-
tween 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.

Applications of quantum theory


Quantum mechanics has had enormous success in explaining many of the fea-
tures of our world. The individual behavior of the subatomic particles that
make up all forms of matter - electrons, protons, neutrons, and so forth - can
A
often only be satisfactorily described using quantum mechanics.Quantum me-
chanics has strongly influenced string theory, a candidate for a theory of every-
thing (see Reductionism). It is also related to statistical mechanics.
Quantum mechanics is important for understanding how individual atoms
combine covalently to form chemicals or molecules. The application of quan-
tum mechanics to chemistry is known as quantum chemistry. (Relativistic)
quantum mechanics can in principle mathematically describe most of chem-
DR
istry. Quantum mechanics can provide quantitative insight into ionic and cova-
lent bonding processes by explicitly showing which molecules are energetically
favorable to which others, and by approximately how much. Most of the cal-
culations performed in computational chemistry rely on quantum mechanics.
Much of modern technology operates at a scale where quantum effects are sig-
nificant. Examples include the laser, the transistor, the electron microscope,
and magnetic resonance imaging. The study of semiconductors led to the in-
vention 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 dis-
tant goal is the development of quantum computers, which are expected to

Quantum mechanics
471

perform certain computational tasks exponentially faster than classical com-


puters. Another active research topic is quantum teleportation, which deals
with techniques to transmit quantum states over arbitrary distances.

FT
Philosophical consequences
Main article: Interpretations of quantum mechanics
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 am-
plitudes and probability distributions took decades to be appreciated.
The Copenhagen interpretation, due largely to the Danish theoretical physicist
Niels Bohr, is the interpretation of quantum mechanics most widely accepted
amongst physicists. According to it, the probabilistic nature of quantum me-
chanics predictions cannot be explained in terms of some other deterministic
theory, and does not simply reflect our limited knowledge. Quantum mechanics
provides probabilistic results because the physical universe is itself probabilistic
rather than deterministic.
A
Albert Einstein, himself one of the founders of quantum theory, disliked this
loss of determinism in measurement. He held that there should be a local
hidden variable theory underlying quantum mechanics and consequently the
present theory was incomplete. He produced a series of objections to the the-
ory, the most famous of which has become known as the EPR paradox. John
Bell showed that the EPR paradox led to experimentally testable differences
between quantum mechanics and local hidden variable theories. Experiments
have been taken as confirming that quantum mechanics is correct and the real
DR
world cannot be described in terms of such hidden variables. "Loopholes" in
the experiments, however, mean that the question is still not quite settled.
See the Bohr-Einstein debates
The Everett many-worlds interpretation, formulated in 1956, holds that all the
possibilities described by quantum theory simultaneously occur in a "multi-
verse" composed of mostly independent parallel universes. This is not accom-
plished by introducing some new axiom to quantum mechanics, but on the con-
trary by removing the axiom of the collapse of the wave packet: All the possible
consistent states of the measured system and the measuring apparatus (includ-
ing 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

Quantum mechanics
472

consistent state contribution to the mentioned superposition, we inhabit. Ev-


erett’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 for us,

FT
making them physically meaningless. This inaccessiblity 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 sys-
tem that was measured originally. This is completely impractical, but even if
one can theoretically do this, it would destroy any evidence that the original
measurement took place (including the physicist’s memory).

History
In 1900, the German physicist Max Planck introduced the idea that energy is
quantized, in order to derive a formula for the observed frequency dependence
of the energy emitted by a black body. In 1905, Einstein explained the photo-
A
electric effect by postulating that light energy comes in quanta called photons.
The idea that each photon had to consist of energy in terms of quanta was a
remarkable achievement as it effectively removed the possibility of black body
radiation attaining infinite energy if it were to be explained in terms of wave
forms only. 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. In 1924, the French physicist Louis de Broglie put
forward his theory of matter waves by stating that particles can exhibit wave
DR
characteristics and vice versa.
These theories, though successful, were strictly phenomenological: there was
no rigorous justification for quantization (aside, perhaps, for Henri Poincaré’s
discussion of Planck’s theory in his 1912 paper Sur la théorie des quanta). 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.
Modern quantum mechanics was born in 1925, when the German physi-
cist Heisenberg developed matrix mechanics and the Austrian physicist
Schrödinger invented wave mechanics and the non-relativistic Schrödinger
equation. Schrödinger subsequently showed that the two approaches were
equivalent.
Heisenberg formulated his uncertainty principle in 1927, and the Copenhagen
interpretation took shape at about the same time. Starting around 1927, Paul

Quantum mechanics
473

Dirac began the process of unifying quantum mechanics with special relativity
by discovering the Dirac equation for the electron. 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

FT
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 Cal Tech, and John 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 the-
ories. Early workers in this area included Dirac, Pauli, Weisskopf, and Jordan.
This area of research culminated in the formulation of quantum electrodynam-
A
ics by Feynman, Dyson, Schwinger, and Tomonaga during the 1940s. Quantum
electrodynamics is a quantum theory of electrons, positrons, and the electro-
magnetic field, and served as a role model for subsequent quantum field theo-
ries.
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
Wilzcek in 1975. Building on pioneering work by Schwinger, Higgs, Goldstone,
Glashow, Weinberg and Salam independently showed how the weak nuclear
DR
force and quantum electrodynamics could be merged into a single electroweak
force.

Founding experiments
• Thomas Young’s double-slit experiment demonstrating the wave nature of
light (c1805)
• Henri Becquerel discovers radioactivity (1896)
• Joseph John Thomson’s cathode ray tube experiments (discovers the elec-
tron 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

Quantum mechanics
474

• 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

FT
are almost uniformly distributed. (1911)
• Professor Walter Ernhart-Plank’s Proton Collapse experiment disproved the
Rutherford model and temporarily cast doubt on the distribution of protons
throughout an atom.
• Otto Stern and Walter 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 1 in the Electron diffraction experiment (1927)
• Clyde L. Cowan and Frederick Reines confirm the existence of the neutrino
in the neutrino experiment (1955)
• Claus Jönsson‘s double-slit experiment with electrons (1961)

See also
• Basics of quantum mechanics • Quantum information
A•


→Measurement in quantum mechanics
Quantum electrochemistry
Quantum chemistry



→Quantum field theory
Quantum thermodynamics
Theoretical chemistry
• Quantum computers

References
• P. A. M. Dirac, The Principles of Quantum Mechanics (1930) – the begin-
ning chapters provide a very clear and comprehensible introduction
DR
• David Griffiths, Introduction to Quantum Mechanics, Prentice Hall, 1995.
ISBN 0-13-111892-7 – A standard undergraduate level text written in an
accessible style.
• Richard P. Feynman, Robert B. Leighton and Matthew Sands (1965). The
Feynman Lectures on Physics, Addison-Wesley. Richard Feynman’s original
lectures (given at CALTECH in early 1962) can also be downloaded as an
MP3 file from www.audible.com 249
• Hugh Everett, Relative State Formulation of Quantum Mechanics, Reviews
of Modern Physics vol 29, (1957) pp 454-462.
• Bryce DeWitt, R. Neill Graham, eds, The Many-Worlds Interpretation of
Quantum Mechanics, Princeton Series in Physics, Princeton University Press
(1973), ISBN 069108131X

249 http://www.audible.com

Quantum mechanics
475

• Albert Messiah, Quantum Mechanics, English translation by G. M. Temmer


of Mécanique Quantique, 1966, John Wiley and Sons, vol. I, chapter IV,
section III.
• Richard P. Feynman, QED: The Strange Theory of Light and Matter – a

FT
popular science book about quantum mechanics and quantum field theory
that contains many enlightening insights that are interesting for the expert
as well
• Marvin Chester, Primer of Quantum Mechanics, 1987, John Wiley, N.Y. IS-
BN 0486428788
• Hagen Kleinert, Path Integrals in Quantum Mechanics, Statistics, Polymer
Physics, and Financial Markets, 3th edition, World Scientific (Singapore,
2004) 250(also available online here 251)
• George Mackey (2004). The mathematical foundations of quantum me-
chanics. Dover Publications. ISBN 0486435172.
• Griffiths, David J. (2004). Introduction to Quantum Mechanics (2nd ed.).
Prentice Hall. ISBN 013805326X.
• Omnes, Roland (1999). Understanding Quantum Mechanics. Princeton
University Press. ISBN 0691004358.
A• J. von Neumann, Mathematical Foundations of Quantum Mechanics,
Princeton University Press, 1955.
• H. Weyl, The Theory of Groups and Quantum Mechanics, Dover Publica-
tions 1950.

Notes
• Note 1: The Davisson-Germer experiment, which demonstrates the wave
DR
nature of the electron 252

External links
General:

• A history of quantum mechanics 253


• A Lazy Layman’s Guide to Quantum Physics 254
• Introduction to Quantum Theory at Quantiki 255

250 http://www.worldscibooks.com/physics/5057.html
251 http://www.physik.fu-berlin.de/~kleinert/b5
252 http://hyperphysics.phy-astr.gsu.edu/hbase/quantum/davger2.html
253 http://www-history.mcs.st-andrews.ac.uk/history/HistTopics/The_Quantum_age_begins.html
254 http://higgo.com/quantum/laymans.htm
255 http://cam.qubit.org/wiki/index.php/Introduction_to_Quantum_Theory

Quantum mechanics
476

• Quantum Physics Made Relatively Simple 256: three video lectures by Hans
Bethe
• Decoherence 257 by Erich Joos

Course material:

FT
• MIT OpenCourseWare: Chemistry 258. See 5.61 259, 5.73 260, and 5.74 261
• MIT OpenCourseWare: Physics 262. See 8.04 263, 8.05 264, and 8.06 265.
• Imperial College Quantum Mechanics Course to Download 266
• A set of downloadable tutorials on Quantum Mechanics, Imperial College 267
• Spark Notes - Quantum Physics 268

FAQs:

• Many-worlds or relative-state interpretation 269


• Measurement in Quantum mechanics 270
• A short FAQ on quantum resonances 271

Media:
A• Everything you wanted to know about the quantum world 272 — archive of
articles from New Scientist magazine.
• " Quantum Trickery: Testing Einstein’s Strangest Theory 273", The New York
Times, December 27, 2005.

Philosophy:

256 http://bethe.cornell.edu/
DR
257 http://www.decoherence.de/
258 http://ocw.mit.edu/OcwWeb/Chemistry/index.htm
259 http://ocw.mit.edu/OcwWeb/Chemistry/5-61Fall-2004/CourseHome/index.htm
260 http://ocw.mit.edu/OcwWeb/Chemistry/5-73Fall-2005/CourseHome/index.htm
261 http://ocw.mit.edu/OcwWeb/Chemistry/5-74Spring-2005/CourseHome/index.htm
262 http://ocw.mit.edu/OcwWeb/Physics/index.htm
263 http://ocw.mit.edu/OcwWeb/Physics/8-04Quantum-Physics-ISpring2003/CourseHome/index.htm
264 http://ocw.mit.edu/OcwWeb/Physics/8-05Fall-2004/CourseHome/index.htm
265 http://ocw.mit.edu/OcwWeb/Physics/8-06Spring-2005/CourseHome/index.htm
266 http://www.imperial.ac.uk/quantuminformation/qi/tutorials
267 http://www3.imperial.ac.uk/portal/page?_pageid=161,482073&_dad=portallive&_schema=PORTALLIVE

#Quantum%20Information%20Tutorials
268 http://www.sparknotes.com/testprep/books/sat2/physics/chapter19section3.rhtml
269 http://www.hedweb.com/manworld.htm
270 http://www.mtnmath.com/faq/meas-qm.html
271 http://www.thch.uni-bonn.de/tc/people/brems.vincent/vincent/faq.html
272 http://www.newscientist.com/channel/fundamentals/quantum-world
273 http://www.nytimes.com/2005/12/27/science/27eins.html?ex=1293339600&en=caf5d835203c3500

&ei=5090

Quantum mechanics
477

• Quantum Mechanics (Stanford Encyclopedia of Philosophy) 274


• David Mermin on the future directions of physics 275
• "Quantum Physics Quackery" 276 by Victor Stenger, Skeptical Inquirer (Jan-
uary/February 1997).

FT
• Crank Dot Net’s quantum physics page 277 — "cranks, crackpots, cooks &
loons on the net"
• Hinduism & Quantum Physics 278

Source: http://en.wikipedia.org/wiki/Quantum_mechanics

Principal Authors: CYD, Lethe, David R. Ingham, Ancheta Wis, Voyajer, Laurascudder, Andris, Ben-
saccount, El C, Anville

Quantum mechanics, philosophy and con-


troversy
AQuantum mechanics has had many detractors including Albert Einstein and
Erwin Schroedinger. Quantum mechanics has had a profound affect on phi-
losophy. Determinism is a philosophical view that the universe is governed by
determinism if given a specific state of the universe at a specific time, the future
state of the universe is fixed as a matter of natural law. The philosophy of de-
terminism was derived from science, from Newton’s laws, and pre-Newtonian
physics, in that the ability to predict future outcomes in the universe (such as
future position of planets) was made possible by science. Quantum mechanics
DR
took away predictability and therefore was a blow to philosophy. However, the
main founder of quantum mechanics, Niels Bohr, is said to have a philosophy
of determinism similar to the rationalization by Immanuel Kant. This article
will attempt, without going into religious implications which are personal mat-
ters, to explain the position of many physicists on quantum mechanics and the
profound effect that quantum mechanics has had on philosophy.

274 http://plato.stanford.edu/entries/qm/
275 http://www.physicstoday.org/pt/vol-54/iss-2/p11.html
276 http://www.csicop.org/si/9701/quantum-quackery.html
277 http://www.crank.net/quantum.html
278 http://www.hinduism.co.za/hinduism.htm

Quantum mechanics, philosophy and controversy


478

Philosophical determinism
The 18th century saw many advances in the domain of science. After Newton,
most scientists agreed on the presupposition that the universe is governed by
(natural) laws that can be discovered and formalized by means of scientific

FT
observation and experiment. This position is known as determinism. Howev-
er, while determinism was the fundamental presupposition of post-Newtonian
physics, it quickly lead philosophers to a tremendous problem: if the universe,
and thus the entire world is governed by natural law, then that means that
human beings are also governed by natural law in their own actions. In other
words, it means that there is no such thing as human freedom. If it is accepted
that everything in the world is governed by natural law, then we must also
accept that it is not possible for us to will our own actions as free individuals;
rather, they must be determined by universal laws of nature. Conversely, if it
is accepted that human beings do have free will, then we must accept that the
world is not entirely governed by natural law. However, if the world is not en-
tirely governed by natural law, then the task of science is rendered impossible:
if the task of science is to discover and formalize the laws of nature, then what
task is left for science if it has been decided that nature is not entirely governed
A
by laws? Thus, there are extremely compelling reasons to want to accept both
free will and determinism. However, the two seem totally irreconcilable.
Immanuel Kant whose work dates towards the end of the 18th Century, at-
tempted to reconcile the seemingly incompatible schools of thought known as
empiricism (e.g., David Hume) and rationalism (e.g., René Descartes). Ac-
cording to the empiricists, the only possible knowledge of the world is the
knowledge that can be obtained by means of perception (inductive reasoning).
DR
Thus, for the empiricists concepts are abstractions that we derive by mentally
comparing several different perceptions and noting some quality shared by all
of them: for example, we see a fire engine, a rubber ball, and a dress, we per-
ceive some quality that is shared by these different objects, and we abstract this
quality from the objects themselves in order to arrive at the concept of the color
red. For Hume and the empiricists, this means that our concepts, such as the
concept of cause and effect, are not actually legitimate properties of the world,
but are rather mental constructs that we produce from repeated observation.
Since we can never actually perceive cause and effect (because it is not an ob-
ject, but rather a relation), we can never obtain certain knowledge of whether
it actually exists. In other words, since we can’t perceive it, we can never be
totally sure that we are not just imagining it. For the rationalists, on the other
hand, the situation is entirely the reverse, and the only certain knowledge is
the knowledge that we derive by means of pure logic (deductive reasoning).
The privileged model of certainty for the rationalists is mathematics. For the

Quantum mechanics, philosophy and controversy


479

rationalists, we can never be certain of any knowledge derived from percep-


tion, since we are capable of perceiving objects that are almost certainly false
in dreams. In other words, since there is no difference between an object per-
ceived while we are awake and the same object perceived during a dream, we

FT
can never derive certainty from perception. Mathematics does not require any
perceivable object in order to arrive at its proofs, because it works in purely
logical relations between concepts.
Kant approached this problem most famously in his major work of epistemolo-
gy, The Critique of Pure Reason (Kritik der reinen Vernunft, 1781). In order to
reconcile these disparate views, Kant found it necessary to split the world into
two completely separate aspects:

• 1. The world as appearance – that is, as it appears to us in our perceptions.


• 2. The world as a thing-in-itself – in other words, independent of all human
perception.

For Kant, all scientific knowledge (which at that time included philosophy too)
refers to the world of/as appearance. The world as a thing-in-itself is, accord-
A
ing to Kant, not a "possible object of experience", and all human (conceptual)
knowledge refers to the world as we experience it. By splitting the world in
the world in this way, Kant was able to offer compelling solutions to some of
the most historically difficult questions faced by philosophy. Most importanly,
it allowed him to offer a solution to the question of free will versus determin-
ism. Kant argued that, in the world of appearances, determinism is the rule. In
other words, according to Kant, in the world of appearances, there is no object
that is not governed by the laws of nature. However, this doesn’t preclude the
possibility that human freedom exists, with the proviso that it exists as a thing-
DR
in-itself. In other words, for Kant, human freedom is not a possible object of
experience, but that doesn’t make it any less real. Even though we can never
perceive human freedom, the mere fact that we can will actions for which we
can find no cause in the world of appearances is enough to make human free-
dom a reasonable assumption. It must remain an assumption, since we cannot
have knowledge of something that is not an object of experience, but it is an
assumption worth making, since it is what makes morality possible.
Thus, Kant was able to offer a coherent answer to the question of how it is pos-
sible for both free will and determinism to apply to the same world, but in order
to do so, he found it necessary to split the world into these two totally sepa-
rate aspects. This method made possible tremendous advances in philosophical
thought throughout the late eighteenth and early nineteenth centuries. How-
ever, it also set strict limits on human knowledge. For Kant, we cannot ’know’

Quantum mechanics, philosophy and controversy


480

freedom or any other thing-in-itself in a rigorously scientific way. Rather, free-


dom is more like a necessary assumption. We can only ’know’ objects as they
appear to us – our knowledge is only knowledge of the world of appearances.
Any claim to have knowledge of objects as they are in themselves is an illegiti-

FT
mate use of the faculties of reason and understanding. This is why, for Kant, it
is impossible to prove the existence of God, of the soul, or of human freedom:
none of these are possible objects of experience. This is not just a historical-
ly specific problem that might be overcome as science advances and we learn
more and more; it is constitutive of all human knowledge. In other words,
even while Kant enabled great leaps forward in philosophical thought, he did
so by introducing a concept of human knowledge as essentially limited, and
essentially fallible.
Although twentieth century scientists left the question of human will to the
philosophers, scientists themselves felt very firmly grounded in the idea that
science could make predictions according to Newton’s laws with regard to ob-
jects in nature. Therefore, determinism was still one of the fundamental axioms
of scientific thought. Even with Einstein’s theory of relativity, determinism was
not seriously challenged. This was all about to change.
A
Consequences of the uncertainty principle
The Uncertainty Principle is a main theory in the physical science of quantum
mechanics that explains the universe at atomic and subatomic scales.
The Uncertainty Principle was developed as an answer to the question: How
does one measure the location of an electron around a nucleus?
In March 1926, working in Niels Bohr’s institute, Werner Heisenberg formulat-
DR
ed the principle of uncertainty thereby laying the foundation of what became
known as the Copenhagen interpretation of quantum mechanics. Heisenberg
had been studying the papers of Paul Dirac and Jordan. Heisenberg discovered
a problem with measurement of basic variables in the equations. His analy-
sis showed that uncertainties, or imprecisions, always turned up if one tried
to measure the position and the momentum of a particle at the same time.
Heisenberg concluded that these uncertainties or imprecisions in the measure-
ments were not the fault of the experimenter, but fundamental in nature and
inherent in quantum mechanics.
The term Copenhagen interpretation of quantum mechanics was often used
interchangeably with and as a synonym for Heisenberg’s Uncertainty Principle
by detractors who believed in fate and determinism and saw the common fea-
tures of the Bohr-Heisenberg theories as a threat. Within the widely but not
universally accepted Copenhagen interpretation of quantum mechanics (i.e. it

Quantum mechanics, philosophy and controversy


481

was not accepted by Einstein or other physicists such as Alfred Lande), the un-
certainty principle is taken to mean that on an elementary level, the physical
universe does not exist in a deterministic form, but rather as a collection of
probabilities, or potentials. For example, the pattern (probability distribution)

FT
produced by millions of photons passing through a diffraction slit can be cal-
culated using quantum mechanics, but the exact path of each photon cannot
be predicted by any known method. The Copenhagen interpretation holds that
it cannot be predicted by any method, not even with theoretically infinitely
precise measurements.
Albert Einstein was not happy with the uncertainty principle, and he challenged
Niels Bohr and Werner Heisenberg with a famous thought experiment (See the
Bohr-Einstein debates for more details).
It is this interpretation that Einstein was questioning when he said "I cannot be-
lieve that God would choose to play dice with the universe." Bohr, who was one
of the authors of the Copenhagen interpretation responded, "Einstein, don’t tell
God what to do." Niels Bohr himself acknowledged that quantum mechanics
and the uncertainty principle were counter-intuitive when he stated, "Anyone
who is not shocked by quantum theory has not understood a single word."
A
The basic debate between Einstein and Bohr (including Heisenberg’s Uncer-
tainty Principle) was that Einstein was in essence saying: "Of course, we can
know where something is; we can know the position of a moving particle if we
know every possible detail, and thereby by extension, we can predict where
it will go." Bohr and Heisenberg were saying the opposite: "There is no way
to know where a moving particle is ever even given every possible detail, and
thereby by extension, we can never predict where it will go."
DR
Einstein was convinced that this interpretation was in error. His reasoning
was that all previously known probability distributions arose from determinis-
tic events. The distribution of a flipped coin or a rolled dice can be described
with a probability distribution (50% heads, 50% tails). But this does not mean
that their physical motions are unpredictable. Ordinary mechanics can be used
to calculate exactly how each coin will land, if the forces acting on it are known.
And the heads/tails distribution will still line up with the probability distribu-
tion (given random initial forces).
Einstein was not adverse to quantum mechanics as a whole, but specifically
with the uncertainty principle itself. As to other basic principles of quantum
mechanics, Einstein whose own general relativity was firmly rooted in field
theory said:
"The de Broglie-Schrödinger method, which has in a certain sense the charac-
ter of a field theory, does indeed deduce the existence of only discrete states,

Quantum mechanics, philosophy and controversy


482

in surprising agreement with empirical facts. It does so on the basis of differ-


ential equations applying a kind of resonance argument." (Albert Einstein, On
Quantum Physics, 1954)
Niels Bohr himself appears to have taken Kant’s view that there are two aspects

FT
of reality, what we can say about and what it is, when Bohr said:
"There is no quantum world. There is only an abstract physical description. It
is wrong to think that the task of physics is to find out how nature is. Physics
concerns what we can say about nature."
In other words, there may be no quantum leaping of electrons as predicted
by quantum mechanics, there may be a definite position of a particle contrary
to the uncertainty principle, but the only way we mere humans can describe
mathematically in a useful way what we see in the real world is to use Quantum
Mechanics. This is because theories are simple models of complex systems. The
universe is too complex to describe without simple models. Because quantum
mechanics is useful and continues to provide sound mathematics when tested,
it is a mainstream theory of the universe at the quantum level. Niels Bohr’s
comment was saying that he himself believed that in all probability the natural
A
world was different than the explanation given by quantum mechanics which
is similar to Kant’s view.
Heisenberg wrote a conversation between himself and Einstein further debat-
ing their different viewpoints as follows:

• Heisenberg: "One cannot observe the electron orbits inside the atom. [...]
but since it is reasonable to consider only those quantities in a theory that
can be measured, it seemed natural to me to introduce them only as entities,
as representatives of electron orbits, so to speak."
DR
• Einstein: "But you don’t seriously believe that only observable quantities
should be considered in a physical theory?"
• "I thought this was the very idea that your relativity theory is based on?"
Heisenberg asked in surprise.
• "Perhaps I used this kind of reasoning," replied Einstein, "but it is nonsense
nevertheless. [...] In reality the opposite is true: only the theory decides
what can be observed."– (translated from "Der Teil und das Ganze" by W.
Heisenberg)

Werner Heisenberg himself said, "‘I myself . . . only came to believe in the
uncertainty relations after many pangs of conscience. . . ." He knew what he
was saying didn’t make sense, but it helped measurements at quantum levels
so much, he did it anyway. Richard Feynman, another major contributor to
quantum theory said, "We have always had a great deal of difficulty under-
standing the world view that quantum mechanics represents. At least I do,

Quantum mechanics, philosophy and controversy


483

because I’m an old enough man that I haven’t got to the point that this stuff
is obvious to me. Okay, I still get nervous with it.... You know how it always
is, every new idea, it takes a generation or two until it becomes obvious that
there’s no real problem. I cannot define the real problem, therefore I suspect

FT
there’s no real problem, but I’m not sure there’s no real problem." He meant
that he understood quantum mechanics very well, but that in 1982 some 50
years later, he still couldn’t reconcile himself to it. That is why Einstein spent
the entire rest of his life trying to disprove the Uncertainty Principle however
there is no other theory to replace quantum mechanics that is so successful at
the quantum level.

Erwin Schrödinger controversy


Later in life, the inventor of wave mechanics of quantum theory, Erwin
Schrödinger began a campaign against the generally accepted quantum de-
scription of wave-particle duality and tried to propose a theory in terms of
waves only. This led him into controversy with other leading physicists since
he rejected mainstream quantum mechanical theory.
Sometimes Schrödinger’s wave equation is erroneously said to give the exact
A
location of the electron and doesn’t need the uncertainty principle. Actually
Schrödinger’s wave equation explains the exact location of a wave. A wave
not being a point particle has a natural integral probability distribution as a
widespread disturbance. So Schrödinger’s equation does in a sense give the
exact location of the electron, however, only in its state of being a widespread
disturbance, a wave. Schrödinger later in life tried to develop a theory that
would show the electron is only a wave and not a particle at all and that
fundamentally the atom is only a wave, thus making the uncertainty principle
DR
obsolete as it was only needed to show the uncertainty of the particle-like
position of the electron and other subatomic particles. This was not a new
theory. The idea that the atom could be explained mathematically as a wave
was introduced in 1922 by Charles Galton Darwin, a physicist, in his paper at
279
. However, the consequences would be that planets, galaxies, human beings
and atoms are completely described as waves of physical disturbance, some
waves being massless as in the case of light and some waves being massive
as the case of the subatomic particles of the atom. Einstein rejected such a
theory when Schrödinger proposed it to him. Einstein followed intuitive lines
of thinking which is why he rejected Heisenberg’s uncertainty principle. There
are difficulties in describing a single wave as having two polarities if the atom
were a single wave and the idea of waves producing spin and magnetic moment
seem hard to overcome, but a solution was proposed in 1927 by Arthur Ruark

279 http://www.pubmedcentral.gov/picrender.fcgi?tool=pmcentrez&blobtype=pdf&artid=1085216

Quantum mechanics, philosophy and controversy


484

in THE IMPULSE MOMENT OF THE LIGHT QUANTUM at 280. However, it is


accepted that quantum mechanics teaches that solid objects only appear solid
due to forces. The atom is mostly space and it is the negative charge of the
electrons that keep atoms from collapsing into each other.

FT
Schrödinger became so disenchanted with the idea of wave-particle duality
that he was known to have said concerning it:
"Let me say at the outset, that in this discourse, I am opposing not a few spe-
cial statements of quantum physics held today (1950s), I am opposing as it
were the whole of it, I am opposing its basic views that have been shaped 25
years ago, when Max Born put forward his probability interpretation, which
was accepted by almost everybody." (Schrödinger Erwin, The Interpretation of
Quantum Physics. Ox Bow Press, Woodbridge, CN, 1995).
"I don’t like it, and I’m sorry I ever had anything to do with it." (Erwin
Schrodinger talking about Quantum Physics)

Comments by other quantum physicists


"This is the third of four lectures on a rather difficult subject – the theory of
A
quantum electrodynamics – and since there are obviously more people here
tonight than there were before, some of you haven’t heard the other two lec-
tures and will find this lecture incomprehensible. Those of you who have heard
the other two lectures will also find this lecture incomprehensible, but you
know that that’s all right: as I explained in the first lecture, the way we have
to describe Nature is generally incomprehensible to us." Richard P. Feynman,
QED, The Strange Theory of Light and Matter, p. 77 [Princeton University
Press, 1985]
DR
"The discomfort that I feel is associated with the fact that the observed perfect
quantum correlations seem to demand something like the "genetic" hypothesis.
For me, it is so reasonable to assume that the photons in those experiments car-
ry with them programs, which have been correlated in advance, telling them
how to behave. This is so rational that I think that when Einstein saw that,
and the others refused to see it, he was the rational man. The other people, al-
though history has justified them, were burying their heads in the sand. I feel
that Einstein’s intellectual superiority over Bohr, in this instance, was enor-
mous; a vast gulf between the man who saw clearly what was needed, and
the obscurantist. So for me, it is a pity that Einstein’s idea doesn’t work. The
reasonable thing just doesn’t work." John Stewart Bell (1928-1990), author of
"Bell’s Theorem" (or "Bell’s Inequality"), quoted in Quantum Profiles, by Jeremy
Bernstein [Princeton University Press, 1991, p. 84]

280 http://www.pubmedcentral.gov/picrender.fcgi?tool=pmcentrez&blobtype=pdf&artid=1085228

Quantum mechanics, philosophy and controversy


485

"Thus the last and most successful creation of theoretical physics, namely quan-
tum mechanics (QM), differs fundamentally from both Newton’s mechanics,
and Maxwell’s e-m field. For the quantities which figure in Quantum Physics’
laws make no claim to describe physical reality itself, but only probabilities

FT
of the occurrence of a physical reality that we have in view. · · · I cannot but
confess that I attach only a transitory importance to this interpretation. I still
believe in the possibility of a model of reality - that is to say, of a theory which
represents things themselves and not merely the probability of their occur-
rence. On the other hand, it seems to me certain that we must give up the
idea of complete localization of the particle in a theoretical model. This seems
to me the permanent upshot of Heisenberg’s principle of uncertainty." (Albert
Einstein, On Quantum Physics, 1954)

External links
• The Center of Quantum Philosophy 281
A
Source: http://en.wikipedia.org/wiki/Quantum_mechanics%2C_philosophy_and_controversy

Quantum mineralogy

Quantum mineralogy is the branch of physics and chemistry that uses fun-
damental (quantum-level) properties of particular elements to describe the
macroscopic properties of minerals containing those elements.
DR
References
• Quantum mineralogy by Bryan C. Chakoumakos in McGraw-Hill Encyclo-
pedia of Science & Technology Online 282

Source: http://en.wikipedia.org/wiki/Quantum_mineralogy

281 http://www.quantumphil.org/
282 http://www.accessscience.com/Encyclopedia/5/56/Est_562950_frameset.html?doi

Quantum mineralogy
486

Quantum phase transition

In physics, a quantum phase transition (QPT) is a phase transition between

FT
different quantum phases (phases of matter at zero temperature). Contrary to
classical phase transitions, quantum phase transitions can be only be accessed
by varying a physical parameter - such as magnetic field or pressure - at abso-
lute zero temperature. The transition describes an abrupt change in the ground
state of a many-body system due to its quantum fluctuations. Such quantum
phase transitions can be first-order phase transition or continuous.
To understand quantum phase transitions, it is useful to contrast them to clas-
sical phase transitions (CPT) (also called thermal phase transitions). A CPT de-
scribes a discontinuity in the thermodynamic properties of a system. It signals
a reorganization of the particles; A canonical example is the freezing transition
of water describing the transition between liquid and ice. The classical phase
transitions are driven by a competition between the energy of a system and the
entropy of its thermal fluctuations. A classical system does not have entropy at
zero temperature and therefore no phase transition can occur.
A
In contrast, even at zero temperature a quantum-mechanical system has quan-
tum fluctuations and therefore can still support phase transitions. As a physical
parameter is varied, quantum fluctuations can drive a phase transition into a
different phase of matter. A canonical quantum phase transition is the well-
studied superconductor/insulator transition in disordered thin films which sep-
arates two quantum phases having different symmetries. Quantum magnets
provide another example of QPT.
DR
Source: http://en.wikipedia.org/wiki/Quantum_phase_transition

Principal Authors: Mikkalai, Taxman, Lankiveil, Commander Keane, Folajimi

Quantum solid

In physics, a quantum solid is a type of solid that is "intrinsically restless", in


the sense that atoms continuously vibrate about their position and exchange
places even at the absolute zero of temperature. The archetypal quantum solid
is low density solid helium.

Quantum solid
487

References
• E.Polturak and N.Gov, Inside a quantum solid, Contemporary Physics 44,
No.2, 145-151, (2003).

FT
Source: http://en.wikipedia.org/wiki/Quantum_solid

Quantum state

A
DR

In quantum mechanics, a quantum state is any possible state in which a quan-


tum mechanical system can be. A fully specified quantum state can be de-
scribed by a state vector, a wavefunction, or a complete set of quantum num-
bers for a specific system. A partially known quantum state, such as an ensem-
ble with some quantum numbers fixed, can be described by a density operator.

Quantum state
488

Bra-ket notation
Paul Dirac invented a powerful and intuitive notation to describe quantum
states, known as bra-ket notation. For instance, one can refer to an |excited
atom> or to |↑i for a spin-up particle, hiding the underlying complexity of the

FT
mathematical description, which is revealed when the state is projected onto
a coordinate basis. For instance, the simple notation |1s> describes the first
hydrogen atom bound state, but becomes a complicated function in terms of
Laguerre polynomials and spherical harmonics when projected onto the ba-
sis of position vectors |r>. The resulting expression Ψ(r)=<r|1s>, which is
known as the wavefunction, is a special representation of the quantum state,
namely, its projection into position space. Other representations, such as pro-
jection into momentum space, are possible. The various representations are
simply different expressions of a single physical quantum state.

Basis states
Any quantum state |ψi can be expressed in terms of a sum of basis states (also
called basis kets) |ki i in the form
A P
|ψi = i ci |ki i

where ci are the coefficients representing the probability amplitude, such that
the absolute square of the probability amplitude, |ci |2 is the probability of a
measurement in terms of the basis states yielding the state |ki i. The normal-
ization condition mandates that the total sum of probabilities is equal to one,
P 2
i |ci | = 1.
DR
The simplest understanding of basis states is obtained by examining the quan-
tum harmonic oscillator. In this system, each basis state |ni has an energy
En = ~ω n + 12 . The set of basis states can be extracted using a construc-


tion operator ↠and a destruction operator â in what is called the ladder oper-
ator method.

Superposition of states
If a quantum mechanical state |ψi can be reached by more than one path, then
|ψi is said to be a linear superposition of states. In the case of two paths, if the
states after passing through path α and path β are
|αi = √1 |0i + √1 |1i,
2 2
and

|βi = √1 |0i − √1 |1i,


2 2

Quantum state
489

then |ψi is defined as the normalized linear sum of these two states. If the two
paths are equally likely, this yields
|ψi = √12 |αi + √12 |βi = √1 ( √1 |0i + √1 |1i) +

FT
2 2 2
√1 ( √1 |0i − √1 |1i) = |0i.
2 2 2

Note that in the states |αi and |βi, the two states |0i and |1i each have a prob-
1
ability of 2 , as obtained by the absolute square of the probability amplitudes,
1 1
which are √2 and ± √2 . In a superposition, it is the probability amplitudes
which add, and not the probabilities themselves. The pattern which results
from a superposition is often called an interference pattern. In the above case,
|0i is said to constructively interfere, and |1i is said to destructively interfere.
For more about superposition of states, see the double-slit experiment.

Pure and mixed states


A pure quantum state is a state which can be described by a single ket vector,
A
or as a sum of basis states. A mixed quantum state is a statistical distribution
of pure states.
The expectation value hai of a measurement A on a pure quantum state is given
by
hai = hψ|A|ψi = i ai hψ|αi ihαi |ψi = i ai |hαi |ψi|2 = i ai P (αi )
P P P

where |αi i are basis kets for the operator A, and P (αi ) is the probability of |ψi
being measured in state |αi i.
DR
In order to describe a statistical distribution of pure states, or mixed state,
the density operator (or density matrix), ρ, is used. This extends quantum
mechanics to quantum statistical mechanics. The density operator is defined
as
P
ρ = s ps |ψs ihψs |

where ps is the fraction of each ensemble in pure state |ψs i. The ensemble
average of a measurement A on a mixed state is given by
[A] = hAi = s ps hψs |A|ψs i = s i ps ai |hαi |ψs i|2 = tr(ρA)
P P P

where 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 over the ensemble of pure states.

Quantum state
490

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

FT
See also
• →Quantum harmonic oscillator
• →Bra-ket notation
• Orthonormal basis
• →Wavefunction
• →Probability amplitude
• Density operator
• Qubit

Source: http://en.wikipedia.org/wiki/Quantum_state

Principal Authors: Cortonin, Fresheneesz, Laussy, Bkalafut, CSTAR


A
Quantum statistical mechanics

Quantum statistical mechanics is the study of statistical ensembles of quan-


tum mechanical systems. A statistical ensemble is described by a density op-
erator 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
DR
under various mathematical formalisms for quantum mechanics. One such for-
malism 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 D X by
R
Exp(X) = R
λ d DX (λ)

assuming, of course that the random variable is integrable or the random vari-
able 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 ofRA defined by
EA (U ) = U
λd E(λ),

Quantum statistical mechanics


491

uniquely determines A and conversely, is uniquely determined by A. E A 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,

FT
we introduce the distribution of A under S which is the probability measure
defined on the Borel subsets of R by
DA (U ) = Tr(EA (U )S).

Similarly, the expected value of A is defined in terms of the probability distri-


bution D A by
R
Exp(A) = R
λ d DA (λ).

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

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


Exp(A) = hψ|A|ψi.
DR
Von Neumann entropy
Of particular significance for describing randomness of a state is the von Neu-
mann entropy of S formally defined by
H(S) = − Tr(S log2 S)
.

Actually the operator S log 2 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

Quantum statistical mechanics


492
 
λ1 0 ··· 0 ···
0
 λ2 ··· 0 · · ·


 ··· 

0 0 ··· λn · · ·

FT
··· ···

and we define
P
H(S) = − i λi log2 λi .

The convention is that 0 log2 0 = 0, 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
1

2(log2 2)2 0 ··· 0 ···
1

 0 3(log2 3)2 ··· 0 · · ·

T =
A


 0 0
···
···
···
1
n(log2 n)2
···



· · ·

T is non-negative trace class and one can show T log 2 T is not trace-class.
Theorem. Entropy is a unitary invariant.
In analogy with classical entropy (notice the similarity in the definitions), H(S )
DR
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
1

n 0 ··· 0
0 1
 n ... 0
 ··· 
1
0 0 ··· n

For such an S, H(S ) = log 2 n. The state S is called the maximally mixed state.
Recall that a pure state is one the form
S = |ψihψ|,

for ψ a vector of norm 1.

Quantum statistical mechanics


493

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.

FT
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 En of H go to +
∞ sufficiently fast, e -r H will be a non-negative trace-class operator for ever
positive r.
The Gibbs canonical ensemble is described by the state
e−βH
S= Tr(e−βH )

where β is such that the ensemble average of energy satisfies


Tr(SH) = E
A
,and
Tr(e−βH ) =
P
n e−βEn

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 Em is
Pe−βE−βE
m
n
.
e
DR
n

Under certain conditions the Gibbs canonical ensemble maximizes the von Neu-
mann 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, 1985.

Source: http://en.wikipedia.org/wiki/Quantum_statistical_mechanics

Principal Authors: CSTAR, Mct mht, Sietse Snel, Phys, E2mb0t

Quantum statistical mechanics


494

Quantum superposition

Quantum superposition is the application of the superposition principle to

FT
quantum mechanics. The superposition principle is the addition of the ampli-
tudes of waves from interference. In quantum mechanics it is the amplitudes
of wavefunctions, or state vectors, that add. It occurs when an object simul-
taneously "possesses" two or more values for an observable quantity (e.g. the
position or energy of a particle).
More specifically, in quantum mechanics, any observable quantity corresponds
to an eigenstate of a Hermitian linear operator. The linear combination of two
or more eigenstates results in quantum superposition of two or more values of
the quantity. If the quantity is measured, the projection postulate states that
the state will be randomly collapsed onto one of the values in the superpo-
sition (with a probability proportional to the square of the amplitude of that
eigenstate in the linear combination).
The question naturally arose as to why "real" (macroscopic, Newtonian) ob-
A
jects and events do not seem to display quantum mechanical features such as
superposition. In 1935, Erwin Schrödinger devised a well-known thought ex-
periment, now known as →Schrödinger’s cat, which highlighted the dissonance
between quantum mechanics and Newtonian physics.
In fact, quantum superposition does result in many directly observable effects,
such as interference peaks from an electron wave in a double-slit experiment.
If two observables correspond to noncommutative operators, they obey an un-
certainty principle and a distinct state of one observable corresponds to a su-
DR
perposition of many states for the other observable.

See also
• →Wave packet
• Quantum computation
• →Penrose Interpretation

Source: http://en.wikipedia.org/wiki/Quantum_superposition

Principal Authors: Reddi, Stevenj, Stevertigo, PopUpPirate, D.328, El C, John187, Retodon8

Quantum superposition
495

Quantum Theory Parallels to Conscious-


ness

FT
Can parallels to certain generic concepts and logical relationships of quantum
theory:

• help theoretically characterize general features of conscious experience, in


particular, nonlocal consciousness?

• transcend the scope of analogy and metaphor to enhance quantitative pre-


diction and provide guidelines for experimental design?

Ordering in random physical processes may be attached to "mindful" medita-


tion via lucid empathy and participatory interest making otherwise stochastic
processes vulnerable to anomalous statistical behavior. (See, for example, the
experiments on "Field Consciousness" discussed in Radin, 1997, Chapter 10,
pp. 157-174).
A
During so-called "unusual", "expanded" or "transcendental" states of conscious-
ness, certain macromolecules or entire cell ensembles in the brain may be men-
tally decoupled from their thermodynamic environment, thus enabling them to
exist in quantum states. By their inherent "long-wavelength" nature, quantum
states are "spread out" beyond their immediate locus. Accordingly, quantum
mind-brain states would be capable of entanglement with similar states in the
mind-brains of other individuals or with labile (stochastic) states of matter in
the remote environment (Schmid, 2000a), (Schmid, 2000c). Labile states of
DR
matter are inherent to the stochastic behavior of inanimate matter, the psy-
chomotorical lability of animate beings, the neuropsychoemotional ambiva-
lence of cognitive beings as well as the ill/pathological/disturbed states of liv-
ing tissue in general.
Such nonlocal states of consciousness may include prayer, meditation, trance,
and dreaming, as well as mystical, out-of-body, and near-death experiences
and, especially, pathological mental states induced by drugs or psychosis. In
fact, "faith", in the sense of an open ("mindful") trusting belief in an inner
connection to other people as well as to the world in general is the core of
spirituality: meaningfulness and belonging.
Intention and decision involve precise cognitive-emotional processes which
may correspond to short-wavelength wavefunctions which are sharply local-
ized. By contrast, "attachment" may reduce potential barriers around certain
macromolecules in the mind-brain, thus relaxing tightly centered (classical)

Quantum Theory Parallels to Consciousness


496

molecular couplings to their biophysical surroundings and leading to the men-


tal quality of "clarity". Subjective "clarity" (lucid empathy, participatory interest
and vivid mental imagery void of expectation) may enable such thermodynam-
ically decoupled states to selectively entangle with others.

FT
The quantum physicist Niels Bohr once said that mental clarity is comple-
mentary to mental precision. How these ideas allow for such phenomena as
precognition, telepathy - encompassing clairvoyance, distant anticipation, re-
mote perception, synchronicity, and the like - psychic healing, and psychoki-
nesis between the members of a fortune teller/event-, percipient/agent-, heal-
er/patient-, or influencer/object-pair is discussed.

References
• Braud, W. G., & Schlitz, M. J. (1991). Consciousness Interactions with re-
mote biological systems: Anomalous intentionality effects. Subtle Energies,
2(1), 1-46.

• Jahn, R. G., & Dunne, B. J. (1986). On the quantum mechanics of con-


Asciousness, with application to anomalous phenomena. Foundations of
Physics, 16(8), 721-772.

• Radin, D. I. (1997). The Conscious Universe: The Scientific Truth of Psychic


Phenomena. San Francisco: HarperEdge.

• Schmid, G. B. (2000a). Das Geheimnis psychogener Todesfälle. intra -


Psychologie und Gesellschaft, 45(September), 14-23.
DR
• Schmid, G. B. (2000b). Tod durch Vorstellungskraft: Das Geheimnis psy-
chogener Todesfälle (1. ed.). Wien-New York: Springer-Verlag.

• Schmid, G. B. (2000c). Tod durch Vorstellungskraft? Die geheimnisvolle


Macht der Gedanken. An der Urania 17 / D-10787 Berlin: URANIA.

• Tittel, W., Brendel, J., Gisin, B., Herzog, T., Zbinden, H., & Gisin, N. (1998).
Experimental demonstration of quantum-correlations over more than 10
kilometers. Physical Review A, 57(5), 3229-3232.

Source: http://en.wikipedia.org/wiki/Quantum_Theory_Parallels_to_Consciousness

Principal Authors: Cholmes75, Omphaloscope, Harlanpaine

Quantum Theory Parallels to Consciousness


497

Quantum tomography

Quantum tomography or quantum state tomography is the process of recon-

FT
structing the quantum state (density matrix) for a source of quantum systems
by measurements on the systems coming from the source. To be able to unique-
ly identify the state, the measurements must be tomographically complete, that
is the measured operators must form an operator basis on the →Hilbert space
of the system.
In quantum process tomography on the other hand, known quantum states are
used to probe a quantum process to find out how the process can be described.

Source: http://en.wikipedia.org/wiki/Quantum_tomography

Principal Authors: V79, Remuel, Conscious

Quantum tunnelling
A
Quantum tunnelling (or tunneling) is the quantum-mechanical effect of tran-
sitioning through a classically-forbidden energy state. It can be generalized to
other types of classically-forbidden transitions as well.
Consider rolling a ball up a hill. If the ball is not given enough velocity, then
it will not roll over the hill. This scenario makes sense from the standpoint
of classical mechanics, but is an inapplicable restriction in quantum mechanics
DR
simply because quantum mechanical objects do not behave like classical objects
such as balls. On a quantum scale, objects exhibit wavelike behavior. For a
quantum particle moving against a potential energy "hill", the wave function
describing the particle can extend to the other side of the hill. This wave
represents the probability of finding the particle in a certain location, meaning
that the particle has the possibility of being detected on the other side of the
hill. This behavior is called tunnelling; it is as if the particle has ’dug’ through
the potential hill.
As this is a quantum and non-classical effect, it can generally only be seen
in nanoscopic phenomena — where the wave nature of particles is more pro-
nounced.
It should be noted that availability of states is necessary for tunneling to occur.
In the above example, the quantum mechanical ball will not appear inside the
hill because there is no available "space" for it to exist, but it can tunnel to

Quantum tunnelling
498

the other side of the hill, where there is free space. Analogously, a particle
can tunnel through the barrier, but unless there are states available within
the barrier, the particle can only tunnel to the other side of the barrier. The
wavefunction describing a particle only expresses the probability of finding the

FT
particle at a location assuming a free state exists.

History and consequences


In the early 1900s, radioactive materials were known to have characteristic
exponential decay rates or half lives. At the same time, radiation emissions
were known to have certain characteristic energies. By 1928, George Gamow
had solved the theory of the alpha decay of a nucleus via tunnelling. Classically,
the particle is confined to the nucleus because of the high energy requirement
to escape the very strong potential. Classically, it takes an enormous amount
of energy to pull apart the nucleus. In quantum mechanics, however, there is
a probability the particle can tunnel through the potential and escape. Gamow
solved a model potential for the nucleus and derived a relationship between
the half-life of the particle and the energy of the emission.
Alpha decay via tunnelling was also solved concurrently by Ronald Gurney and
A
Edward Condon. Shortly thereafter, both groups considered whether particles
could also tunnel into the nucleus.
After attending a seminar by Gamow, Max Born recognized the generality of
quantum-mechanical tunnelling. He realised that tunnelling phenomena was
not restricted to nuclear physics, but was a general result of quantum mechan-
ics that applies to many different systems. Today the theory of tunnelling is
even applied to the early cosmology of the universe.
DR
Quantum tunnelling was later applied to other situations, such as the cold
emission of electrons, and perhaps most importantly semiconductor and su-
perconductor physics. Phenomena such as field emission, important to flash
memory, are explained by quantum tunnelling. Tunnelling is a source of major
current leakage in Very-large-scale integration (VLSI) electronics, and results
in the substantial power drain and heating effects that plague high-speed and
mobile technology.
Another major application is in electron-tunnelling microscopes (see scanning
tunnelling microscope) which can resolve objects that are too small to see us-
ing conventional microscopes. Electron tunnelling microscopes overcome the
limiting effects of conventional microscopes (optical aberrations, wavelength
limitations) by scanning the surface of an object with tunnelling electrons.

Quantum tunnelling
499

Semiclassical calculation
Let us consider the time-independent →Schrödinger equation for one particle,
in one dimension, under the influence of a hill potential V (x).
2 2
~ d

FT
− 2m dx2
Ψ(x) + V (x)Ψ(x) = EΨ(x)

d2 2m
dx2
Ψ(x) = ~2
(V (x) − E) Ψ(x)

Now let us recast the wave function Ψ(x) as the exponential of a function.
Ψ(x) = eΦ(x)

Φ00 (x) + Φ0 (x)2 = 2m


~2
(V (x) − E)

Now let us separate Φ0 (x) into real and imaginary parts.


Φ0 (x) = A(x) + ıB(x)

A0 (x) + A(x)2 − B(x)2 = 2m


(V (x) − E)
A ~2

B 0 (x) − 2A(x)B(x) = 0

Next we want to take the semiclassical approximation to solve this. That means
we expand each function as a power series in ~. From the equations we can
already see that the power series must start with at least an order of ~−1 to
satisfy the real part of the equation. But as we want a good classical limit, we
DR
also want to start with as high a power of Planck’s constant as possible.
A(x) = ~1 ∞ i
P
i=0 ~ Ai (x)

1 P∞ i B (x)
B(x) = ~ i=0 ~ i

The constraints on the lowest order terms are as follows.


A0 (x)2 − B0 (x)2 = 2m (V (x) − E)

A0 (x)B0 (x) = 0

If the amplitude varies slowly as compared to the phase, we set A0 (x) = 0 and
get

Quantum tunnelling
500
p
B0 (x) = ± 2m (E − V (x))

Which is obviously only valid when you have more energy than potential -
classical motion. After the same procedure on the next order of the expansion

FT
we get
R p 2m
ı dx (E−V (x))+θ
~2
Ψ(x) ≈ C e √ 2m
[4] 2 (E−V (x))
~

On the other hand, if the phase varies slowly as compared to the amplitude,
we set B0 (x) = 0 and get
p
A0 (x) = ± 2m (V (x) − E)

Which is obviously only valid when you have more potential than energy -
tunnelling motion. Grinding out the next order of the expansion yields
R p 2m R p 2m
+ dx (V (x)−E) − dx (V (x)−E)
C+ e ~2 +C− e ~2
Ψ(x) ≈ √
[4] 2m (V(x)−E)
~2
A
It is apparent from the denominator, that both these approximate solutions
are bad near the classical turning point E = V (x). What we have are the
approximate solutions away from the potential hill and beneath the potential
hill. Away from the potential hill, the particle acts similarly to a free wave
- the phase is oscillating. Beneath the potential hill, the particle undergoes
exponential changes in amplitude.
In a specific tunnelling problem, Rwe p
might already suspect that the transition
− dx 2m (V (x)−E)
amplitude be proportional to e ~2 and thus the tunnelling be
DR
exponentially dampened by large deviations from classically permitable mo-
tion.
But to be complete we must find the approximate solutions everywhere and
match coefficients to make a global approximate solution. We have yet to
approximate the solution near the classical turning points E = V (x).
Let us label a classical turning point x1 . Now because we are near E = V (x1 ),
we can easily expand 2m ~2
(V (x) − E) in a power series.
2m
~2
(V (x) − E) = U1 (x − x1 ) + U2 (x − x1 )2 + · · ·

2m
Let us only approximate to linear order ~2
(V (x) − E) = U1 (x − x1 )
2
d
dx2
Ψ(x) = U1 (x − x1 )Ψ(x)

Quantum tunnelling
501

This differential equation looks deceptively simple. It takes some trickery to


transform this into a Bessel equation. The solution is as follows.
√   √ 1
  √ 1

Ψ(x) = x − x1 C+ 1 J+ 1 23 U1 (x − x1 ) 3 + C− 1 J− 1 23 U1 (x − x1 ) 3
3 3 3 3

FT
Hopefully this solution should connect the far away and beneath solutions.
Given the 2 coefficients on one side of the classical turning point, we should
be able to determine the 2 coefficients on the other side of the classical turning
point by using this local solution to connect them. We should be able to find a
relationship between C, θ and C+ , C− .
Fortunately the Bessel function solutions will asymptote into sine, cosine and
exponential functions in the proper limits. The relationship can be found as
follows.
C+ = 21 C cos θ − π4


π

C− = −C sin θ − 4
A
Now we can easily construct global solutions and solve tunnelling problems.
outgoing 2
C
The transmission coefficient, C
, for a particle tunnelling through a
incoming
single potential barrier is found to be
R x2 p 2m
−2 dx (V (x)−E)
e x1 ~2
T =  R x2 p 2m 2
−2 dx (V (x)−E)
1+ 14 e x1 ~2
DR
Where x1 , x2 are the 2 classical turning points for the potential barrier. If
we take the classical limit of all other physical parameters much larger than
Planck’s constant, abbreviated as ~ → 0, we see that the transmission co-
efficient correctly goes to zero. This classical limit would have failed in the
unphysical, but much simpler to solve, situation of a square potential.

Quantum tunnelling
502



A
See also
→Josephson effect
→SQUID
FT
DR
• Tunnel diode
• WKB approximation
• Scanning tunnelling microscope

References
• Razavy, Mohsen (2003). Quantum Theory of Tunneling. World Scientific.
ISBN 9812380191.
• Griffiths, David J. (2004). Introduction to Quantum Mechanics (2nd ed.).
Prentice Hall. ISBN 013805326X.
• Liboff, Richard L. (2002). Introductory Quantum Mechanics. Addison-
Wesley. ISBN 0805387145.

Quantum tunnelling
503

External links

Source: http://en.wikipedia.org/wiki/Quantum_tunnelling

FT
Principal Authors: Light current, C h fleming, Yguff88, GregRM, Folajimi, Mako098765, Nobbie,
Shaddack, Asr1, Oleg Alexandrov

Quantum vibration

A quantum vibration is a vibration of a chemical bond in a molecule that


must be treated quantum mechanically. The low-lying vibration energy states
can be described as states of the quantum harmonic oscillator, and as higher
vibrational states, near the bond disassociation limit, as Morse oscillators.
A molecule can vibrate in many ways and each of them we can call a vibra-
tional mode. The vibrations can be seen with IR spectroscopy for example. As a
help to calculate the number of vibrational modes, it’s convenient to determine
A
the number of degrees of freedom available to vibration. As a generalization
any molecule consisting of N atoms will have 3N freedoms for translational
motion. 3 degrees are translational freedoms. There are also 3 degrees of ro-
tational freedom for non-linear molecules and 2 degrees of rotational freedom
for linear molecules. This leaves 3N-5 degrees of vibrational freedom for linear
molecules and 3N-6 degrees of vibrational freedom for non-linear molecules.
As an example H 2O, a non-linear molecule, will have 3*3-6 = 3 degrees of
vibrational freedoms, or modes.
DR
Source: http://en.wikipedia.org/wiki/Quantum_vibration

Principal Authors: Wolf530, Salsb, Cypa, Grendelkhan, Conscious

Quantum vibration
504

Quantum well

A quantum well is a potential well that confines particles, which were orig-

FT
inally free to move in three dimensions, to two dimensions, forcing them to
occupy a planar region. The effects of quantum confinement take place when
the quantum well thickness becomes comparable at the de Broglie wavelength
of the carriers (generally electrons and holes), leading to energy levels called
"energy subbands", i.e., the carriers can only have discrete energy values.

Fabrication
Quantum wells are formed in semiconductors by having a material, like gallium
arsenide sandwiched between two layers of a material with a wider bandgap,
like aluminium arsenide. These structures can be grown by molecular beam
epitaxy or chemical vapor deposition with control of the layer thickness down
to monolayers.

Applications
A
Because of their quasi-two dimensional nature, electrons in quantum wells
have a sharper density of states than bulk materials. As a result quantum
wells are in wide use in diode lasers. They are also used to make HEMTs (High
Electron Mobility Transistors), which are used in low-noise electronics
By doping either the well itself, or preferably, the barrier of a quantum well
with donor impurities, a two-dimensional electron gas (abbreviated 2DEG) can
be formed. This quasi-two dimensional system has interesting properties at
DR
low temperature. One such property is the quantum Hall effect, seen at high
magnetic fields. Acceptor dopants can lead to a two-dimensional hole gas.

See also
• →Particle in a box
• Quantum wire
• Quantum dot
• Quantum-well intermixing (QWI)

Source: http://en.wikipedia.org/wiki/Quantum_well

Principal Authors: Jaraalbe, Tantalate, Rkuchta, DV8 2XL, Dobromila

Quantum well
505

Quantum Zeno effect

The quantum Zeno effect is a quantum mechanical phenomenon first de-

FT
scribed by George Sudarshan and Baidyanaith Misra of the University of Texas
in 1977. It describes the situation that an unstable particle, if observed con-
tinuously, will never decay. This occurs because every measurement causes the
wavefunction to "collapse" to a pure eigenstate of the measurement basis.
In general, the Zeno effect can be de ned as class of phenomena when a tran-
sition is suppressed by some interaction which allows the interpretation of the
nal state in terms of "a transition has not yet occurred" or "a transition already
occurred" ( 283). In quantum mechanics, such an interaction is called “measure-
ment” because its result can be interpreted in terms of classical mechanics.
Frequent measurement prohibits the transition. Various versions of the Zeno
e ect fall into the de nition above.
Given a system in a state A, which is the eigenstate of some measurement oper-
ator. Say the system under free time evolution will decay with a certain prob-
Aability into state B. If measurements are made periodically, with some finite
interval between each one, at each measurement, the wavefunction collapses
to an eigenstate of the measurement operator. Between the measurements,
the system evolves away from this eigenstate into a superposition state of the
states A and B. When the superposition state is measured, it will again collapse,
either back into state A as in the first measurement, or away into state B. The
probability that it will collapse back into the same state A is higher if the sys-
tem has had less time to evolve away from it. In the limit as the time between
measurements goes to zero, the probability of a collapse back to the original
DR
state A goes to one. Hence, the system doesn’t evolve from A to B.
In reality, collapse of the wavefunction is not a discrete, instantaneous event. A
measurement could be approximated by strongly coupling the quantum system
to the noisy thermal environment for a brief period of time. The time it takes
for the wavefunction to "collapse" is related to the decoherence time of the
system when coupled to the environment. The stronger the coupling is, and the
shorter the decoherence time, the faster it will collapse. So in the decoherence
picture, the quantum Zeno effect corresponds to the limit where a quantum
system is continuously coupled to the environment, and where that coupling is
infinitely strong, and where the "environment" is an infinitely large source of
thermal randomness.

283 http://annex.jsap.or.jp/OSJ/opticalreview/TOC-Lists/vol12/12e0363tx.htm

Quantum Zeno effect


506

Experimentally, strong suppression of the evolution of a quantum system due


to environmental coupling has been observed in a number of microscopic sys-
tems. One such experiment was performed in October 1989 by Itano, Heinzen,
Bollinger and Wineland at NIST ( PDF 284). Approximately 5000 9Be + ions were

FT
stored in a cylindrical Penning trap and laser cooled to below 250mK. A res-
onant RF pulse was applied which, if applied alone, would cause the entire
ground state population to migrate into an excited state. After the pulse was
applied, the ions were monitored for photons emitted due to relaxation. The
ion trap was then regularly "measured" by applying a sequence of ultraviolet
pulses, during the RF pulse. As expected, the ultraviolet pulses suppressed
the evolution of the system into the excited state. The results were in good
agreement with theoretical models.
The quantum Zeno effect takes its name from Zeno’s arrow paradox, which is
the argument that since an arrow in flight does not move during any single
instant, it couldn’t possibly be moving overall.

See also
A• Interference

Source: http://en.wikipedia.org/wiki/Quantum_Zeno_effect

Principal Authors: Tim Starling, Seth Ilys, Domitori, Pjacobi, Gerd Breitenbach

Quasistability
DR
Quasistability is the local stability of a system at a local minimum of a poten-
tial. The local minimum usually is not a global true minimum of the potential.
The quasistable state may decay to a global minimum state via quantum me-
chanic effect.

Source: http://en.wikipedia.org/wiki/Quasistability

284 http://www.boulder.nist.gov/timefreq/general/pdf/858.pdf

Quasistability
507

QWiki

Qwiki 285 is a quantum physics wiki devoted to the collective creation of con-

FT
tent that is technical and useful to practicing scientists. The site is nominally
centered around quantum physics, but all scientists are invited to contribute,
including – but not limited to – computer scientists, control theorists, electri-
cal engineers, and mathematicians. More specifically, this site is designed for
people who post content to the arXiv and quant-ph.

Source: http://en.wikipedia.org/wiki/QWiki

Principal Authors: Melaen, Mellery, JanusDC

Range criterion
AIn quantum mechanics, in particular quantum information, the Range criteri-
on is a necessary condition that a state must satisfy in order to be separable.
In other words, it is a separability criterion.

The result
Consider a quantum mechanical system composed of n subsystems. The state
space H of such a system is the tensor product of those of the subsystems, i.e.
H = H1 ⊗ · · · ⊗ Hn .
DR
For simplicity we will assume throughout that all relevant state spaces are finite
dimensional.
The criterion reads as follows: If ρ is a separable mixed state acting on H, then
the range of ρ is spanned by a set of product vectors.

Proof
In general, if a matrix M is of the form M = i vi vi∗ , it is obvious that the
P

range of M, Ran(M), is contained in the linear span of {vi }. On the other


hand, we can also show vi lies in Ran(M), for all i. Assume w.l.o.g. i = 1.
We can write M = v1 v1∗ + T , where T is Hermitian and positive semidefinite.
There are two possibilities:
1) span{v1 } ⊂Ker(T). Clearly, in this case, v1 ∈ Ran(M).

285 http://’’’qwiki’’’.caltech.edu

Range criterion
508

2) Notice 1) is true if and only if Ker(T) ⊥ ⊂ span{v1 }⊥ , where ⊥ denotes


orthogonal compliment. By Hermiticity of T, this is the same as Ran(T) ⊂
span{v1 }⊥ . So if 1) does not hold, the intersection Ran(T) ∩ span{v1 } is
nonempty, i.e. there exists some complex number α such that T w = αv1 . So

FT
M w = hw, v1 iv1 + T w = (hw, v1 i + α)v1 .

Therefore v1 lies in Ran(M).


Thus Ran(M) coincides with the linear span of {vi }. The range criterion is a
special case of this fact.
A density matrix ρ acting on H is separable if and only if it can be written as
P ∗ ⊗ · · · ⊗ ψ ψ∗
ρ = i ψ1,i ψ1,i n,i n,i

∗ is a (un-normalized) pure state on the j -th subsystem. This is


where ψj,i ψj,i
also
P ∗ ⊗ · · · ⊗ ψ ∗ ).
ρ = i (ψ1,i ⊗ · · · ⊗ ψn,i )(ψ1,i n,i
A
But this is exactly the same form as M from above, with the vectorial product
state ψ1,i ⊗ · · · ⊗ ψn,i replacing vi . It then immediately follows that the range
of ρ is the linear span of these product states. This proves the criterion.

References
• P. Horodecki, "Separability Criterion and Inseparable Mixed States with Pos-
itive Partial Transposition", Physics Letters A 232, (1997).
DR
Source: http://en.wikipedia.org/wiki/Range_criterion

Relativistic particle

A relativistic particle is a particle moving with a speed close to the speed of


light, such that effects of special relativity are important for the description of
its behavior.
Massless particles (e.g., photons) are always moving at the speed of light,
therefore they are always relativistic.
Massive particles are relativistic when their kinetic energy is comparable or
greater than the energy mc2 corresponding to their rest mass. (This condition

Relativistic particle
509

implies that their speed is close to the speed of light.) Such relativistic particles
are generated in particle accelerators, and are naturally occurring in cosmic ra-
diation. In astrophysics, jets of relativistic plasma are produced by the centers
of active galaxies and quasars.

FT
A charged relativistic particle crossing the interface of two media with different
dielectric constants emits transition radiation. This is exploited in the transition
radiation detectors of high-velocity particles.
See also:

• Special relativity
• Relativistic wave equations
• Lorentz factor
• Relativistic mass
• Relativistic plasma
• Relativistic jet
• Relativistic beaming
A
Source: http://en.wikipedia.org/wiki/Relativistic_particle

Resonance

This article is about resonance in physics. For other senses of this term, see
resonance (disambiguation).
DR
In physics, resonance is the tendency of a system to oscillate with higher am-
plitude when the frequency of its oscillations matches the system’s natural fre-
quency of vibration (its resonant frequency) than it does at other frequencies.

Examples
Examples are the acoustic resonances of musical instruments, the tidal reso-
nance of the Bay of Fundy, orbital resonance as exemplified by some moons
of the solar system’s gas giants, the resonance of the basilar membrane in the
biological transduction of auditory input, and resonance in electrical circuits.
A resonant object, whether mechanical, acoustic, or electrical, will probably
have more than one resonant frequency (especially harmonics of the strongest
resonance). It will be easy to vibrate at those frequencies, and more difficult
to vibrate at other frequencies. It will "pick out" its resonant frequency from

Resonance
510

a complex excitation, such as an impulse or a wideband noise excitation. In


effect, it is filtering out all frequencies other than its resonance.
See also: center frequency

FT
Theory
For a linear oscillator with a resonant frequency Ω, the intensity of oscillations
I when the system is driven with a driving frequency ω is given by:
Γ
I(ω) ∝ 2
2 .
(ω−Ω) +( Γ2 )
2

The intensity is defined as the square of the amplitude of the oscillations. This
is a Lorentzian function, and this response is found in many physical situations
involving resonant systems. Γ is a parameter dependent on the damping of the
oscillator, and is known as the linewidth of the resonance. Heavily damped os-
cillators tend to have broad linewidths, and respond to a wider range of driving
frequencies around the resonant frequency. The linewidth is inversely propor-
tional to the Q factor, which is a measure of the sharpness of the resonance.
A
Quantum mechanics
A resonance is a quantum state whose mean energy lies above the fragmenta-
tion threshold of a system and is associated with:

• a pronounced variation of the cross sections if the fragmentation energy lies


in the neighbourhood of the energy of the resonance (energy-dependent
definition) - The width of this neighbourhood is called the width of the
resonance.
DR
• an exponential decay of the system when the system has a mean ener-
gy close to the resonance energy (time-dependent definition, i.e. in time-
resolved spectroscopy) - The lifetime (or inverse of the exponent of the ex-
ponential signal) of the resonance is proportional to the inverse of its width.
Resonances are usually classified into shape and Feshbach resonances or in-
to Breit-Wigner and →Fano resonances.

Quantum field theory


In quantum field theory, resonance is an unstable particle/bound state. It is
characterized by a complex pole off the real line in the S-matrix (which hap-
pens to be analytic). A sharp resonance is a resonance with a sharp peak in
the S-matrix (which corresponds to a long lifetime compared to the reciprocal
of its mass) while a broad resonance is a resonance with a spread out peak
(which corresponds to a short lifetime relative to the reciprocal of its mass). If

Resonance
511

a resonance is too broad, it might not be considered as a particle at all even if


it has a complex pole (far from the real line).
See also relativistic Breit-Wigner distribution
If the resonance happens to be a "fundamental particle" (i.e. described by a

FT
"fundamental field" of its own), it shows up as a complex pole off the real line
in the 2-point connected correlation function (i.e. the propagator).

’Old Tacoma Narrows’ bridge failure


The Old Tacoma Narrows Bridge has been popularized in physics text books
as a classical example of resonance, but this description is misleading. It is
more correct to say that it failed due to the action of self-excited forces, by
an aeroelastic phenomenon known as flutter. Robert H. Scanlan, father of the
field of bridge aerodynamics, wrote an article about this misunderstanding 286.

See also
• Center frequency
A•

Driven harmonic motion
Formant
• Harmonic oscillator
• Impedance
• Q factor
• Resonator
• Schumann resonance
• Simple harmonic motion
DR
• Tuned circuit
• Wave
• Gluonic vacuum field

Reference

External links
• Lectures in Physics 287 - Resonance from an energetic perspective
• RMCybernetics - Resonance 288 Resonance Research.

286 K. Billah and R. Scanlan (1991), Resonance, Tacoma Narrows Bridge Failure, and Undergraduate

Physics Textbooks, American Journal of Physics, 59(2), 118–124 (PDF)(http://www.ketchum.org


/billah/Billah-Scanlan.pdf)
287 http://www.vias.org/physics/bk3_02_04.html
288 http://www.rmcybernetics.com/research/resonance/resonance.htm

Resonance
512

• Greene, Brian, " Resonance in strings 289". The Elegant Universe, NOVA
(PBS)
• Hyperphysics section on resonance concepts 290
• A short FAQ on quantum resonances 291

FT
• Resonance versus resonant 292
• YouTube 293 - Video of the effects of resonance on rice

Source: http://en.wikipedia.org/wiki/Resonance

Principal Authors: Omegatron, Heron, Jitse Niesen, Michael Hardy, Hyacinth, DrBob, N.MacInnes,
Phys

Ring wave guide

In quantum mechanics, the ring wave guide starts from the one dimensional,
time independent →Schrödinger equation:
A ~
− 2m
2
∇2 ψ = Eψ

Using polar coordinates on the 1 dimensional ring, the wave function depends
only on the angular coordinate, and so
1 ∂2
∇2 = r2 ∂θ2

Requiring that the wave function be periodic in θ with a period 2 π (from the
DR
demand that the wave functions be single-valued functions on the circle), and
that they be normalized leads to the conditions
R 2π 2
0 |ψ(θ)| dθ = 1 ,

and
ψ(θ) = ψ(θ + 2 π)

Under these conditions, the solution to the Schrodinger equation is given by

289 http://www.pbs.org/wgbh/nova/elegant/resonance.html
290 http://hyperphysics.phy-astr.gsu.edu/hbase/sound/rescon.html#c1
291 http://www.thch.uni-bonn.de/tc/people/brems.vincent/vincent/faq.html
292 http://users.ece.gatech.edu/~mleach/misc/resonance.html
293 http://www.youtube.com/watch?v=Zkox6niJ1Wc

Ring wave guide


513
r

ψ(θ) = √1 e±i ~ 2mEθ

The energy eigenvalues E are quantized because of the periodic boundary con-
ditions, and they are required to satisfy

FT
r
√ r

e±i ~ 2mEθ = e±i ~ 2mE(θ+2π) , or

r

e±i2π ~ 2mE = 1 = ei2πn

This leads to the energy eigenvalues


n2 ~2
E= 2mr2
where n = 0, 1, 2, 3, . . .

The full wave functions are, therefore


ψ(θ) = √1 e±inθ

Quantum states found:


n = 0:
Aψ is a constant function, and E = 0. This represents a stationary particle
(no angular momentum spinning around the ring).

n = 1:
~2
E= 2mr2

and
DR
ψ(θ) = √1 e±iθ

This produces two independent states that have the same energy level (de-
generacy) and can be linearly combined arbitrarily; instead of exp(± · · ·)
one can choose the sine and cosine functions. These two states represent
particles spinning around the ring in clockwise and counterclockwise direc-
tions. The angular momentum is ±~.

n = 2 (and higher):
the energy level is proportional to n2 , the angular momentum to n. There
are always two (degenerate) quantum states.

Ring wave guide


514

Except for the case n = 0, there are two quantum states for every value of n
(corresponding to e±inθ ). Therefore there are 2n+1 states with energies less
than an energy indexed by the number n.

FT
Application
In organic chemistry, aromatic compounds contain atomic rings, such as ben-
zene rings (the Kekulé structure) consisting of five or six, usually carbon,
atoms. So does the surface of "buckyballs" (buckminsterfullerene). These
molecules are exceptionally stable.
The above explains why the ring behaves like a circular wave guide. The excess
(valency) electrons spin around in both directions.
To fill all energy levels up to n requires 2 × (2n + 1) electrons, as electrons have
additionally two possible orientations of their spins.
The rule that 4n + 2 excess electrons in the ring produces an exceptionally
stable ("aromatic") compound, is known as the Hückel’s rule.
A
Source: http://en.wikipedia.org/wiki/Ring_wave_guide

Principal Authors: Pfalstad, Linas, Oleg Alexandrov, Waveguy, Wik

Ritz method

In physics, the Ritz method is a variational method named after Walter Ritz.
It can be applied in quantum mechanical problems to provide an upper-bound
DR
on the ground state energy.
As with other variational methods, a trial wave function is tested on the sys-
tem. This trial function contains one or more adjustable parameters, which are
varied to find a lowest energy configuration.
It can be shown that the ground state energy, E0 , satisfies an inequality:
E0 < Ψ∗ ĤΨ dτ
R

that is, the ground-state energy is less than this value. The trial wave-function
will always give an expectation value larger than the ground-energy (or at
least, equal to it).
If the trial wave function is known to be orthogonal to the ground state, then
it will provide a boundary for the energy of some excited state.

Ritz method
515

The Ritz ansatz function (trial function - i.e., the assumed form of the
eigenfunctions) is a linear combination of N known basis functions {Ψi },
parametrized by unknown coefficients:
Φ= N
P
i=1 ci Ψi .

FT
With a known hamiltonian, we can write its expected value as

PN PN PN PN ∗
ci Ψi |Ĥ| ci Ψi c c H
j=1 i j ij A
ε =
Pi=1N PNi=1 = Pi=1N PN ∗
≡ B .
cΨ|
i=1 i i

i=1 i i i=1
c c S
j=1 i j ij

The basis functions are usually not orthogonal, so that the overlap matrix S
is has nonzero diagonal elements. Either {ci } or c∗i (the conjugation of the


first) can be used to minimze the expectation value. For instance, by making
the partial derivatives of ε over c∗i zero, the following equality is obtained


for every k = 1,2,...,N:


PN
∂ε c (Hkj −εSkj )
j=1 j
∗ = B = 0,
∂c
k

which leads to a set of N secular equations:


APN 
j=1 cj Hkj − εSkj = 0 for k = 1, 2, ..., N .


In the above equations, energy ε and the coefficients cj are unknown. With
respect to c, this is a homogeneous set of linear equations, which has a solution
when the determinant of the coefficients to these unknowns is zero:

det Hkj − εSkj = 0,
DR
which in turn is true only for N values of ε. Furthermore, since the hamilto-
nian is a hermitian operator. matrix H is also hermitian and the values of εi
will be real. The lowest value among εi (i=1,2,..,N), ε0 , will be the best ap-
proximation to the ground state for the basis functions used. The remaining
N-1 energies are estimates of excited state energies. An approximation for the

wave function of state i can be obtained by finding the coefficients cj from
the corresponding secular equation.

Source: http://en.wikipedia.org/wiki/Ritz_method

Principal Authors: Karol Langner, Klemen Kocjancic, Nimur, Emersoni

Ritz method
516

Rutherford model

The Rutherford model of the atom was devised by Ernest Rutherford around

FT
1911 after he performed scattering experiments which showed that the →Plum
pudding model of the atom was incorrect. In the Rutherford model, an atom
is made up of a nucleus surrounded by a cloud of orbiting electrons. However,
the Rutherford model did not attribute any structure to the orbiting electrons.
The Rutherford model of the atom was soon superseded by the Bohr atom,
which used some of the early quantum mechanical results to give structure to
the orbiting electrons.
The Rutherford model was very important because it proposed the concept of
the nucleus. After the discovery of the Rutherford model, the study of the atom
branched into two separate fields, nuclear physics which studies the nucleus of
the atom, and atomic physics which studies the structure of the orbiting elec-
trons. In the Rutherford model the nucleus consisted of protons and embedded
electrons, this was however proven false later.
ASee also
• Atomic nucleus

External Links
• World Of Atoms - "Rutherford’s Model" 294
• Rutherford’s Model 295
DR
• Rutherford’s Model 296

Source: http://en.wikipedia.org/wiki/Rutherford_model

294 http://library.thinkquest.org/C0122360/full/1-1_e.html
295 http://www.rwc.uc.edu/koehler/biophys/7a.html
296 http://www2.kutl.kyushu-u.ac.jp/seminar/MicroWorld1_E/Part2_E/P25_E/Rutherford_model_E.htm

Rutherford model
517

Rutherford scattering

In physics, Rutherford scattering is a phenomenon that was explained by

FT
Ernest Rutherford in 1911, and led to the development of the orbital theory of
the atom. It is now exploited by the materials analytical technique Rutherford
backscattering. Rutherford scattering is also sometimes referred to as Coulomb
scattering because it relies on static electric (Coulomb) forces. A similar pro-
cess probed the insides of nuclei in the 1960s, called deep inelastic scattering.
The discovery was made by Hans Geiger and Ernest Marsden in 1909 when
they performed the gold foil experiment under the direction of Rutherford, in
which they fired a beam of alpha particles (helium nuclei) at layers of gold
leaf only a few atoms thick. At the time of the experiment, the atom was
thought to be analogous to a plum pudding (as proposed by J.J. Thomson),
with the negative charges (the plums) found throughout a positive sphere (the
pudding). If the plum-pudding model were correct, the positive “pudding”,
being more spread out than in the current model of a concentrated nucleus,
would not be able to exert such large coulombic forces, and the alpha particles
A
should only be deflected by small angles as they pass through.
However, the intriguing results showed that around 1 in 8000 alpha particles
were deflected by very large angles (over 90◦ ). From this, Rutherford con-
cluded that the majority of the mass was concentrated in a minute, positively
charged region (the nucleus) surrounded by electrons. When a (positive) alpha
particle approached sufficiently close to the nucleus, it was repelled strongly
enough to rebound at high angles. The small size of the nucleus explained
the small number of alpha particles that were repelled in this way. Rutherford
DR
showed, using the method below, that the size of the nucleus was about 10 -14
m.

Details of calculating nuclear size


For head on collisions between alpha particles and the nucleus, all the kinetic
1
energy ( 2 mv 2 ) of the alpha particle is turned into potential energy and the
particle is at rest. The distance from the centre of the alpha particle to the
centre of the nucleus (b ) at this point is a maximum value for the radius, if it
is evident from the experiment that the particles have not hit the nucleus.
Applying the inverse-square law between the charges on the electron and nu-
cleus, one can write:
1 2 1 q1 q2
2 mv = 4π0 · b
Rearranging:

Rutherford scattering
518

1 2q1 q2
b= 4π0 · mv 2
For an alpha particle:

• m (mass) = 6.7×10 -27 kg

FT
• q 1 = 2×(1.6×10 -19) C
• q 2 (for gold) = 79×(1.6×10 -19) C
• v (initial velocity) = 2×10 7 m/s

Substituting these in gives the value of about 2.7×10 -14 m. (The true radius is
about 7.3×10 -15 m.)

See also:
• Coulomb collision

References
• E. Rutherford, The Scattering of α and β Particles by Matter and the Struc-
A ture of the Atom 297, Philosophical Magazine. Series 6, vol. 21. May 1911
• H. Geiger and E. Marsden, On a Diffuse Reflection of the α-Particles 298,
Proceedings of the Royal Society, 1909 A vol. 82, p. 495-500

Source: http://en.wikipedia.org/wiki/Rutherford_scattering

Principal Authors: Sodium, Linas, Michael Hardy, Jll, HenkvD, Art Carlson, Awolf002, Andre Engels
DR
Rydberg formula

The Rydberg formula (Rydberg-Ritz formula) is used in atomic physics for


determining the full spectrum of light emission from hydrogen, later extended
to be useful with any element by use of the Rydberg-Ritz combination principle.

297 http://fisica.urbenalia.com/arts/structureatom.pdf
298 http://dbhs.wvusd.k12.ca.us/webdocs/Chem-History/GM-1909.html

Rydberg formula
519

1
λvac

Where
= RH Z 2 1
n21
− 1
n22

FT
The spectrum is the set of wavelengths of photons emitted when electrons jump
between discrete energy levels, "shells" around the atom of a certain chemical
element. This discovery was later to provide motivation for the creation of
quantum physics.
The formula was invented by the Swedish physicist Johannes Rydberg and
presented on November 5, 1888.
A
Rydberg formula for hydrogen


λvac is the wavelength of the light emitted in vacuum,


DR
RH is the Rydberg constant for hydrogen,

n1 and n2 are integers such that n1 < n2 ,

Z is the atomic number, which is 1 for hydrogen.

By setting n1 to 1 and letting n2 run from 2 to infinity, the spectral lines known
as the Lyman series converging to 91nm are obtained, in the same manner:
n1 n2 Name Converge toward
1 2 → ∞ Lyman series 91nm
2 3 → ∞ Balmer series 365nm
3 4 → ∞ Paschen series 821nm
4 5 → ∞ Brackett series 1459nm
5 6 → ∞ Pfund series 2280nm

Rydberg formula
520

6 7 → ∞ Humphreys series 3283nm

The Lyman series is in the ultraviolet while the Balmer series is in the visible
and the Paschen, Brackett, Pfund, and Humphreys series are in the infrared.

FT
Rydberg formula for any hydrogen-like element
The formula above can be extended for use with any hydrogen-like chemical
elements.
 
1 2 1 − 1
λvac = RZ n 2
n 2
1 2

where
λvac is the wavelength of the light emitted in vacuum;

R is the Rydberg constant for this element;

Z is the atomic number, i.e. the number of protons in the atomic nucleus
of this element;
A
n1 and n2 are integers such that n1 < n2 .

It’s important to notice that this formula can be applied only to hydrogen-like,
also called hydrogenic atoms chemical elements, i.e. atoms with only one
electron on external system of orbitals. Examples would include He +, Li 2+,
Be 3+ etc.
DR
History
By 1890, Rydberg had discovered a formula describing the relation between the
wavelengths in lines of alkali metals and found that the Balmer equation was
a special case. Although the Rydberg formula was later found to be imprecise
with heavier atoms, it is still considered accurate for all the hydrogen series and
for alkali metal atoms with a single valency electron orbiting well clear of the
inner electron core. By 1906, Lyman had begun to analyze the hydrogen series
of wavelengths in the ultraviolet spectrum named for him that were already
known to fit the Rydberg formula.
Rydberg simplified his calculations by using the ‘wavenumber’ (the number of
waves occupying a set unit of length) as his unit of measurement. He plotted
the wavenumbers of successive lines in each series against consecutive integers
which represented the order of the lines in that particular series. Finding that

Rydberg formula
521

the resulting curves were similarly shaped, he sought a single function which
could generate all of them when appropriate constants were inserted.

References

FT
Mike Sutton, “Getting the numbers right – the lonely struggle of Rydberg”
Chemistry World, Vol. 1, No. 7, July 2004.

See also
Rydberg-Ritz combination principle

Source: http://en.wikipedia.org/wiki/Rydberg_formula

Principal Authors: Nixdorf, Voyajer, Laurascudder, Chris Roy, Xerxes314, Tantalate, W.marsh, Wibbly-
wobbly

Scattering channel
A
In scattering theory, a scattering channel is a quantum state of the colliding
system before or after the collision (t → ±∞). The →Hilbert space spanned
by the states before collision (in states) is equal to the ones spanned by the
states after collision (out states) which are both →Fock spaces if there is a
mass gap. This is the reason why the S matrix which maps the in states onto
the out states must be unitary. The scattering channel are also called scattering
asymptotes. The Møller operators are mapping the scattering channels onto the
DR
corresponding states which are solution of the →Schrödinger equation taking
the interaction Hamiltonian into account. The Møller operators are isometric.
See also: LSZ formalism

Source: http://en.wikipedia.org/wiki/Scattering_channel

Principal Authors: Phys, OpenToppedBus

Scattering channel
522

Scattering theory

In mathematics and physics, scattering theory is a framework for studying

FT
and understanding the scattering of waves and particles. Prosaically, wave
scattering corresponds to the collision and scattering of a wave with some
material object, for instance sunlight scattered by rain drops to form a rain-
bow. Examples of particle scattering includes the motion of billiard balls or the
→Rutherford scattering of alpha particles by gold nucleii. More precisely, scat-
tering consists of the study of how solutions of partial differential equations,
propagating freely "in the distant past", come together and interact with one
another or with a boundary condition, and then propagate way "to the distant
future".
The direct scattering problem is the problem determining the distribution of
scattered radiation/particle flux basing on the characteristics of the scatterer.
The inverse scattering problem is the problem of determining the characteris-
tics of an object (its shape, internal constitution, etc.) from measurement data
A
of radiation or particles scattered from the object.
Since its early statement for radiolocation, the problem has found vast num-
ber of applications, such as echolocation, geophysical survey, nondestructive
testing, medical imaging and quantum field theory, to name just a few.

In theoretical physics
In mathematical physics, scattering theory is a framework for studying and
understanding the interaction or scattering of solutions to partial differential
DR
equations. In acoustics, the differential equation is the wave equation, and
scattering studies how its solutions, the sound waves, scatter from solid objects
or propagate through non-uniform media (such as sound waves, in sea water,
coming from a submarine). In the case of classical electrodynamics, the differ-
ential equation is again the wave equation, and the scattering of light or radio
waves is studied. In quantum mechanics and particle physics, the equations are
those of QED, QCD and the Standard Model, the solutions of which correspond
to fundamental particles. In quantum chemistry, the solutions correspond to
atoms and molecules, governed by the Schroedinger equation.

Elastic and inelastic scattering


The example of scattering in quantum chemistry is particularly instructive, as
the theory is reasonably complex while still having a good foundation on which

Scattering theory
523

to build an intuitive understanding. When two atoms are scattered off one an-
other, one can understand them as being the bound state solutions of some dif-
ferential equation. Thus, for example, the hydrogen atom corresponds to a so-
lution to the Schroedinger equation with an inverse-square law central poten-

FT
tial. The scattering of two hydrogen atoms will disturb the state of each atom,
resulting in one or both becoming excited, or even ionized. Thus, collisions can
be either elastic (the internal quantum states of the particles are not changed)
or inelastic (the internal quantum states of the particles are changed). From
the experimental viewpoint the observable quantity is the cross section. From
the theoretical viewpoint the key quantity is the S matrix.

Topics in physics
According to the optics classification of the Optical Society of America this field
consists of the following topics:
• Aerosol and cloud effects • Long-wave scattering • Rayleigh Scattering
• Atmospheric scattering • Mie theory • Scattering from rough surfaces
• Backscattering • Multiple scattering • Stimulated scattering
• Diffusion • Scattering measurements • Stimulated Brillouin scattering
• Extinction • Brillouin scattering • Stimulated Raman scattering



A Index measurements
Inverse scattering
Linewidth



Molecular scattering
Particle scattering
Raman scattering


Scintillation
Turbid media

The mathematical framework


In mathematics, scattering theory deals with a more abstract formulation of the
same set of concepts. For example, if a differential equation is known to have
some simple, localized solutions, and the solutions are a function of a single
DR
parameter, that parameter can take the conceptual role of time. One then asks
what might happen if two such solutions are set up far away from each other,
in the "distant past", and are made to move towards each other, interact (under
the constraint of the differential equation) and then move apart in the "future".
The scattering matrix then pairs solutions in the "distant past" to those in the
"distant future".
Solutions to differential equations are often posed on manifolds. Frequently,
the means to the solution requires the study of the spectrum of an operator
on the manifold. As a result, the solutions often have a spectrum that can be
identified with a →Hilbert space, and scattering is described by a certain map,
the S matrix, on Hilbert spaces. Spaces with a discrete spectrum correspond to
bound states in quantum mechanics, while a continuous spectrum is associated
with scattering states. The study of inelastic scattering then asks how discrete
and continuous spectra are mixed together.

Scattering theory
524

An important, notable development is the inverse scattering transform, central


to the solution of many exactly solvable models.

References

FT
• Lectures of the European school on theoretical methods for electron and
positron induced chemistry, Prague, Feb. 2005 299

Source: http://en.wikipedia.org/wiki/Scattering_theory

Principal Authors: Vb, Pflatau, Charles Matthews, David R. Ingham, Fuhghettaboutit

Schrödinger equation

In physics, the Schrödinger equation, proposed by the Austrian physicist Er-


win Schrödinger in 1925, describes the space- and time-dependence of quan-
A
tum mechanical systems. It is of central importance to the theory of quantum
mechanics, playing a role analogous to Newton’s second law in classical me-
chanics.
In the mathematical formulation of quantum mechanics, each system is associ-
ated with a complex →Hilbert space such that each instantaneous state of the
system is described by a unit vector in that space. This state vector encodes the
probabilities for the outcomes of all possible measurements applied to the sys-
tem. As the state of a system generally changes over time, the state vector is a
DR
function of time. The Schrödinger equation provides a quantitative description
of the rate of change of the state vector.
Using Dirac’s bra-ket notation, the definition of energy results in the time
derivative operator: at time t by |ψ (t)i. The Schrödinger equation is

H(t) |ψ (t)i = i~ ∂t |ψ (t)i

where i is the imaginary unit, t is time, ∂/∂t is the partial derivative with
respect to t, ~ is the reduced Planck’s constant (Planck’s constant divided by
2π), ψ(t) is the wave function, and H (t) is the Hamiltonian (a self-adjoint
operator acting on the state space).
The Hamiltonian describes the total energy of the system. As with the force
occurring in Newton’s second law, its exact form is not provided by the

299 http://www.mif.pg.gda.pl/homepages/slawek/epic/sem.htm

Schrödinger equation
525

Schrödinger equation, and must be independently determined based on the


physical properties of the system.

Time-independent Schrödinger equation

FT
For many real-world problems the energy distribution does not change with
time, and it is useful to determine how the stationary states vary with position
x (independent of the time t). The Schrödinger equation is often introduced
without bra-ket notation in the following ways:
300
One dimensional time-independent :
2
~ d ψ(x)
2
− 2m (dx)2
+ U (x)ψ(x) = Eψ(x)

301
3-dimensional time-independent :
~2
− 2m ∇2 ψ(r) + U (r)ψ(r) = Eψ(r)

For every time-independent Hamiltonian, H, there exists a set of quantum


states, |ψn i, known as energy eigenstates, and corresponding real numbers
AEn satisfying the eigenvalue equation
H |ψn (x)i = En |ψn (x)i .

Such a state possesses a definite total energy, whose value En is the eigen-
value of the state vector with the Hamiltonian. This eigenvalue equation is
referred to as the time-independent Schrödinger equation. Self-adjoint op-
erators such as the Hamiltonian have the property that their eigenvalues are
always real numbers, as we would expect since the energy is a physically ob-
DR
servable quantity.
On inserting the time-independent Schrödinger equation into the full
Schrödinger equation, we get

i~ ∂t |ψn (t)i = En |ψn (t)i .

It is easy to solve this equation. One finds that the state vectors of the energy
eigenstates change by only a complex phase:
|ψ (t)i = e−iEt/~ |ψ (0)i .

300 http://vergil.chemistry.gatech.edu/notes/quantrev/node8.html
301 http://vergil.chemistry.gatech.edu/notes/quantrev/node9.html

Schrödinger equation
526

Energy eigenstates are convenient to work with because their time-dependence


is so simple; that is why the time-independent Schrödinger equation is so use-
ful. We can always choose a set of instantaneous energy eigenstates whose
state vectors {|ni} form a basis for the state space. Then any state vector

FT
|ψ (t)i can be written as a linear superposition of energy eigenstates:
P P 2
|ψ (t)i = n cn (t) |ni , H |ni = En |ni , n |cn (t)| = 1.

(The last equation enforces the requirement that |ψ (t)i, like all state vectors,
must be a unit vector.) Applying the Schrödinger equation to each side of the
first equation, and using the fact that the energy basis vectors are by definition
linearly independent, we readily obtain
i~ ∂c n
∂t = En cn (t) .

Therefore, if we know the decomposition of |ψ (t)i into the energy basis at time
t = 0, its value at any subsequent time is given simply by
|ψ (t)i = n e−iEn t/~ cn (0) |ni .
P
A
Schrödinger wave equation
The state space of certain quantum systems can be spanned with a position
basis. In this situation, the Schrödinger equation may be conveniently refor-
mulated as a partial differential equation for a wavefunction, a complex scalar
field that depends on position as well as time. This form of the Schrödinger
equation is referred to as the Schrödinger wave equation.
Elements of the position basis are called position eigenstates. We will consider
DR
only a single-particle system, for which each position eigenstate may be denot-
ed by |ri, where the label r is a real vector. This is to be interpreted as a state
in which the particle is localized at position r. In this case, the state space is
the space of all square-integrable complex functions.

The wavefunction
We define the wavefunction as the projection of the state vector |ψ (t)i onto
the position basis:
ψ (r, t) ≡ hr|ψ (t)i .

Since the position eigenstates form a basis for the state space, the integral over
all projection operators is the identity operator:
|ri hr| d3 r = I.
R

Schrödinger equation
527

This statement is called the resolution of the identity. With this, and the fact
that kets have unit norm, we can show that where ψ (r, t)∗ denotes the complex
conjugate of ψ (r, t). This important result tells us that the absolute square of
the wavefunction, integrated over all space, must be equal to 1:

FT
|ψ (r, t)|2 d3 r = 1.
R

We can thus interpret the absolute square of the wavefunction as the prob-
ability density for the particle to be found at each point in space. In other
words, |ψ (r, t)|2 d3 r is the probability, at time t, of finding the particle in the
infinitesimal region of volume d3 r surrounding the position r.
We have previously shown that energy eigenstates vary only by a complex
phase as time progresses. Therefore, the absolute square of their wavefunc-
tions do not change with time. Energy eigenstates thus correspond to static
probability distributions.

Operators in the position basis


Any operator A acting on the wavefunction is defined in the position basis by
AAψ (r, t) ≡ hr|A|ψ (t)i .

The operators A on the two sides of the equation are different things: the one
on the right acts on kets, whereas the one of the left acts on scalar fields. It is
common to use the same symbols to denote operators acting on kets and their
projections onto a basis. Usually, the kind of operator to which one is referring
is apparent from the context, but this is a possible source of confusion.
Using the position-basis notation, the Schrödinger equation can be written as
DR

Hψ (r, t) = i~ ∂t ψ (r, t) .

This form of the Schrödinger equation is the Schrödinger wave equation. It


may appear that this is an ordinary differential equation, but in fact the Hamil-
tonian operator typically includes partial derivatives with respect to the posi-
tion variable r. This usually leaves us with a difficult linear partial differential
equation to solve.

Non-relativistic Schrödinger wave equation


In non-relativistic quantum mechanics, the Hamiltonian of a particle can be
expressed as the sum of two operators, one corresponding to kinetic energy
and the other to potential energy. The Hamiltonian of a particle with no electric
charge and no spin in this case is:

Schrödinger equation
528
h i
~2
Hψ (r, t) = (T + V ) ψ (r, t) = − 2m ∇2 + V (r) ψ (r, t) = i~ ∂ψ
∂t (r, t)

where

FT
p2
T = 2m is the kinetic energy operator,

m is the mass of the particle,

p = ~i ∇ is the momentum operator,

V = V (r) is the potential energy operator,

V is a real scalar function of the position operator r,

∇ is the gradient operator, and


A ∇2 is the Laplace operator.

This is a commonly encountered form of the Schrödinger wave equation,


though not the most general one. The corresponding time-independent equa-
tion is
h 2 i
~
− 2m ∇2 + V (r) ψ (r) = Eψ (r) .
DR
The relativistic generalisations of this wave equation are the Dirac equation,
→Klein-Gordon equation, Proca equation, Maxwell equations etc, depending
on spin and mass of the particle. See relativistic wave equations for details.

Probability currents
In order to describe how probability density changes with time, it is acceptable
to define probability current or probability flux. The probability flux represents
a flowing of probability across space.
For example, consider a Gaussian probability curve centered around x0 , imag-
ine that x0 moving in a speed v toward the right. Then one may say that the
probability is flowing toward right, i.e., there is a probability flux directed to
the right.
The probability flux j is defined as:
~ 1 ~
j= m · 2i (ψ ∗ ∇ψ − ψ∇ψ ∗ ) = m Im (ψ ∗ ∇ψ)

Schrödinger equation
529

and measured in units of (probability)/(area × time) = r -2t -1.


The probability flux satisfies a quantum continuity equation, i.e.:

(x, t) + ∇ · j = 0

FT
∂t P

where P (x, t) is the probability density and measured in units of (probabili-


ty)/(volume) = r -3. This equation is the mathematical equivalent of probabili-
ty conservation law.
It is easy to show that for a plane wave,
|ψi = Aeikx e−iωt

the probability flux is given by


j (x, t) = |A|2 k~
m.

Solutions of the Schrödinger equation


A
Analytical solutions of the time-independent Schrödinger equation can be ob-
tained for a variety of relatively simple conditions. These solutions provide
insight into the nature of quantum phenomena and sometimes provide a rea-
sonable approximation of the behavior of more complex systems (e.g., in sta-
tistical mechanics, molecular vibrations are often approximated as harmonic
oscillators). Several of the more common analytical solutions include:

• The free particle


DR
• The particle in a box
• The finite potential well
• The particle in a ring
• The particle in a spherically symmetric potential
• The quantum harmonic oscillator
• The hydrogen atom or hydrogen-like atom
• The ring wave guide
• The particle in a one-dimensional lattice (periodic potential)

For many systems, however, there is no analytic solution to the Schrödinger


equation. In these cases, one must resort to approximate solutions. Some of
the common techniques are:

Schrödinger equation
530

• Perturbation theory
• The variational principle underpins many approximate methods (like the
popular Hartree-Fock method which is the basis of the post Hartree-Fock
methods)

FT
• Quantum Monte Carlo methods
• Density functional theory
• The WKB approximation
• discrete delta-potential method

See also
• →Schrödinger picture
• Basic quantum mechanics
• Quantum number
• Principal quantum number
• Azimuthal quantum number
• Magnetic quantum number
• Spin quantum number
A• Dirac equation

References
• E. Schrödinger, Ann. Phys. (Leipzig) 489 (1926) p.79
• E. Schrödinger, Phys. Rev. 28 (1926) p. 1049

Modern reviews
DR
• David J. Griffiths (2004). Introduction to Quantum Mechanics (2nd ed.).
Prentice Hall. ISBN 013805326X.

External links
• Linear Schrödinger Equation 302 at EqWorld: The World of Mathematical
Equations.

• Nonlinear Schrödinger Equation 303 at EqWorld: The World of Mathematical


Equations.

302 http://eqworld.ipmnet.ru/en/solutions/lpde/lpde108.pdf
303 http://eqworld.ipmnet.ru/en/solutions/npde/npde1403.pdf

Schrödinger equation
531

• The Schrödinger Equation in One Dimension 304 as well as the directory of


the book 305.
306
All about 3D schrodinger Equation

FT
Source: http://en.wikipedia.org/wiki/Schr%C3%B6dinger_equation

Principal Authors: CYD, MathKnight, Passw0rd, Michael Hardy, Fresheneesz, Alex valavanis, Nom-
monomanac, Camembert, Voyajer, Linas

Schrödinger picture

In quantum mechanics, a state function is a linear combination (a superposi-


tion) of eigenstates. In the Schrödinger picture, all of these eigenstates are
constantly rotating through time.
This rotation is not in any ordinary spatial sense. Each eigenvector has an
amplitude which is a complex number. This amplitude is a coefficient which
A
multiplies one of the basis vectors. The complex coefficient has a magnitude
and a direction. Therefore a state function is a linear combination of basis
vectors, each one multiplied by a complex coefficient which has a magnitude
and a direction in the complex plane. These coefficients can be thought of as
phasors.
In the Schrödinger picture, these phasor coefficients are constantly rotating in a
circle through time. The rotation operator which causes their rotation is called
the propagator. The time evolution of a Schrödinger wave function can be
DR
effected mathematically by multiplying the wave function with the propagator.
The propagator effects a simultaneous rotation of all the phasor coefficients of
all the (infinite) basis vectors which form the state function.
Let |ψe (0)i represent an energy eigenstate at time 0. Then the rotation of the
phasor coefficient of this eigenstate through time can be described by:
|ψe (t)i = e−iHt/~ |ψe (0)i.

where e−iHt/~ is a rotation in the complex plane, and H is the scalar Hamilto-
nian. Taking the time derivative of |ψe (t)i yields

304 http://www.colorado.edu/UCB/AcademicAffairs/ArtsSciences/physics/TZD/PageProofs1/TAYL07-203-

247.I.pdf
305 http://www.colorado.edu/UCB/AcademicAffairs/ArtsSciences/physics/TZD/PageProofs1/
306 http://hyperphysics.phy-astr.gsu.edu/hbase/hframe.html

Schrödinger picture
532

d −i
dt |ψe (t)i = ~ H|ψe (t)i

which is the Schrödinger equation for time evolution.


Thus, in the Schrödinger formulation of quantum mechanics, all unperturbed

FT
state functions are time-harmonic. State functions in the Schrödinger picture
are never entirely static, they are always undulating. This is why state func-
tions in the Schrödinger formulation are called wavefunctions. It reveals the
undulatory nature of matter: the wave-particle duality. (Actually, wavefunc-
tions also are also undulatory in space, independently of time.)
The alternative to the Schrödinger picture is to switch to a rotating reference
frame, which is itself being rotated by the propagator. Since the undulatory
rotation is now being assumed by the reference frame itself, an undisturbed
state function appears to be truly static. This is the →Heisenberg picture.
See also interaction picture.

Further reading

APrinciples of Quantum Mechanics by R. Shankar, Plenum Press.

Source: http://en.wikipedia.org/wiki/Schr%C3%B6dinger_picture

Principal Authors: AugPi, Pen of bushido, Cyp, NawlinWiki, Lethe

Schrödinger’s cat
DR
Schrödinger’s cat is a seemingly paradoxical thought experiment devised by
Erwin Schrödinger that attempts to illustrate the incompleteness of an early in-
terpretation of quantum mechanics when going from subatomic to macroscopic
systems. The experiment proposes:
A cat is placed in a sealed box. Attached to the box is an apparatus con-
taining a radioactive atomic nucleus and a canister of poison gas. This
apparatus is separated from the cat in such a way that the cat can in no
way interfere with it. The experiment is set up so that there is exactly a
50% chance of the nucleus decaying in one hour. If the nucleus decays, it
will emit a particle that triggers the apparatus, which opens the canister
and kills the cat. If the nucleus does not decay, then the cat remains alive.
According to quantum mechanics, the unobserved nucleus is described as a
superposition (meaning it exists partly as each simultaneously) of "decayed

Schrödinger’s cat
533

A FT
Figure 43 Schrödinger’s Cat: If the nucleus in the bottom left decays, the geiger counter on its
right will sense it and trigger the release of the gas. In one hour, there is a 50% chance that the
nucleus will decay, and therefore that the gas will be released and kill the cat.

nucleus" and "undecayed nucleus". However, when the box is opened the
experimenter sees only a "decayed nucleus/dead cat" or an "undecayed
nucleus/living cat."
DR
The question is: when does the system stop existing as a mixture of states and
become one or the other? (See basis function.) The purpose of the experiment
is to illustrate a paradox; as Schrödinger wrote, "The (wavefunction) for the
entire system (has) the living and the dead cat (pardon the expression) mixed
307
or smeared out in equal parts". Because we cannot get along without making
classical approximations, quantum mechanics is incomplete without some rules
to relate the classical and quantum descriptions. One way of looking at this
connection is to say that the wavefunction collapses and the cat becomes dead
or remains alive instead of a mixture of both.
The point of view that this thought experiment most clearly refutes is that the
laws of physics are different for experiments than for other interactions. In the
case of the cat dying, a necropsy would show a time of death that would be
before the opening of the box. The geiger counter, in moving to one outcome

Schrödinger’s cat
534

or the other, is, in effect, the "observer", the same as a human observing the
outcome with senses.
The original article appeared in the German magazine Naturwissenschaften
("Natural Sciences") in 1935: E. Schrödinger: "Die gegenwärtige Situation in

FT
der Quantenmechanik" ("The present situation in quantum mechanics"), Natur-
wissenschaften, 48, 807, 49, 823, 50, 844 (November 1935). It was intended
as a discussion of the EPR article published by Einstein, Podolsky and Rosen
in the same year. Apart from introducing the cat, Schrödinger also coined the
term "entanglement" (German: Verschränkung) in his article.
Albert Einstein was impressed; in a letter to Schrödinger 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 reality
- reality as something independent of what is experimentally established.
Their interpretation is, however, refuted most elegantly by your system of
radioactive atom + amplifier + charge of gun powder + cat in a box, in
which the psi-function of the system contains both the cat alive and blown
A
to bits. Nobody really doubts that the presence or absence of the cat is
something independent of the act of observation.

Nowadays, the mainstream interpretation is that the triggering of the device is


the actual observation that collapses the wave function.

Copenhagen interpretation
In the Copenhagen interpretation, a system stops being a superposition of
DR
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. Some interpret the exper-
iment 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 observa-
tion performed does the wave function collapse into one of the two states.
More intuitively, some feel that the "observation" is taken when a particle from
the nucleus hits the detector. Recent developments in quantum physics show
that measurements of quantum phenomena taken by non-conscious "observers"
(such as a wiretap) most definitely alter the quantum state of the phenomena
from the point of view of conscious observers reading the wiretap, lending
support to this idea.

Schrödinger’s cat
535

A precise rule is that probability enters at the point where the classical ap-
proximation is first used to describe the system - almost by tautology, as the
classical approximation is just a simplification of the quantum mathematics,
and so must introduce imprecision in the measurement, which can be viewed

FT
as probability. Note, however, that this only applies to descriptions of the sys-
tem, not the system itself.
Under Copenhagen, the amount of uncertainty for a complex quantum system
is predicted by quantum decoherence. Particles which exchange photons (and
possibly other atomic or subatomic particles) become entangled with each oth-
er from the point of view of an observer, meaning that these particles can only
be described accurately with reference to each other, which decreases the total
uncertainty of those particles from the point of view of our observer. By the
time one has reached "macroscopic" levels - such as a cat, which is made up of
a number of atomic particles almost too large to express with words - so many
particles have become entangled with each other so as to decrease the uncer-
tainty to almost zero. (Quantum effects in huge collections of particles are
only seen in very rare, and often man-made, situations, such as a Bose-Einstein
condensate). Thus, at least from the point of view of the observer, any improb-
A
ability regarding the cat as a system of quantum particles has disappeared due
to the massive amount of entanglement between all of the particles that make
it up, meaning that the cat does not truly exist as both alive and dead at the
same time, at least from the point of view of any observer viewing the cat.
It is interesting to note that even before observation was noted to be funda-
mentally distinct from consciousness through experimentation, the experiment
always contained at least two "observers" - the physicist and the cat. Even had
the physicist been unaware of the cat’s state in the hypothetical experiment,
DR
one would have had to posit that the cat, at least, would have been quite sure
of its status (at least, as long as the gas had not yet ended its ability to "ob-
serve"). However, since "observation" has been shown by experiment to have
nothing to do with consciousness - or at the very least, any traditional defini-
tion of consciousness - most conjecture along these lines probably falls under
the "interesting but physically irrelevant" category.
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 measure-
ment are themselves treated classically. This is surely wrong: Physicists and
their apparatus must be governed by the same quantum mechanical rules

Schrödinger’s cat
536

that govern everything else in the universe. But these rules are expressed
in terms of a wavefunction (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?

FT
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 mechan-
ics. The Copenhagen rules clearly work, so they have to be accepted. But
this leaves the task of explaining them by applying the deterministic equa-
tion for the evolution of the wavefunction, the Schrödinger equation, to
observers and their apparatus.

Everett many-worlds interpretation & consistent


histories
In the many-worlds interpretation of quantum mechanics, which does not sin-
A
gle out observation as a special process, both states persist, but decoherent
from each other. When an observer opens the box, he becomes entangled with
the cat, so observer-states corresponding to the cat being alive and dead are
formed, and each can have no interaction with the other. The same mecha-
nism 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.
In other words, when the box is opened, the universe (or at least the part of the
DR
universe containing the observer and cat) is split into two separate universes,
one containing an observer looking at a box with a dead cat, one containing an
observer looking at a box with a live cat.

Practical applications
The experiment is a purely theoretical one, and the machine proposed does not
exist.
This has some practical use in quantum computing and quantum cryptography.
It is possible to send light that is in a superposition of states down a fiber optic
cable. Placing a wiretap in the middle of the cable which intercepts and re-
transmits the transmission will collapse the wavefunction (in the Copenhagen
interpretation, "perform an observation") and cause the light to fall into one
state or another. By performing statistical tests on the light received at the
other end of the cable, one can tell whether it remains in the superposition of

Schrödinger’s cat
537

states or has already been observed and retransmitted. In principle, this allows
the development of communication systems that cannot be tapped without
the tap being noticed at the other end. This experiment (which can be per-
formed, although a workable quantum cryptographic communications system

FT
which can transmit large quantities of data has not yet been constructed) also
illustrates that "observation" in the Copenhagen interpretation has nothing to
do with consciousness, in that a perfectly unconscious wiretap will cause the
statistics at the end of the wire to be different.
In quantum computing, the phrase "cat state" often refers to the special entan-
glement of qubits where the qubits are in an equal superposition of all being 0
and all being 1, i.e. |00...0i + |11...1i.
A variant of the Schrödinger’s Cat experiment known as the quantum suicide
machine has been proposed by cosmologist Max Tegmark. It examines the
Schrödinger’s Cat experiment from the point of view of the cat, and argues
that this may be able to distinguish between the Copenhagen interpretation
and many worlds. Another variant on the experiment is Wigner’s friend.
Physicist Stephen Hawking once exclaimed, "When I hear of Schrödinger’s cat, I
A
reach for my gun," paraphrasing German playwright and Nazi "Poet Laureate",
Hanns Johst’s famous phrase "Wenn ich ’Kultur’ höre, entsichere ich meinen
Browning! " ("When I hear the word ’culture’, I release the safety on my Brown-
ing!")
In fact, Hawking and many other physicists are of the opinion that the "Copen-
hagen School" interpretation of quantum mechanics unduly stresses the role of
the observer. A final consensus on this point among physicists seems still to be
out of reach.
DR
Related humor

Figure 44 Another joke about Schrödinger’s cat and


the <BLINK> tag in HTML.

• Curiosity may have killed the cat, Schrödinger only killed half of it.

• Many pet cats have been named "Schrödinger" as an allusion to


Schrödinger’s thought experiment.

Schrödinger’s cat
538

• In 1982 Cecil Adams, in his column The Straight Dope, wrote a concise and
humourous description of the thought experiment, and Einstein’s refutation
of same, in the form of an epic poem. "The story of Schroedinger’s cat (an
epic poem)" 308

FT
• "Schrödinger’s Cat" is a novel by Robert Anton Wilson in which the Many-
Worlds Interpretation is explored. Wilson refers to this model as the
"Everett-Wheeler-Graham Model (after its postulators)". In his DVD biog-
raphy entitled "Maybe Logic", Wilson suggests that our general misconcep-
tions and confusion about quantum mechanics may stem from the use of
logical constructs that are inadequate for describing the universe. See Gen-
eral Semantics.

• The term "Schrödinger’s Terrorist" has been used to semi-humorously la-


bel terrorists whose status as living or dead is unknown and/or subject to
contradictory rumors such as Osama bin Laden.

• A "Schrödinger’s Date" is a meeting between two people that may or may


not be a date, and whose status cannot be determined. Only the odds of it
A being one or the other can be estimated.

• Novelist Douglas Adams posits an amusing anecdote in his novel Dirk Gen-
tly’s Holistic Detective Agency, wherein Schrödinger’s cat is neither alive
nor dead when the box is opened, and is, in fact, not there. It is later re-
vealed that the cat simply became bored with the experiment and wandered
off. They should have used a rat. A rat indeed obediently stays put until
the cat shows up.
DR
• The manga version of Hellsing features a catboy character named
Schrödinger, who has the ability to be "everywhere and nowhere" at once.

• There’s also the mock headline "Schrödinger’s cat found half-alive: quan-
tum theory a mistake!"

• From the syndicated comic strip, "Brewster Rockit": "Schrödinger’s cat-litter


box. If you don’t observe it, you won’t have to change it."

• In Stargate SG-1, Carter gives a friend from a race of highly advanced physi-
cists a tabby cat called Schrödinger as a gesture of good will.

• In in the pilot episode of Sliders, Quinn has a pet cat named Schrödinger.

308 http://www.straightdope.com/classics/a1_122.html

Schrödinger’s cat
539

• In the NetHack computer game, one occasionally finds a quantum mechanic


with a box with a cat named "Schrödinger’s Cat" inside. It is dead in half of
the instances.

• Schrödinger’s Fridge 309, from the comic Bob the Angry Flower.

FT
• In The Last Hero by Terry Pratchett, Albert tries to explain the Schrödinger’s
Cat experiment to Death, who doesn’t quite grasp the concept of not know-
ing if something is alive or dead. Death finally gives up the experiment,
remarking, "I don’t hold with cruelty to cats." Also, in Lords and Ladies,
Pratchett describes the results of the experiment in three possible states:
"Alive", "Dead", and "Bloody Furious."

• Tears for Fears Track on their Saturnine Martial & Lunatic Album is called
Schrödinger’s cat.

• The online retailer ThinkGeek sells a t-shirt that has the words
"Schrödinger’s cat is dead" on the front and "Schrödinger’s Cat is not dead"
on the back.
A• Cartoon: You see a picture of a refrigerator with a post-it-note saying "I’m
out of town for a few days. Please be so nice and feed my cat. Thanks.
Schrödinger".

See also
• Schroedinbug
• →Schrödinger’s cat in fiction
DR
• →Double-slit experiment
• Quantum suicide

External links
• Erwin Schrödinger, The Present Situation in Quantum Mechanics
(Translation) 310
• A Lazy Layman’s Guide to Quantum Physics 311
• Quantum Mechanics and Schrodinger’s Cat 312
• The many worlds of quantum mechanics 313

309 http://www.angryflower.com/schrod.gif
310 http://www.tu-harburg.de/rzt/rzt/it/QM/cat.html
311 http://www.higgo.com/quantum/laymans.htm
312 http://www.harrymaugans.com/2006/05/03/in-search-of-schrodingers-cat/
313 http://www.sankey.ws/qm.html

Schrödinger’s cat
540

Source: http://en.wikipedia.org/wiki/Schr%C3%B6dinger%27s_cat

Principal Authors: David R. Ingham, DV8 2XL, Commonbrick, Sam Spade, David R. Ingharn, Florian-
Marquardt, Mako098765, Jcfolsom, CYD, Faseidrnan

FT
Schrödinger’s cat in fiction

→Schrödinger’s cat is a seemingly paradoxical thought experiment devised by


Erwin Schrödinger that attempts to illustrate the incompleteness of the theory
of quantum mechanics when going from subatomic to macroscopic systems. In
1935, Schrödinger published an essay describing the conceptual problems in
quantum mechanics. A brief paragraph in this essay described the cat paradox:
One can even set up quite ridiculous cases. A cat is penned up in a steel
chamber, along with the following diabolical 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 one hour
Aone 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 which 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 first atomic decay would have
poisoned it. The Psi function for the entire system would express this by
having in it the living and the dead cat (pardon the expression) mixed or
smeared out in equal parts.
DR
Introduction
It was not long before science-fiction writers picked up this evocative concept,
often using it in a humorous vein. Several have taken the thought experiment
a step further, pointing out or extra complications which might arise should the
experiment actually be performed. For example, in his novel American Gods,
Neil Gaiman has a character observe, "if they don’t ever open the box to feed it
it’ll eventually just be two different kinds of dead." Likewise, Terry Pratchett’s
Lords and Ladies adds the issue of a third possible state, in the case of Greebo,
"Bloody Furious". Douglas Adams describes an attempt to enact the experiment
in Dirk Gently’s Holistic Detective Agency. By using clairvoyance to see inside
the box, it was found that the cat was neither alive nor dead, but missing, and
Dirk’s services were employed in order to recover it.

Schrödinger’s cat in fiction


541

On a somewhat more serious level, Ian Stewart’s novel Flatterland (a sequel


to Flatland ) attempts to explain many concepts in modern mathematics and
physics through the device of having a young female Flatlander explore other
parts of the "Mathiverse". Schrödinger’s Cat is just one of the many strange

FT
Mathiverse denizens she and her guide meet; the cat is still uncertain whether
it is alive or dead, long after it left the box. Her guide, the Space Hopper, reas-
sures the Cat with a modern view of quantum decoherence. Ursula K. Le Guin
wrote a story entitled "Schrödinger’s Cat" in 1974 (reprinted in The Compass
Rose, published in 1982), which also deals with decoherence. Greg Egan’s
novel Quarantine, billed as "a story of quantum catastrophe", features an al-
ternate solution to the paradox: in Egan’s version of quantum mechanics, the
wave function does not collapse naturally. Only certain living things — human
beings among them — collapse the wave function of things they observe. Hu-
mans are therefore highly dangerous to other lifeforms which require the full
diversity of uncollapsed wavefunctions in order to survive.
As Egan notes, Schrödinger’s hypothetical cat is one of the most familiar illus-
trations of quantum-mechanical oddities. In Quarantine, a physicist asks the
narrator, an ex-cop and private investigator, if he has ever heard of "the quan-
A
tum measurement problem". The narrator is naturally confused, but when
asked if he’s heard of Schrödinger’s Cat, he replies, "Of course."
Fiction writers have confined other animals besides cats in such contraptions.
Dan Simmons’ novel Endymion begins with hero Raul Endymion trapped in
a Schrödinger-style box. In the fortieth-anniversary Doctor Who audio dra-
ma "Zagreus" (2003), the Doctor is locked in a lead-lined box also containing
cyanide in an effort to explain his situation of being neither dead nor alive.
Afterwards, the Doctor does mention that he had met Schrödinger’s Cat.
DR
In addition, the name "Schrödinger" has itself become an inside joke, often
employed to elicit a chuckle from those familiar with physics. In an episode
of the 1940s radio drama "Maddox’s Bedtime Stories for Kid Geniuses" entitled
"Little Eve and Gogo", the main character is a teenage female scientist named
Kat Schrödinger. Cats named Schrödinger appear in the television series Sliders
and Stargate SG-1, for example, as well as in Carol Hill’s feminist science fiction
classic "The Eleven Million Mile High Dancer". The animated series Futurama,
several of whose production team had advanced degrees in science and math,
includes many jokes of this sort; in the episode "A Clone of My Own", a brief
shot reveals a nightclub called "Schrödinger’s Kit-Kat Club" (also an allusion
to Cabaret). Robert A. Heinlein’s novel The Cat Who Walks Through Walls
features a cat named Pixel, affectionately termed "Schrödinger’s Cat" due to
his ability to be wherever his favorite person is. Pixel’s ability to walk through
walls is due to the fact that he does not know that it is impossible.

Schrödinger’s cat in fiction


542

Other assorted examples


Books and stories

FT
• The Schrödinger’s Cat trilogy is the name commonly given to a trilogy of
science fiction/conspiracy theory novels written by Robert Anton Wilson. It
consists of The Universe Next Door, The Trick Top Hat and The Homing
Pigeons.
• In The Last Hero, another of Terry Pratchett’s Discworld novels, Death’s
surly assistant Albert tries to explain to Death the concept of "Quantum"
with an object lesson similar to "Schrödinger’s Cat." Death waves it off,
saying, I DON’T HOLD WITH CRUELTY TO CATS (since, of course, Death
speaks entirely in capital letters).
• In The Coming of the Quantum Cats, by Frederik Pohl, a man meets his
many corresponding selves as he travels into the multiverse.
• In the manga Hellsing, there is a catboy named Schrödinger. He looks
human, save for the furry ears perched on his head. This being claims to
be everywhere and nowhere at once, and as such, has the ability to teleport
A

and to survive contact-range gunshots through the skull.
Steve Martin’s 1998 book Pure Drivel includes a piece entitled
"Schrödinger’s Cat", which presents a summary of the theory, followed
by several fictitious, nonsensical theories, including "Wittgenstein’s Ba-
nana", "Apollo’s Non-Apple Non-Strudel", and "Chef Boyardee’s Bungee
Cord" (which begins, "A bungee cord is hooked at one end to a neutrino,
while the other end is hooked to a vibraphone...")
• Saturday by Ian McEwan refers to Schrödinger’s Cat.
DR
• Schrödinger’s Cat is discussed in "The Summer of Love" by Lisa Mason.
• A two-headed cat, one head being conscious the other sleeping or presum-
ably dead, appears in one of Christopher Stasheff’s series of "Her Majesties
Wizard" novels, along with other thought experiment beings, the most no-
table being Maxwell’s Demon.
• George Alec Effinger’s novelette, "Schrödinger’s Kitten" (1988), received
both the Hugo and the Nebula Award. Schrodinger’s Kitten was published
in book form in 1992.
• In " Hopdinger’s Cat 314", a short story by the Roaming Janitors, Colin demon-
strates that Schrödinger’s cat itself is the problem with the experiment,
since it, of course, knows whether it is alive or dead. (Contrast the cat
in Stewart’s Flatterland, which lacks even this information.)

314 http://www.fictionpress.com/read.php?storyid=1822969

Schrödinger’s cat in fiction


543

• Schrödinger’s cat also appears in Nano by John Robert Marlow, which in-
cludes the theoretical experiment as well as a cat that is quite real.
• The manga Ah! My Goddess features a "Schrödinger’s Whale". This whale
has the ability to travel through space, though it cannot stay at the same

FT
space for more than three days, as that would make its existence certain.
This rule makes it difficult for such whales to find companions and mates.
• In the DC Comics series Animal Man, there was a storyline in which Animal
Man found himself displaced in time. To explain his situation to his friend,
Nowhere Man, he proposed "Schrodinger’s Pizza" — one with peppers, and
one without.

Audiovisual media

• In the episode of NUMB3RS entitled "Identity Crisis", Charlie re-phrases


Schrödinger’s Cat to illustrate the point that the tests used to dis-
prove/prove a murder suspect’s identity are built on a false premise.
• In the anime Master of Mosquiton, there is a discussion between
Schrödinger and the vampire Saint-Germain, wherein Germain claims he
can control fate. To prove this, Germain predicts the cast of a die. Just as
A the die is about to land as predicted, Schrödinger’s pet cat bats the die so
that it lands on a different number.
• Schrödinger’s cat is referenced in the movie John Carpenter’s Prince Of
Darkness.
• A Dutch short movie from 1990 is called Schrödingers kat. A teacher ex-
plains the theory to his class. At the same time a man called Schrödinger
(played by Bruno Ganz) dies of a heart attack in his apartment, while neigh-
bours assume he is still alive and send him Christmas cards. 315
DR
• A track on the Tears for Fears album Saturnine Martial & Lunatic (1996) is
called "Schrödinger’s Cat".
• Schrödinger’s cat is referenced in the TV series Stargate SG1

Games

• In the computer game NetHack, monsters known as quantum mechanics


may carry a chest containing Schrödinger’s Cat. When opened, there is a
50% chance of finding it dead and a 50% chance of it jumping out alive.
• In both Shin Megami Tensei: Digital Devil Saga games, Schrödinger is the
name of a cat that always seems to be at the right place at the right time.
• In Wild ARMs 3, the Schrödinger family owns a talking flying cat named
Shady. A scene shows Shady retrieving a gem only to trigger an invisible

315 http://www.imdb.com/title/tt0210966/

Schrödinger’s cat in fiction


544

barrier around him. While Maya Schrödinger explains about Shady’s claus-
trophobia, Shady screams "Whaaaaaaaaaaaah! No poison gas, please!"
• In the MMORPG Kingdom of Loathing, the PvP betting game known as "the
Money Making Game" has one message that uses Schrödinger’s cat as a

FT
contest. The betting player guesses whether the cat is dead or alive when
the box is opened.
• In the roleplaying game supplement GURPS Infinite Worlds, Schrödinger’s
Cat is used to illustrate the Many-worlds interpretation. Infinite Worlds,
as the name suggests, is a science-fiction setting where player-controlled
characters can physically travel to the alternate realities proposed by many-
worlds theory through the science of Parachronics.

Webcomics

• A cat named Schrödinger appears as a semi-recurring character in Checker-


board Nightmare, a webcomic by Kristofer Straub. Schrödinger can see all
possible states of existence at once, and as a result is very much insane, as
detailed in his first appearance 316.
A• The webcomic Two Lumps has an episode 317 about Schrödinger’s cat.
"They’re talking about schreddin’ cats and puttin’ ’em in a box!!!"
• Stephen Notley’s Bob the Angry Flower cartoon strip "Schrödinger’s
Fridge" 318 stated that it could not be determined if beer was in the fridge
or not. Although Bob whipped open the fridge door and was observing
his "frickin’ head off" the beers would not actualize. Bob quickly cursed
Quantum Physics.
• A "Schrödinger Box" (which no one is allowed to open) is used in Agnostica
celebrations in the comic strip Nukees.
DR
Reference
• ↑ E. Schrödinger, Die gegenwartige Situation in der Quantenmechanik,
Naturwissenschaftern. 23: pp. 807-812; 823-823, 844-849. (1935). En-
glish translation: John D. Trimmer, Proceedings of the American Philosoph-
ical Society 124, pp. 323-38 (1980), reprinted in Quantum Theory and
Measurement, p. 152 (1983).

316 http://checkerboardnightmare.com/d/20010319.html
317 http://www.twolumps.net/d/20040714.html
318 http://www.angryflower.com/schrod.gif

Schrödinger’s cat in fiction


545

External links
• Poem by Cecil Adams 319
• Stephen Notley’s Bob the Angry Flower: Schrödinger’s Fridge 320

FT
• "The Adventures of Schrodinger the Cat" comic strip 321

Source: http://en.wikipedia.org/wiki/Schr%C3%B6dinger%27s_cat_in_fiction

Schwinger’s variational principle

In Schwinger’s variational approach to quantum field theory, introduced by


Julian Schwinger, the quantum action is an operator. This is unlike the func-
tional integral (path integral) approach where the action is a classical function-
al.
Suppose we have a complete set of commuting operators (or anticommuting
Afor fermions)  and another set B̂. Let |A> be the eigenstate of  with eigen-
value A and similarly for |B>. There is some ambiguity in the phase, but that
can be taken care of in the quantum action S AB associated with  and B̂.
Suppose also we have not just one model of quantum mechanics or quantum
field theory but a whole family of them, varying smoothly. So, |A> and |B> are
"different" for each model in the family. S AB also varies smoothly. Schwinger’s
variational principle tells us
δ < A|B >= i < A|δSAB |B >.
DR
Source: http://en.wikipedia.org/wiki/Schwinger%27s_variational_principle

Principal Authors: Phys, Joyous!, Conscious, Matt McIrvin, Aranel

319 http://www.straightdope.com/classics/a1_122.html
320 http://www.angryflower.com/schrod.gif
321 http://glp.customer.netspace.net.au/schrod/

Schwinger’s variational principle


546

Selection rule

In physics, especially in the context of quantum mechanics, a selection rule

FT
is a condition constraining the physical properties of the initial system and the
final system that is necessary for a process to occur with a nonzero probability.
See also: angular momentum coupling
In many cases, a transition involves the emission of radiation, that is, a pho-
ton is emitted. In general, electric (charge) radiation or magnetic (current,
magnetic moment radiation) can be classified into multipoles E (electric) or
M (magnetic) of order 2, e.g. E1, E2, E3 for electric dipole, quadrupole or
octupole. The radiation field will be a sum of the multipole contributions;
however, usually one or two multipoles dominate.
The emitted particle carries away an angular momentum , which for the photon
must be at least 1, since it is a vector particle (i.e., it has J P = 1 -). Thus there is
no E0 (electric monopoles) or M0 (magnetic monopoles) radiation (the latter
is forbidden because magnetic monopoles do not seem to exist).
A
Since the total angular momentum has to be conserved during the transition,
we have that
Ji = Jf + λ

p
where kλk = λ(λ + 1) ~ , and its z-projection is given by λz = µ ~. The
corresponding quantum numbers , µ must satisfy
|Ji − Jf | ≤ λ ≤ Ji + Jf
DR
and
µ = Mi − Mf .

Parity is also preserved. For electric multipole transitions


π(Eλ) = πi πf = (−1)λ

while for magnetic multipoles


π(Mλ) = πi πf = (−1)λ+1 .

Thus, parity does not change for E-even or M-odd multipoles, while it changes
for E-odd or M-even multipoles.

Selection rule
547

These considerations generate different sets of transitions rules depending on


the multipole order and type. The expression forbidden transitions is often
used; this does not mean that these transitions cannot occur, only that they are
electric-dipole forbidden. These transitions are perfectly possible, they merely

FT
occur at a lower rate. If the rate for an E1 transition is non-zero, the transition
is said to be permitted; if it is zero, then M1, E2, etc. transitions can still
produce radiation, albeit with much lower transitions rates. These are the
so-called forbidden transitions. The transition rate decreases by a factor of
approximately 10 -3 from one multipole to the next one, so the lowest multipole
transitions are most likely to occur.
Semi-forbidden transitions (resulting in so called intercombination lines) are
electric dipole (E1) transitions for which the selection rule that the spin does
not change is violated. This is a result of the failure of LS coupling.

Summary table
Electric Magnetic Electric Magnetic Electric Magnetic
dipole (E1) dipole (M1) quadrupole quadrupole octupole octupole
(E2) (M2) (E3) (M3)
ARigorous
rules
(1) ∆J = 0, ±1
(J = 0 6↔ 0)
∆J = 0, ±1, ±2 ∆J = 0, ±1, ±2, ±3
(J = 0 6↔ 0, 1; 12 6↔(021 )6↔ 0, 1, 2; 12 6↔ 12 , 32 ; 1 6↔
(2) ∆MJ = 0, ±1 ∆MJ = 0, ±1, ±2 ∆MJ = 0, ±1, ±2, ±3
(3) πf = −πi πf = πi πf = −πi πf = πi
LS coupling (4) One electron No electron None or one One electron One electron One electron
jump∆l = jump∆l = electron jump∆l = jump∆l = jump∆l =
±1 0, jump∆l = ±1 ±1, ±3 0, ±2
∆n = 0 0, ±2
(5) If ∆S = If ∆S = If ∆S = If ∆S =
DR
∆L = 0,0∆L
±1 = 0 ∆L = 0, ±1, ±2 ∆L = 0, ±1, ±2, ±3
0 (L = 0 6↔ 0) 0 (L = 0 6↔ 0, 1) 0 (L= 0 6↔ 0, 1, 2; 1 6↔ 1)
Inter- (6) If ∆S = If ∆S = If ∆S = If ∆S = If ∆S =
mediate ±1∆L = 0, ±1, ±2 ∆L = 0, ±1,
∆L = 0, ±1 ∆L = 0, ∆L
±1, = 0, ±1,
coupling
±1 (L = 0 6↔ 0)
±2, ±3 ±2, ±3, ±4 ±2
±1 (L = 0 6↔ 0) ±1 (L = 0 ±1
6↔ (L
0, 1)
= 0 6↔ 0)

External links
• National Institue of Standards and Technology 322
• Lecture notes from The University of Sheffield 323

322 http://physics.nist.gov/Pubs/AtSpec/node17.html
323 http://www.shef.ac.uk/physics/teaching/phy332/atomic_physics3.pdf

Selection rule
548

Source: http://en.wikipedia.org/wiki/Selection_rule

Principal Authors: Joriki, Gaius Cornelius, Laurascudder, Jag123, Passw0rd

FT
Semiclassical

In physics, the adjective semiclassical has different precise meanings depend-


ing on the context. All these meanings usually refer to some approximation,
limit or situation that combines quantum and classical aspects in a given prob-
lem. The plurality of meanings comes from the fact that the passage from
quantum to classical mechanics is generally a very difficult task. Some of the
possible significations are the following:
First, semiclassical approximation may refer to quantum-mechanical calcu-
lations that are obtained by considering a small perturbation of a classical
calculation, for example the WKB approximation in non-relativistic quantum
mechanics or the loop expansion or the instanton methods in quantum field
A
theory. In quantum field theory, a semiclassical correction arises from one-loop
Feynman diagrams. The semiclassical effective action is
h i
Γ[φ] = S[φ] + 21 T r ln S (2) [φ] + ...

Second, in the context of open quantum systems and measurement theory,


where one considers the dynamics of a given quantum system in interaction
with an environment, the semiclassical regime may refer to the situation in
DR
which the wavefunction of the system is approximately peaked around the
solution of the corresponding classical equations of motion. Corrections to the
classical trajectory and the dispersion of the solution around the mean value
are usually considered.
Third, semiclassical gravity is the approximation to the yet unknown theory
of quantum gravity in which one treats matter fields as being quantum and the
gravitational field as being classical. The classical Einstein equations are com-
puted with the expectation value of the quantum matter fields in the classical
background. Semiclassical gravity has applications in black hole physics and
physical cosmology.
A semiclassical approximation is any high frequency approximation (or "high
energy approximation"), less extreme than classical mechanics, that is used to
approximate quantum mechanics.

Semiclassical
549

Source: http://en.wikipedia.org/wiki/Semiclassical

Principal Authors: Daniel Arteaga, QFT, Lumidek, Scottbeck, Wahoofive

FT
Separable states

In quantum mechanics, separable quantum states are states without quantum


entanglement.

Separable pure states


For simplicity, the following assumes all relevant state spaces are finite dimen-
sional. First, consider separability for pure states.
Let H1 and H2 be quantum mechanical state spaces, that is, finite dimensional
→Hilbert spaces with basis states {|ai i}ni=1 and {|bj i}m
j=1 , respectively. By a
postulate of quantum mechanics, the state space of the composite system is
given by the tensor product
H1 ⊗ H2
A N
with base states {|ai i |bj i}, or in more compact notation {|ai bj i}. From the
very definition of the tensor product, any vector of norm 1, i.e. a pure state of
the composite system, can be written as
|ψi = Σi,j ci,j |ai i ⊗ bj i = Σi,j ci,j |ai bj i

If a pure states |ψi ∈ H1 ⊗ H2 that can be written in the form |ψi = |ψ1 i ⊗ |ψ2 i
DR
where |ψi i is a pure state of the i-th subsystem, it is said to be separable.
Otherwise it is called entangled. Formally, the embedding of a product of states
into the product space is given by the Segre embedding. That is, a quantum-
mechanical pure state is separable if and only if it is in the image of the Segre
embedding.
A standard
 example of an (un-normalized) entangled state is
1
0
0 ∈ H ⊗ H
|ψi =  

where H is the Hilbert space of dimension 2. We see that when a system is


in an entangled pure state, it is not possible to assign states to its subsystems.
This will be true, in the appropriate sense, for the mixed state case as well.

Separable states
550

The above discussion can be extended to the case of when the state space is
infinite dimensional with virtually nothing changed.

Separability for mixed states

FT
Consider the mixed state case. A mixed state of the composite system is de-
scribed by a density matrix ρ acting on H1 ⊗ H2 . ρ is separable if there exist
pk ≥ 0, {ρk1 } and {ρk2 } which are mixed states of the respective subsystems
such that
ρ = k pk ρk1 ⊗ ρk2
P

where
P
k pk = 1.

Otherwise ρ is called an entangled state. We can assume without loss of gen-


erality in the above expression that {ρk1 } and {ρk2 } are all rank-1 projections,
that is, they represent pure ensembles of the appropriate subsystems. It is clear
from the definition that the family of separable states is a convex set.
A
Notice that, again from the definition of the tensor product, any density ma-
trix, indeed any matrix acting on the composite state space, can be trivially
written in the desired form, if we drop the requirement that {ρk1 } and {ρk2 } are
P
themselves states and k pk = 1.
In the language of quantum communication, a separable state can be created
from any other state using local actions and classical communication while an
entangled state cannot.
DR
When the state spaces are infinite dimensional, density matrices are replaced
by positive trace class operators with trace 1, and a state is separable if it can
be approximated, in trace norm, by states of the above form.

Extending to the multipartite case


The above discussion generalizes easily to the case of a quantum system con-
sisting of more than two subsystems. Let a system have n subsystems and have
state space H = H1 ⊗ · · · ⊗ Hn . A pure state |ψi ∈ H is separable if it takes the
form
|ψi = |ψ1 i ⊗ · · · ⊗ |ψn i.

Similarly, a mixed state ρ acting on H is separable if it is a convex sum


ρ = k pk ρk1 ⊗ · · · ρkn .
P

Separable states
551

Or, in the infinite dimensional case, ρ is separable if it can be approximated in


the trace norm by states of the above form.

Separability criterion

FT
The problem of deciding whether a state is separable in general is sometimes
called the separability problem in quantum information. It is considered to be
a difficult problem. A feeling for this difficulty can be obtained if we attempt
to solve the problem by employing the obvious naive approach, for a fixed di-
mension. We see that, using the naive techniqie, the problem quickly becomes
intractable, even for low dimensions. Thus more sophisticated formulations
are required. The separability problem is a subject of current research.
A separability criterion is a necessary condition a state must satisfy to be sepa-
rable. In the low dimensional (2 X 2 and 2 X 3 ) cases, the →Peres-Horodecki
criterion is actually a necessary and sufficient condition for separability. Other
separability criterions include the →Range criterion and Reduction criterion.

Characterization via algebraic geometry


A
Quantum mechanics may be modelled on a projective Hilbert space, and the
categorical product of two such spaces is the the Segre embedding. In the
bipartite case, a quantum state is separable if and only if it lies in the image of
the Segre embedding.

See also
• →Entanglement witness
DR
Source: http://en.wikipedia.org/wiki/Separable_states

Principal Authors: Mct mht, Matthew Mattic, Vegalabs, Oleg Alexandrov, Keenan Pepper

Separable states
552

Shelter Island Conference

The first Shelter Island Conference on the Foundations of Quantum Mechan-

FT
ics was held from June 2-4, 1947 at the Ram’s Head Inn in Shelter Island, New
York. The most famous participant, J. Robert Oppenheimer, deemed it the most
successful scientific meeting he had ever attended. A relatively young Richard
Feynman would later observe, "There have been many conferences in the world
since, but I’ve never felt any to be as important as this." The conference cost
$850.
Shelter Island was the first major opportunity since Pearl Harbor and the Man-
hattan Project for the leaders of the American physics community to escape the
paranoia of war. As Julian Schwinger would later recall, "It was the first time
that people who had all this physics pent up in them for five years could talk to
each other without somebody peering over their shoulders and saying, ’Is this
cleared?’ "

Organization
A
The conference was conceived by Duncan MacInnes, a scientist studying elec-
trochemistry at the Rockefeller Institute for Medical Research. Once the pres-
ident of the New York Academy of Sciences, MacInnes had already organized
a number of small scientific conferences. However, he believed that the later
conferences had suffered from a bloated attendance, and over this issue, he
resigned from the Academy in January 1945. That fall, he approached the
National Academy of Sciences (NAS) with the idea of a series of 2–3 day con-
ferences limited to 20–25 people. Frank Jewett, the head of the NAS, liked
DR
the idea; he envisioned a "meeting at some quiet place where the men could
live together intimately", possibly "at an inn somewhere", and suggested that
MacInnes focus on a couple of pilot programs. MacInnes’ first choice was "The
Nature of Biopotentials", a subject close to his own heart; the second would be
"The Postulates of Quantum Mechanics", which later became "Foundations of
Quantum Mechanics".
K. K. Darrow, a theoetical physicist at Bell Labs and secretary of the Amer-
ican Physical Society, offered his help in organizing the quantum mechanics
conference. The two decided to emulate the success of the early Solvay Con-
gresses, and they consulted with Léon Brillouin, who had some experience in
that area. In turn, Brillouin suggested consulting Wolfgang Pauli, the recent
Nobel medalist at the Institute for Advanced Study at Princeton.

Shelter Island Conference


553

In January 1946, MacInnes, Darrow, Brillouin, and Pauli met in New York and
exchanged letters. Pauli was enthusiastic about the topic, but he was primar-
ily interested in bringing together the international physics community after
the ordeal of the war. He suggested a large conference, including many old-

FT
er, foreign physicists, much to MacInnes’ chagrin. With Jewett’s encourage-
ment, MacInnes asked Pauli for suggestions of "younger men" such as John
Archibald Wheeler, explaining that the Rockefeller Foundation would support
only a small conference. Pauli and Wheeler replied that MacInnes’ conference
might be merged with Niels Bohr’s conference on Wave Mechanics in Denmark
in 1947; they pointed out that the Niels Bohr Institute had close ties with the
Rockefeller Foundation anyway. Darrow wrote to Wheeler that Bohr’s confer-
ence was a poor replacement because it would draw few Americans. Finally,
Shelter Island was explicitly an American conference.
In the coming months, Wheeler...

Lamb shift
The muon
AParticipants
• Hans Bethe • Duncan MacInnes • Bruno Rossi
• David Bohm • Robert Eugene Marshak • Julian Schwinger
• Gregory Breit • John von Neumann • Robert Serber
• Karl K. Darrow • Arnold Nordsieck • Edward Teller
• Herman Feshbach • J. Robert Oppenheimer • George Uhlenbeck
• Richard Feynman • Abraham Pais • John Hasbrouck van Vleck
• Hendrik Kramers • Linus Pauling • Victor Frederick Weisskopf
DR
• Willis Lamb • Isidor Isaac Rabi • John Archibald Wheeler

See also
• Solvay Conference

References
Primary sources

• Bethe, Hans (August 15, 1947). "The Electromagnetic Shift of Energy


Levels" 324. Phys. Rev. 72 (4): 339–341.
• Lamb and Retherford (August 1, 1947). "Fine Structure of the Hydrogen
Atom by a Microwave Method" 325. Phys. Rev. 72 (3): 241–243.

324 http://link.aps.org/abstract/PR/v72/p339

Shelter Island Conference


554

• Marshak and Bethe (September 15, 1947). "On the Two-Meson


Hypothesis" 326. Phys. Rev. 72 (6): 506–509.
• Weisskopf, Victor (September 15, 1947). "On the Production Process of
Mesons" 327. Phys. Rev. 72 (6): 510.

FT
Reviews

• Robert Crease and Charles Mann (1996). The second creation: Makers
of the revolution in twentieth-century physics. Rutgers UP. ISBN 0-8135-
2177-7.
• Schweber, Silvan (1985). R. Jackiw, N. Khuri, S. Weinberg, E. Witten "A
Short History of Shelter Island I". Shelter Island II: Proceedings of the 1983
Shelter Island Conference on Quantum Field Theory and the Fundamental
Problems of Physics, 301–343, Cambridge, MA: MIT Press. ISBN 0-262-
10031-2.
• Schweber, Silvan (1986). "Feynman and the visualization of space-time
processes" 328. Rev. Mod. Phys. 58 (2): 449–508.
• Schweber, Silvan (1994). QED and the men who made it: Dyson, Feynman,
A•
Schwinger, and Tomonaga. Princeton UP. ISBN 0-691-03685-3.
Smith, Richard (1996). Notes on the 1947 Shelter Island Conference and
Its Participants 329. Retrieved on January 17, 2006.

External links
• The Shelter Island Conference 330 from the National Academy of Sciences
• Excerpt from 1983 issue 331 of Physics Today
DR
• This Month in Physics History, June 2000 332 from The American Physical
Society
• Shelter island conference 333 from Issues in Science and Technology Summer
1997, provided by ProQuest

325 http://link.aps.org/abstract/PR/v72/p241
326 http://link.aps.org/abstract/PR/v72/p506
327 http://link.aps.org/abstract/PR/v72/i6/p510/s1
328 http://prola.aps.org/abstract/RMP/v58/i2/p449_1
329 http://www.sherryart.com/nano/shelter.html
330 http://www7.nationalacademies.org/archives/shelterisland.html
331 http://feynman.physics.lsa.umich.edu/~mduff/talks/1983%20-%20Shelter%20Island%20Conference

/1983%20-%20Shelter%20Island%20Conference.pdf
332 http://www.aps.org/apsnews/0600/060006.cfm
333 http://www.findarticles.com/p/articles/mi_qa3622/is_199707/ai_n8765069/print

Shelter Island Conference


555

Source: http://en.wikipedia.org/wiki/Shelter_Island_Conference

Single particle reconstruction

FT
In physics, in the area of microscopy, single particle reconstruction is a tech-
nique in which large numbers of images (10,000 - 1,000,000) of ostensibly
identical individual molecules or macromolecular assemblies are combined to
produce a 3 dimensional reconstruction. This is a complementary technique
to crystallography of biological molecules. As molecules/assembies become
larger, it becomes more difficult to prepare high resolution crystals. For sin-
gle particle reconstruction, the opposite is true. Larger objects actually im-
prove the resolution of the final structure. In single particle reconstruction,
the molecules/assemblies in solution are prepared in a thin layer of vitreous
(glassy) ice, then imaged on an electron cryomicroscope (see Transmission
electron microscopy). Images of individual molecules/assemblies are then se-
lected from the micrograph and then a complex series of algorithms is applied
A
to produce a full volumetric reconstruction of the molecule/assembly. In the
1990’s this technique was limited to roughly 2 nm resolution, providing on-
ly gross features of the objects being studied. However, recent improvements
in both microscope technology as well as available computational capabilities
now make 0.5 nm resolution possible.

References
• The National Center for Macromolecular Imaging, Houston Texas USA 334
DR
Source: http://en.wikipedia.org/wiki/Single_particle_reconstruction

334 http://ncmi.bcm.tmc.edu

Single particle reconstruction


556

Slater determinant

A Slater determinant (named after the American physicist John C. Slater) is

FT
an expression in quantum mechanics for the wavefunction of a many-fermion
system, which by construction satisfies the Pauli principle.
The Slater determinant arises from the consideration of a wavefunction for a
collection of electrons. The wavefunction for each individual electron is known
as a spin-orbital, χ(x), where x indicates the position and spin of the electron.

Two-particle case
The simplest way to approximate the wavefunction of a many-particle system
is to take the product of properly chosen one-electron wavefunctions of the
individual particles. For the two-particle case, we have
Ψ(x1 , x2 ) = χ1 (x1 )χ2 (x2 )

This expression occurs in Hartree theory and is known as a Hartree product.


A
However, it is not satisfactory for fermions, such as electrons, because the
wavefunction is not antisymmetric. An antisymmetric wavefunction can be
mathematically described as follows:
Ψ(x1 , x2 ) = −Ψ(x2 , x1 )

Therefore the Hartree product does not satisfy the Pauli principle. This problem
can be overcome by taking a linear combination of both Hartree products
DR
Ψ(x1 , x2 ) = √1 {χ1 (x1 )χ2 (x2 ) − χ1 (x2 )χ2 (x1 )}
2

where the coefficient is a normalization factor. This wavefunction is antisym-


metric and no longer distinguishes between electrons. Moreover, it also goes to
zero if any two wavefunctions or two electrons are the same. This is equivalent
to satisfying the Pauli exclusion principle.

Generalization to the Slater determinant


The expression can be generalised to any number of fermions by writing it as
a determinant. For an N-electron
system, the Slater determinant is defined as
χ1 (x1 ) χ1 (x2 ) · · · χ1 (xN )

χ2 (x1 ) χ2 (x2 ) · · · χ2 (xN )
Ψ(x1 , x2 , . . . , xN ) = √1N ! . .. ..

..

. .

χN (x1 ) χN (x2 ) · · · χN (xN )

Slater determinant
557

The linear combination of Hartree products for the two-particle case can clear-
ly be seen as identical with the Slater determinant for N =2. It can be seen
that the use of (Slater) determinants assures an antisymmetrized function on

FT
the outset, symmetric functions are automatically rejected. In the same way,
the use of Slater determinants assures the obeying of the Pauli principle; the
determinant will vanish if any of the two spin-orbitals are identical, for this
leads to two identical columns.
A single Slater determinant is used as an approximation to the electronic wave-
function in Hartree-Fock theory. In more accurate theories (such as configura-
tion interaction and MCSCF), a linear combination of Slater determinants is
needed.

Source: http://en.wikipedia.org/wiki/Slater_determinant

Principal Authors: Karada, Karol Langner, Charles Matthews, Syntax, CiaPan


A
Spin-1/2

In quantum mechanics, spin is an intrinsic property of all elementary parti-


cles. Fermions, the particles that make up ordinary matter, have half-integer
spin. An important special case of this are the spin-1/2 particles. All known
elementary particles that are fermions have spin 1/2.

Overview
DR
Particles which are spin-1/2 include the electron, the proton, the neutron, the
neutrino, and the quarks. Spin-1/2 objects have peculiar dynamics which are
not accurately described using classical physics; they are among the simplest
systems which require quantum mechanics to describe them. As such, the study
of the behavior of spin-1/2 systems forms a central part of undergraduate and
graduate-level instruction in quantum mechanics.

General properties
Spin-1/2 objects are all Fermions, a fact explained by the spin statistics theo-
rem, and satisfy the Pauli exclusion principle. Spin-1/2 particles can have a per-
manent magnetic moment along the direction of their spin, and this magnetic
moment gives rise to electromagnetic interactions that depend on the spin.
One such effect that was important in the discovery of spin is the →Zeeman
effect.

Spin-1/2
558

Unlike in more complicated quantum mechanical systems, the spin of a spin-


1/2 particle can be expressed as a linear combination of just two eigenstates, or
eigenspinors. These are traditionally labeled spin up and spin down. Because
of this the quantum mechanical spin operators can be represented as simple

FT
2x2 matrices, as opposed to the infinite dimensional matrices commonly need-
ed to represent operators like energy or position. These matrices are called the
Pauli matrices.
Raising and lowering operators can be constructed for spin 1/2 objects; these
obey the same commutation relations as other angular momentum operators.

Connection to the uncertainty principle


One consequence of the generalized uncertainty principle is that the spin pro-
jection operators (which measure the spin along a given direction like x, y,
or z), cannot be measured simultaneously. Physically, this means that it is ill
defined what axis a particle is spinning about. A measurement of the z compo-
nent of spin destroys any information about the x and y components that might
previously have been obtained.
A
Stern-Gerlach experiment
When a spin-1/2 particle with non-zero magnetic moment like an electron is
placed in an inhomogenous magnetic field, it experiences a force. This acts to
separate out particles in the spin up state from particles in the spin down state.
This is the idea behind the →Stern-Gerlach experiment.

Mathematical Description
DR
The quantum state of the spin of a spin-1/2 particle can be described by a
complex-valued vector with two components called a two-component spinor.
When spinors are used to describe the quantum states, quantum mechanical
operators are represented by 2 x 2, complex-valued Hermitian matrices.
For example, the spin projection operator Sz effects a measurement of the spin
in the z direction.
 
1 0
Sz = ~2 σz = ~
2 0 −1

Sz operator has two eigenvalues, of ± ~2 , which correspond to the eigenvectors


 
1
= sz = + 21 = |↑i

0

Spin-1/2
559
 
0
= sz = − 21 = |↓i

1

These vectors form a complete basis for the →Hilbert space describing the

FT
spin-1/2 particle. Thus, linear combinations of these two states can represent
all possible states of the spin.

See also
• Spin
• Spinor
• Fermions
• Pauli matrices

References
Griffiths, David J. (2005) Introduction to Quantum Mechanics(2nd ed.). Upper
Saddle River, NJ: Pearson Prentice Hall. ISBN 0-13-111892-7.
A
Source: http://en.wikipedia.org/wiki/Spin-1/2

Spin-orbital

In quantum mechanics, a spinorbital is a one-particle wavefunction taking


both the position and spin angular momentum of a particle as its parameters.
DR
The spinorbital of a single electron, for example, is a complex-valued function
of four real variables: the three scalars used to define its position, and a fourth
scalar, m s, which can be either +1/2 or -1/2:
χ(x, y, z, ms )

We can also write it more compactly as a function of a position vector ~r =


(x, y, z) and the quantum number m s:
χ(~r, ms ).

For a general particle with spin s, m s can take values between -s to s in integer
steps. The electron has s=1/2.
A spinorbital is usually normalized, such that the probability of finding the
particle anywhere in space with any spin is equal to 1:

Spin-orbital
560
Ps 3~ |χ(~r, ms )|2 = 1.
R
ms =−s ∞ d r

From a normalized spinorbital, one can calculate the probability that the par-
ticle is in an arbitrary volume of space V and has an arbitrary spin ms :

FT
P (V, ms ) = V d3~r |χ(~r, ms )|2 .
R

Source: http://en.wikipedia.org/wiki/Spin-orbital

Squashed entanglement

Squashed entanglement, also called CMI entanglement (CMI can be pro-


nounced "see me"), is an information theoretic measure of quantum entangle-
ment for a bipartite quantum system. If %A,B is the density matrix of a sys-
tem (A, B) composed of two subsystems A and B, then the CMI entanglement
A
ECM I of system (A, B) is defined by
Eq.(1) ECM I (%A,B ) = 1
2 % min
A,B,Λ ∈K
S(A : B|Λ),

where K is the set of all density matrices %A,B,Λ for a tripartite system (A, B, Λ)
such that %A,B = trΛ (%A,B,Λ ). Thus, CMI entanglement is defined as an ex-
tremum of a functional S(A : B|Λ) of %A,B,Λ . We define S(A : B|Λ), the quan-
tum Conditional Mutual Information (CMI), below. A more general version
of Eq.(1) replaces the “min" (minimum) in Eq.(1) by an “inf" (infimum).
DR
Motivation for definition of CMI entanglement
CMI entanglement has its roots in classical (non-quantum) information theory,
as we explain next.
Given any two random variables A, B, classical information theory defines the
mutual information, a measure of correlations, as
Eq.(2) H(A : B) = H(A) + H(B) − H(A, B) .

For three random variables A, B, C, it defines the CMI as


H(A : B|Λ) = H(A|Λ) + H(B|Λ) − H(A, B|Λ)
Eq.(3)
= H(Λ) − H(A, Λ) − H(B, Λ) + H(A, B, Λ)
.

Squashed entanglement
561

It can be shown that H(A : B|Λ) ≥ 0.


Now suppose %A,B,Λ is the density matrix for a tripartite system (A, B, Λ). We
will represent the partial trace of %A,B,Λ with respect to one or two of its sub-
systems by %A,B,Λ with the symbol for the traced system erased. For example,

FT
%A,B = traceΛ (%A,B,Λ ). One can define a quantum analogue of Eq.(2) by
Eq.(4) S(A : B) = S(%A ) + S(%B ) − S(%A,B ) ,

and a quantum analogue of Eq.(3) by


Eq.(5) S(A : B|Λ) = S(%Λ ) − S(%A,Λ ) − S(%B,Λ ) + S(%A,B,Λ ) .

It can be shown that S(A : B|Λ) ≥ 0. This inequality is often called the strong-
subadditivity property of quantum entropy.
Consider three random variables A, B, Λ with probability distribution
PA,B,Λ (a, b, λ), which we will abbreviate as P (a, b, λ). For those special
P (a, b, λ) of the form
Eq.(6) P (a, b, λ) = P (a|λ)P (b|λ)P (λ) ,
A
it can be shown that H(A : B|Λ) = 0. Probability distributions of the form
Eq.(6) are in fact described by the Bayesian network shown in Fig.1.
DR

One can define a classical CMI entanglement by


Eq.(7) ECM I (PA,B ) = min H(A : B|Λ),
PA,B,Λ ∈K

where K is the set of all probability distributions PA,B,Λ in three random vari-
P
ables A, B, Λ, such that λ PA,B,Λ (a, b, λ) = PA,B (a, b) for all a, b. Because,
given a probability distribution PA,B , one can always extend it to a probability
distribution PA,B,Λ that satisfies Eq.(6), it follows that the classical CMI entan-
glement, ECM I (PA,B ), is zero for all PA,B . The fact that ECM I (PA,B ) always

Squashed entanglement
562

vanishes is an important motivation for the definition of ECM I (%A,B ). We want


a measure of quantum entanglement that vanishes in the classical regime.
Suppose wλ for λ = 1, 2, ..., dim(Λ) is a set of non-negative numbers that add
up to one, and |λ > for λ = 1, 2, ..., dim(Λ) is an orthonormal basis for the

FT
Hilbert space associated with a quantum system Λ. Suppose %λA and %λB , for
λ = 1, 2, ..., dim(Λ) are density matrices for the systems A and B, respectively.
It can be shown that the following density matrix
%A,B,Λ = λ %λA %λB wλ |λ >< λ|
P
Eq.(8)

satisfies S(A : B|Λ) = 0. Eq.(8) is the quantum counterpart of Eq.(6). Trac-


ing the density matrix of Eq.(8) over Λ, we get %A,B = λ %λA %λB wλ , which
P

is a separable state. Therefore, ECM I (%A,B ) given by Eq.(1) vanishes for all
separable states.
Next suppose |ψA,Bλ > for λ = 1, 2, ..., dim(Λ) are some states in the Hilbert
space associated with a quantum system (A, B). Let K be the set of density
matrices defined previously for Eq.(1). Define Ko to be the set of all density
matrices %A,B,Λ that are elements of K and have the special form %A,B,Λ =
P
A λ λ
λ |ψA,B >< ψA,B |wλ |λ >< λ| . It can be shown that if we replace in Eq.(1)
the set K by its proper subset Ko , then Eq.(1) reduces to the definition of
entanglement of formation for mixed states, as given in Ben96. K and Ko
represent different degrees of knowledge as to how %A,B,Λ was created. K
represents total ignorance.

History
Classical CMI, given by Eq.(3), first entered information theory lore, shortly
DR
after Shannon’s seminal 1948 paper and at least as early as 1954 in McG54.
The quantum CMI, given by Eq.(5), was first defined by Cerf and Adami in
Cer96. However, it appears that Cerf and Adami did not realize the relation
of CMI to entanglement or the possibility of obtaining a measure of quantum
entanglement based on CMI; this can be inferred, for example, from a later
paper, Cer97, where they try to use S(A|B) instead of CMI to understand
entanglement. The first paper to explicitly point out a connection between
CMI and quantum entanglement appears to be Tuc99.
The final definition Eq.(1) of CMI entanglement was first given by Tucci in a
series of 6 papers. (See, for example, Eq.(8) of Tuc02 and Eq.(42) of Tuc01a).
In Tuc00b, he pointed out the classical probability motivation of Eq.(1), and
its connection to entanglement of formation for mixed states. In Tuc01a, he
presented an algorithm and computer program, based on the Arimoto-Blahut
method of information theory, for calculating CMI entanglement numerically.

Squashed entanglement
563

In Tuc01b, he calculated CMI entanglement analytically, for a mixed state of


two qubits.
In Hay03, Hayden, Jozsa, Petz and Winter explored the connection between
quantum CMI and separability.

FT
Since CMI entanglement reduces to entanglement of formation if one mini-
mizes over Ko instead of K, one expects that CMI entanglement inherits many
desirable properties from entanglement of formation. As first shown in Ben96,
entanglement of formation does not increase under LOCC (Local Operations
and Classical Communication). In Chr03, Christandl and Winter showed that
CMI entanglement also does not increase under LOCC, by adapting Ben96 ar-
guments about entanglement of formation. In Chr03, they also proved many
other interesting inequalities concerning CMI entanglement, and explored its
connection to other measures of entanglement. The name squashed entan-
glement first appeared in Chr03. In Chr05, Christandl and Winter calculated
analytically the CMI entanglement of some interesting states.
In Ali03, Alicki and Fannes proved the continuity of CMI entanglement.

References
A• Ali03 R. Alicki, M. Fannes, “Continuity of quantum mutual information",
quant-ph/0312081 335
• Ben96 C.H. Bennett, D.P. DiVincenzo, J.A. Smolin, W.K. Wootters, “Mixed
State Entanglement and Quantum Error Correction", quant-ph/9604024 336
• Cer96 N. J. Cerf, C. Adami, “Quantum Mechanics of Measurement", quant-
ph/9605002 337
• Cer97 N.J. Cerf, C. Adami, R.M. Gingrich, “Quantum conditional operator
DR
and a criterion for separability", quant-ph/9710001 338
• Chr03 M. Christandl, A. Winter, “Squashed Entanglement - An Additive
Entanglement Measure", quant-ph/0308088 339
• Chr05 M. Christandl, A. Winter, “Uncertainty, Monogamy, and Locking of
Quantum Correlations", quant-ph/0501090 340
• Chr06 M. Christandl, Ph.D. Thesis, quant-ph/0604183 341

335 http://arxiv.org/abs/quant-ph/0312081
336 http://arxiv.org/abs/quant-ph/quant-ph/9604024
337 http://arxiv.org/abs/quant-ph/9605002
338 http://arxiv.org/abs/quant-ph/9710001
339 http://arxiv.org/abs/quant-ph/0308088
340 http://arxiv.org/abs/quant-ph/0501090
341 http://arxiv.org/abs/quant-ph/0604183

Squashed entanglement
564

• Hay03 P. Hayden, R. Jozsa, D. Petz, A. Winter, “Structure of states


which satisfy strong subadditivity of quantum entropy with equality" quant-
ph/0304007 342
• McG54 W.J. McGill, “Multivariate Information Transmission", IRE Trans.

FT
Info. Theory 4(1954) 93-111.
• Tuc99 R.R. Tucci, “Quantum Entanglement and Conditional Information
Transmission", quant-ph/9909041 343
• Tuc00a R.R. Tucci,“Separability of Density Matrices and Conditional Infor-
mation Transmission", quant-ph/0005119 344
• Tuc00b R.R. Tucci, “Entanglement of Formation and Conditional Informa-
tion Transmission", quant-ph/0010041 345
• Tuc01a R.R. Tucci, “Relaxation Method For Calculating Quantum Entangle-
ment", quant-ph/0101123 346
• Tuc01b R.R. Tucci, “Entanglement of Bell Mixtures of Two Qubits", quant-
ph/0103040 347
• Tuc02 R.R. Tucci, “Entanglement of Distillation and Conditional Mutual
Information", quant-ph/0202144 348
ASource: http://en.wikipedia.org/wiki/Squashed_entanglement

Squeezed coherent state

In physics, a squeezed coherent state is every state in the →Hilbert space of


quantum mechanics that saturates the uncertainty principle that is the product
DR
of the corresponding two operators takes on its minimum value:
~
∆x∆p = 2

The simplest such state is the ground state |0i of the quantum harmonic oscil-
lator. The next simple class of states that satisfies this identity are the family of
coherent states |αi.

342 http://arxiv.org/abs/quant-ph/0304007
343 http://arxiv.org/abs/quant-ph/9909041
344 http://arxiv.org/abs/quant-ph/0005119
345 http://arxiv.org/abs/quant-ph/0010041
346 http://arxiv.org/abs/quant-ph/0101123
347 http://arxiv.org/abs/quant-ph/0103040
348 http://arxiv.org/abs/quant-ph/0202144

Squeezed coherent state


565

Mathematical definition
The most general wave function that satisfies the identity above is the
squeezed coherent state (we work in units with ~ = 1)
(x−x )2
 

FT
ψ(x) = C exp − 2w20 + ip0 x
0

where C, x0 , w0 , p0 are constants (a normalization constant, the center of the


wavepacket, its width, and its average momentum). The new feature relative
to a coherent state is the free value of the width w0 , which is the reason why
the state is called "squeezed".
The squeezed state above is an eigenstate of a linear operator
x̂ + ip̂w02

and the corresponding eigenvalue equals x0 + ip0 w02 . In this sense, it is a


generalization of the ground state as well as the coherent state.

Examples of squeezed coherent states


A
Depending on at which phase the state’s quantum noise is reduced one can dis-
tinguish amplitude-squeezed and phase-squeezed states or general quadrature
squeezed states. If no coherent excitation exists the state is called a squeezed
vacuum. The figures below give a nice visual demonstration of the close con-
nection between squeezed states and Heisenbergs uncertainty relation: Dimin-
ishing the quantum noise at a specific quadrature (phase) of the wave has
as a direct consequence an enhancement of the noise of the complementary
quadrature, that is the field at the phase shifted by π/2.
DR
From the top:

• Vacuum state
• Squeezed vacuum state
• Phase-squeezed state
• arbitrary squeezed state
• Amplitude-squeezed state

As can be seen at once in contrast to the coherent state the quantum noise
is not independent of the phase of the light wave anymore. A characteristic
broadening and narrowing of the noise during one oscillation period can be
observed. The wave packet of a squeezed state is defined by the square of the
wave function introduced in the last paragraph. They correspond to the prob-
ability distribution of the electric field strength of the light wave. The moving
wave packets display an oscillatory motion combined with the widening and

Squeezed coherent state


566

A FT
Figure 45 Figure 1: Measured quan-
tum noise of the electric field of differ-
ent squeezed states in dependence of
the phase of the light field. For the first
two states a 3π-interval is shown, for
the last three states, belonging to a dif-
DR
ferent set of measurements it is a 4π-
interval. (source: link 1 and ref. 3)

narrowing of their distribution: The "breathing" of the wave packet. For an


amplitude-squeezed state, the most narrow distribution of the wave packet is
reached at the field maximum, resulting in an amplitude that is defined more
precisely than the one of a coherent state. For a phase-squeezed state the most
narrow distribution is reached at field zero, resulting in an average phase value
that is better defined than the one of a coherent state.
In phase space quantum mechanical uncertainties can be depicted by Wigner
distributions. The intensity of the light wave, its coherent excitation is given by
the displacement of the Wigner distribution from the origin. A change in the
phase of the squeezed quadrature results in a rotation of the distribution.

Squeezed coherent state


567

A FT
Figure 46 Figure 2: Oscillating wave
packets of the five states.

Photon number distributions and phase distribu-


tions of squeezed states
The squeezing angle, that is the phase with minimum quantum noise, has a
DR
large influence on the photon number distribution of the light wave and its
phase distribution as well.
For amplitude squeezed light the photon number distribution is usually nar-
rower than the one of a coherent state of the same amplitude resulting in sub-
poissonian light, whereas its phase distribution ist wider. The opposite ist true
for the phase-squeezed light, which displays a large intensity (photon number)
noise but a narrow phase distribution.
For the squeezed vacuum state the photon number distribution displays odd-
even-oscillations. This can be explained by the mathematical form of the
squeezing operator, that resembles the operator for two-photon generation and
annihilation processes. Photons in a squeezed vacuum state are more likely to
appear in pairs.

Squeezed coherent state


568

A FT
Figure 47 Figure 3: Wigner functions of the five
states. The ripples are due to experimental inaccu-
racies.

Experimental realizations of squeezed coherent


states
There has been a whole variety of successful demonstrations of squeezed states.
The most prominent ones were experiments with light fields using lasers and
non-linear optics. But squeezed states have also been realized via motional
DR
states of an ion in a trap, phonon states in crystal lattices or atom ensembles.
Even macroscopic oscillators were driven into classical motional states that
were very similar to squeezed coherent states.

Applications
Squeezed states of the light field can be used to enhance precision measure-
ments. For example phase-squeezed light can improve the phase read out of in-
terferometric measurements (see for example gravitational waves). Amplitude-
squeezed light can improve the read out of very weak spectroscopic signals.
Various squeezed coherent states, generalized to the case of many degrees of
freedom, are used in various calculations in quantum field theory, for example
Unruh effect and Hawking radiation (generally: particle production in curved
backgrounds).

Squeezed coherent state


569

A FT
Figure 48 Figure 1: Measured
quantum noise of the electric
field of different squeezed states
in dependence of the phase of the
light field. For the first two states
a 3π-interval is shown, for the
DR
last three states, belonging to a
different set of measurements it
is a 4π-interval. (source: link 1
and ref. 3)

See also
• Quantum optics

External links
• An introduction to quantum optics of the light field 349

349 http://gerdbreitenbach.de/gallery

Squeezed coherent state


570

A
References
FT
Figure 49 Figure 2: Os-
cillating wave packets of
the five states.

• Loudon, Rodney, The Quantum Theory of Light (Oxford University Press,


DR
2000), [ISBN 0198501773]
• D.F. Walls and G.J. Milburn, Quantum Optics, Springer Berlin 1994
• G. Breitenbach, S. Schiller, and J. Mlynek, "Measurement of the quantum
states of squeezed light", Nature, 387, 471 (1997) 350

Source: http://en.wikipedia.org/wiki/Squeezed_coherent_state

Principal Authors: Gerd Breitenbach, TimBentley, Gaius Cornelius, Alai, Matt McIrvin

350 http://www.exphy.uni-duesseldorf.de/Publikationen/1997/N387/471z.htm

Squeezed coherent state


571

A
SQUID
FT
Figure 50 Figure 3: Wigner functions of
the five states. The ripples are due to ex-
perimental inaccuracies.

For other uses, see Squid (disambiguation).

SQUIDs, or Superconducting Quantum Interference Devices, are used to


DR
measure extremely small magnetic fields; they are currently the most sensitive
such devices (magnetometers) known, with noise levels as low as 3 fT·Hz - /2.
1

While a typical fridge magnet is ten thousand microteslas, some processes in


animals produce very small magnetic fields; typically sized between a nanotesla
(a million femto-tesla or fT) and a microtesla (1000 nanotesla), and SQUIDs
are well suited to studying these.

History and Design


The DC SQUID was invented in 1964 by Robert Jaklevic, John Lambe, Arnold
Silver, and James Mercereau of Ford Research Labs after B. D. Josephson pos-
tulated the Josephson junction in 1962 and the first Josephson Junction was
made by John Rowell and Philip Anderson at Bell Labs in 1963. The RF SQUID
was invented in 1965 by James Edward Zimmerman and Arnold Silver at Ford.

SQUID
572

A FT
Figure 51 Figure 1: Measured
quantum noise of the electric
field of different squeezed states
in dependence of the phase of the
light field. For the first two states
a 3π-interval is shown, for the
DR
last three states, belonging to a
different set of measurements it
is a 4π-interval. (source: link 1
and ref. 3)

There are two main types of SQUID, DC and RF (or AC). RF SQUIDs have only
one Josephson junction whereas DC SQUIDs have two or more junctions. This
makes DC SQUIDs more difficult and expensive to produce, but DC SQUIDs are
much more sensitive.
Most SQUIDs are fabricated from lead or pure niobium. The lead is usually in
the form of an alloy with 10% gold or indium, as pure lead is unstable when its
temperature is repeatedly changed. The base electrode of the SQUID is made
of a very thin niobium layer, formed by deposition, and the tunnel barrier is
oxidised onto this niobium surface. The top electrode is a layer of lead alloy
deposited on top of the other two, forming a sandwich arrangement.

SQUID
573

A FT
Figure 52 Figure 2: Os-
cillating wave packets of
the five states.

More recently developed "High Temperature" SQUIDS are made of a substance


called YBCO (chemical formula YBa 2Cu 3O 7-x), and are cooled by liquid nitro-
gen which is cheaper and more easily handled than liquid helium. They are less
DR
sensitive than conventional "Low Temperature" SQUIDS but many applications
do not require the extreme sensitivity of the LT SQUID.
The basic principle of operation is closely linked to flux quantisation. This is the
phenomenon that the favoured states for a loop of superconductor are those
where the flux inside is a multiple of the flux quantum.

Uses for SQUIDs


The extreme sensitivity of SQUIDs make them ideal for studies in biology. Mag-
netoencephalography (MEG), for example, uses measurements from an array
of SQUIDs to make inferences about neural activity inside brains. Because
SQUIDs can operate at acquisition rates much higher than the highest temporal
frequency of interest in the signals emitted by the brain (kHz), MEG achieves
good temporal resolution. Another application is the scanning SQUID micro-
scope, which uses a SQUID immersed in liquid helium as the probe. The use of
SQUIDs in oil prospecting, earthquake prediction and geothermal energy sur-
veying is becoming more widespread as superconductor technology develops;

SQUID
574

A FT
Figure 53 Figure 3: Wigner functions of
the five states. The ripples are due to ex-
perimental inaccuracies.

they are also used as precision movement sensors in a variety of scientific ap-
plications, such as the detection of gravity waves. Four SQUIDs are currently
employed on Gravity Probe B in order to test the limits of the theory of general
relativity.

SQUIDs in Fiction
The science fiction writer William Gibson made reference to SQUIDs in his story
DR
Johnny Mnemonic, where a genetically engineered ex-military dolphin uses a
SQUID implant to read a memory device in the title character’s brain. SQUIDs
are also referenced in the film Strange Days, where they are used to record and
play back human memories, which are exchanged on the black market.

See also
Spallation Neutron Source Superconducting RF Cavities

External links
• Dr. John Bland’s explanation 351.

351 http://www.cmp.liv.ac.uk/frink/thesis/thesis/node47.html

SQUID
575

A FT
Figure 54 Figure 4: Measured photon number
distributions for an amplitude-squeezed state,
a coherent state, and a phase squeezed state.
Bars refer to theory, dots to experimental values.
(source: link 1 and ref. 2)
DR

Figure 55 Figure 5: Pegg-Barnett phase distribu-


tion of the three states.

Source: http://en.wikipedia.org/wiki/SQUID

Principal Authors: Gene Nygaard, Lostart, TwoOneTwo, Cspalletta, Tzartzam, Glenn, Slicky,
Frencheigh, McE, Maximus Rex

SQUID
576

A FT
Figure 56 Figure 4: Measured photon
number distributions for an amplitude-
squeezed state, a coherent state, and
a phase squeezed state. Bars refer
to theory, dots to experimental values.
(source: link 1 and ref. 2)
DR

Figure 57 Figure 5: Pegg-Barnett phase


distribution of the three states.

Stark effect

In atomic physics, the Stark effect is the splitting and shift of a spectral line into
several components in the presence of an electric field. The amount of splitting
itself is called the Stark shift. It is analogous to the →Zeeman effect where a
spectral line is split into several components in the presence of a magnetic field.

Stark effect
577

A FT
Figure 58 Figure 4: Measured photon
number distributions for an amplitude-
squeezed state, a coherent state, and
a phase squeezed state. Bars refer
to theory, dots to experimental values.
(source: link 1 and ref. 2)
DR

Figure 59 Figure 5: Pegg-Barnett phase


distribution of the three states.

The Stark effect is responsible for the pressure broadening (Stark broadening)
of spectral lines by charged particles.

History
The effect is named after Johannes Stark, who discovered it in 1913. It was
independently discovered in the same year by the Italian physicist Antonino
Lo Surdo, and is thus sometimes called the Stark-Lo Surdo effect. Earlier,

Stark effect
578

A
(source: link 1 and ref. 3)
FT
Figure 60 Figure 4: Measured photon number distributions for a squeezed-vacuum state.

unsuccessful, attempts to compute the magnitude of the effect, and to discov-


er the perturbation, had been made by Voigt in 1899. In 1916, Epstein and
Schwarzschild were able to perform computations using the Bohr model of the
atom to exactly fit the magnitude of the Stark effect in hydrogen. In 1920,
Hendrik Kramers was able to perform calculations within the Bohr model to
estimate the relative intensities of the lines in the line pattern.
DR
While first-order perturbation effects for the Stark effect in hydrogen are in
agreement for the Bohr model and the quantum-mechanical theory of the
atom, higher order effects are not. Measurements of the Stark effect under
high field strengths confirmed the correctness of the quantum theory over the
Bohr model.

Mechanism
The effect arises because of the interaction between the electric dipole moment
of an electron with an external electric field. If the electric field is uniform over
the length scale of the atom, then the perturbing Hamiltonian is of the form
~ = eEz ẑ.
H 1 = p~ · E

The first order energy shift of the state |ψm i due to the perturbation is given
by ∆Em = eEz hψm | ẑ |ψm i (see Perturbation theory). Since the unperturbed

Stark effect
579

A
Figure 61
(SQUID).
FT
A prototype of a Semiconductor Superconducting Quantum Interference Device
DR

Figure 62 The inner workings of a early Super-


conducting Quantum Interference Device (SQUID).

states may be degenerate, we normally need to use the eigenvectors of H 1

Stark effect
580

A FT
Figure 63 Second order Stark shifts in hydrogen. Magnetic quantum number: m = 1, parabol-
ic quantum number: (n1 - n2) = -n+4,-n+2,...,n-2,n-4. States with higher principle quantum
number n experience a greater energy shift for a given electric field strength.actual size

when calculating the energy shifts. The effect of H 1 is therefore to lift this
degeneracy, which is observed experimentally as a splitting of spectral lines.

Quantum-Confined Stark Effect


In a semiconductor heterostructure, where a small bandgap material is sand-
DR
wiched between two layers of a larger bandgap material, the Stark effect can be
dramatically enhanced by bound excitons. This is due to the fact that the elec-
tron and hole which form the exciton are pulled in opposite directions by the
applied electric field, but they remain confined in the smaller bandgap materi-
al, so the exciton is not merely pulled apart by the field. The quantum-confined
Stark effect is widely used for semiconductor-based optical modulators, partic-
ularly for optical fiber communications.

References
• Voigt, Annalen der Physik, 69, 297 (1899), and 4, 197 (1901).
• Epstein, Annalen der Physik, 50, 489 (1916).
• Schwarzschild, Sitzber. Berliner Akad., (1916) p. 548.

Stark effect
581

• Kramers, Danske Vidensk. Selsk. Skrifter (8), III, 3, 287. (1920), and
Zeitschrift fur Physik, 3. 169 (1920).
• E. U. Condon and G. H. Shortley (1935). The Theory of Atomic Spectra.
Cambridge University Press. ISBN 521-09209-4. (Chapter 17 provides a

FT
comprehensive treatment, as of 1935.)

Source: http://en.wikipedia.org/wiki/Stark_effect

Principal Authors: Linas, Laurascudder, Lionelbrits, Piil, That Guy, From That Show!

Stationary state

In quantum mechanics, a stationary state is an eigenstate of a Hamiltonian.


It is called stationary because, as an eigenstate, it is not subject to change or
decay (to a lower energy state) over time.
In practice, stationary states are never truly "stationary" for all time. Rather,
A
they refer to the eigenstate of a Hamiltonian where small perturbative effects
have been ignored. The language allows one to discuss the eigenstates of the
unperturbed Hamiltonian, whereas the perturbation will eventually cause the
stationary state to decay. The only true stationary state is the ground state.

Ground state
The ground state of a quantum mechanical system is its lowest-energy state.
An excited state is any state with energy greater than the ground state. The
DR
ground state of a quantum field theory is usually called the vacuum state or
the vacuum.
If more than one ground state exists, they are said to be degenerate. Many sys-
tems have degenerate ground states, for example, the hydrogen atom. It turns
out that degeneracy occurs whenever a nontrivial unitary operator commutes
with the Hamiltonian of the system.
According to the third law of thermodynamics, a system at absolute zero tem-
perature exists in its ground state; thus, its entropy is determined by the de-
generacy of the ground state. Many systems, such as a perfect crystal lattice,
have a unique ground state and therefore have zero entropy at absolute zero
(because ln(1) = 0).
The condition of an atom, ion, or molecule, when all of its electrons are in their
lowest possible energy levels, is called, not excited. When an atom is in its
ground state, its electrons fill the lowest energy orbitals completely before they

Stationary state
582

begin to occupy higher energy orbitals, and they fill subshells in accordance
with Hund’s rule (usually!).

See also

FT
• Quantum number
• Quantum mechanic vacuum or vacuum state
• Virtual particle

Source: http://en.wikipedia.org/wiki/Stationary_state

Stern-Gerlach experiment

In quantum mechanics, the Stern-Gerlach experiment, named after Otto


Stern and Walther Gerlach, is a celebrated experiment in 1920 on deflection of
A
particles, often used to illustrate basic principles of quantum mechanics. It can
be used to demonstrate that electrons and atoms have intrinsically quantum
properties, and how measurement in quantum mechanics affects the system
being measured.

Basic theory and description


DR

Figure 64 Classical model of spin-


ning particle.

Otto Stern and Walther Gerlach devised an experiment that would determine
whether particles had any intrinsic angular momentum. If we imagine a clas-
sical system, such as the earth orbiting the sun, then the earth has angular
momentum from both its orbit around the sun and the orbit around its axis (its
spin). We are seeking to determine whether individual particles like electrons
have any "spin" angular momentum. To do this, we consider the electron to be
like a classical dipole with two halves of charge spinning quickly. If this parti-
cle is in a magnetic field, it will begin to precess because of the torque that the
magnetic field is exerting on the dipole (see Torque-induced precession).

Stern-Gerlach experiment
583

If the particle is travelling in a homogeneous magnetic field, then the forces


exerted on opposite ends of the dipole cancel each other out and the motion of
the particle is unaffected. If we are conducting the experiment using electrons,
then we must compensate for the tendency of any charged particle to curl in its

FT
path through a magnetic field (see cyclotron motion). This can be done easily
using an electric field of appropriate magnitude and oriented transverse to the
charged particle’s path. We can then ignore the fact that electrons are charged.
The Stern-Gerlach experiment can be conducted using electrically neutral par-
ticles and the same conclusion is reached. We are concerning ourselves with
angular momentum only, not any electrostatic phenomena.

A
Figure 65 Basic elements of the Stern-Gerlach experiment.
DR
If our particle is travelling through an inhomogeneous magnetic field, then the
force on one end of the dipole will be slightly greater than the opposing force
on the other end of the dipole. This leads to the particle being deflected in
the inhomogeneous magnetic field. The direction in which the particles are
deflected is typically called the "z" direction.
If our particles are classical, "spinning" particles, then we would expect the
distribution of their spin angular momentum vectors to be truly random and
thus each particle would be deflected up or down by a different amount. Thus,
we would expect an even distribution on the screen of our detector. Instead,
we find that the particles passing through the device are deflected either up or
down by a specific amount! This can only mean that spin angular momentum
is quantized, i.e. it can only take on discrete values. There is not a continuous
distribution of possible angular momenta!

Stern-Gerlach experiment
584

A Figure 66
FT
Spin values for fermions.

Electrons are spin-1/2 particles. These have only two possible spin values,
called spin-up and spin-down. The exact value of their spin is +h/2 or -h/2.
What does this quantity imply about the nature of these particles? If this value
arises as a result of the particles rotating the way a planet rotates, then the
individual particles would have to be spinning impossibly fast! The speed of
rotation would be in excess of the speed of light and thus impossible. We
DR
conclude that the spin angular momentum has nothing to do with rotation and
is a purely quantum mechanical phenomenon. That is why it is sometimes
known as the "intrinsic angular momentum."
For electrons, two possible values for spin exist, as well as for the proton and
the neutron, which are composite particles made up of three quarks each,
which are themselves spin 1/2 particles. Other particles may have a different
number of possible values. Delta baryons (∆ ++, ∆ +, ∆ 0, ∆ -), for example,
are spin-3/2 particles and have four possible values for spin angular momen-
tum. Vector mesons, as well as photons, W and Z bosons and gluons are spin-1
particles and have three possible values for spin angular momentum.
To describe the experiment with spin-1/2 particles mathematically, it is easiest
to use Dirac’s bra-ket notation. As the particles pass through the Stern-Gerlach
device, they are "being observed." The act of observation in quantum mechanics
is equivalent to measuring them. Our observation device is the detector and

Stern-Gerlach experiment
585

in this case we can observe one of two possible values, either spin up or spin
down. These are described by the angular momentum quantum number j,
which can take on one of the two possible allowed values, either +h/2 or -
h/2. The act of observing (measuring) corresponds to the operator J z. In

FT
mathematical terms,
E E
|ψi = c1 ψj=+ ~ + c2 ψj=− ~

2 2

The constants c 1 and c 2 are complex numbers. The square of their absolute
values determines the probability of the state |ψ> being found with one of
the two possible values for j. The constants must also be normalized so the
probability of finding the wavefunction in one of either state is unity. Here we
know that the probability of finding the particle in each state is 0.5. Therefore
we also the values of the constants c 1 and c 2. These are
c1 = √1
2

c2 = √1
2
A
Sequential experiments
If we combine some Stern-Gerlach apparati we can clearly see that
they do not act as simple selectors, but alter the states observed
(as in light polarization), according to quantum mechanics laws:
DR

History
The Stern-Gerlach experiment was performed in Frankfurt, Germany in 1920
by Otto Stern and Walther Gerlach. At the time, Stern was an assistant to
Max Born at the University of Frankfurt’s Institute for Theoretical Physics, and

Stern-Gerlach experiment
586

A Figure 67 FT
A plaque at the Frankfurt institute commemorating the experiment

Gerlach was an assistant at the same university’s Institute for Experimental


Physics.
At the time of the experiment, the most prevalent model for describing the
atom was the →Bohr model, which described electrons as going around the
positively-charged nucleus only in certain discrete atomic orbitals or energy
levels. Since the electron was quantized to be only in certain positions in space,
DR
the separation into distinct orbits was referred to as space quantization.

Impact
The Stern-Gerlach experiment, had one of the biggest impacts on modern
physics:

• In the decade that followed, scientists showed using similar techniques,


that the nucleus of some atoms also have quantized angular momentum.
It is the interaction with the spin of the electron that is responsible for the
hyperfine structure of the spectroscopic lines.

• In the thirties, using an extended version of the S-G apparatus, Isidor Rabi
and colleagues showed that by using a varying magnetic field, one can force
the magnetic momentum to go from one state to the other. The series of
experiments culminated in 1937 when they discovered that state transitions

Stern-Gerlach experiment
587

could be induced using time varying fields or RF fields. The so called Rabi
oscillation is the working mechanism for the Magnetic Resonance Imaging
equipment found in hospitals.

• Later Norman F. Ramsey, modified the Rabi apparatus to increase the inter-

FT
action time with the field. The extreme sensitivity due to frequency of the
radiation makes this very useful for keeping accurate time, and is still used
today in atomic clocks.

• In the early sixties, Ramsey and Daniel Kleppner used a S-G system to pro-
duce a beam of polarized hydrogen as the source of energy for the Hydrogen
Maser, which is still one of the most popular atomic clocks.

• The direct observation of the spin is the most direct proof of quantization
and in quantum mechanics.

External links
• Stern-Gerlach Experiment Java Applet Animation 352
A• Detailed explanation of the Stern-Gerlach Experiment 353

References
• Friedrich, Bretislav and Herschbach, Dudley. "Stern and Gerlach: How
a Bad Cigar Helped Reorient Atomic Physics" 354 Physics Today, December
2003.
DR
Source: http://en.wikipedia.org/wiki/Stern-Gerlach_experiment

Principal Authors: BeardedPhysicist, Michael Hardy, Creidieki, Sonett72, Theresa knott

352 http://www.if.ufrgs.br/~betz/quantum/SGPeng.htm
353 http://galileo.phys.virginia.edu/classes/252/Angular_Momentum/Angular_Momentum.html
354 http://www.physicstoday.org/vol-56/iss-12/p53.html

Stern-Gerlach experiment
588

Subatomic particle

A subatomic particle is a particle smaller than an atom: it may be elementary

FT
or composite. Particle physics and nuclear physics concern themselves with the
study of these particles, their interactions, and matter made up of them which
do not aggregate into atoms.
These particles include atomic constituents such as electrons, protons, and
neutrons (protons and neutrons are actually composite particles, made up of
quarks), as well as other particles such as photons and neutrinos which are
produced copiously in the sun. However, most of the particles that have been
discovered and studied are not encountered under normal earth conditions;
they are produced in cosmic rays and during scattering processes in particle
accelerators.
A
DR
Figure 68 Helium atom (schematic)
Showing two protons (red), two neutrons (green) and two electrons
(yellow).

Dividing an atom
The study of electrochemistry led G. Johnstone Stoney to postulate the exis-
tence of the electron (denoted e -) in 1874 as a constituent of the atom. It
was observed in 1897 by J. J. Thomson. Subsequent speculation about the
structure of atoms was severely constrained by the 1907 experiment of Ernest
Rutherford which showed that the atom was mostly empty space, and almost
all its mass was concentrated into the (relatively) tiny atomic nucleus. The
development of the quantum theory led to the understanding of chemistry in

Subatomic particle
589

terms of the arrangement of electrons in the mostly empty volume of atoms.


Protons (p +) were known to be the nucleus of the hydrogen atom. Neutrons
(n) were postulated by Rutherford and discovered by James Chadwick in 1932.
The word nucleon denotes both the neutron and the proton.

FT
Electrons, which are negatively charged, have a mass of 1/1836 of a hydrogen
atom, the remainder of the atom’s mass coming from the positively charged
proton. The atomic number of an element counts the number of protons. Neu-
trons are neutral particles with a mass almost equal to that of the proton.
Different isotopes of the same nucleus contain the same number of protons but
differing numbers of neutrons. The mass number of a nucleus counts the total
number of nucleons.
Chemistry concerns itself with the arrangement of electrons in atoms and
molecules, and nuclear physics with the arrangement of protons and neutrons
in a nucleus. The study of subatomic particles, atoms and molecules, their
structure and interactions, involves quantum mechanics and quantum field the-
ory (when dealing with processes that change the number of particles). The
study of subatomic particles per se is called particle physics. Since many par-
ticles need to be created in high energy particle accelerators or cosmic rays,
A
sometimes particle physics is also called high energy physics.

Classification of subatomic particles


Symmetries play a very important role in the physics of subatomic particles
by providing intrinsic quantum numbers which are used to classify particles.
Poincare symmetry, which is the full symmetry of special relativity, is enjoyed
by any Hamiltonian which describes these particles. Hence all particles have
DR
the following quantum numbers —

• the mass (m) of the particle,


• its spin (J): all particles with integer values of spin are called bosons, those
with half-integer spins are called fermions.
• its intrinsic parity (P), which is a multiplicative quantum number.

In addition, some particles may have a definite C-parity (C). Particles may also
carry other quantum numbers related to internal symmetries, such as charges
and flavour quantum numbers.
Corresponding to every particle there exists an antiparticle. Every additive
quantum number of a particle is reversed in sign for the antiparticle. Equality
of the masses and lifetimes of particle and antiparticle follows in local quan-
tum field theories through CPT symmetry, and hence tests of these equalities
constitute important tests of this symmetry.

Subatomic particle
590

Elementary particles
A full classification of subatomic particles involves understanding the funda-
mental forces that they are subject to: the electromagnetic, weak and strong
forces. In the modern unified quantum field theory of these three forces, called

FT
the standard model, the elementary particles are

• spin J = 1 particles called gauge bosons. These include


• photons, which are carriers of the electromagnetic force,
• W bosons and Z bosons which mediate the weak forces, and
• gluons, which carry the strong force.
• spin J = 1/2 fermions which constitute all matter in the universe and
come in two varieties—
• leptons such as the electron, muon, tau lepton, the three corresponding
neutrinos (these are called six flavours of leptons), and their antiparti-
cles. These are affected essentially only by the weak and electromagnet-
ic forces. The former allow flavour changes (for example, from a muon
to an electron)
• quarks which come in six other flavours, and are affected by all three
A forces unified into the standard model. The weak interactions cause
flavour changes.
• spin J = 0 (and P = +1) Higgs boson which is responsible for the
masses of the quarks, leptons, W and Z bosons. This remains to be
actually seen in experiments; a major purpose of the Large Hadron
Collider (LHC) is to search for this particle.

Conjectures and predictions


DR
Further structures beyond the standard model are often invoked. In particu-
lar, there is a search for a theory that unifies the standard model with gravity.
There is strong evidence that when such a theory is found it will include gravi-
tons (constrained to have spin J = 2), to mediate this fourth fundamental
interaction. A further structure called supersymmetry is often invoked, al-
though direct experimental evidence for it is lacking. Supersymmetric ex-
tensions of the standard model would contain a bosonic partner for each
of the fermions described above (called selectrons, smuons, staus, sneu-
trinos, squarks), and a fermionic partner for each boson (called gauginos
and Higgsinos). Supersymmetric extensions which include a theory of
gravity (called supergravity) also involve a partner of the graviton, called
the gravitino, which has spin J = 3/2. In many versions of these theories
there are extra bosons called axions with J = 0 and P = -1. Relic particles
are postulated to be remnants of the early cosmological expansion of the
Big Bang.

Subatomic particle
591

There were attempts to build theories which posited that the elementary par-
ticles in the standard model are actually composites built out of really ele-
mentary particles variously called preons, rishons or quinks. However, these
theories are so strongly constrained by experimental data now that they are al-

FT
most ruled out. Extended supersymmetric theories have also been postulated;
these allow particles such as leptoquarks, which transmute leptons into quarks.

Composite particles
All observed subatomic composite particles are called hadrons. All bosonic
hadrons are called mesons and all fermionic hadrons are baryons. The most
well-known baryons are the constituents of atomic nuclei called protons and
neutrons, and collectively named nucleons. The quark model of hadrons posits
that mesons are built out of a quark and an antiquark, whereas a baryon is
made up of three quarks. As of 2005, searches for exotic hadrons are currently
under way.

History
J. J. Thomson discovered electrons in 1897. In 1905 Albert Einstein demon-
A
strated the physical reality of the photons which were postulated by Max Planck
in order to solve the problem of black body radiation in thermodynamics.
Ernest Rutherford discovered in 1907 in the gold foil experiment that the atom
is mainly empty space, and that it contains a heavy but small atomic nucleus.
The early successes of the quantum theory involved explaining properties of
atoms in terms of their electronic structure. The proton was soon identified as
the nucleus of hydrogen. The neutron was postulated by Rutherford follow-
ing his discovery of the nucleus, but was discovered by James Chadwick much
DR
later, in 1932. Neutrinos were postulated in 1931 by Wolfgang Pauli (and
named by Enrico Fermi) to be produced in beta decays (the weak interaction)
of neutrons, but were not discovered till 1956. Pions were postulated by Hide-
ki Yukawa as mediators of the strong force which binds the nucleus together.
The muon was discovered in 1936 by Carl D. Anderson, and initially mistaken
for the pion. In the 1950s the first kaons were discovered in cosmic rays. The
development of new particle accelerators and particle detectors in the 1950s
led to the discovery of a huge variety of hadrons, prompting Wolfgang Pauli’s
remark: "Had I foreseen this, I would have gone into botany". The classification
of hadrons through the quark model in 1961 was the beginning of the golden
age of modern particle physics, which culminated in the completion of the uni-
fied theory called the standard model in the 1970s. The discovery of the gauge
bosons through the 1980s, and the verification of their properties through the
1990s is considered to be an age of consolidation in particle physics. Among
the standard model particles the existence of the Higgs boson remains to be

Subatomic particle
592

verified—this is seen as the primary physics goal of the accelerator called the
Large Hadron Collider in CERN. All particles found till now fit into the standard
model.

See also

FT
• Poincare symmetry, CPT invariance, spin statistics theorem, bosons and
fermions.
• Particle physics, list of particles, the quark model and the standard model.

External links
• particleadventure.org: The Standard Model 355
• particleadventure.org: Particle chart 356
• University of California: Particle Data Group 357
• Annotated Physics Encyclopædia: Quantum Field Theory 358
• Jose Galvez: Chapter 1 Electrodynamics (pdf) 359
ASource: http://en.wikipedia.org/wiki/Subatomic_particle

Principal Authors: Bambaiah, Xerxes314, Ahoerstemeier, Glenn, Bevo, Doug Bell, LittleDan, GraemeL

Superdense coding
DR
Superdense coding is a technique used in quantum information theory to send
two bits of classical information using only one qubit, with the aid of entangle-
ment.

The idea
Suppose Alice would like to send classical information to Bob using qubits.
Alice would encode the classical information in a qubit and send it to Bob.
After receiving the qubit, Bob recovers the classical information via measure-
ment. The question is: how much classical information can be transmitted per
qubit? Since non-orthogonal quantum states can not be distinguished reliably,

355 http://particleadventure.org/particleadventure/frameless/standard_model.html
356 http://particleadventure.org/particleadventure/frameless/chart.html
357 http://pdg.lbl.gov/
358 http://web.mit.edu/redingtn/www/netadv/qft.html
359 http://jgalvez.home.cern.ch/jgalvez/School/pdf/LM-WeakIteractions.pdf

Superdense coding
593

one would guess that Alice can do no better than one classical bit per qubit.
Indeed this bound on efficiency has been proven formally. Thus there is no
advantage gained in using qubits instead of classical bits. However, with the
additional assumption that Alice and Bob share an entangled state, two classi-

FT
cal bits per qubit can be achieved. The term superdense refers to this doubling
of efficiency. We next describe this prodedure.

The result
Crucial to the procedure is the shared entangled state between Alice and Bob,
and the property of entangled state that a (maximally) entangled states can be
transformed into another such state via local manipulation.
Suppose parts of a Bell state, say
|Ψ+ i = √1 (|0iA ⊗ |1iB + |1iA ⊗ |0iB )
2

is distributed to Alice and Bob. The first subsystem, denoted by subscript A,


belongs to Alice and the second, B, system to Bob. By only manipulating her
particle locally, Alice can transform the composite system into any one of the
A
Bell states (this is not so surprising, since entanglement can not be broken
using local operations):

• Obviously, if Alice does nothing, the system remains in the state |Ψ+ i.

• If Alice sends her particle through the unitary gate


 
0 1
σ1 =
1 0
DR
(notice this is one of the Pauli matrices), the total two-particle system now is
in state
(σ1 ⊗ I)|Ψ+ i = |Φ+ i.

• If σ1 is replaced by σ3 , the initial state |Ψ+ i is transformed into |Ψ− i.

• Similarly, if Alice applies iσ2 ⊗ I to the system, the resulting state is |Φ− i

So, depending on the message she would like to send, Alice performs one of the
four local operations given above and sends her qubit to Bob. By performing
a projective measurement in the Bell basis on the two particle system, Bob
decodes the desired message.

Superdense coding
594

Notice that if some mischievous person, Carol, intercepts Alice’s qubit en route
to Bob, all that is obtained by Carol is part of an entangled state. No useful
information whatsoever is gained by Carol, unless she can interact with Bob’s
qubit.

FT
References
• C. Bennett and S.J. Wiesner. Communication via one- and two-particle
operators on Einstein-Podolsky-Rosen states. Phys. Rev. Lett., 69:2881,
1992 360

Source: http://en.wikipedia.org/wiki/Superdense_coding

Superselection sector
AA superselection sector is a concept used in quantum mechanics.
One of the insights of quantum mechanics is that not all self-adjoint operators
are observables.
Suppose we are given an operator unital *-algebra A and an observable uni-
tal *-subalgebra O (The observables would then correspond to the self-adjoint
elements of O). A reducible unitary representation of O is decomposable into
the direct sum of inequivalent irreducible unitary representations of O (I’ll ex-
plain why they have to be inequivalent in a moment). Each irrep is called a
DR
superselection sector. →Observables map a state in each irrep into another
state in the same irrep. The relative phase of a superposition of nonzero states
from different irreps is not observable (the expectation values of the observ-
ables can’t distinguish between them). In the density state formulation where
states are positive linear functionals of O where the unit of O is mapped to 1
(the unit, 1 O is "intuitively" an observable, a trivial one, no doubt), this would
correspond to a mixed state; and in fact, all possible values for the relative
phase would give rise to the same state. (In the density state formulation, the
direct sum of two or more equivalent irreps would give rise to exactly the same
states as a single rep alone, so by Occam’s razor, we can cut off all but one
irrep)

360 http://prola.aps.org/abstract/PRL/v69/i20/p2881_1

Superselection sector
595

Symmetries often give rise to superselections sectors; but this is not the only
reason for the occurrence of superselection sectors. In the study of decoher-
ence, for example, if we restrict ourselves to observations in a local region, we
can have approximate superselection sectors.

FT
Say a group G acts upon A, and H is a unitary rep of A, and also a unitary rep
of G such that for all g in G, a in A and |ψi in H,
g [a|ψi] = g [a] g [|ψi]

(i.e. the representation of H, as a unitary rep of A is a G -intertwiner).


O is an invariant subalgebra of A under G (all observables are invariant un-
der G, but not every self-adjoint operator invariant under G is necessarily an
observable). H decomposes into superselection sectors, each of which is the
direct product of an irrep of G and a rep of O. The irrep of G acts trivially
under O. Using the same Occam’s razor argument as above, we can reduce it
to a rep of O alone. However, we can still keep the irrep of G as a label for the
superselection sector, if we wish.
Actually, insisting that H is a rep of G is unnecessarily restrictive. To take an
A
example, let G be the Lorentz group and A be an operator algebra rep of G. H
could be a Hilbert space of A which contains, among other things states with
an odd number of fermions. This would mean H is not a rep of the Lorentz
group, although it IS a rep of its double cover.
Or let’s say G is a group with an extension K. A is again an algebra rep of G.
(Any rep of G can be turned into a rep of K) Then, it’s possible to have a unitary
rep of A, H which is a unitary rep of K but not G.
Actually, we can be more general than that. Replacing G with a Lie algebra, Lie
DR
superalgebra or a Hopf algebra would still work. See algebra representation of
a Lie superalgebra, unitary representation of a star Lie superalgebra, algebra
representation of a Hopf algebra and representation of a Hopf algebra.

Examples
A simple example would be a quantum mechanical particle confined to a closed
loop (i.e. a periodic line of period L ). The superselection sectors are labeled by
an angle θ between 0 and 2π. All the wave functions within a single superse-
lection sector satisfy
ψ (x + L) = eiθ ψ (x)

This would be the →Aharonov-Bohm effect, if we introduced a locally flat con-


nection A.
Source: http://en.wikipedia.org/wiki/Superselection_sector

Superselection sector
596

Principal Authors: Phys, Kebes, Charles Matthews, Sam Hocevar, Woohookitty

Supersymmetric quantum mechanics

FT
In theoretical physics, supersymmetric quantum mechanics is an area of re-
search where mathematical concepts from high-energy physics are applied to
the seemingly more prosaic field of quantum mechanics.

Introduction
Understanding the consequences of supersymmetry has proven mathematical-
ly daunting, and it has likewise been difficult to develop theories that could
account for symmetry breaking, i.e., the lack of observed partner particles of
equal mass. To make progress on these problems, physicists developed super-
symmetric quantum mechanics, an application of the SUSY superalgebra to
quantum mechanics as opposed to quantum field theory. It was hoped that
studying SUSY’s consequences in this simpler setting would lead to new un-
A
derstanding; remarkably, the effort created new areas of research in quantum
mechanics itself.
For example, as of 2004 students are typically taught to "solve" the hydrogen
atom by a laborious process which begins by inserting the Coulomb potential
into the →Schrödinger equation. After a considerable amount of work using
many differential equations, the analysis produces a recursion relation for the
Laguerre polynomials. The final outcome is the spectrum of hydrogen-atom
energy states (labeled by quantum numbers n and l ). Using ideas drawn from
DR
SUSY, the final result can be derived with significantly greater ease, in much
the same way that operator methods are used to solve the harmonic oscillator.
Oddly enough, this approach is analogous to the way Erwin Schrödinger first
solved the hydrogen atom. Of course, he did not call his solution supersym-
metric, as SUSY was thirty years in the future—but it is still remarkable that the
SUSY approach, both older and more elegant, is taught in so few universities.
The SUSY solution of the hydrogen atom is only one example of the very gen-
eral class of solutions which SUSY provides to shape-invariant potentials, a
category which includes most potentials taught in introductory quantum me-
chanics courses.
SUSY quantum mechanics involves pairs of Hamiltonians which share a partic-
ular mathematical relationship, which are called partner Hamiltonians. (The
potential energy terms which occur in the Hamiltonians are then called part-
ner potentials.) An introductory theorem shows that for every eigenstate of

Supersymmetric quantum mechanics


597

one Hamiltonian, its partner Hamiltonian has a corresponding eigenstate with


the same energy (except possibly for zero energy eigenstates). This fact can be
exploited to deduce many properties of the eigenstate spectrum. It is analogous
to the original description of SUSY, which referred to bosons and fermions. We

FT
can imagine a "bosonic Hamiltonian", whose eigenstates are the various bosons
of our theory. The SUSY partner of this Hamiltonian would be "fermionic",
and its eigenstates would be the theory’s fermions. Each boson would have a
fermionic partner of equal energy—but, in the relativistic world, energy and
mass are interchangeable, so we can just as easily say that the partner particles
have equal mass.
SUSY concepts have provided useful extensions to the WKB approximation. In
addition, SUSY has been applied to non-quantum statistical mechanics through
the Fokker-Planck equation, showing that even if the original inspiration in
high-energy particle physics turns out to be a blind alley, its investigation has
brought about many useful benefits.

The SUSY QM superalgebra


In fundamental quantum mechanics, we learn that an algebra of operators is
A
defined by commutation relations among those operators. For example, the
canonical operators of position and momentum have the commutator [x,p]=i.
(Here, we use "natural units" where Planck’s constant is set equal to 1.) A more
intricate case is the algebra of angular momentum operators; these quantities
are closely connected to the rotational symmetries of three-dimensional space.
To generalize this concept, we define an anticommutator, which relates opera-
tors the same way as an ordinary commutator, but with the opposite sign:
DR
{A, B} = AB + BA.

If operators are related by anticommutators as well as commutators, we say


they are part of a Lie superalgebra. Let’s say we have a quantum system de-
scribed by a Hamiltonian H and a set of N self-adjoint operators Q i. We shall
call this system supersymmetric if the following anticommutation relation is
valid for all i, j = 1, . . . , N :
{Qi , Qj } = Hδij .

If this is the case, then we call Q i the system’s supercharges.

Supersymmetric quantum mechanics


598

Example
Let’s look at the example of a one-dimensional nonrelativistic particle with a
2D (i.e., two state) internal degree of freedom called "spin" (it’s not really
spin because "real" spin is a property of 3D particles). Let b be an operator

FT
which transforms a "spin up" particle into a "spin down" particle. Its adjoint
b † then transforms a spin down particle into a spin up particle; the operators
are normalized such that the anticommutator {b,b †}=1. And of course, b 2=0.
Let p be the momentum of the particle and x be its position with [x,p]=i. Let
W (the "superpotential") be an arbitrary complex analytic function of x and
define the supersymmetric operators
Q1 = 21 (p − iW )b + (p + iW † )b†
 

i
(p − iW )b − (p + iW † )b†
 
Q2 = 2

Note that Q 1 and Q 2 are self-adjoint. Let the Hamiltonian


(p+={W })2 <{W }2 <{W }0 †
H = {Q1 , Q1 } = {Q2 , Q2 } = 2 + 2 + 2 (bb − b† b)
A
where W’ is the derivative of W. Also note that {Q 1,Q 2}=0. This is nothing
other than N = 2 supersymmetry. Note that ={W } acts like an electromagnetic
vector potential.
Let’s also call the spin down state "bosonic" and the spin up state "fermionic".
This is only in analogy to quantum field theory and should not be taken literally.
Then, Q 1 and Q 2 maps "bosonic" states into "fermionic" states and vice versa.
Let’s reformulate this a bit:
DR
Define
Q = (p − iW )b

and of course,
Q† = (p + iW † )b†

{Q, Q} = {Q† , Q† } = 0

and
{Q† , Q} = 2H

Supersymmetric quantum mechanics


599

An operator is "bosonic" if it maps "bosonic" states to "bosonic" states and


"fermionic" states to "fermionic" states. An operator is "fermionic" if it maps
"bosonic" states to "fermionic" states and vice versa. Any operator can be ex-
pressed uniquely as the sum of a bosonic operator and a fermionic operator.

FT
Define the supercommutator [,} as follows: Between two bosonic operators or
a bosonic and a fermionic operator, it is none other than the commutator but
between two fermionic operators, it is an anticommutator.
Then, x and p are bosonic operators and b, b† , Q and Q† are fermionic opera-
tors.
Let’s work in the →Heisenberg picture where x, b and b† are functions of time.
Then,
[Q, x} = −ib

[Q, b} = 0

[Q.b† } = dx
dt − i<{W }
A[Q† , x} = ib†

[Q† , b} = dx
dt + i<{W }

[Q† , b† } = 0

This is nonlinear in general: i.e., x(t), b(t) and b† (t) do not form a linear
DR
SUSY representation because <{W } isn’t necessarily linear in x. To avoid this
problem, define the self-adjoint operator F = <{W }. Then,
[Q, x} = −ib

[Q, b} = 0

[Q.b† } = dx
dt − iF

[Q, F } = − db
dt

[Q† , x} = ib†

Supersymmetric quantum mechanics


600

[Q† , b} = dx
dt + iF

[Q† , b† } = 0

FT
db†
[Q† , F } = dt

and we see that we have a linear SUSY representation.


Now let’s introduce two "formal" quantities, θ; and θ̄ with the latter being the
adjoint of the former such that
{θ, θ} = {θ̄, θ̄} = {θ̄, θ} = 0

and both of them commute with bosonic operators but anticommute with
fermionic ones.
Next, we define a construct called a superfield:
f (t, θ̄, θ) = x(t) − iθb(t) − iθ̄b† (t) + θ̄θF (t)
Af is self-adjoint, of course. Then,
[Q, f } = ∂
∂θ f

− iθ̄ ∂t f,

[Q† , f } = ∂
∂ θ̄
f ∂
− iθ ∂t f.

Refererences
DR
Supersymmetric quantum mechanics on arxiv.org 361

Source: http://en.wikipedia.org/wiki/Supersymmetric_quantum_mechanics

Principal Authors: Phys, Anville, Agentsoo, Phil Boswell, Charles Matthews

361 http://xstructure.inr.ac.ru/x-bin/theme2.py?arxiv=hep-th&level=2&index1=19

Supersymmetric quantum mechanics


601

Thermal de Broglie wavelength

In physics, the Thermal de Broglie wavelength is defined for a free ideal gas

FT
of massive particles in equilibrium as:
q
h2
Λ = 2πmkT

where

• h is Planck’s constant
• m is the mass of a gas particle
• k is Boltzmann’s constant
• T is the Temperature of the gas

The thermal de Broglie wavelength is roughly the average de Broglie wave-


length of the gas particles in an ideal gas at the specified temperature. We can
take the average interparticle spacing in the gas to be approximately (V/N) 1/3
where V is the volume and N is the number of particles. When the thermal
A
de Broglie wavelength is much smaller than the interparticle distance, the gas
can be considered to be a classical or Maxwell-Boltzmann gas. On the other
hand, when the thermal de Broglie wavelength is on the order of, or larger than
the interparticle distance, quantum effects will dominate and the gas must be
treated as a →Fermi gas or a →Bose gas, depending on the nature of the gas
particles. The critical temperature is the transition point between these two
regimes, and at this critical temperature, the thermal wavelength will be ap-
proximately equal to the interparticle distance. That is, the quantum nature of
DR
the gas will be evident for
V
N Λ3
≤1

and in this case the gas will obey →Bose-Einstein statistics or →Fermi-Dirac
statistics, whichever is appropriate. On the other hand, for
V
N Λ3
>> 1

the gas will obey →Maxwell-Boltzmann statistics.

Massless particles
For a massless particle, the thermal wavelength may be defined as:
ch
Λ=
2kT π 1/3

Thermal de Broglie wavelength


602

where c is the speed of light. As with the massive thermal wavelength, this
is of the order of the average wavelength of the particles in the gas. This is
derived from the more general definition of the thermal wavelength due to Yan
(Yan 2000) described below.

FT
General definition of the thermal wavelength
A general definition of the thermal wavelength for an ideal quantum gas in
any number of dimensions and for a generalized relationship between energy
and momentum (dispersion relationship) has been given by Yan (Yan 2000). It
is of practical importance, since there are many experimental situations with
different dimensionality and dispersion relationships. If n is the number of
dimensions, and the relationship between energy (E) and momentum (p) is
given by:
E = aps

where a and s are constants, then the thermal wavelength is defined as:
 h i1/n
a 1/s Γ(n/2+1)
Λ = √hπ kT
A Γ(n/s+1)

where Γ is the Gamma function. For example, in the usual case of massive
particles in a 3-D gas we have n=3 , and E=p 2/2m which gives the above
results for massive particles. For massless particles in a 3-D gas, we have n=3 ,
and E=pc which gives the above results for massless particles.

References
DR
• Zijun Yan, "General thermal wavelength and its applications", Eur. J. Phys.
21 (2000) 625-631. http://www.iop.org/EJ/article/0143-0807/21/6/314
/ej0614.pdf

Source: http://en.wikipedia.org/wiki/Thermal_de_Broglie_wavelength

Principal Authors: PAR, Feezo, Charles Matthews, VivaEmilyDavies

Thermal de Broglie wavelength


603

Topological order

In physics, topological order is a new kind of order (a new kind of organi-

FT
zation of particles) in a quantum state that is beyond the Landau symmetry-
breaking description. It cannot be described by local order parameters and long
range correlations. However, topological orders can be described by a new set
of quantum numbers, such as ground state degeneracy, quasiparticle fractional
statistics, edge states, topological entropy, etc. Roughly speaking, topological
order is a non-local quantum entanglement in quantum states. States with
different topological orders can change into each other only through a phase
transition.
A large class of topological orders is realized through a mechanism called
string-net condensation. This class of topological orders is described and clas-
sified by a beautiful mathematical theory — tensor category theory. One finds
that string-net condensation can generate infinitely many different types of
topological orders, which may indicate that there are many different new types
of materials remaining to be discovered.
A
Background
Although all matter is formed by atoms, matter can have very different proper-
ties and appear in very different forms, such as solid, liquid, superfluid, mag-
net, etc. According to condensed matter physics and the principle of emer-
gence, the different properties of materials originate from the different ways in
which the atoms are organized in the materials. Those different organizations
of the atoms (or other particles) are formally called the orders in the materials.
DR
Atoms can organize in many ways which lead to many different orders and
many different types of materials. With so many different orders, we need a
general understanding of the orders. Landau symmetry-breaking theory pro-
vides such a general understanding. It points out that different orders real-
ly correspond to different symmetries in the organizations of the constituent
atoms. As a material changes from one order to another order (i.e., as the ma-
terial undergoes a phase transition), what happens is that the symmetry of the
organization of the atoms changes.
For example, atoms have a random distribution in a liquid, so a liquid remains
the same as we displace it by an arbitrary distance. We say that a liquid has
a continuous translation symmetry. After a phase transition, a liquid can turn
into a crystal. In a crystal, atoms organize into a regular array (a lattice). A
lattice remains unchanged only when we displace it by a particular distance, so
a crystal has only discrete translation symmetry. The phase transition between

Topological order
604

a liquid and a crystal is a transition that changes the continuous translation


symmetry of the liquid to the discrete symmetry of the crystal. Such change in
symmetry is called symmetry breaking. The essence of the difference between
liquids and crystals is therefore that the atoms have different symmetries in the

FT
two phases.
Landau symmetry-breaking theory is a very successful theory. For a long time,
physicists believed that Landau symmetry-breaking theory describes all possi-
ble orders in materials, and all possible (continuous) phase transitions.
However, in last twenty years, it has become more and more apparent that Lan-
dau symmetry-breaking theory may not describe all possible orders. In 1989,
physicists introduced chiral spin state in an attempt to explain high tempera-
ture superconductivity. At first people still wanted to use Landau symmetry-
breaking theory to describe the chiral spin state. They identified the chiral spin
state as a state that breaks the time reversal and parity symmetries, but not the
spin rotation symmetry. However, it was quickly realized that there are many
different chiral spin states that have exactly the same symmetry, so symmetry
alone was not enough to characterize different chiral spin states. This means
that the chiral spin states contain a new kind of order that is beyond symme-
A
try description. This new kind of order was named topological order. A new
quantum number, ground state degeneracy, was introduced to characterize the
different topological orders in chiral spin states. (The name "topological order"
is motivated by the low energy effective theory of the chiral spin states, which
is a topological quantum field theory.)
But experiments soon indicated that chiral spin states do not describe high-
temperature superconductors, and the theory of topological order became a
theory with no experimental realization. However, the similarity between chi-
DR
ral spin states and quantum Hall states allows one to use the theory of topolog-
ical order to describe different quantum Hall states. Just like chiral spin states,
different quantum Hall states all have the same symmetry and are beyond the
Landau symmetry-breaking description. One finds that the different orders in
different quantum Hall states can indeed be described by topological orders,
so the topological order does have experimental realizations.

Applications
The materials described by Landau symmetry-breaking theory have had a sub-
stantial impact on technology. For example, Ferromagnetic materials that break
spin rotation symmetry can be used as the media of digital information storage.
A hard drive made by ferromagnetic materials can store gigabytes of informa-
tion on it. Liquid crystals that break the rotational symmetry of molecules
find wide application in display technology. Nowadays one can hardly find a

Topological order
605

household without a liquid crystal display somewhere in it. Crystals that break
translation symmetry lead to well defined electronic bands which in turn allow
us to make semiconducting devices such as transistors. Topologically ordered
states are a new class of materials that are even richer than symmetry breaking

FT
states. This may suggest an exciting potential for applications.
One theorized application would be to use topologically ordered states as me-
dia for quantum computing. A topologically ordered state is a state with com-
plicated non-local quantum entanglement. The non-locality means that the
quantum entanglement in a topologically ordered state is distributed among
many different particles. As a result, the pattern of quantum entanglements
cannot be destroyed by local perturbations. This significantly reduces the effect
of decoherence. This suggests that if we use different quantum entanglements
in a topologically ordered state to encode quantum information, the informa-
tion may last much longer. The quantum information encoded by the topologi-
cal quantum entanglements can also be manipulated by dragging the topolog-
ical defects around each other. This process may provide a physical appartus
for performing quantum computations. Therefore, topologically ordered states
may provide natural media for both quantum memory and quantum compu-
A
tation. Such realizations of quantum memory and quantum computation may
potentially be made fault tolerant.

Potential impact
Why is topological order important? Landau symmetry-breaking theory is a
cornerstone of condensed matter physics. It used to define the territory of con-
densed matter research. The existence of topological order appears to indicate
that nature is much richer than Landau symmetry-breaking theory has so far
DR
indicated. Some suggest a potential for topological order (or more precisely,
string-net condensation) to provide a unified origin for photons, electrons and
other elementary particles in our universe.

References
Fractional quantum Hall states:

• Two-Dimensional Magnetotransport in the Extreme Quantum Limit, D. C.


Tsui and H. L. Stormer and A. C. Gossard, Phys. Rev. Lett., 48, 1559 (1982)
• Anomalous Quantum Hall Effect: An Incompressible Quantum Fluid with
Fractionally Charged Excitations, R. B. Laughlin, Phys. Rev. Lett., 50, 1395
(1983)

Chiral spin states:

Topological order
606

• Equivalence of the resonating-valence-bond and fractional quantum Hall


states, V. Kalmeyer and R. B. Laughlin, Phys. Rev. Lett., 59, 2095 (1987)
• Chiral Spin States and Superconductivity, Xiao-Gang Wen, F. Wilczek and
A. Zee, Phys. Rev., B39, 11413 (1989)

FT
Topological order:

• Quantum field theory and the Jones polynomial, E. Witten, Comm. Math.
Phys., 121, 351 (1989)
• Vacuum Degeneracy of Chiral Spin State in Compactified Spaces, Xiao-Gang
Wen, Phys. Rev. B, 40, 7387 (1989)
• Topological Orders in Rigid States, Xiao-Gang Wen, Int. J. Mod. Phys., B4,
239 (1990)
• Off-diagonal long-range order, oblique confinement, and the fractional
quantum Hall effect, S. M. Girvin and A. H. MacDonald, Phys. Rev. Lett.,
58, 1252 (1987)
• Effective-Field-Theory Model for the Fractional Quantum Hall Effect, S. C.
Zhang and T. H. Hansson and S. Kivelson, Phys. Rev. Lett., 62, 82 (1989)
• Fractional Statistics and the Quantum Hall Effect, D. Arovas and J. R. Schri-


Aeffer and F. Wilczek, Phys. Rev. Lett., 53, 722 (1984)
Gapless Boundary Excitations in the FQH States and in the Chiral Spin
States, Xiao-Gang Wen, Phys. Rev. B, 43, 11025 (1991)

String-net condensation:

• Photons and electrons as emergent phenomena, Michael A. Levin, Xiao-


Gang Wen, Rev. Mod. Phys., 77, 871 (2005)
DR
• String-net condensation: A physical mechanism for topological phases,
Michael Levin, Xiao-Gang Wen, Phys. Rev. B, 71, 045110 (2005)

Quantum computing:

• Fault-tolerant quantum computation by anyons, A. Yu. Kitaev Ann. Phys.


(N.Y.), 303, 2 (2003)
• Topological quantum computation, Michael H. Freedman, Alexei Kitaev,
Michael J. Larsen, and Zhenghan Wang, Bull. Amer. Math. Soc., 40, 31
(2003)
• Topological quantum memory, Eric Dennis, Alexei Kitaev, Andrew Landahl,
and John Preskill, J. Math. Phys., 43, 4452 (2002)
• Proposed Experiments to probe the Non-Abelian nu=5/2 Quantum Hall
State, Ady Stern and Bertrand I. Halperin, Phys. Rev. Lett., 96, 016802
(2006)

Topological order
607

Source: http://en.wikipedia.org/wiki/Topological_order

FT
Topological quantum number

In physics, a topological quantum number is any quantity, in a physical theo-


ry, that takes on only one of a discrete set of values, due to topological consid-
erations. Most commonly, topological quantum numbers are topological invari-
ants associated with topological defects or soliton-type solutions of some set of
differential equations modeling a physical system, as the solitons themselves
owe their stability to topological considerations. The specific "topological con-
siderations" are usually due to the appearance of the fundamental group or a
higher-dimensional homotopy group in the description of the problem, quite
often because the boundary, one which the boundary conditions are specified,
has a non-trivial homotopy group that is preserved by the differential equa-
tions. The topological quantum number of a solution is sometimes called the
A
winding number of the solution.
Recent ideas about the nature of phase transitions indicates that topological
quantum numbers, and their associated solitons, can be created or destroyed
during a phase transition.

Particle physics
In particle physics, an example is given by the Skyrmion, for which the baryon
number is a topological quantum number. The origin comes from the fact that
DR
the isospin is modelled by SU(2), which is isomorphic to the 3-sphere S3 . By
taking real three-dimensional space, and closing it with a point at infinity, one
also gets a 3-sphere. Solutions to Skyrme’s equations in real three dimensional
space map a point in "real" (physical; Euclidean) space to a point on the 3-
manifold SU(2). Topologically distinct solutions "wrap" the one sphere around
the other, such that one solution, no matter how it is deformed, cannot be
"unwrapped" without creating a discontinuity in the solution. In physics, such
discontinuities are associated with infinite energy, and are thus not allowed.
In the above example, the topological statement is that the 3rd homotopy
group of the three sphere is
π3 (S3 ) = Z

and so the baryon number can only take on integer values.

Topological quantum number


608

A generalization of these ideas is found in the Wess-Zumino-Witten model.

Exactly solvable models


Additional examples can be found in the domain of exactly solvable models,

FT
such as the sine-Gordon equationand the Korteweg-de Vries equation. The one-
dimensional sine-Gordon equation makes for a particularly simple example, as
the fundamental group at play there is
π1 (S1 ) = Z

and so is literally a winding number: a circle can be wrapped around a circle


an integer number of times.

Solid state physics


In solid state physics, certain types of crystalline dislocations, such as screw
dislocations, can be described by topological solitons. An example includes
screw-type dislocations associated with Germanium whiskers.
A
See also
• Inverse scattering transform

Source: http://en.wikipedia.org/wiki/Topological_quantum_number

Principal Authors: Linas, MarSch, Conscious, Charles Matthews, Kymara


DR
Transformation theory (quantum mechan-
ics)

The term transformation theory refers to a procedure used by P. A. M. Dirac


in his early formulation of quantum theory, from around 1927.
The term is related to the famous wave-particle duality, according to which a
particle (a "small" physical object) may display either particle or wave aspects,
depending on the observational situation. Or, indeed, a variety of intermediate
aspects, as the situation demands.
This "transformation" idea also refers to the changes a physical object may
undergo in the course of time, whereby it may "move" between "positions" in
its Hilbert "space".

Transformation theory (quantum mechanics)


609

Remaining in full use today, it would be regarded as a topic in the mathematics


of →Hilbert space, although technically speaking it is somewhat more general
in scope. While the terminology is reminiscent of motion in ordinary space,
the Hilbert space of a quantum object is more general, and holds its entire

FT
quantum state.

Source: http://en.wikipedia.org/wiki/Transformation_theory_%28quantum_mechanics%29

Principal Authors: Charles Matthews

T-symmetry

T-symmetry is the symmetry of physical laws under a time-reversal


transformation—
T : t 7→ −t.
A
The universe is not symmetric under time reversal, although in restricted con-
texts one may find this symmetry. Physicists distinguish time asymmetries that
are intrinsic to the dynamic laws of nature, and those that are due to the initial
conditions of our universe. The T-asymmetry of the weak nuclear force is of
the first kind, while the T-asymmetry of the second law of thermodynamics is
of the second kind.
Physicists also discuss the time-reversal invariance of local and/or macroscopic
descriptions of physical systems, independent of the invariance of the underly-
DR
ing microscopic physical laws. For example, Maxwell’s equations with material
absorption or Newtonian mechanics with friction are not time-reversal invari-
ant at the macroscopic level where they are normally applied, even if they are
invariant at the microscopic level when one includes the atomic motions into
which the "lost" energy is translated.

T-symmetry
610

A FT
Macroscopic phenomena: the second law of ther-
modynamics
Our daily experience shows that T-symmetry does not hold for the behavior
of bulk materials. Of these macroscopic laws, most notable is the second law
DR
of thermodynamics. Many other phenomena, such as the relative motion of
bodies with friction, or viscous motion of fluids, reduce to this, because the
underlying mechanism is the dissipation of usable energy (for example, kinetic
energy) into heat.
Is this time-asymmetric dissipation really inevitable? This question has been
considered by many physicists, often in the context of Maxwell’s demon. The
name comes from a thought experiment described by James Clerk Maxwell in
which a microscopic demon guards a gate between two halves of a room. It
only lets slow molecules into one half, only fast ones into the other. By even-
tually making one side of the room cooler than before and the other, hotter, it
seems to reduce the entropy of the room, and reverse the arrow of time. Many
analyses have been made of this; all show that when the entropy of room and
demon are taken together, this total entropy does increase. Modern analyses
of this problem have taken into account Claude E. Shannon’s relation between

T-symmetry
611

entropy and information. Many interesting results in modern computing are


closely related to this problem— reversible computing, quantum computing
and physical limits to computing, are examples. These seemingly metaphysical
questions are today, in these ways, slowly being converted to the stuff of the

FT
physical sciences.
The consensus nowadays hinges upon the Boltzmann-Shannon identification
of the logarithm of phase space volume with negative of Shannon information,
and hence to entropy. In this notion, a fixed initial state of a macroscopic
system corresponds to relatively low entropy because the coordinates of the
molecules of the body are constrained. As the system evolves in the presence of
dissipation, the molecular coordinates can move into larger volumes of phase
space, becoming more uncertain, and thus leading to increase in entropy.
However, one can equally well imagine a state of the universe in which the
motions of all of the particles at one instant were the reverse (strictly, the CPT
reverse). Such a state would then evolve in reverse, so presumably entropy
would decrease (Loschmidt’s paradox). Why is ’our’ state preferred over the
other?
A
One position is to say that the constant increase of entropy we observe hap-
pens only because of the initial state of our universe. Other possible states of
the universe (for example, a universe at heat death equilibrium) would actu-
ally result in no increase of entropy. In this view, the apparent T-asymmetry of
our universe is a problem in cosmology: why did the universe start with a low
entropy? This view, if it remains viable in the light of future cosmological ob-
servation, would connect this problem to one of the big open questions beyond
the reach of today’s physics— the question of initial conditions of the universe.
DR
Microscopic phenomena: time reversal invariance
Since most systems are asymmetric under time reversal, it is interesting to ask
whether there are any phenomena which do have this symmetry. In classical
mechanics, a velocity v reverses under the operation of T, but an accelera-
tion does not. Therefore, one models dissipative phenomena through terms
which are odd in v. However, delicate experiments in which known sources
of dissipation are removed reveal that the laws of mechanics are time reversal
invariant.
However, the motion of a charged body in a magnetic field, B involves the
velocity through the Lorentz force term v×B, and might seem at first to be
asymmetric under T. A closer look assures us that B also changes sign under
time reversal. This happens because a magnetic field is produced by an electric
current, J, which reverses sign under T. Thus, the motion of classical charged
particles in electromagnetic fields is also time reversal invariant. (Despite this,

T-symmetry
612

it is still useful to consider the time-reversal non-invariance in a local sense


when the external field is held fixed, as when the magneto-optic effect is ana-
lyzed. This allows one to analyze the conditions under which optical phenom-
ena that locally break time-reversal, such as Faraday isolators, can occur.) The

FT
laws of gravity also seem to be time reversal invariant in classical mechanics.
In physics one separates the laws of motion, ie, kinematics, from the laws of
force, called dynamics. Following the classical kinematics of Newton’s laws
of motion, the kinematics of quantum mechanics is built in such a way that
it presupposes nothing about the time reversal symmetry of the dynamics. In
other words, if the dynamics is invariant, then the kinematics will allow it to
remain invariant; if the dynamics is not, then the kinematics will also show
this. The structure of the quantum laws of motion are richer, and we examine
these next.

Time reversal in quantum mechanics


A
DR

This section contains a discussion of the three most important properties of


time reversal in quantum mechanics; namely,

T-symmetry
613

• that it must be represented as an anti-unitary operator,


• that it protects non-degenerate quantum states from having an electric
dipole moment,
• that it has two-dimensional representations with the property T 2 = -1.

FT
The strangeness of this result is clear if one compares it with parity. If parity
transforms a pair of quantum states into each other, then the sum and differ-
ence of these two basis states are states of good parity. Time reversal does not
behave like this. It seems to violate the theorem that all Abelian groups be rep-
resented by one dimensional irreducible representations. The reason it does
this, is that it is represented by an anti-unitary operator. It thus opens the way
to spinors in quantum mechanics.

Anti-unitary representation of time reversal


Eugene Wigner showed that a symmetry operation S of a Hamiltonian is rep-
resented, in quantum mechanics either by an unitary operator, S = U, or an
antiunitary one, S = UK where U is unitary, and K denotes complex conju-
gation. For parity (physics) one has PxP = -x and PpP = -p, where x and p
A
are the position and momentum operators. In canonical quantization, one has
the commutator [x, p] = ih /2π, where h is the Planck’s constant. This com-
mutator is invariant if P is chosen to be unitary, ie, PiP = i. Such an argument
can be attempted for time reversal, T. one has TxT = x and TpT = -p, and the
commutator is invariant only if T is chosen to be anti-unitary, ie, TiT = -i. For
a particle with spin, one can use the representation
T = e−iπSy /~ K,
DR
where S y is the y-component of the spin, to find that TJT = -J.

Electric dipole moments


This has an interesting consequence on the electric dipole moment (EDM) of
any particle. The EDM is defined through the shift in the energy of a state when
it is put in an external electric field: ∆e = d·E + E ·δ·E, where d is called the
EDM and δ, the induced dipole moment. One important property of an EDM
is that the energy shift due to it changes sign under a parity transformation.
However, since d is a vector, its expectation value in a state |ψ> it must be
proportional to <ψ|J |ψ>. Thus, under time reversal, an invariant state must
have vanishing EDM. In other words, a non-vanishing EDM signals both P and
T symmetry-breaking.

T-symmetry
614

It is interesting to examine this argument further, since one feels that some
molecules, such as water, must have EDM irrespective of whether T is a sym-
metry. This is correct: if a quantum system has degenerate ground states which
transform into each other under parity, then time reversal need not be broken

FT
to give EDM.
Experimentally observed bounds on the electric dipole moment of the nucle-
on currently set stringent limits on the violation of time reversal symmetry in
the strong interactions, and their modern theory: quantum chromodynamics.
Then, using the CPT invariance of a relativistic quantum field theory, this puts
strong bounds on strong CP violation.

Kramer’s theorem
For T, which is an anti-unitary Z 2 symmetry generator
T 2 = UKUK = U U * = U (U T ) -1 = Φ,

where Φ is a diagonal matrix of phases. As a result, U = ΦU T and U T = U Φ,


showing that
A U = Φ U Φ.

This means that the entries in Φ are ±1, as a result of which one may have ei-
ther T 2 = ±1. This is specific to the anti-unitarity of T. For an unitary operator,
such as the parity, any phase is allowed.
Next, take a Hamiltonian invariant under T. Let |a> and T |a> be two quantum
states of the same energy. Now, if T 2 = -1, then one finds that the states are or-
thogonal: a result which goes by the name of Kramer’s theorem. This implies
DR
that if T 2 = -1, then there is a two-fold degeneracy in the state. This result
in non-relativistic quantum mechanics presages the spin statistics theorem of
quantum field theory.
→Quantum states which give unitary representations of time reversal, ie, have
T 2=1, are characterized by a multiplicative quantum number, sometimes
called the T-parity.
Time reversal transformation for fermions in quantum field theories can be rep-
resented by an 8-component spinor 362 in which the above mentioned T-parity
can be a complex number with unit radius. The CPT invariance is not a theo-
rem but a better to have propert in these class of theories.

362 http://arxiv.org/abs/hep-th/0010074

T-symmetry
615

Time reversal of the known dynamical laws


Unsolved problems in physics: Why are some of the fundamental forces sym-
metric under time reversal whereas others are not?
The study of particle physics has culminated in a codification of the basic laws

FT
of dynamics into the standard model. This is formulated as a quantum field
theory which has CPT symmetry, ie, the laws are invariant under simultaneous
operation of time reversal, parity and charge conjugation. However, time re-
versal itself is seen not to be a symmetry (this is usually called CP violation).
There are two possible origins of this asymmetry, one through the mixing of
different flavours of quarks in their weak decays, the second through a direct
CP violation in strong interactions. The first is seen in experiments, the second
is strongly constrained by the non-observation of the EDM of a neutron.

See also
• The second law of thermodynamics and Maxwell’s demon (also Loschmidt’s
paradox).
A•


Applications to reversible computing and quantum computing, including
limits to computing.
The standard model of particle physics, CP violation, the CKM matrix and
the strong CP problem
• Neutrino masses, CPT invariance and tests of CPT violation.

References and external links


• Maxwell’s demon: entropy, information, computing, edited by H.S.Leff and
DR
A.F. Rex (IOP publishing, 1990) [ISBN 0750300574]
• Maxwell’s demon, 2: entropy, classical and quantum information, edited by
H.S.Leff and A.F. Rex (IOP publishing, 2003) [ISBN 0750307595]
• The emperor’s new mind: concerning computers, minds, and the laws
of physics, by Roger Penrose (Oxford university press, 2002) [ISBN
0192861980]
• CP violation, by I.I. Bigi and A.I. Sanda (Cambridge University Press, 2000)
[ISBN 0521443490]
• Particle Data Group on CP violation 363
• the Babar 364 experiment in SLAC
• the BELLE 365 experiment in KEK

363 http://pdg.lbl.gov/2004/reviews/cpviolrpp.pdf
364 http://www-public.slac.stanford.edu/babar/
365 http://belle.kek.jp

T-symmetry
616

• the KTeV 366 experiment in Fermilab


• the CPLEAR 367 experiment in CERN

FT
Source: http://en.wikipedia.org/wiki/T-symmetry

Principal Authors: Bambaiah, Roadrunner, Michael Hardy, Pcarbonn, Jheald

Two interfering electron wave-packets

A
DR
Figure 69

The green plane is the x-y-plane, where two (non-interacting) electron wave-
packets meet. The vertical direction shows the real part of ψ (x, y). The semi-
transparent white plane in the top shows the density of detection probability,
i.e, |ψ (x, y) |2 , as blue spots. The blue in the middle is the same again.
Before interfering, both electrons have circular detection probabilities. During
the interference, you can see lines, where there are strong wave movements
and others, with no motion in between. The lines without wave motion are

366 http://kpasa.fnal.gov:8080/public/ktev.html
367 http://cplear.web.cern.ch/cplear/Welcome.html

Two interfering electron wave-packets


617

called knot-lines. As you see, the detection probability is zero on knots. After
the interference, both electrons move as if they never had seen the other one.
The smearing out of electrons is their usual behavior due to dispersion. It is
independent of interference.

FT
See also
• →Double-slit experiment
• →Wave-particle duality
• Light

Source: http://en.wikipedia.org/wiki/Two_interfering_electron_wave-packets

Principal Authors: Linas, Uyanga, SimonP, Conscious, Nikki chan

Ultraviolet catastrophe
A
The ultraviolet catastrophe, also called the Rayleigh-Jeans catastrophe, was
a prediction of early 20th century classical physics that an ideal black body at
thermal equilibrium will emit radiation with infinite power. As experimental
observation showed this to be clearly false, it was one of the first clear indica-
tions of problems with classical physics. The solution to this problem led to the
development of an early form of quantum mechanics.
The term "ultraviolet catastrophe" was first used in 1911 by Paul Ehrenfest,
DR
although the concept goes back to 1905; the word "ultraviolet" refers to the fact
that the problem appears in the short wavelength region of the electromagnetic
spectrum. Since the first appearance of the term, it has also been used for other
predictions of a similar nature, e.g. in quantum electrodynamics (also used in
those cases: ultraviolet divergence).
The ultraviolet catastrophe results from the equipartition theorem of classical
statistical mechanics which states that all modes (degrees of freedom) of a
system at equilibrium have an average energy of kT /2. According to classical
electromagnetism, the number of electromagnetic modes in a 3-dimensional
cavity, per unit frequency, is proportional to the square of the frequency. This
therefore implies that the radiated power per unit frequency should follow
the Rayleigh-Jeans law, and be proportional to frequency squared. Thus, both
the power at a given frequency and the total radiated power go to infinity as
higher and higher frequencies are considered: this is clearly an impossibility, a

Ultraviolet catastrophe
618

point that was made independently by Einstein and by Lord Rayleigh and Sir
James Jeans in the year 1905. Einstein pointed out that the difficulty could be
avoided by making use of a hypothesis put forward five years earlier by Max
Planck. Planck had postulated that electromagnetic energy did not follow the

FT
classical description, but could only oscillate or be emitted in discrete packets
of energy proportional to the frequency (as given by Planck’s law). This has the
effect of reducing the number of possible modes with a given energy at high
frequencies in the cavity described above, and thus the average energy at those
frequencies by application of the equipartition theorem. The radiated power
eventually goes to zero at infinite frequencies, and the total predicted power
is finite. The formula for the radiated power for the idealized system (black
body) was in line with known experiments, and came to be called Planck’s law
of black body radiation. Based on past experiments, Planck was also able to
determine the value of its parameter, now called Planck’s constant. The packets
of energy later came to be called photons, and played a key role in the quantum
description of electromagnetism.
Many popular histories of physics, as well as a number of textbooks, present an
incorrect version of the history of the Ultraviolet Catastrophe. In this version,
A
the "catastrophe" was first noticed by Planck, who developed his formula in re-
sponse. In fact Planck never concerned himself with this aspect of the problem,
because he did not believe that the equipartition theorem was fundamental -
his motivation for introducing "quanta" was entirely different. It was several
years later that physicists realized that Planck’s law resolved a fundamental
crisis of classical physics.

References
DR
• Kroemer, Herbert; Kittel, Charles (1980). Thermal Physics (2nd ed.). W. H.
Freeman Company. ISBN 0716710889. (See Chapter 4)
• Cohen-Tannoudji, Claude; Diu, Bernard; Laloë, Franck (1977). Quantum
Mechanics: Volume One. Hermann, Paris. 624–626
• Kragh, Helge. "Max Planck: The reluctant revolutionary" 368. Physics World.
December 2000

External Links
• Ultraviolet Catastrophe and Schrodinger’s Cat 369

368 http://www.physicsweb.org/articles/world/13/12/8/1
369 http://www.harrymaugans.com/2006/05/03/in-search-of-schrodingers-cat/

Ultraviolet catastrophe
619

Source: http://en.wikipedia.org/wiki/Ultraviolet_catastrophe

Principal Authors: Stevenj, Roadrunner, Jonathan F, Pcarbonn, Rparson, RTC

FT
Uncertainty principle

In quantum physics, the Heisenberg uncertainty principle or just Uncertainty


principle (sometimes also the Heisenberg indeterminacy principle - a name
given to it by Niels Bohr) states that one cannot measure values (with arbitrary
precision) of certain conjugate quantities, which are pairs of observables of a
single elementary particle. These pairs include the position and momentum.
Mathematics provides a positive lower bound for the product of the uncertain-
ties of measurements of the conjugate quantities. The uncertainty principle is
one of the cornerstones of quantum mechanics and was discovered by Werner
Heisenberg in 1927.
The Uncertainty principle follows from the mathematical definition of opera-
tors in quantum mechanics; it is represented by a set of theorems of functional
A
analysis. It is often confused with the observer effect.

Overview
The concept of probability distributions pervades the science of measurement.
Until the beginning of the discovery of quantum physics, it was thought that the
only uncertainty in measurement was caused by the limitations of a measuring
tool’s precision. But it is now understood that no treatment of any scientific
subject, experiment, or measurement is said to be accurate without disclosing
DR
the nature of the probability distribution (sometimes called the error) of the
measurement. Uncertainty is the characterization of the relative narrowness
or broadness of the distribution function applied to a physical observation.
Illustrative of this is an experiment in which a particle is prepared in a def-
inite state and two successive measurements are performed on the particle.
The first one measures the particle’s position and the second immediately after
measures its momentum. Each time the experiment is performed, some value
x is obtained for position and some value p is obtained for momentum. De-
pending upon the precision of the instrument taking the measurements, each
successive measurement of the positions and momenta respectively should be
nearly identical, but in practice they will exhibit some deviation due to con-
straints of measurement using a real world instrument that is not infinitely

Uncertainty principle
620

precise. However, Heisenberg showed that, even in theory with a hypotheti-


cal infinitely precise instrument, no measurement could be made to arbitrary
accuracy of both the position and the momentum of a physical object.
The Heisenberg uncertainty principle (developed in an essay published in

FT
1927) provides a quantitative relationship between the uncertainties of the
hypothetical infinitely precise measurements of p and x as measured by the
sizes of their distributions in the following way: If the particle state is such
that the first measurement yields a dispersion of values ∆x, then the second
measurement will have a distribution of values whose dispersion ∆p is at least
inversely proportional to ∆x. For the limiting case, the constant of proportion-
ality is derivable using commutator arithmetic. It is equal to Planck’s constant
divided by 4π.
This stipulates that the product of the uncertainties in position and momentum
is equal to or greater than about 10−35 joule-seconds. Therefore, the product of
the uncertainties only becomes significant for regimes where the uncertainty in
position or momentum measurements is small. Thus, the uncertainty principle
governs the observable nature of atoms and subatomic particles while its effect
on measurements in the macroscopic world is negligible and can be usually
A
ignored.
The Heisenberg uncertainty relations are a theoretical bound over all mea-
surements. They hold for so-called ideal measurements, sometimes called von
Neumann measurements. They hold even more so for non-ideal or Landau
measurements.

Wave-particle duality and the relationship to the


DR
uncertainty principle
A fundamental consequence of the Heisenberg Uncertainty Principle is that no
physical phenomena can be (to arbitrary accuracy) described as a "classic point
particle" or as a wave but rather the microphysical situation is best described
in terms of wave-particle duality. The uncertainty principle, as initially con-
sidered by Heisenberg, is concerned with cases in which neither the wave nor
the point particle descriptions are fully and exclusively appropriate, such as a
particle in a box with a particular energy value. Such systems are character-
ized neither by one unique "position" (one particular value of distance from
a potential wall) nor by one unique value of momentum (including its direc-
tion). Any observation that determines either a position or a momentum of
such a waveparticle to arbitrary accuracy - known as wavefunction collapse -
is subject to the condition that the width of the wavefunction collapse in po-
sition, multiplied by the width of the wavefunction collapse in momentum, is

Uncertainty principle
621

constrained by the principle to be greater than or equal to Planck’s constant


divided by 4π.
Every measured particle in quantum mechanics exhibits wavelike behaviour
so there is an exact, quantitative analogy between the Heisenberg uncertainty

FT
relations and properties of waves or signals. For example, in a time-varying
signal such as a sound wave, it is meaningless to ask about the frequency spec-
trum at a single moment in time because the measure of frequency is the mea-
sure of a repetition recurring over a period of time. In order to determine the
frequencies accurately, the signal needs to be sampled for a finite (non-zero)
time. This necessarily implies that time precision is lost in favor of a more ac-
curate measurement of the frequency spectrum of a signal. This is analogous to
the relationship between momentum and position, and there is an equivalent
formulation of the uncertainty principle which states that the uncertainty of
energy of a wave (directly proportional to the frequency) is inversely propor-
tional to the uncertainty in time with a constant of proportionality identical to
that for position and momentum.

Common incorrect explanation of the uncertainty principle


A
The uncertainty principle in quantum mechanics is sometimes erroneously ex-
plained by claiming that the measurement of position necessarily disturbs a
particle’s momentum. Heisenberg himself may have initially offered explana-
tions which suggested this view. That this disturbance does not describe the
essence of the uncertainty principle in current theory has been demonstrat-
ed above. The fundamentally non-classical characteristics of the uncertainty
measurements in quantum mechanics were clarified due to the EPR paradox
which arose from Einstein attempting to show flaws in quantum measurements
DR
that used the uncertainty principle. Instead of Einstein succeeding in showing
uncertainty was flawed, Einstein guided researchers to examine more close-
ly what uncertainty measurements meant and led to a more refined under-
standing of uncertainty. Prior to the publication of the EPR paper in 1935, a
measurement was often visualized as a physical disturbance inflicted directly
on the measured system, being sometimes illustrated as a thought experiment
called Heisenberg’s microscope. 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 ex-
planations, which are still encountered in popular expositions of quantum me-
chanics, 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. Heisenberg’s original argument
used the ’old’ quantum theory (namely, the Einstein-deBroglie relations) and
provided a heuristic argument that the position and momentum observables

Uncertainty principle
622

were not simultaneously observable with infinite precision. The more modern
uncertainty relations deal with independent measurements being done on an
ensemble of systems.

FT
Formulation and characteristics
Measurements of position and momentum taken in several identical copies of a
system in a given state will vary according to known probability distributions.
This is the fundamental postulate of quantum mechanics.
If we compute the uncertainty ∆x of the position measurements and the stan-
dard deviation ∆p of the momentum measurements, then
~
∆x∆p ≥ 2

where
~ is the reduced Planck’s constant (Planck’s constant divided by 2π).

Heisenberg did not just use any arbitrary number to describe the minimum
standard deviation between position and momentum of a particle. Heisenberg
A
knew that particles behaved like waves and he knew that the energy of any
wave is the frequency multiplied by Planck’s constant. In a wave, a cycle is
defined by the return from a certain position to the same position such as from
the top of one crest to the next crest. This actually is equivalent to a circle of
360 degrees, or 2π radians. Therefore, dividing h by 2π describes a constant
that when multiplied by the frequency of a wave gives the energy of one radian.
Heisenberg took 1/2 of ~ as his standard deviation. This can be written as ~
over 2 as above or it can be written as h/(4π). Normally one will see ~ over 2
DR
as this is simpler.
Two years earlier in 1925 when Heisenberg had developed his matrix mechan-
ics the difference in position and momentum were already showing up in the
formula. In developing matrix mechanics Heisenberg was measuring ampli-
tudes of position and momentum of particles such as the electron that have a
period of 2π, like a cycle in a wave, which are called Fourier series variables.
When amplitudes of position and momentum are measured and multiplied to-
gether, they give intensity. However, Heisenberg found that when the position
and momentum were multiplied together in that respective order or in the
reverse order, there was a difference between the two calculated intensities
of h/(2π). In other words, the two quantities position and momentum did
not commute. In 1927, to develop the standard deviation for the uncertainty
principle, Heisenberg took the gaussian distribution or bell curve for the im-
precision in the measurement of the position q of a moving electron to the

Uncertainty principle
623

corresponding bell curve of the measured momentum p. That gave the mini-
mum standard deviation to be 1/2 of h/(2π), or, ~/2.
In some treatments, the "uncertainty" of a variable is taken to be the small-
est width of a range which contains 50% of the values, which, in the case of

FT
normally distributed variables, leads to a larger lower bound of h/(2π) for the
product of the uncertainties. Note that this inequality allows for several pos-
sibilities: the state could be such that x can be measured with high precision,
but then p will only approximately be known, or conversely p could be sharply
defined while x cannot be precisely determined. In yet other states, both x and
p can be measured with "reasonable" (but not arbitrarily high) precision.

Common observables which obey the uncertainty principle


An uncertainty relation arises between any two observable quantities that can
be defined by non-commuting operators. This means that the uncertainty prin-
ciple arises in measuring the position and the velocity of an object, or in mea-
suring the position and momentum of an object.

• The most common one is the uncertainty relation between position and
A momentum of a particle in space:
~
∆xi ∆pi ≥ 2

• The uncertainty relation between two orthogonal components of the total


angular momentum operator of a particle is as follows:
~
∆Ji ∆Jj ≥ 2 |hJk i|
DR
where i, j, k are distinct and J i denotes angular momentum along the x i
axis.

One of the theorems


Theorem. For arbitrary symmetric operators A : H → H and B : H → H, and
any element x of H such that A B x and B A x are both defined (so that in
particular, A x and B x are also defined), then
hBAx|xihx|BAxi = hABx|xihx|ABxi = |hBx|Axi|2 ≤ ||Ax||2 ||Bx||2

This is an immediate consequence of the Cauchy-Bunyakovski-Schwarz in-


equality.

Uncertainty principle
624

Consequently, the following general form of the uncertainty principle, first


pointed out in 1930 by Howard Percy Robertson and (independently) by Er-
win Schrödinger, holds:
1
4 |h(AB − BA)x|xi|2 ≤ ||Ax||2 ||Bx||2 .

FT
This inequality is called the Robertson-Schrödinger relation.
The operator A B - B A is called the commutator of A, B and is denoted [A, B ].
It is defined on those x for which A B x and B A x are both defined.
From the Robertson-Schrödinger relation, the following Heisenberg uncer-
tainty relation is immediate:
Suppose A and B are two observables which are identified to self-adjoint (and
in particular symmetric) operators. If B A ψ and A B ψ are defined then

∆ψ A ∆ψ B ≥ 12 h[A, B]iψ

where
hXiψ = hψ|Xψi
A
is the operator mean of observable X in the system state ψ and
q
∆ψ X = hX 2 iψ − hXi2ψ

is the operator standard deviation of observable X in the system state ψ


The above definitions of mean and standard deviation are defined formally
in purely operator-theoretic terms. The statement becomes more meaningful
DR
however, once we note that these actually are the mean and standard deviation
for the measured distribution of values. See quantum statistical mechanics.
It may be evaluated not only for pairs of conjugate operators (e.g. those defin-
ing measurements of distance and of momentum, or of duration and of energy)
but generally for any pair of Hermitian operators. There is also an uncertain-
ty relation between the field strength and the number of particles which is
responsible for the phenomenon of virtual particles.
Note that it is possible to have two non-commuting self-adjoint operators A
and B which share an eigenvector ψ, in this case ψ represents a pure state in
which it is predictable with probability one what the result of measuring A or
B will be in spite of their not being simultaneously measurable.

Uncertainty principle
625

Energy, time and further generalizations


General arguments, connected with the theory of relativity, point out that
seemingly a relation like the following should exist:

FT
∆E∆t ≥ ~.
But its correct mathematical formulation was given 370 only in 1945 by L. I.
Mandelshtam and I. E. Tamm.

History and interpretations


Main article: →Interpretation of quantum mechanics
The Uncertainty Principle was developed as an answer to the question: How
does one measure the location of an electron around a nucleus?
In the summer of 1922 Heisenberg met Niels Bohr, the founding father of
quantum mechanics, and in September 1924 Heisenberg went to Copenhagen,
where Bohr had invited him as a research associate and later as his assistant.
In 1925 Werner Heisenberg laid down the basic principles of a complete quan-
tum mechanics. In his new matrix theory he replaced classical commuting
A
variables with non-commuting ones. Heisenberg’s paper marked a radical de-
parture from previous attempts to solve atomic problems by making use of
observable quantities only. He wrote in a 1925 letter, "My entire meagre efforts
go toward killing off and suitably replacing the concept of the orbital paths
that one cannot observe." Rather than struggle with the complexities of three-
dimensional orbits, Heisenberg dealt with the mechanics of a one-dimensional
vibrating system, an anharmonic oscillator. The result was formulae in which
quantum numbers were related to observable radiation frequencies and inten-
sities. In March 1926, working in Bohr’s institute, Heisenberg formulated the
DR
principle of uncertainty thereby laying the foundation of what became known
as the Copenhagen interpretation of quantum mechanics.
Albert Einstein was not happy with the uncertainty principle, and he challenged
Niels Bohr and Werner Heisenberg with a famous thought experiment (See the
Bohr-Einstein debates for more details): we fill a box with a radioactive materi-
al which randomly emits radiation. The box has a shutter, which is opened and
immediately thereafter shut by a clock at a precise time, thereby allowing some
radiation to escape. So the time is already known with precision. We still want
to measure the conjugate variable energy precisely. Einstein proposed doing
this by weighing the box before and after. The equivalence between mass and
energy from special relativity will allow you to determine precisely how much
energy was left in the box. Bohr countered as follows: should energy leave,

370 http://daarb.narod.ru/mandtamm-eng.html

Uncertainty principle
626

then the now lighter box will rise slightly on the scale. That changes the posi-
tion of the clock. Thus the clock deviates from our stationary reference frame,
and again by special relativity, its measurement of time will be different from
ours, leading to some unavoidable margin of error. In fact, a detailed analysis

FT
shows that the imprecision is correctly given by Heisenberg’s relation.
The term Copenhagen interpretation of quantum mechanics was often used
interchangeably with and as a synonym for Heisenberg’s Uncertainty Principle
by detractors who believed in fate and determinism and saw the common fea-
tures of the Bohr-Heisenberg theories as a threat. Within the widely but not
universally accepted Copenhagen interpretation of quantum mechanics (i.e. it
was not accepted by Einstein or other physicists such as Alfred Lande), the un-
certainty principle is taken to mean that on an elementary level, the physical
universe does not exist in a deterministic form—but rather as a collection of
probabilities, or potentials. For example, the pattern (probability distribution)
produced by millions of photons passing through a diffraction slit can be cal-
culated using quantum mechanics, but the exact path of each photon cannot
be predicted by any known method. The Copenhagen interpretation holds that
it cannot be predicted by any method, not even with theoretically infinitely
A
precise measurements.
It is this interpretation that Einstein was questioning when he said "I cannot be-
lieve that God would choose to play dice with the universe." Bohr, who was one
of the authors of the Copenhagen interpretation responded, "Einstein, don’t tell
God what to do." Niels Bohr himself acknowledged that quantum mechanics
and the uncertainty principle were counter-intuitive when he stated, "Anyone
who is not shocked by quantum theory has not understood a single word."
The basic debate between Einstein and Bohr (including Heisenberg’s Uncer-
DR
tainty Principle) was that Einstein was in essence saying: "Of course, we can
know where something is; we can know the position of a moving particle if we
know every possible detail, and thereby by extension, we can predict where
it will go." Bohr and Heisenberg were saying the opposite: "There is no way
to know where a moving particle is ever even given every possible detail, and
thereby by extension, we can never predict where it will go."
Einstein was convinced that this interpretation was in error. His reasoning
was that all previously known probability distributions arose from determinis-
tic events. The distribution of a flipped coin or a rolled dice can be described
with a probability distribution (50% heads, 50% tails). But this does not mean
that their physical motions are unpredictable. Ordinary mechanics can be used
to calculate exactly how each coin will land, if the forces acting on it are known.
And the heads/tails distribution will still line up with the probability distribu-
tion (given random initial forces).

Uncertainty principle
627

Einstein assumed that there are similar hidden variables in quantum mechanics
which underlie the observed probabilities and that these variables, if known,
would show that there was what Einstein termed "local realism", a description
opposite to the uncertainty principle, being that all objects must already have

FT
their properties before they are observed or measured. For the greater part
of the twentieth century, there were many such hidden variable theories pro-
posed, but in 1964 John Bell theorized the Bell inequality to counter them,
which postulated that although the behavior of an individual particle is ran-
dom, it is also correlated with the behavior of other particles. Therefore, if
the uncertainty principle is the result of some deterministic process in which a
particle has local realism, it must be the case that particles at great distances
instantly transmit information to each other to ensure that the correlations
in behavior between particles occur. The interpretation of Bell’s theorem ex-
plicitly prevents any local hidden variable theory from holding true because it
shows the necessity of a system to describe correlations between objects. The
implication is, if a hidden local variable is the cause of particle 1 being at a
position, then a second hidden local variable would be responsible for particle
2 being in its own position - and there is no system to correlate the behavior
A
between them. Experiments have demonstrated that there is correlation. In the
years following, Bell’s theorem was tested and has held up experimentally time
and time again, and these experiments are in a sense the clearest experimental
confirmation of quantum mechanics. It is worth noting that Bell’s theorem only
applies to local hidden variable theories; non-local hidden variable theories can
still exist (which some, including Bell, think is what can bridge the conceptual
gap between quantum mechanics and the observable world).
Whether Einstein’s view or Heisenberg’s view is true or false is not a directly
DR
empirical matter. One criterion by which we may judge the success of a sci-
entific theory is the explanatory power it gives us, and to date it seems that
Heisenberg’s view has been the better at explaining physical subatomic phe-
nomena.

The uncertainty principle in popular culture


The uncertainty principle is stated in popular culture in many ways, for exam-
ple by stating that it is impossible to know both where an electron is and where
it is going at the same time. This is roughly correct, although it fails to mention
an important part of the Heisenberg principle, which is the quantitative bounds
on the uncertainties.
The uncertainty principle is frequently, but incorrectly, confused with the "ob-
server effect", wherein the observation of an event changes the event. The

Uncertainty principle
628

observer effect is an important effect in many fields, from electronics to psy-


chology and social science.
In some science fiction stories, a device to circumvent the uncertainty principle
is called a Heisenberg compensator, most famously in Star Trek for use on the

FT
transporter; however, it is not clear what compensating means.
In Stephen Donaldson’s Gap Cycle science fiction book series, one of the char-
acters postulates a socio-political version of the uncertainty principle: namely,
that by determining his precise "location" in the current political landscape,
he is prevented from simultaneously calculating the likely direction of political
events in the near future.
In software programming, a Heisenbug is a software error that disappears or
alters its characteristics when it is researched.
The Heisenberg Principle was referenced in "Prophecy", an episode of Stargate
SG-1.

Humor
The unusual nature of Heisenberg’s uncertainty principle, and its distinctive
A
name, has made it the source of several jokes.
It is said that a popular item of graffiti at the physics department of university
campuses is the slogan "Heisenberg may have been here."
In another uncertainty principle joke, a quantum physicist is stopped on the
highway by a police officer who asks "Do you know how fast you were going,
sir?", to which the physicist responds, "No, but I know exactly where I am!".
The biggest flop since the Edsel... The Heisenbergmobile. The problem was
that when you look at the speedometer you got lost.
DR
The character, Phillip Richbourg, of Night Court got into a heated debate with
Judge Harry about this very principle. He argued the "ball theory", claiming
that if you roll a ball down a hill, and you know exactly when it has been
exactly 10sec, and you can measure distance exactly, I don’t see how you can’t
know where the ball is. It was very comical and fed in with the show’s theme.
A reference to the uncertainty principle is made in an animation on a popular
online animations site, albinoblacksheep.com . Jokes with Einstein 5 refers to
a failed experiment to undermine this principle.

See also
• →Quantum indeterminacy
• Basics of quantum mechanics
• Correspondence principle

Uncertainty principle
629

References
Journal articles

FT
• W. Heisenberg, "Über den anschaulichen Inhalt der quantentheoretischen
Kinematik und Mechanik", Zeitschrift für Physik, 43 1927, pp. 172-198.
English translation: J. A. Wheeler and H. Zurek, Quantum Theory and Mea-
surement Princeton Univ. Press, 1983, pp. 62-84.
• L. I. Mandelshtam, I. E. Tamm " The uncertainty relation between energy
and time in nonrelativistic quantum mechanics 371", Izv. Akad. Nauk SSSR
(ser. fiz.) 9, 122-128 (1945). English translation: J. Phys. (USSR) 9,
249-254 (1945).
• G. Folland, A. Sitaram, "The Uncertainty Principle: A Mathematical Survey",
Journal of Fourier Analysis and Applications, 1997 pp 207-238.

External links
• Stanford Encyclopedia of Philosophy entry 372
A•

aip.org: Quantum mechanics 1925-1927 - The uncertainty principle 373
Eric Weisstein’s World of Physics - Uncertainty principle 374
• Schrödinger equation from an exact uncertainty principle 375
• John Baez on the time-energy uncertainty relation 376
• Beating the uncertainty principle in finite-parameter systems 377

Source: http://en.wikipedia.org/wiki/Uncertainty_principle
DR
Principal Authors: CSTAR, Voyajer, Light current, ScienceApologist, Fwappler, Stevenj, Linas, CarlHe-
witt, AxelBoldt

371 http://daarb.narod.ru/mandtamm-eng.html
372 http://plato.stanford.edu/entries/qt-uncertainty/
373 http://www.aip.org/history/heisenberg/p08.htm
374 http://scienceworld.wolfram.com/physics/UncertaintyPrinciple.html
375 http://arxiv.org/abs/quant-ph/0102069
376 http://math.ucr.edu/home/baez/uncertainty.html
377 http://www.seeingwithsound.com/freqtime.htm

Uncertainty principle
630

Unitarity

In mathematics and physics, unitarity is the property of an operator (or a

FT
matrix) that is unitary.
In physics, the requirement of unitarity of the evolution operator or the S-
matrix (the evolution operator from t = −∞ to t = +∞) is essential for the
physical interpretation of any complete theory. These operators must preserve
the squared length of the original vector in the →Hilbert space simply because
the physical interpretation of this squared length, according to the basic prin-
ciples of quantum mechanics, is the total probability of all possible alternatives
of the evolution - and the total probability of all alternatives must always be
equal to 100 percent. (In various approximate theories, non-unitary evolu-
tion operators can represent a nonzero probability of "other" alternatives that
we are not interested in - for example, a reduced total probability of different
states of a kaon represents its decay.)
The evolution operator is automatically unitary if it can be calculated from a
A
Hamiltonian that is Hermitian in a consistent theory.

Source: http://en.wikipedia.org/wiki/Unitarity

Principal Authors: Lumidek, MathMartin, Charles Matthews, Beland, Keenan Pepper

Unitarity bound
DR
In theoretical physics, a unitarity bound is any inequality that follows from
the unitarity of the evolution operator, i.e. from the statement that probabili-
ties are numbers between 0 and 1 whose sum is conserved. Unitarity implies,
among other things, the optical theorem. According to the optical theorem,
the imaginary part of a probability amplitude Im(M) of the forward scatter-
ing is related to the total cross section, up to some numerical factors. Because
|M |2 for the forward scattering process is one of the terms that contributes to
the total cross section, it cannot exceed the total cross section i.e. Im(M). The
inequality
|M |2 ≤ Im(M )

Unitarity bound
631

implies that the complex number M must belong to a certain disk in the com-
plex plane. Similar unitarity bounds imply that the amplitudes and cross sec-
tion can’t increase too much with energy or they must decrease as quickly as a
certain formula dictates.

FT
Source: http://en.wikipedia.org/wiki/Unitarity_bound

Variational method (quantum mechanics)

The variational method is, in quantum mechanics, one way of finding ap-
proximations to the lowest energy eigenstate or ground state. The basis for
this method is the variational principle.

Introduction
Suppose we are given a →Hilbert space and a Hermitian operator over it called
the Hamiltonian H. Ignoring complications about continuous spectra, we look
A
at the discrete spectrum of H and the corresponding eigenspaces of each eigen-
value (see spectral theorem for Hermitian operators for the mathematical
background):
P
1 = λ∈Spec(H) |ψλ ihψλ |

with
hψα |ψβ i = δαβ
DR
and
E
Ĥ|ψλ = λ|ψλ i.

Physical states are normalized, meaning that their norm is equal to 1. Once
again ignoring complications involved with a continuous spectrum of H, sup-
pose it is bounded from below and that its greatest lower bound is E 0. Suppose
also that we know the corresponding state |ψ>. The expectation value of H is
then
P



hψ|H|ψi = λ1 ,λ2 ∈Spec(H) ψ|ψλ1 ψλ1 |H|ψλ2 ψλ2 |ψ

P 2 P 2
= λ∈Spec(H) λ| hψλ |ψi | ≥ λ∈Spec(H) E0 | hψλ |ψi | = E0

Variational method (quantum mechanics)


632

Ansatz
Obviously, if we were to vary over all possible states with norm 1 trying to
minimize the expectation value of H, the lowest value would be E 0 and the cor-
responding state would be an eigenstate of E 0. Varying over the entire Hilbert

FT
space is usually too complicated for physical calculations, and a subspace of the
entire Hilbert space is chosen, parametrized by some (real) differentiable pa-
rameters α i (i =1,2..,N). The choice of the subspace is called the ansatz. Some
choices of ansatzes lead to better approximations than others, therefore the
choice of ansatz is important.
Let’s assume there is some overlap between the ansatz and the ground state
(otherwise, it’s a bad ansatz). We still wish to normalize the ansatz, so we
have the constraints
hψ(αi )|ψ(αi )i = 1

and we wish to minimize


ε(αi ) hψ(αi )|H|ψ(αi )i.
A
This, in general, is not an easy task, since we are looking for a global minimum
and finding the zeroes of the partial derivatives of  over α i is not sufficient. If
ψ (α i) is expressed as a linear combination of other function (α i being the ce-
officients), as in the Ritz method, there is only one minimum and the problem
is straightforward. There are other, non-linear methods, however, such as the
Hartree-Fock method, that are also not characterized by a multitude of minima
and are therefore comfortable in calcualtions.
DR
There is an additional complication in the calculations described. As  tends
toward E 0 in minimization calculations, there is no guarantee that the corre-
sponding trial wavefunctions will tend to the actual wavefunction. This has
been demonstrated by calculations using a modified harmonic oscillator as a
model system, in which an exactly solvable system is approached using the
variational method. A wavefunction different from the exact one is obtained
by use of the method described above.

See also
• Hartree-Fock method
• →Ritz method

Source: http://en.wikipedia.org/wiki/Variational_method_%28quantum_mechanics%29

Variational method (quantum mechanics)


633

Principal Authors: Karol Langner, Vb, Gparker, MarSch, Agentsoo

Variational perturbation theory

FT
In mathematics, variational perturbation theory is a mathematical method to
convert divergent power series in a small expansion parameter, say
s= ∞ n
P
n=0 an g ,

into a convergent series in powers


s= ∞ ω n
P
n=0 bn /(g ) ,

where ω is a critical exponent. This is possible with the help of variational


parameters, which are determined by optimization order by order in g.
Variational perturbation theory is an important mathematical tool in the theory
of critical phenomena. It has led to the most accurate predictions of critical
A
exponents.

See also
• →Variational method (quantum mechanics)

References
DR
• Hagen Kleinert, Path Integrals in Quantum Mechanics, Statistics, Polymer
Physics, and Financial Markets, 3. Auflage, World Scientific (Singapore,
2004) 378 (readable online here 379) (see Chapter 15)
• Hagen Kleinert and Verena Schulte-Frohlinde, Critical Properties of φ 4-
Theories, World Scientific (Singapur, 2001) 380; Paperback ISBN 981-02-
4658-76 (readable online here 381) (see Chapter 19)

Source: http://en.wikipedia.org/wiki/Variational_perturbation_theory

Principal Authors: Nimur, GangofOne, Alektzin, Charles Matthews

378 http://www.worldscibooks.com/physics/5057.html
379 http://www.physik.fu-berlin.de/~kleinert/b5
380 http://www.worldscibooks.com/physics/4733.html
381 http://www.physik.fu-berlin.de/~kleinert/b8

Variational perturbation theory


634

Wavefunction

This article discusses the concept of a wavefunction as it relates to quantum

FT
mechanics. The term has a significantly different meaning when used in the
context of classical mechanics or classical electromagnetism.

Definition
The modern usage of the term wavefunction refers to any vector or function
which describes the state of a physical system. Typically, a wavefunction is
either:

• a complex vector with finitely many components


 
c1
~  .. 
ψ= .
cn
A,

• a complex vector with infinitely many components


 
c1
 .. 
~= .
 
ψ  cn 
 
..
DR
.
,

• or a complex function of one or more real variables (a "continuously in-


dexed" complex vector)

ψ(x1 , . . . xn ).

In all cases, the wavefunction provides a complete description of the associated


physical system.

Wavefunction
635

Interpretation
The physical interpretation of the wavefunction is context dependent. Several
examples are provided below, followed by a detailed discussion of the three
cases described above.

FT
One particle in one spatial dimension
The spatial wavefunction associated with a particle in one dimension is a com-
plex function ψ(x) defined over the real line. The positive function |ψ|2 is
interpreted as the probability density associated with the particle’s position.
That is, the probability of a measurement of the particle’s position yielding a
value in the interval [a, b] is given by
Rb
Pab = a |ψ(x)|2 dx.

This leads to the normalization condition


R∞ 2
−∞ |ψ(x)| dx = 1 .

since, clearly, the probability of the particle being somewhere on the line must
A
be 1.

One particle in three spatial dimensions


The three dimensional case is analogous to the one dimensional case; the wave-
function is a complex function ψ(x, y, z) defined over three dimensional space,
and its complex square is interpreted as a three dimensional probability density
function:
PR = R |ψ(x, y, z)|2 dV
R
DR
The normalization condition is likewise
|ψ(x, y, z)|2 dV = 1
R

where the preceding integral is taken over all space.

Two distinguishable particles in three spatial dimensions


In this case the wavefunction is a complex function of six spatial variables,
ψ(x1 , y1 , z1 , x2 , y2 , z2 ) , and |ψ|2 is the joint probability density associated with
the positions of both particles. Thus the probability that a measurement of the
positions of both particles indicates particle one is in region R and particle two
is region S is

Wavefunction
636

|ψ|2 dV2 dV1


R R
PR,S = R S

where dV1 = dx1 dy1 dz1 , and similarly for dV2 .


The normalization condition is then:

FT
|ψ(x, y, z)|2 dV2 dV1 = 1
R

where the preceding integral is taken over the full range of all six variables.
Given a wave function of ψ of a systems consisting of two (or more) particles,
it is in general not possible to assign a definite wavefuction to a single-particle
subsystem. In other words, the particles in the system can be entangled.

One particle in one dimensional momentum space


The wavefunction for a one dimensional particle in momentum space is a com-
plex function ψ(p) defined over the real line. The quantity |ψ|2 is interpreted
as a probability density function in momentum space:
Rb
Pab = a |ψ(p)|2 dp
A
As in the position space case, this leads to the normalization condition:
R∞ 2
−∞ |ψ(p)| dp = 1.

Spin 1/2
The wavefunction for a spin 1/2 particle (ignoring its spatial degrees of free-
dom) is acolumn vector
~ = c1
DR
ψ
c2
.

The meaning of the vector’s components depends on the basis, but typically
c1 and c2 are respectively the coefficients of spin up and spin down in the z
direction. In Dirac notation this is:
|ψi = c1 | ↑z i + c2 | ↓z i

The values |c1 |2 and |c2 |2 are then respectively interpreted as the probability
of obtaining spin up or spin down in the z direction when a measurement of
the particle’s spin is performed. This leads to the normalization condition
|c1 |2 + |c2 |2 = 1 .

Wavefunction
637

Interpretation
A wavefunction describes the state of a physical system by expanding it in terms
of other states of the same system. We shall denote the state of the system
under consideration as |ψi and the states into which it is being expanded as

FT
|φi i. Collectively the latter are referred to as a basis or representation. In what
follows, all wavefunctions are assumed to be normalized.

Finite vectors
A wavefunction which is a vector ψ ~ with n components describes how to ex-
press the state of the physical system |ψi as the linear combination of finitely
many basis elements |φi i, where i runs from 1 to n. In particular the equation
 
c1
~  .. 
ψ= .
cn
,
A
which is a relation between column vectors, is equivalent to
|ψi = ni=1 ci |φi i,
P

which is a relation between the states of a physical system. Note that to pass
between these expressions one must know the basis in use, and hence, two col-
umn vectors with the same components can represent two different states of a
system if their associated basis states are different. An example of a wavefunc-
tion which is a finite vector is furnished by the spin state of a spin-1/2 particle,
DR
as described above.
The physical meaning of the components of ψ ~ is given by the wavefunction
collapse postulate:
If the states |φi i have distinct, definite values, λi , of some dynamical vari-
able (e.g. momentum, position, etc) and a measurement of that variable is
performed on a system in the state

P
|ψi = i ci |φi i

then the probability of measuring λi is |ci |2 , and if the measurement yields


λi , the system is left in the state |φi i.

Wavefunction
638

Infinite vectors
The case of an infinite vector with a discrete index is treated in the same man-
ner a finite vector, except the sum is extended over all the basis elements.
Hence 

FT
c1
 .. 
~= .
 
ψ  cn 
 
..
.

is equivalent to
P
|ψi = i ci |ψi i,

where it is understood that the above sum includes all the components of ψ. ~
The interpretation of the components is the same as the finite case (apply the
collapse postulate).
A
Continuously indexed vectors (functions)
In the case of a continuous index, the sum is replaced by an integral; an ex-
ample of this is the spatial wavefunction of a particle in one dimension, which
expands the physical state of the particle, |ψi, in terms of states with definite
position, |xi. Thus
R∞
|ψi = −∞ ψ(x)|xi dx.
DR
Note that |ψi is not the same as ψ(x) . The former is the actual state of the
particle, whereas the latter is simply a wavefunction describing how to express
the former as a superposition of states with definite position. In this case the
base states themselves can be expressed as
R∞
|x0 i = −∞ δ(x − x0 )|xi dx

and hence the spatial wavefunction associated with |x0 i is δ(x − x0 ) .

Formalism
Given an isolated physical system, the allowed states of this system (i.e. the
states the system could occupy without violating the laws of physics) are part
of a →Hilbert space H. Some properties of such a space are
1. If |ψi and |φi are two allowed states, then

Wavefunction
639

a|ψi + b|φi

is also an allowed state, provided |a|2 + |b|2 = 1. (This condition is due to


normalisation.)

FT
2. There is always an orthonormal basis of allowed states of the vector
space H.

The wavefunction associated with a particular state may be seen as an expan-


sion of the state in a basis of H. For example,
{| ↑z i, | ↓z i}

is a basis for the space associated with the spin of a spin-1/2 particle and
consequently the spin state of any such particle can be written uniquely as
a| ↑z i + b| ↓z i.
A
Sometimes it is useful to expand the state of a physical system in terms of states
which are not allowed, and hence, not in H. An example of this is the spacial
wavefunction associated with a particle in one dimension which expands the
state of the particle in terms of states with definite position.
Every Hilbert space H is equipped with an inner product. Physically, the nature
of the inner product is contingent upon the kind of basis in use. When the basis
is a countable set {|φi i} , and orthonormal, i.e.
hφi |φj i = δij .
DR
Then an arbitrary vector |ψi can be expressed as
P
|ψi = i ci |φi i

where ci = hφi |ψi.


If one chooses a "continuous" basis as, for example, the position or coordi-
nate basis consisting of all states of definite position {|xi}, the orthonormality
condition holds similarly:
hx|x0 i = δ(x − x0 ).

We have the analogous identity


hx| ψ(x0 )|x0 i dx0 = ψ(x0 )δ(x − x0 ) dx0 = ψ(x).
R R

Wavefunction
640

See also
• →Wave packet
• Boson - particles with symmetric wavefunction under permutation (i.e.

FT
switching positions)
• Fermion - particles with antisymmetric wavefunction under permutation
• →Quantum mechanics
• →Schrödinger equation
• →Normalisable wavefunction

References
• Griffiths, David J. (2004). Introduction to Quantum Mechanics (2nd ed.).
Prentice Hall. ISBN 013805326X.

Source: http://en.wikipedia.org/wiki/Wavefunction

Principal Authors: CSTAR, Joshua Barr, CygnusPius, Oleg Alexandrov, Dhc529


A
Wave packet
The neutrality of this article is disputed.
Please see the discussion on the talk page.

In physics, a wave packet is an envelope or packet containing an arbitrary


number of wave forms. In quantum mechanics the wave packet is ascribed a
DR
special significance: it is interpreted to be a "probability wave" describing the
probability that a particle or particles in a particular state will have a given
position and momentum.
By applying the →Schrödinger equation in quantum mechanics it is possible to
deduce the time evolution of a system, similar to the process of the Hamiltonian
formalism in classical mechanics. The wave packet is a mathematical solution
to the Schrödinger equation. The square of the area under the wave packet
solution is interpreted to be the probability density of finding the particle in a
region.
In the coordinate representation of the wave (such as the Cartesian coordi-
nate system) the position of the wave is given by the position of the packet.
Moreover, the narrower the wave packet, and therefore the better defined the
position of the wave packet, the larger the uncertainty in the momentum of the
wave. This tradeoff is known as the Heisenberg uncertainty principle.

Wave packet
641

Background
In the early 1900s it became apparent that classical mechanics had some major
failings. Isaac Newton originally proposed the idea that light came in discrete
packets which he called "corpuscles", but the wave-like behavior of many light

FT
phenomena quickly led scientists to favor a wave description of electromag-
netism. It wasn’t until the 1930s that the particle nature of light really began
to be widely accepted in physics. The development of quantum mechanics
— and its success at explaining confusing experimental results — was at the
foundation of this acceptance.
One of the most important concepts in the formulation of quantum mechanics
is the idea that light comes in discrete bundles called photons. The energy of
light is a discrete function of frequency:
E = nhf

The energy is an integer, n, multiple of Planck’s constant, h, and frequency, f.


This resolved a significant problem in classical physics, called the ultraviolet
catastrophe.
A
The ideas of quantum mechanics continued to be developed throughout the
20th century. The picture that was developed was of a particulate world, with
all phenomena and matter made of and interacting with discrete particles;
however, these particles were described by a probability wave. The interac-
tions, locations, and all of physics would be reduced to the calculations of
these probability amplitude waves. The particle-like nature of the world was
significantly confirmed by experiment, while the wave-like phenomena could
be characterized as consequences of the wave packet nature of particles.
DR
Mathematics of wave packets
As an example, consider wave solutions to the following wave equation:
∂2u
∂t2
= c2 ∇2 u

where c is the speed of the wave’s propagation in a given medium. The wave
equation has plane-wave solutions
u(x, t) = eik·x−iωt

ω
where |k| = c.
To simplify, consider only waves propagating in one dimension. Then the gen-
eral solution is

Wave packet
642

u(x, t) = Aeikx−iωt + Be−ikx−iωt

A wave packet is a localized disturbance that results from the sum of many
different wave forms. If the packet is strongly localized, more frequencies are

FT
needed to allow the constructive superposition in the region of localization and
destructive superposition outside the region. From the basic solutions in the
one dimension, a general form of a wave packet can be expressed as
R∞
f (x, t) = √1 −∞ A(k)eikx−iω(k)t dk.


The factor 1/ 2π comes from Fourier transform conventions. The amplitude
A(k) contains the coefficients of the linear superposition of the plane wave
solutions. These coefficients can in turn be expressed as a function of f (x, t)
evaluated at t = 0:
R∞
A(k) = √1 −∞ f (x, 0)e−ikx dk.

Quantum mechanical waves


A
A quantum mechanical wave in its most salient and simple form is a solution
to a differential equation. It is a bridge that provides mathematical insight into
physical problems. The solutions of these mathematical models are postulated
in quantum mechanics to provide the possible or observable outcomes for any
experiment. Unfortunately, the wave packet combined with the probabilistic
nature of measurement leads to some peculiarities.
First, the solutions do not provide answers about single experiments. These
solutions can only be confirmed by measurement, and measurement is proba-
DR
bilistic in nature. Experimental confirmation of a prediction of quantum me-
chanics is only given in the outcomes of repeated similar experiments. The
first mistake that many people make in thinking about quantum mechanical
predictions is in thinking about only single experiments. In reality, quantum
mechanics has little to say about the observations made in single experiments
unless the system is prepared in an energy eigenstate.
Next, the quantum mechanical wave is a representation of a particle. The wave
carries the information about particle position and momentum and also any
other observable that can be derived from position and momentum. However,
there is no reason to believe that a quantum wave actually is a particle. In the
words of Dirac, the wave expresses information. A quantum mechanical wave
can be nothing more than a mathematical model.

Wave packet
643

Nevertheless, it is reasonable to think about experiments in terms of quantum


mechanical intuition. It is this blurring of the line between mathematical mod-
eling and the quantum picture of the world that so often leads to confusion.
For the purposes of the uninitiated it would be much safer to consider only

FT
the model, and leave the intuition to those who are better acquainted with the
mathematical intricacies of quantum mechanics.

Superposition
One of the most common classes of problems discussed in a quantum mechan-
ics are interference phenomena. These interference phenomena apparently
arise from the self-interaction of particles and the wave-like nature of these
interactions. Such self-interaction is enabled by the principle of superposition.
A particle does not negotiate any single path through a diffraction grating; its
probability wave actually coincidently traverses all possible paths. Ultimately it
is the act of measurement that collapses the wave packet to the single observed
outcome.

The collapse of the wave packet


A
The superposition principle of quantum mechanics allows any solution to the
Schrödinger equation to be composed of a linear combination of any number
of possible states in a complete set of commuting observables. The act of mea-
surement typically collapses these superimposed states into a single outcome.
For instance, the state of an electron passing through a double slit is most cor-
rectly described as a combination of each of the possible individual paths it
might take, resulting in the famous double slit diffraction pattern. If, howev-
er, a measurement is made at the slits, then the double slit diffraction pattern
DR
disappears because the electron now travels through only a single slit.
Quantum mechanics places no constraints on how we interpret the collapse of
the wave packet. We might say that during the act of observation the electron
suddenly and probabilistically jumps into the state we measure. On the other
hand we might suppose, as many early physicists, including Albert Einstein,
erroneously did, that the particle was in the observed state all along. This
is the classical interpretation and does not agree with experimental evidence.
Because none of the modern interpretations can be falsified experimentally,
quantum mechanics provides no way of knowing how or why the wave packet
collapsed, and what if any significance the event holds.
Pointedly, Dirac intimates that this question has less significance than many
people assume. The importance of quantum mechanics lies in the predictions
that it makes, and the abject success of those predictions, culminating in nearly
overwhelming experimental confirmation of the theory. The fact that we don’t

Wave packet
644

really know how to interpret the collapse of the wave packet at this time is
irrelevant, what is important for physics is that the theory works at all.

Metaphysical claims

FT
Many discussions about the collapse of the wave packet and quantum super-
position occur in metaphysical and speculative fiction circles, primarily based
on attempts to draw analogies between the language used in quantum physics
with the world of the layman. Common themes are the many-worlds interpre-
tation and the related multiverse, proof of the existence of God, teleportation,
faster-than-light travel, and proof of human super-consciousness.
Generally, like most fiction, these themes sacrifice scientific rigor in favor of
evocative concepts, as the physics of quantum mechanics is not easily described
relative to human experience.

References
• Jackson, J.D. (1975). Classical Electrodynamics (2nd Ed.). New York: John
Wiley & Sons, Inc. ISBN 0-471-43132-X
A
Source: http://en.wikipedia.org/wiki/Wave_packet

Principal Authors: John187, Laurascudder, Art LaPella, CLW, Hansm

Wave-particle duality
DR
In physics, wave-particle duality holds that light and matter exhibit properties
of both waves and of particles. It is a central concept of quantum mechanics.
The idea is rooted in a debate over the nature of light and matter dating back
to the 1600s, when competing theories of light were proposed by Christiaan
Huygens and Isaac Newton.
The photon was the first entity that was seen to exhibit these dualistic proper-
ties. And so wave-particle duality is often stated like this: "A photon sometimes
acts like a wave, and sometimes acts like a particle, but not at the same time."
However, this is slightly misleading, because a photon always acts like both to
varying degrees. For example, when shooting single photons through a slit, a
detector can detect each photon when it hits a photosensitive screen (its posi-
tion is recorded) - but over time, the detector will detect the same diffraction

Wave-particle duality
645

pattern as it would if the photons were given off all in one burst. This is be-
cause any given trajectory the photon could take has a certain probability that
is dictated by the properties of an electromagnetic wave.
Once it was realised that all particles exhibit wave-particle duality this lead

FT
rapidly into the development of "new" quantum mechanics, superseding the
old Bohr atomic planetary model. The new quantum mechanics incorporated
wave-particle duality into the core of the formalism, where it remains to this
day. Through the work of Albert Einstein, Louis de Broglie, Arthur Compton
and many others, it is now accepted that all objects have both wave and particle
nature (though this phenomenon is only detectable on small scales, such as
with atoms), and that quantum mechanics provides the over-arching theory
resolving this paradox.

History
At the close of the 19th century, the case for atomic theory, that matter was
made of particulate objects or atoms, was well established. Electricity, first
thought to be a fluid, was understood to consist of particles called electrons,
as demonstrated by J.J. Thomson by his research into the work of Rutherford,
A
who had investigated using cathode rays that an electrical charge would actu-
ally travel across a vacuum from cathode to anode. In brief, it was understood
that much of nature was made of particles. At the same time, waves were well
understood, together with wave phenomena such as diffraction and interfer-
ence. Light was believed to be a wave, as Thomas Young’s double-slit experi-
ment and effects such as Fraunhofer diffraction had clearly demonstrated the
wave-like nature of light.
DR
But as the 20th century turned, problems had emerged with this viewpoint.
The photoelectric effect, as analyzed in 1905 by Albert Einstein, demonstrated
that light also possessed particle-like properties, further confirmed with the
discovery of the Compton effect in 1923. Later on, the diffraction of electrons
would be predicted and experimentally confirmed, thus showing that electrons
must have wave-like properties in addition to particle properties.
This confusion over particle versus wave properties was eventually resolved
with the advent and establishment of quantum mechanics in the first half of
the 20th century, which ultimately explained wave-particle duality. It pro-
vided a single unified theoretical framework for understanding that all matter
can behave in both a wave-like and a particle-like fashion in the appropriate
circumstances. Quantum mechanics holds that every particle in nature, be it
a photon, electron or atom, is described by a solution to a differential equa-
tion, most typically, the Schroedinger equation. The solutions to this equation
are known as wave functions, as they are inherently wave-like in their form.

Wave-particle duality
646

They can diffract and interfere, leading to the wave-like phenomena that are
observed. Yet also, the wave functions are interpreted as describing the proba-
bility of finding a particle at a given point in space. Thus, if one is looking for
a particle, one will find one, with a probability density given by the square of

FT
the magnitude of the wave function.
One does not observe the wave-like quality of everyday objects because the as-
sociated wavelengths of people-sized objects are exceedingly small. The wave-
length is given essentially as the inverse of the size of the object, with the factor
given by Planck’s constant h, an extremely small number.

Huygens and Newton; earliest theories of light


The earliest comprehensive theory of light was advanced by Christiaan Huy-
gens, who proposed a wave theory of light, and in particular demonstrated
how waves might interfere to form a wave-front, propagating in a straight line.
However, the theory had difficulties in other matters, and was soon overshad-
owed by Isaac Newton’s corpuscular theory of light. That is, Newton proposed
that light consisted of small particles, with which he could easily explain the
phenomenon of reflection. With considerably more difficulty, he could also ex-
A
plain refraction through a lens, and the splitting of sunlight into a rainbow by
a prism.
Because of Newton’s immense intellectual stature, his theory went essentially
unchallenged for over a century, with Huygens’ theories all but forgotten. With
the discovery of diffraction in the early 19th century, the wave theory was
revived, and so by the advent of the 20th century, a scientific debate over
waves vs. particles had already been thriving for a very long time.
DR
Fresnel, Maxwell, and Young
In the early 1800s, the double-slit experiments by Young and Fresnel provided
evidence for Huygens’ theories: these experiments showed that when light is
sent through a grid, a characteristic interference pattern is observed, very sim-
ilar to the pattern resulting from the interference of water waves; the wave-
length of light can be computed from such patterns. Maxwell, during the late-
1800s, explained light as the propagation of electromagnetic waves with the
Maxwell equations. These equations were verified by experiment, and Huy-
gens’ view became widely accepted.

Wave-particle duality
647

Einstein and photons


In 1905, Albert Einstein provided a remarkable explanation of the photoelectric
effect, a hitherto troubling experiment which the wave theory of light seemed
incapable of explaining. He did so by postulating the existence of photons,

FT
quanta of light energy with particulate qualities.
In the photoelectric effect, it was observed that shining a light on certain met-
als would lead to an electric current in a circuit. Presumably, the light was
knocking electrons out of the metal, causing them to flow. However, it was
also observed that while a dim blue light was enough to cause a current, even
the strongest, brightest red light caused no current at all. According to wave
theory, the strength or amplitude of a light wave was in proportion to its bright-
ness: a bright light should have been plenty strong enough to create a large
current. Yet, oddly, this was not so.
Einstein explained this conundrum by postulating that the electrons were
knocked free of the metal by incident photons, with each photon carrying an
amount of energy E that was related to the frequency, ν of the light by
E = hν ,
A
where h is Planck’s constant (6.626 x 10 -34 J seconds). Only photons of a high-
enough frequency, (above a certain threshold value) could knock an electron
free. For example blue light, but not red light, had sufficient energy to free
an electron from the metal. More intense light above the threshold frequen-
cy could release more electrons, but no amount of light below the threshold
frequency could release an electron.
DR
Einstein was awarded the Nobel Prize in Physics in 1921 for his theory of the
photoelectric effect.

De Broglie
In 1924, Louis-Victor de Broglie formulated the de Broglie hypothesis, claiming
that all matter has a wave-like nature; he related wavelength, (lambda), and
momentum, p:
h
λ= p

This is a generalization of Einstein’s equation above since the momentum of a


photon is given by p = E / c where c is the speed of light in vacuum, and = c
/ ν.
De Broglie’s formula was confirmed three years later for electrons (which have
a rest-mass) with the observation of electron diffraction in two independent

Wave-particle duality
648

experiments. At the University of Aberdeen, George Paget Thomson passed a


beam of electrons through a thin metal film and observed the predicted in-
terference patterns. At Bell Labs Clinton Joseph Davisson and Lester Halbert
Germer guided their beam through a crystalline grid.

FT
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.

Wave nature of large objects


Similar experiments have since been conducted with neutrons and protons.
Among the most famous experiments are those of Estermann and Otto Stern in
1929. Authors of similar recent experiments with atoms and molecules claim
that these larger particles also act like waves.
A dramatic series of experiments emphasizing the action of gravity in relation
to wave-particle duality were conducted in the 1970’s using the neutron in-
terferometer. Neutrons, subatomic particles in atomic nuclei, provide much of
the mass of a nucleus and thus of ordinary matter. Neutrons are fermions, and
A
thus obey the Pauli Exclusion Principle. In the neutron interferometer, they act
as quantum-mechanical waves directly subject to the force of gravity. While
the results were not surprising since gravity was known to act on everything
- even deflecting light on large scales and acting on photons as well on small-
er scales (the Pound-Rebka falling photon experiment), the self-interference of
the quantum mechanical wave of a massive fermion in a gravitational field had
never been experimentally confirmed before.
In 1999, the diffraction of C 60 fullerenes by researchers from the Univer-
DR
sity of Vienna was reported 1. Fullerenes are rather large and massive ob-
jects, having an atomic mass of about 720. The de Broglie wavelength is
2.5 picometers, whereas the diameter of the molecule is about 1 nanometer,
i.e. about 400 times larger. As of 2005, this is the largest object for which
quantum-mechanical wave-like properties have been directly observed in far-
field diffraction. The experimenters have assumed the arguments of wave-
particle duality and have assumed the validity of de Broglie’s equation in their
argument. In 2003 the Vienna group has meanwhile also demonstrated the
wave-nature of tetraphenylporphyrin 4 - a flat biodye with an extension of about
2 nm and a mass of 614 amu. For this demonstration they employed a near-
field Talbot Lau interferometer 2,3. In the same interferometer they also found
interference fringes for C60F48, a fluorinated buckyball with a mass of about
1600 amu, composed of 108 atoms 4. Large molecules are already so complex
that they give experimental access to some aspects of the quantum-classical
interface, i.e. to certain decoherence mechanisms 5,6.

Wave-particle duality
649

Whether objects heavier than the Planck mass (about the weight of a large bac-
terium) have a de Broglie wavelength is theoretically unclear and experimen-
tally unreachable. The wavelength would be smaller than the Planck length,
a scale at which current theories of physics may break down or need to be

FT
replaced by more general ones.

Theoretical sketch and remarks on philosophical in-


quiry
The wave-particle paradox is resolved in the theoretical framework of quantum
mechanics. This framework is deep and complex and therefore impossible to
adequately summarize in brief.
Every particle in nature can be described as a superposition of solutions to
a differential equation. The most basic is the Schroedinger equation, but this
does not include any relativistic effects and is not generally realistic. The Klein-
Gordon Equation is a relativistic version of the Schroedinger equation that is
applicable to spin-0 particles, and the Dirac Equation is the relativistic version
of Schroedinger’s equation for spin-1/2 particles. The solutions to these equa-
A
tions contain oscillatory mathematical components and are hence inherently
wave-like in nature. In practice these are referred to as wave functions and
they can describe diffractions, interferance with one another or themselves,
and otherwise accurately predict observed wave-like phenomena such as is de-
scribed in the double-slit experiment.
Wave functions are often interpreted as describing the probability of finding
their corresponding particle at a given point in space at a given time. For
example, upon setting up an experiment involving a moving particle, one can
DR
’look’ for that particle to arrive at some particular location using a detection ap-
paratus set up at that location. While quantum behavior follows well-defined
deterministic equations (such as the wave function), the solutions to these
equations are probabilistic. The probability of the detector detecting the parti-
cle is calculated by taking the integral of the product of the wave function and
its complex conjugate. While the wave function can be thought of as smeared
out in space, in practice the detector will always either *see* or *not see* the
entire particle in question; it will never see a fractional piece of the particle, like
two-thirds of an electron. Hence the strange duality: The particle propagates
in space in a distributed, probabilistic wavelike fashion but arrives at a detector
as a localized, complete corpuscle. This paradoxical conceptual framework has
some explanations in the forms of the Copenhagen interpretation, Path Integral
Formulation, or the Many Worlds Interpretation. It is important to realize that
all of these interpretations are equivalent and result in the same predictions
even though they offer widely different philosophical interpretations.

Wave-particle duality
650

A more mathematically concise formulation of the wave function is to treat


it as a ket. Essentially each quantum object can be described by making use
of an infinite dimensional →Hilbert space. The wave functions, or kets reside
in the Hilbert space, and are eigenfunctions of an eigenequation in which an

FT
operator acts on the ket to return the ket multiplied by a scalar eigenvalue. In
this sense, the quantum system is neither a particle nor a wave. It is described
by an abstract ket that will appear to behave as one or the other depending on
what kind of observation you are making at the time.
While quantum mechanics makes astoundingly accurate predictions about the
outcomes of such experiments, its philosophical meaning is still sought after
and debated. This debate has evolved as a broadening of the original strug-
gles to comprehend wave-particle duality. What does it mean for a proton, to
behave both as a particle and as a wave? How can an antimatter electron be
mathematically equivalent to a regular electron moving backwards in time un-
der certain circumstances, and what implications does this have for our experi-
ence of time as one-directional? How can a particle seemingly teleport through
a barrier while soccer balls regularly fail to pass through cement walls? The im-
plications of these facets of quantum mechanics continue to puzzle many who
A
delve into the subject. The discussion currently can be investigated further
under headings of local realism and quantum measurement. As to the assump-
tion of (nonlocal) hidden variables in a hypothesized sub-quantum domain,
see Bohm interpretation, or quantum cybernetics, respectively.
In John Cramer’s transactional interpretation of QM, the probabilistic or parti-
cle aspect of matter is de-emphasized, in favor of a more comprehensive use of
waves to explain all the same phenomena. This approach is also used in Carver
Mead’s Collective Electrodynamics approach to quantum electron phenomena.
DR
Whether these approaches lead to verifiably distinct physical predictions re-
mains to be seen, but they at least provide an alternative philosophy to the
Copenhagen interpretation, a way to understand QM and entanglement with-
out Einstein’s so-called "spooky action at a distance". They substitute instead
the non-causality of an advanced wave to allow a mutual wave transaction
between events at zero interval in spacetime, as in Wheeler–Feynman electro-
dynamics.
Some physicists intimately associated with the historical struggle to arrive at
the rules of quantum mechanics have viewed these philosophical debates on
wave-particle duality and related matters as attempts to impose human expe-
rience on the quantum (microscopic) world. Since by its nature this world is
completely non-intuitive, quantum theory (they would assert) must be learned
on its own terms independent of experience-based human intuition. The sci-
entific merit of searching too deeply for a ’meaning’ to quantum mechanics

Wave-particle duality
651

is thereby suspect; Bell’s theorem and experiments it inspires provide a good


example of such testing of the foundations of quantum mechanics. From a
physics viewpoint, the inability of a new quantum philosophy to satisfy the
testability criterion or alternatively the inability to find a flaw in the predictive

FT
power of the existing theory reduces to a null proposition, perhaps even risking
degeneration into pseudoscience.

Applications
Wave-particle duality is exploited in electron microscopy, where the small
wavelengths associated with the electron can be used to view objects much
smaller than what is visible using visible light.

See also
• This entry is still rather rudimentary. It would be very desirable to add a
information on the huge amount of literature on atomic beam interferome-
try, Bose Einstein Condensation, decoherence experiments as well as inter-
esting papers on diffraction of cold molecules and several recent electron
A interference studies.
• Afshar experiment
• Arago spot
• Hanbury-Brown and Twiss effect
• →Scattering theory

References
DR
• Note 1: Arndt, Markus, O. Nairz, J. Voss-Andreae, C. Keller, G. van der
Zouw, A. Zeilinger (14 October 1999). "Wave-particle duality of C60" 382.
Nature 401: 680-682.
• Note 2: Clauser, John F., S. Li (1994). "Talbot von Lau interefometry with
cold slow potassium atoms." 383. Phys. Rev. A 49: R2213-17.
• Note 3: Brezger, Björn, Lucia Hackermüller, Stefan Uttenthaler, Julia
Petschinka, Markus Arndt and Anton Zeilinger (2002). "Matter-wave in-
terferometer for large molecules" 384. Phys. Rev. Lett. 88: 100404.
• Note 4: Hackermüller, Lucia, Stefan Uttenthaler, Klaus Hornberger, Elisa-
beth Reiger, Björn Brezger, Anton Zeilinger and Markus Arndt (2003). "The

382 http://www.nature.com/cgi-taf/DynaPage.taf?file=/nature/journal/v401/n6754/full/401680a0_fs.

html&content_filetype=pdf
383 http://ojps.aip.org
384 http://ojps.aip.org/getpdf/servlet/GetPDFServlet?filetype=pdf&id=PRLTAO000088000010100404000001

&idtype=cvips

Wave-particle duality
652

wave nature of biomolecules and fluorofullerenes" 385. Phys. Rev. Lett. 401:
680-682.
• Note 5: Hornberger, Klaus, Stefan Uttenthaler,Björn Brezger, Lucia Hack-
ermüller, Markus Arndt and Anton Zeilinger (2003). "Observation of Colli-

FT
sional Decoherence in Interferometry" 386. Phys. Rev. Lett. 90: 160401.
• Note 6: Hackermüller, Lucia, Klaus Hornberger, Björn Brezger, Anton
Zeilinger and Markus Arndt (2004). "Decoherence of matter waves by ther-
mal emission of radiation" 387. Nature 427: 711-714.
• R. Nave. Wave-Particle Duality 388. (Web page) HyperPhysics. Georgia State
University, Department of Physics and Astronomy. Retrieved on December
12, 2005.
• Markus Arndt (2006). Interferometry and decoherence experiments with
large molecules 389. (Web page) University of Vienna. Retrieved on May 6,
2006.

Source: http://en.wikipedia.org/wiki/Wave-particle_duality

Principal Authors: Linas, StuRat, Cdang, Pizza Puzzle, Srleffler


A
Wien’s displacement law

Wien’s displacement law is a law of physics that states that there is an inverse
relationship between the wavelength of the peak of the emission of a black
body and its temperature.
DR
b
λmax = T

where
λmax is the peak wavelength in meters,

T is the temperature of the blackbody in kelvins (K), and

385 http://www.nature.com/cgi-taf/DynaPage.taf?file=/nature/journal/v401/n6754/full/401680a0_fs.

html&content_filetype=pdf
386 http://ojps.aip.org/getpdf/servlet/GetPDFServlet?filetype=pdf&id=PRLTAO000090000016160401000001

&idtype=cvips
387 http://www.nature.com/cgi-taf/DynaPage.taf?file=/nature/journal/v427/n6976/full/nature02276_fs.

html&content_filetype=PDF
388 http://hyperphysics.phy-astr.gsu.edu/hbase/mod1.html
389 http://www.quantum.at/typo/index.php?id=177&tx_jppageteaser_pi1=35

Wien’s displacement law


653

A FT
Figure 70 The wavelength corresponding to the peak emission in various black body
spectra as a function of temperature

b is a constant of proportionality, called Wien’s displacement constant, in


kelvin-meters.
DR
The value of this constant (CODATA 2002 recommended) is b = 2.8977685 ×
10−3 ± 5.1 × 10−9 m · K
For optical wavelengths, it is often more convenient to use the nanometer in
place of the meter as the unit of measure. In this case, the constant becomes
2.8977685 × 10 6 kelvin-nanometers.

Explanation
Fundamentally, Wien’s law states that the hotter an object is, the shorter the
wavelength at which it will emit most of its radiation. For example, the surface
temperature of the Sun is 5780 K. Using Wien’s law, this temperature corre-
sponds to a peak emission at a wavelength of 500 nm. This wavelength is
fairly in the middle of the visual spectrum (see for example the article color),
because of the spread resulting in white light. Due to the Rayleigh scattering of

Wien’s displacement law


654

blue light by the atmosphere this white light is separated somewhat, resulting
in a blue sky and a yellow sun.
A lightbulb has a glowing wire with a somewhat lower temperature, resulting
in yellow light, and something that is "red hot" is again a little less hot.

FT
The law is named for Wilhelm Wien, who formulated the relationship in 1893
based on empirical data.

Frequency form
In terms of frequency f (in hertz), Wien’s displacement law becomes
αk
fmax = h T ≈ (5.879 × 1010 Hz/K) · T

where
α ≈ 2.821439... is a constant resulting from the numerical solution of the
maximization equation,

k is Boltzmann’s constant,
A
h is Planck’s constant, and

T is temperature (in kelvin).

Because the spectrum resulting from →Planck’s law of black body radiation
takes a different shape in the frequency domain from that of the wavelength
domain, the frequency location of the peak emission does not correspond to the
DR
peak wavelength using the simple relationship between frequency, wavelength,
and the speed of light.

Derivation
Wilhelm Wien formulated this law, in 1893, based entirely on empirical obser-
vations, prior to the development of →Planck’s law of black body radiation.
With the benefit of hindsight, however, it is now possible to derive Wien’s law
as a direct consequence of Planck’s more general expression.
From Planck’s law, we know that the spectrum of black body radiation is
8πhc 1
u(λ) = λ5 ehc/λkT −1

The value of λ for which this function is maximized is sought. To find it, we
differentiate u(λ) with respect to λ and set it equal to zero

Wien’s displacement law


655
 
∂u hc ehc/λkT 1 5
∂λ = 8πhc kT λ7 (ehc/λkT −1)2
− λ6 ehc/λkT −1
=0

hc 1
λkT 1−e−hc/λkT −5=0

FT
If we define
hc
x≡ λkT

then
x
1−e−x −5=0

This equation cannot be solved in terms of elementary functions. It can be


solved in terms of Lambert’s Product Log function but an exact solution is not
important in this derivation. One can easily find the numerical value of x
x = 4.965114231744276 . . . (dimensionless)

Solving for the wavelength λ in units of nanometers, and using units of kelvins
A
for the temperature yields:
λmax = hc 1
= 2.89776829...×106 nm·K
.
kx T T

The frequency form of Wien’s displacement law is derived using similar meth-
ods, but starting with Planck’s law in terms of frequency instead of wavelength.

External links
DR
• Eric Weisstein’s World of Physics 390
• PlanetPhysics 391

Source: http://en.wikipedia.org/wiki/Wien%27s_displacement_law

Principal Authors: Metacomet, Lir, PAR, Vsmith, Patrick

390 http://scienceworld.wolfram.com/physics/WiensDisplacementLaw.html
391 http://planetphysics.org/?op=getobj&from=objects&id=20

Wien’s displacement law


656

Wigner-Eckart theorem

The Wigner-Eckart theorem is a theorem of representation theory and quan-

FT
tum mechanics allowing operators to be transformed from one basis to another.
These transformations involve the use of →Clebsch-Gordan coefficients.

References
• Eric W. Weisstein, Wigner-Eckart theorem 392 at MathWorld.
• Wigner-Eckart theorem 393

Source: http://en.wikipedia.org/wiki/Wigner-Eckart_theorem

Principal Authors: Spiralhighway, Magnus Manske, Charles Matthews, Hillman, Oleg Alexandrov

Wigner quasi-probability distribution


A
See also Wigner distribution, a disambiguation page.

A topic in Quantum theory


<> 394

The Wigner quasi-probability distribution was introduced by Eugene Wigner


in 1932 to study quantum corrections to classical statistical mechanics. The
DR
goal was to replace the wavefunction that appears in Schrödinger’s equation
with a probability distribution in phase space. It was independently derived
by Hermann Weyl in 1931 as the symbol of the density matrix in representa-
tion theory in mathematics. It was once again derived by J. Ville in 1948 as
a quadratic (in signal) representation of the local time-frequency energy of a
signal. It is also known as the "Wigner function," "Wigner-Weyl transformation"
or the "Wigner-Ville distribution". It has applications in statistical mechan-
ics, quantum chemistry, quantum optics, classical optics and signal analysis in
diverse fields such as electrical engineering seismology, biology, and engine
design.
A classical particle has a definite position and momentum and hence, is rep-
resented by a point in phase space. When one has a collection (ensemble)

392 http://mathworld.wolfram.com/Wigner-EckartTheorem.html
393 http://electron6.phys.utk.edu/qm2/modules/m4/wigner.htm

Wigner quasi-probability distribution


657

of particles, the probability of finding a particle at a certain position in phase


space is given by a probability distribution. This is not true for a quantum parti-
cle due to the uncertainty principle. Instead, one can create a quasi-probability
distribution, which necessarily does not satisfy all the properties of a normal

FT
probability distribution. For instance, the Wigner distribution can go negative
for states which have no classical model (and hence, it can be used to identify
non-classical states).
The Wigner distribution P (q, p) is defined as:
1
R∞ ∗ 2ipy/~
P (x, p) = π~ −∞ dy ψ (x + y)ψ(x − y)e

where ψ is the wavefunction and x and p are position and momentum but could
be any conjugate variable pair. (ie. real and imaginary parts of the electric field
or frequency and time of a signal). It is symmetric in x and p:
1
R∞ ∗ −2ixq/~
P (x, p) = π~ −∞ dq φ (p + q)φ(p − q)e

where φ is the Fourier transform of ψ.


In the case of a mixed state:
A 1
P (x, p) = π~
R∞
−∞ dy hx − y|ρ̂|x + yie
2ipy/~

where ρ is the density matrix.

Mathematical properties
1. P (x, p) is real
2. The x and p probability distributions are given by the marginals:
DR
• Typically the trace of ρ is equal to 1.
• 1. and 2. imply the P(x,p) is negative somewhere, with the exception of the
coherent state (and mixtures of coherent states) and the squeezed vacuum
state.

3. P (x, p) has the following reflection symmetries:

• Time symmetry: ψ(x) → ψ(x)∗ ⇒ P (x, p) → P (x, −p)


• Space symmetry: ψ(x) → ψ(−x) ⇒ P (x, p) → P (−x, −p)

4. P (x, p) is Galilei-invariant:

• It is not Lorentz invariant.

5. The equation of motion for each point in the phase space is classical in the
absence of forces: 6. State overlap is calculated as:

Wigner quasi-probability distribution


658

A FT
Figure 71 Figure 1: The Wigner quasi-
probability distribution for a) the vacuuum
b) An n = 1 →Fock state (e.g. a single pho-
ton) c) An n = 5 Fock state.

7. Operators and expectation values (averages) are calculated as follows:


DR
8. In order that P (x, p) represent physical (positive) density matrices: where
|θ> is a pure state.

Uses of the Wigner function outside quantum me-


chanics
• In the modelling of optical systems such as telescopes or fibre telecommu-
nications devices, the Wigner function is used to bridge the gap between
simple ray tracing and the full wave analysis of the system. Here p/~ is
replaced with k =|k |sinθ≈|k |θ in the small angle (paraxial) approximation.
In this context, the Wigner function is the closest one can get to describing
the system in terms of rays at position x and angle θ while still including
the effects of interference. If it becomes negative at any point then simple
ray-tracing will not suffice to model the system.

Wigner quasi-probability distribution


659


A FT
Figure 72 Figure 2: A contour plot of the Wigner-Ville distribution for a chirped pulse of light.
The plot makes it obvious that the frequency is a linear function of time.

In signal analysis, a time-varying electrical signal, mechanical vibration, or


sound wave are represented by a Wigner function. Here, x is replaced with
the time and p/~ is replaced with the angular frequency ω=2πf, where f is
the regular frequency.
DR
• In ultrafast optics, short laser pulses are characterized with the Wigner func-
tion using the same f and t substitutions as above. Pulse defects such as
chirp (the change in frequency with time) can be visualized with the Wign-
er function. See Figure 2.

• In quantum optics, x and p/~ are replaced with the X and P quadratures,
the real and imaginary components of the electric field (see coherent state).
The plots in Figure 1 are of quantum states of light.

Measurements of the Wigner function


• Tomography
• Homodyne detection
• FROG Frequency-resolved optical gating

Wigner quasi-probability distribution


660

Other related quasi-probability distributions


The Wigner distribution was the first quasi-probability distribution but many
more followed with various advantages:

FT
• Glauber P representation
• Husimi Q representation

Historical note
As the introduction shows, the formula for the Wigner function was indepen-
dently derived many times in different contexts. In fact, apparently Wigner
was unaware that even within the context of quantum theory, it had been in-
troduced previously by Heisenberg and Dirac. However, they missed its sig-
nificance as they believed it was only an approximation to the true quantum
description of a system such as the atom. Incidentally, Dirac would later be-
come Wigner’s brother-in-law. See references.

See also

AHeisenberg group

References
• E.P. Wigner, "On the quantum correction for thermodynamic equilibrium",
Phys. Rev. 40 (June 1932) 749-759.
• H. Weyl, Z. Phys. 46, 1 (1927).
• H. Weyl, Gruppentheorie und Quantenmechanik (Leipzig: Hirzel)(1928).
DR
• H. Weyl, The Theory of Groups and Quantum Mechanics (Dover, New York,
1931).
• J. Ville, "Théorie et Applications de la Notion de Signal Analytique", Cables
et Transmission, 2A: (1948) 61-74.
• W. Heisenberg, "Über die inkohärente Streuung von Röntgenstrahlen",
Physik. Zeitschr. 32, 737-740 (1931).
• P.A.M. Dirac, "Note on exchange phenomena in the Thomas atom", Proc.
Camb. Phil. Soc. 26, 376-395 (1930).
• C. Zachos, D. Fairlie, and T. Curtright, Quantum Mechanics in Phase Space
( World Scientific, Singapore, 2005).

Source: http://en.wikipedia.org/wiki/Wigner_quasi-probability_distribution

Principal Authors: J S Lundeen, Michael Hardy, Linas, Btyner, Mjb

Wigner quasi-probability distribution


661

Work function

The work function is the minimum energy (usually measured in electron volts)

FT
needed to remove an electron from a solid to a point immediately outside the
solid surface. Here "immediately" means that the final electron position is far
from the surface on the atomic scale but still close to the solid on the macro-
scopic scale. Work function is an important property of metal. The magnitude
of work function is usually about a half of the ionization energy of a free atom
of the metal.

Work Function and Surface Effect


Work function W of a metal is closely related to its Fermi energy level F yet
the two quantities are not exactly the same. This is due to the surface effect of a
real-world solid: a real-world solid is not infinitely extended with electrons and
ions repeatedly filling every primitive cell over all Bravais lattice sites. Neither
can one simply take a set of Bravais lattice sites {R} inside the geometrical
region V which the solid occupies and then fill undistorted charge distribution
A
basis into every primitive cells of {R} . Indeed, the charge distribution in
those cells near the surface will be distorted significantly from that in a cell of
an ideal infinite solid, resulting in an effective surface dipole distribution, or,
sometimes both a surface dipole distribution and a surface charge distribution.
It can be proved that if we define work function as the minimum energy need-
ed to remove an electron to a point immediately out of the solid, the effect
of the surface charge distribution can be neglected, leaving only the surface
dipole distribution. Let the potential energy difference across the surface due
DR
to effective surface dipole be WS . And let F be the →Fermi energy calculated
for the finite solid without considering surface distortion effect, when taking
the convention that the potential at r → ∞ is zero. Then, the correct formula
for work function is:
W = −F + WS
Where F is negative, which means that electrons are bound in the solid.

Example
For example, Caesium has ionization energy 3.9 eV and work function 1.9 eV.

Work function
662

Photoelectric work function


The work function is the minimum energy that must be given to an electron to
liberate it from the surface of a particular metal. In the photoelectric effect if
a photon with an energy greater than the work function is incident on a metal

FT
photoelectric emission occurs. Any excess energy is given to the electron as
kinetic energy.
Photoelectric work function:
φ=hf 0,
where h is Planck’s constant and f 0 is the minimum (threshold) frequency of
the photon required for photoelectric emission.

Thermionic work function


The work function is also important in the theory of thermionic emission. Here
the electron gains its energy from heat rather than photons. In this case, as for
an electron escaping from the heated negatively-charged filament of a vacuum
tube, the work function may be called the thermionic work function. Tungsten
is a very common metal for vacuum tube elements, with a work function of
A
approximately 4.5 eV.
The thermionic work function depends on the orientation of the crystal and will
tend to be smaller for metals with an open lattice, larger for metals in which
the atoms are closely packed. The range is about 1.5–6 V. It is somewhat higher
on dense crystal faces than open ones.

Applications
DR
In electronics the work function is important for design of the metal-
semiconductor junction in Schottky diodes and for design of vacuum tubes.
The work function is also important in the theory of thermionic emission, here
the electron gains its energy from heat rather than photons. In this case, as for
example that of an electron escaping from the heated negatively-charged fila-
ment of a vacuum tube, the work function may be called the thermionic work
function. Tungsten is a very common metal for vacuum tube elements, with a
work function of approximately 4.5 eV.
It depends on the orientation of the crystal and will tend to be smaller for
metals with an open lattice, larger for metals in which the atoms are closely
packed. The range is about 1.5–6 eV. It is somewhat higher on dense crystal
faces than open ones.

Work function
663

See also
• free energy for the Helmholtz free energy equation, which is the thermody-
namic work, note that this work is not related to electron emission and is

FT
thus not directly related to the work function.
• Electron affinity. See NEA cathode 395 for an application to condensed mat-
ter.

Reference
Solid State Physics, by Ashcroft and Mermin. Thomson Learning, Inc, 1976

External links
• Work functions of common metals 396
• Work functions of various metals for the photoelectric effect 397
• Contains some work functions for metals 398 and the corresponding heat re-
quired to eject them by thermionic emission
• Some work functions 399
A
Source: http://en.wikipedia.org/wiki/Work_function

Principal Authors: Omegatron, Jonathunder, Arnero, RTC, Pacaro

Zeeman effect
DR
The Zeeman effect (IPA [ze m n]) is the splitting of a spectral line into several
components in the presence of a magnetic field. It is analogous to the →Stark
effect, the splitting of a spectral line into several components in the presence
of an electric field.

395 http://www.slac.stanford.edu/cgi-wrap/getdoc/slac-pub-8355.pdf
396 http://www.pulsedpower.net/Info/WorkFunctions.htm
397 http://hyperphysics.phy-astr.gsu.edu/hbase/tables/photoelec.html
398 http://www.fnrf.science.cmu.ac.th/theory/linac/Linac%20Basic%20Concepts.html
399 http://www.physchem.co.za/Light/Particles.htm#Work%20function

Zeeman effect
664

Introduction
In most atoms, there exist several electronic configurations that have the same
energy, so that transitions between different pairs of configurations correspond
to a single line.

FT
The presence of a magnetic field breaks the degeneracy, since it interacts in a
different way with electrons with different quantum numbers, slightly modify-
ing their energies. The result is that, where there were several configurations
with the same energy, now there are different energies, which give rise to sev-
eral very close spectral lines.

A
Without a magnetic field, configurations a, b and c have the same energy, as
do d, e and f. The presence of a magnetic field splits the energy levels. A line
DR
produced by a transition from a, b or c to d, e or f now will be several lines
between different combinations of a, b, c and d, e, f. Not all transitions will be
possible – see transition rules.
Since the distance between the Zeeman sub-levels is proportional with the
magnetic field, this effect was used by astronomers to measure the magnet-
ic field of the Sun and other stars.
There is also an anomalous Zeeman effect that appears on transitions where
the net spin of the electrons is not 0, the number of Zeeman sub-levels being
even instead of odd if there’s an uneven number of electrons involved. It was
called "anomalous" because the electron spin had not yet been discovered, and
so there was no good explanation for it, at the time that Zeeman observed the
effect.

Zeeman effect
665

If the magnetic field strength is too high, the effect is no longer linear; at
even higher field strength, electron coupling is disturbed and the spectral lines
rearrange. This is called Paschen-Back effect.
The Zeeman effect is named after the Dutch physicist Pieter Zeeman.

FT
Theoretical presentation
The total Hamiltonian of an atom in a magnetic field is:
H = H0 + H1 = H0 +
P ~ ·S
ξ(r~α )L ~ − P µ~α · B
~
α α

where H0 is the unperturbed Hamiltonian of the atom, and the sums over α
are sums over the electrons in the atom. The term
~ ·S
ξ(r~α )L ~

is the LS-coupling for each electron (indexed by α) in the atom. The sum
vanishes if there is only one electron. The magnetic coupling
~ = µB (gL L
µ~α · B ~ + gS S)
~ ·B~
~
A
is the energy due to the magnetic moment µ of the α-th electron. It can be
written as sum of contributions of the orbital angular momentum and of spin
angular momentum, with each multiplied by the gyroscopic or Landé g-factor.
By projecting the vector quantities onto the z-axis, the Hamiltonian may be
written as
H = H0 + ξ(r)L ~ + µB (gL Lz + gs Sz )Bz ≈ Hat + µB (Jz + Sz )Bz
~ ·S
~
DR
where the approximation results from taking the g-factors are gL = 1 and
gS ≈ 2. The summation over the electrons was omitted for readability. Here,
Jz = Lz + Sz is the total angular momentum, and the LS-coupling term has
been folded into H0 .
The size of the interaction term H ’ is not always small, and can induce large
effects on the system. In the Paschen-Back effect, described below, H ’ cannot
be treated as a perturbation, as its magnitude is comparable to or larger than
the unperturbed system Hat . The H ’ term does not commute with Hat . In
particular, Sz doesn’t commute with the spin-orbit interaction in Hat .

Zeeman effect
666

Strong Field (Paschen-Back effect)


To simplify the solution, it is useful to assume that [Hat , Sz ] = 0, so that Lz and
Sz have a set of common eigenfunctions with respect to Hat . This allows the
expectation values of Lz and Sz to be easily evaluated on a general state |Ai:

FT
 
Hat + Bz~µB (Lz + 2Sz ) |Ai = (Eat + Bz µB (ml + 2ms )|Ai

The above may be read as implying that the LS-coupling is completely broken
by the external field. The system re-arranges substantially according to the
Bz field. The ml and ms are still "good" quantum numbers. This implies that
the selection rules obtained from ∆S = 0, ∆L = ±1 are still very likely for
the system. In particular, apart from the line splittings one might normally
expect, only three spectral lines will be visible, corresponding to the ∆m = ±1
transition rule. The splitting depends upon the l level being considered. The
spectral lines depend on the transition frequencies, that is, on the difference of
energy.

See also
A• →Stark effect

References
Historical
• Condon, E. U.; G. H. Shortley (1935). The Theory of Atomic Spectra. Cam-
bridge University Press. ISBN 521-09209-4. (Chapter 16 provides a com-
DR
prehensive treatment, as of 1935.)
• Zeeman, P. (1897). "(Title unknown)". Phil.Mag. 43: 226.
• Zeeman, P. (11 February 1897). "The Effect of Magnetisation on the Nature
of Light Emitted by a Substance" 400. Nature 55: 347.

Modern
• Forman, Paul (1970). "Alfred Landé and the anomalous Zeeman Effect,
1919-1921". Historical Studies in the Physical Sciences 2: 153—261.
• Griffiths, David J. (2004). Introduction to Quantum Mechanics (2nd ed.).
Prentice Hall. ISBN 013805326X.
• Liboff, Richard L. (2002). Introductory Quantum Mechanics. Addison-
Wesley. ISBN 0805387145.

400 http://dbhs.wvusd.k12.ca.us/webdocs/Chem-History/Zeeman-effect.html

Zeeman effect
667

Source: http://en.wikipedia.org/wiki/Zeeman_effect

Principal Authors: Linas, AstroNomer, Bogdangiusca, Voyajer, Laurascudder, Nsh

A FT
DR
DR 668

AFT
669

GNU Free Documentation License


Version 1.2, November 2002

Copyright (C) 2000,2001,2002 Free Software Foundation, Inc. 51 Franklin St, Fifth Floor, Boston, MA 02110-1301
USA Everyone is permitted to copy and distribute verbatim copies of this license document, but changing it is not
allowed.

FT
0. PREAMBLE
The purpose of this License is to make a manual, textbook, or other functional and useful document "free" in the sense
of freedom: to assure everyone the effective freedom to copy and redistribute it, with or without modifying it, either
commercially or noncommercially. Secondarily, this License preserves for the author and publisher a way to get credit
for their work, while not being considered responsible for modifications made by others.
This License is a kind of "copyleft", which means that derivative works of the document must themselves be free in the
same sense. It complements the GNU General Public License, which is a copyleft license designed for free software.
We have designed this License in order to use it for manuals for free software, because free software needs free
documentation: a free program should come with manuals providing the same freedoms that the software does. But
this License is not limited to software manuals; it can be used for any textual work, regardless of subject matter
or whether it is published as a printed book. We recommend this License principally for works whose purpose is
instruction or reference.

1. APPLICABILITY AND DEFINITIONS


This License applies to any manual or other work, in any medium, that contains a notice placed by the copyright
holder saying it can be distributed under the terms of this License. Such a notice grants a world-wide, royalty-free
license, unlimited in duration, to use that work under the conditions stated herein. The "Document", below, refers to
any such manual or work. Any member of the public is a licensee, and is addressed as "you". You accept the license if
you copy, modify or distribute the work in a way requiring permission under copyright law.
A
A "Modified Version" of the Document means any work containing the Document or a portion of it, either copied
verbatim, or with modifications and/or translated into another language.
A "Secondary Section" is a named appendix or a front-matter section of the Document that deals exclusively with the
relationship of the publishers or authors of the Document to the Document’s overall subject (or to related matters)
and contains nothing that could fall directly within that overall subject. (Thus, if the Document is in part a textbook of
mathematics, a Secondary Section may not explain any mathematics.) The relationship could be a matter of historical
connection with the subject or with related matters, or of legal, commercial, philosophical, ethical or political position
regarding them.
The "Invariant Sections" are certain Secondary Sections whose titles are designated, as being those of Invariant Sec-
tions, in the notice that says that the Document is released under this License. If a section does not fit the above
definition of Secondary then it is not allowed to be designated as Invariant. The Document may contain zero Invariant
Sections. If the Document does not identify any Invariant Sections then there are none.
DR
The "Cover Texts" are certain short passages of text that are listed, as Front-Cover Texts or Back-Cover Texts, in the
notice that says that the Document is released under this License. A Front-Cover Text may be at most 5 words, and a
Back-Cover Text may be at most 25 words.
A "Transparent" copy of the Document means a machine-readable copy, represented in a format whose specification is
available to the general public, that is suitable for revising the document straightforwardly with generic text editors
or (for images composed of pixels) generic paint programs or (for drawings) some widely available drawing editor,
and that is suitable for input to text formatters or for automatic translation to a variety of formats suitable for input
to text formatters. A copy made in an otherwise Transparent file format whose markup, or absence of markup, has
been arranged to thwart or discourage subsequent modification by readers is not Transparent. An image format is not
Transparent if used for any substantial amount of text. A copy that is not "Transparent" is called "Opaque".
Examples of suitable formats for Transparent copies include plain ASCII without markup, Texinfo input format, LaTeX
input format, SGML or XML using a publicly available DTD, and standard-conforming simple HTML, PostScript or PDF
designed for human modification. Examples of transparent image formats include PNG, XCF and JPG. Opaque formats
include proprietary formats that can be read and edited only by proprietary word processors, SGML or XML for which
the DTD and/or processing tools are not generally available, and the machine-generated HTML, PostScript or PDF
produced by some word processors for output purposes only.
The "Title Page" means, for a printed book, the title page itself, plus such following pages as are needed to hold, legibly,
the material this License requires to appear in the title page. For works in formats which do not have any title page as
such, "Title Page" means the text near the most prominent appearance of the work’s title, preceding the beginning of
the body of the text.
670

A section "Entitled XYZ" means a named subunit of the Document whose title either is precisely XYZ or contains XYZ
in parentheses following text that translates XYZ in another language. (Here XYZ stands for a specific section name
mentioned below, such as "Acknowledgements", "Dedications", "Endorsements", or "History".) To "Preserve the Title"
of such a section when you modify the Document means that it remains a section "Entitled XYZ" according to this
definition.
The Document may include Warranty Disclaimers next to the notice which states that this License applies to the

FT
Document. These Warranty Disclaimers are considered to be included by reference in this License, but only as regards
disclaiming warranties: any other implication that these Warranty Disclaimers may have is void and has no effect on
the meaning of this License.

2. VERBATIM COPYING
You may copy and distribute the Document in any medium, either commercially or noncommercially, provided that
this License, the copyright notices, and the license notice saying this License applies to the Document are reproduced
in all copies, and that you add no other conditions whatsoever to those of this License. You may not use technical
measures to obstruct or control the reading or further copying of the copies you make or distribute. However, you may
accept compensation in exchange for copies. If you distribute a large enough number of copies you must also follow
the conditions in section 3.
You may also lend copies, under the same conditions stated above, and you may publicly display copies.

3. COPYING IN QUANTITY
If you publish printed copies (or copies in media that commonly have printed covers) of the Document, numbering
more than 100, and the Document’s license notice requires Cover Texts, you must enclose the copies in covers that
carry, clearly and legibly, all these Cover Texts: Front-Cover Texts on the front cover, and Back-Cover Texts on the
back cover. Both covers must also clearly and legibly identify you as the publisher of these copies. The front cover
must present the full title with all words of the title equally prominent and visible. You may add other material on the
covers in addition. Copying with changes limited to the covers, as long as they preserve the title of the Document and
A
satisfy these conditions, can be treated as verbatim copying in other respects.
If the required texts for either cover are too voluminous to fit legibly, you should put the first ones listed (as many as
fit reasonably) on the actual cover, and continue the rest onto adjacent pages.
If you publish or distribute Opaque copies of the Document numbering more than 100, you must either include a
machine-readable Transparent copy along with each Opaque copy, or state in or with each Opaque copy a computer-
network location from which the general network-using public has access to download using public-standard network
protocols a complete Transparent copy of the Document, free of added material. If you use the latter option, you
must take reasonably prudent steps, when you begin distribution of Opaque copies in quantity, to ensure that this
Transparent copy will remain thus accessible at the stated location until at least one year after the last time you
distribute an Opaque copy (directly or through your agents or retailers) of that edition to the public.
It is requested, but not required, that you contact the authors of the Document well before redistributing any large
number of copies, to give them a chance to provide you with an updated version of the Document.
DR
4. MODIFICATIONS
You may copy and distribute a Modified Version of the Document under the conditions of sections 2 and 3 above,
provided that you release the Modified Version under precisely this License, with the Modified Version filling the role
of the Document, thus licensing distribution and modification of the Modified Version to whoever possesses a copy of
it. In addition, you must do these things in the Modified Version:

A. Use in the Title Page (and on the covers, if any) a title distinct from that of the Document, and from those of
previous versions (which should, if there were any, be listed in the History section of the Document). You may use
the same title as a previous version if the original publisher of that version gives permission.
B. List on the Title Page, as authors, one or more persons or entities responsible for authorship of the modifications
in the Modified Version, together with at least five of the principal authors of the Document (all of its principal
authors, if it has fewer than five), unless they release you from this requirement.
C. State on the Title page the name of the publisher of the Modified Version, as the publisher.
D. Preserve all the copyright notices of the Document.
E. Add an appropriate copyright notice for your modifications adjacent to the other copyright notices.
F. Include, immediately after the copyright notices, a license notice giving the public permission to use the Modified
Version under the terms of this License, in the form shown in the Addendum below.
G. Preserve in that license notice the full lists of Invariant Sections and required Cover Texts given in the Document’s
license notice.
H. Include an unaltered copy of this License.
671

I. Preserve the section Entitled "History", Preserve its Title, and add to it an item stating at least the title, year, new
authors, and publisher of the Modified Version as given on the Title Page. If there is no section Entitled "History" in
the Document, create one stating the title, year, authors, and publisher of the Document as given on its Title Page,
then add an item describing the Modified Version as stated in the previous sentence.
J. Preserve the network location, if any, given in the Document for public access to a Transparent copy of the Docu-
ment, and likewise the network locations given in the Document for previous versions it was based on. These may

FT
be placed in the "History" section. You may omit a network location for a work that was published at least four
years before the Document itself, or if the original publisher of the version it refers to gives permission.
K. For any section Entitled "Acknowledgements" or "Dedications", Preserve the Title of the section, and preserve in the
section all the substance and tone of each of the contributor acknowledgements and/or dedications given therein.
L. Preserve all the Invariant Sections of the Document, unaltered in their text and in their titles. Section numbers or
the equivalent are not considered part of the section titles.
M. Delete any section Entitled "Endorsements". Such a section may not be included in the Modified Version.
N. Do not retitle any existing section to be Entitled "Endorsements" or to conflict in title with any Invariant Section.
O. Preserve any Warranty Disclaimers.

If the Modified Version includes new front-matter sections or appendices that qualify as Secondary Sections and contain
no material copied from the Document, you may at your option designate some or all of these sections as invariant.
To do this, add their titles to the list of Invariant Sections in the Modified Version’s license notice. These titles must be
distinct from any other section titles.
You may add a section Entitled "Endorsements", provided it contains nothing but endorsements of your Modified
Version by various parties–for example, statements of peer review or that the text has been approved by an organization
as the authoritative definition of a standard.
You may add a passage of up to five words as a Front-Cover Text, and a passage of up to 25 words as a Back-Cover
Text, to the end of the list of Cover Texts in the Modified Version. Only one passage of Front-Cover Text and one
of Back-Cover Text may be added by (or through arrangements made by) any one entity. If the Document already
A
includes a cover text for the same cover, previously added by you or by arrangement made by the same entity you
are acting on behalf of, you may not add another; but you may replace the old one, on explicit permission from the
previous publisher that added the old one.
The author(s) and publisher(s) of the Document do not by this License give permission to use their names for publicity
for or to assert or imply endorsement of any Modified Version.

5. COMBINING DOCUMENTS
You may combine the Document with other documents released under this License, under the terms defined in section
4 above for modified versions, provided that you include in the combination all of the Invariant Sections of all of the
original documents, unmodified, and list them all as Invariant Sections of your combined work in its license notice,
and that you preserve all their Warranty Disclaimers.
The combined work need only contain one copy of this License, and multiple identical Invariant Sections may be
DR
replaced with a single copy. If there are multiple Invariant Sections with the same name but different contents, make
the title of each such section unique by adding at the end of it, in parentheses, the name of the original author or
publisher of that section if known, or else a unique number. Make the same adjustment to the section titles in the list
of Invariant Sections in the license notice of the combined work.
In the combination, you must combine any sections Entitled "History" in the various original documents, forming
one section Entitled "History"; likewise combine any sections Entitled "Acknowledgements", and any sections Entitled
"Dedications". You must delete all sections Entitled "Endorsements."

6. COLLECTIONS OF DOCUMENTS
You may make a collection consisting of the Document and other documents released under this License, and replace
the individual copies of this License in the various documents with a single copy that is included in the collection,
provided that you follow the rules of this License for verbatim copying of each of the documents in all other respects.
You may extract a single document from such a collection, and distribute it individually under this License, provided
you insert a copy of this License into the extracted document, and follow this License in all other respects regarding
verbatim copying of that document.

7. AGGREGATION WITH INDEPENDENT WORKS


A compilation of the Document or its derivatives with other separate and independent documents or works, in or on
a volume of a storage or distribution medium, is called an "aggregate" if the copyright resulting from the compilation
is not used to limit the legal rights of the compilation’s users beyond what the individual works permit. When the
Document is included in an aggregate, this License does not apply to the other works in the aggregate which are not
themselves derivative works of the Document.
672

If the Cover Text requirement of section 3 is applicable to these copies of the Document, then if the Document is less
than one half of the entire aggregate, the Document’s Cover Texts may be placed on covers that bracket the Document
within the aggregate, or the electronic equivalent of covers if the Document is in electronic form. Otherwise they must
appear on printed covers that bracket the whole aggregate.

8. TRANSLATION

FT
Translation is considered a kind of modification, so you may distribute translations of the Document under the terms
of section 4. Replacing Invariant Sections with translations requires special permission from their copyright holders,
but you may include translations of some or all Invariant Sections in addition to the original versions of these Invariant
Sections. You may include a translation of this License, and all the license notices in the Document, and any Warranty
Disclaimers, provided that you also include the original English version of this License and the original versions of
those notices and disclaimers. In case of a disagreement between the translation and the original version of this
License or a notice or disclaimer, the original version will prevail.
If a section in the Document is Entitled "Acknowledgements", "Dedications", or "History", the requirement (section 4)
to Preserve its Title (section 1) will typically require changing the actual title.

9. TERMINATION
You may not copy, modify, sublicense, or distribute the Document except as expressly provided for under this License.
Any other attempt to copy, modify, sublicense or distribute the Document is void, and will automatically terminate
your rights under this License. However, parties who have received copies, or rights, from you under this License will
not have their licenses terminated so long as such parties remain in full compliance.

10. FUTURE REVISIONS OF THIS LICENSE


The Free Software Foundation may publish new, revised versions of the GNU Free Documentation License from time
to time. Such new versions will be similar in spirit to the present version, but may differ in detail to address new
problems or concerns. See http://www.gnu.org/copyleft/.
Each version of the License is given a distinguishing version number. If the Document specifies that a particular
A
numbered version of this License "or any later version" applies to it, you have the option of following the terms and
conditions either of that specified version or of any later version that has been published (not as a draft) by the Free
Software Foundation. If the Document does not specify a version number of this License, you may choose any version
ever published (not as a draft) by the Free Software Foundation.
DR
673

List of Figures
Please note that images have various different licenses and usage constraints.
License details for a specific image can be found at:

FT
http://en.wikipedia.org/wiki/Image:IMAGENAME.JPG

For each image we list: description, file name, licence, user who added the image (if available), url
for further information (if available).

Note that license names are abbreviated like GFDL, PD, fairuse, etc. Please have a look at the cor-
responding page at Wikipedia for further information about the images and details of the associated
licences.

aharonov-bohm.png 5
Bohratommodel.png, 14
Bose_Einstein_condensate.png, PD-USGov-NIST 23
BoseEinsteinGas1.png, PD-self 37
Coherent_noise_compare3.jpg, GFDL 55
coherent_state_wavepacket.jpg 56
wigner_function_coherent_state.jpg 57
photon_numbers_coherent_state.jpg 58
Coherent_state2.png, GFDL 59
A Compton scattering diagram.png, PD-self
Renormalized-vertex.png, GFDL
Fringespos.png, PD
68
76
106
FD_e_mu.jpg, PD-ineligible 131
FD_kT_e.jpg, PD-ineligible 132
FD_e_kT.jpg, PD-ineligible 134
Atom.png, GFDL 197
Vortex.jpg, fairusereview 198
Holography-reconstruct.png, GFDL 201
Human brain NIH.jpg, PD-USGov-HHS-NIH 202
DR
Epithelial-cells.jpg, GFDL 204
Common clownfish.jpg, FormerFeaturedPicture 205
Constructive_and_Destructive_Interference.gif, self2|GFDL|cc-by-sa-2.5,2.0,1.0 227
NASA_Hydrogen_spectrum.jpg, PD-USGov-NASA 228
Gallery_SineWave_Generation.jpg, NoRightsReserved 231
Bohr-planetary-atom-model.jpg, GFDL-self 233
Bohr_atomic_wave.jpg, GFDL-self 237
Josephson_junction_IV.png, PD-USGov 255
Bohratommodel.png, 270
1D normal modes.gif, PD-user|Régis Lachaume|Lachaume 307
parity_1drep.png 325
1-D Box.png 333
Three_paths_from_A_to_B.png, GFDL 349
Penrose_Interpretation.jpg, PD 360
Photoelectric_effect.png, GFDL 372
bbs.jpg 381
Plum_pudding_atom.svg, cc-sa-1.0 388
state_discrimination_proj.png 395
state_discrimination_POVM.png 397
674

Witten.jpg, Promotional 431


Schr-harmonic.png, Cc-by-sa-2.0 443
QHarmonicOscillator.png, GFDL 444
HAtomOrbitals.png, GFDL-en 460
Schroedingerscat2.jpg, GFDL-self 533

FT
Schrödinger’s cat - BLINK tag.gif, NowCommonsThis 537
noise_squeezed_states.jpg 566
wave_packet_squeezed_states.jpg 567
wigner_function_squeezed_states.jpg 568
noise_squeezed_states.jpg 569
wave_packet_squeezed_states.jpg 570
wigner_function_squeezed_states.jpg 571
noise_squeezed_states.jpg 572
wave_packet_squeezed_states.jpg 573
wigner_function_squeezed_states.jpg 574
photon_numbers_squeezed_coherent_states_subpoisson.jpg 575
phase_distribution_squeezed_coherent_states_subpoisson.jpg 575
photon_numbers_squeezed_coherent_states_subpoisson.jpg 576
phase_distribution_squeezed_coherent_states_subpoisson.jpg 576
photon_numbers_squeezed_coherent_states_subpoisson.jpg 577
phase_distribution_squeezed_coherent_states_subpoisson.jpg 577
photon_numbers_squeezed_vacuum.jpg 578
Squid prototype2.jpg, NoRightsReserved 579
A
Squid prototype.jpg, NoRightsReserved
Stark splitting in hydrogen.png, PD-self
Spin_up.gif, GFDL
579
580
582
Stern-Gerlach experiment.PNG, GFDL 583
Quantum_projection_of_S_onto_z_for_spin_half_particles.PNG, GFDL 584
SternGerlach2.jpg, GFDL 586
Atom.png, GFDL 588
Interfering Electron Wave Packets animated.gif, GFDL 616
bbs.jpg 653
Wigner_functions.jpg, PD 658
DR
Chirpedpulse.jpg, PD 659
675

Index

1/f noise 409, 409 645, 647

FT
11 february 666 alfred clebsch 52
14 october 651 alfred lande 481, 626
19th century 8, 15, 645 alfred landé 666
20th century 15, 104, 430, 458, 460, 641, algebra of physical space 98, 102
645 algebra representation of a hopf algebra 595
21 cm line 192 algebra representation of a lie superalgebra
25 april 319 595
3-body problem 270 alice and bob 455
3-sphere 607 alkali metal 187
4-vector 264 alloy 572
5th century bc 9 alpha decay 498
6-j symbol 54 alpha particle 10, 163, 199, 312, 517, 522
6th century bc 9 aluminium 163
čech cohomology 17 aluminium arsenide 504
647 american gods 540
american physical society 552, 554
a ammonia 370
a. elgart 3 amplitude 412, 641, 647
A
a. s. wightman 280
abelian 300, 613
abelian group 323, 324
analog model of gravity 85
analysis 196
analytic 510
abraham-lorentz force 434 ancient india 9
a brief history of time 194 ancilla 398
absolute value 177, 334, 403, 468 angle 174
absolute zero 22, 138, 486, 581 angular frequency 87, 156
absorption spectrum 459 angular momentum 11, 16, 52, 54, 67, 170,
abstract 197 190, 241, 259, 260, 260, 262, 262, 263,
abuse of notation 44 327, 344, 344, 345, 460, 463, 466, 558,
accuracy 620 582, 597
DR
accurate 619 angular momentum coupling 54, 546
acoustic metric 85 angular momentum operator 182, 183
acoustic resonance 509 animal 571
acoustics 412, 522 anime 543
action 160 annalen der physik 373, 387
action-angle coordinates 16 annihilate 438
action principle 468 annihilation operator 56, 259
activity 33, 137, 288 anode 154
adhesion 23 anomaly (physics) 356
adjoint 79, 81, 445 ansatz 632
adjoint representation 78 antibaryon 120
aether 438 anticommutator 82, 597, 599
affine structure 354 anticommute 239
afshar experiment 66, 222, 298, 651 antiderivation 356
ah! my goddess 543 antilinear 81
aharonov-bohm effect 50, 595 antimatter 119
albert abraham michelson 192 antiparticle 118, 301, 353, 438, 589, 590
albert einstein 10, 11, 23, 34, 62, 213, 226, antisymmetric 640
230, 247, 372, 375, 382, 386, 451, 460, antisymmetric tensor 262
462, 471, 477, 534, 591, 625, 643, 645, antiunitary 315, 613
676

antonino lo surdo 577 b


applet 379 b-bbar oscillation 303
approximation 144, 533 b. d. josephson 571
arago spot 651 baidyanaith misra 505
arnold sommerfeld 11, 131 balmer’s formula 235

FT
aromatic 514 balmer series 128
arthur compton 86, 87, 645 banach space 41, 124, 175, 180
arthur holly compton 67 bandgap 504
arxiv 507 baryon 117, 120, 129, 312, 327, 591
arxiv:hep-th/9912205 435 baryon number 607
asher peres 215, 363 baryons 584
as of 2005 591, 648 basic quantum mechanics 284, 530
as of 2006 213 basics of quantum mechanics 628
associated legendre function 189 basilar membrane 509
associated legendre polynomials 54 basis function 533
astrodynamics 264 bayesian network 561
astronomy 256 bay of fundy 509
astrophysics 509 bcs theory 421
asymptote 521 becquerel 312
asymptotic freedom 78 behaviour 196
atom 8, 12, 14, 15, 67, 104, 117, 127, 153, bell’s inequalities 273, 454
163, 182, 187, 191, 248, 262, 311, 382, bell’s inequality 414
388, 405, 421, 458, 459, 459, 460, 461, bell’s theorem 218, 222, 484, 651
A
464, 467, 474, 517, 519, 522, 568, 586,
588, 591, 645, 645, 664
atomic clock 193
bell inequality 627
bell labs 86, 552, 571, 648
bell state 593
atomic coherence 27 bell test experiments 207, 415, 454
atomic nucleus 312, 464, 516, 520, 532, benzene 514
588, 591 berlin 405
atomic number 188, 520, 589 berry phase 6
atomic orbital 187, 377, 586 bertrand’s theorem 262
atomic physics 14, 191, 312, 516, 518, 576 bessel function 345
atomic theory 645 beta decay 93, 311, 591
DR
atomic units 299 bethe 434
atomism 8, 9 big bang 71, 194, 225, 590
atoms 124, 197, 225, 407 big crunch 195
attachment 495 billiard balls 522
audio system measurements 409 binomial coefficient 32, 137
auger 12 binomial theorem 321
auger electron 211 biochemistry 8
auger electron spectroscopy 13 biological 410
augustin fresnel 226 biological molecule 555
autocatalytic set 205 biology 656
automorphism 315 biopotential 552
automorphism group 180 black-body 229
avlis 193 black-body radiation 375
avogadro’s number 424 black body 380, 380, 472, 617, 652
avoided crossing 98 black body radiation 473, 591, 618
axiom 411, 430 blackbody spectrum 269
axion 590 black hole 21, 84, 92, 93, 379, 548
azimuth 188, 344 bloch’s theorem 339
azimuthal quantum number 241, 248, 530 bloch sphere 404, 453
bloch wave 139
677

bob the angry flower 544, 545 brillouin zone 139


bogolibov transformation 84 brownian motion 10, 158, 162, 348
bohm interpretation 199, 219, 222, 295, 650 brst 356
bohr 191, 472 bruno ganz 543
bohr-einstein debates 222, 465, 471, 481, bryce dewitt 474

FT
625 buckyball 648
bohr-sommerfeld quantization 413 buckyballs 514
bohr atom 516 buddhism 203, 206
bohr model 11, 88, 163, 164, 187, 190, 389, buddhist 9
586 bijective 178, 315
bohr model of the atom 153, 578
bohr radius 16, 73, 183, 189 c
boltzmann’s constant 31, 32, 34, 90, 132, c*-algebra 363
137, 166, 285, 288, 407, 601, 654 c* algebra 272
boltzmann constant 139, 140, 419 c-parity 589
boltzmann distribution 165, 287 c-symmetry 301
borel functional calculus 169, 419, 491 c. k. ogden 203
borel set 390 caesium 193, 661
boris podolsky 455 calculus 174
born-oppenheimer approximation 98, 115, caltech 318
116, 299, 391 canonical angular momentum 49
bose-einstein 290 canonical commutation relation 445
bose-einstein condensate 30, 34, 35, 38, 61, canonical commutation relations 277, 447
A
535
bose-einstein statistics 24, 34, 138, 291, 383,
601
canonical coordinates 48, 74
canonical ensemble 289
canonical one-form 126
bose gas 24, 27, 141, 601 canonical partition function 493
bosenova 26 canonical quantization 50, 77, 79, 269, 271,
boson 30, 55, 62, 73, 80, 81, 118, 118, 120, 326, 421, 425
122, 424, 433, 589, 592, 640 canonical transformation technique 267
bosons 22, 24, 27, 30, 34, 133, 286 capacitor 145
bottom quark 301 car algebra 82
boundary condition 148, 343, 513, 522 carbon 514
DR
boundary conditions 607 cargo cult science 318
boundary value problem 311 carl d. anderson 591
bounded operator 125 carl sagan 192
bounded set 179 carl wieman 22, 24
bound particle 365 carol hill 541
bound state 13, 169, 175, 345, 364, 488, cartesian coordinates 345
510, 523, 523 cartesian coordinate system 196, 282, 640
box 532 carver mead 650
bra-ket 168, 584 casimir effect 438, 438
bra-ket notation 52, 81, 113, 144, 174, 179, casimir invariant 263
211, 271, 281, 284, 298, 353, 415, 473, cat 297, 532
488, 490, 524 categorical product 551
brackett series 128 cathode 154
bragg 86, 87 cathode ray 10, 645
brahman 206 cat state 537
brain 202, 495, 573 cauchy sequence 175
brane 122 ccr algebra 50, 81
bravais lattice 661 cecil adams 538
brian david josephson 253 center frequency 510, 511
brian greene 123 central extension 323
678

central force 260, 260, 261, 262, 263 classical physics 47, 173, 269, 283, 323, 557,
central force field 345 617
central potential 523 claude cohen-tannoudji 25
centripetal force 18 claude e. shannon 610
cern 122, 592, 616 claus jönsson 104, 474

FT
cgs 5 clebsch-gordan coefficients 656
chain rule 370 clifford algebra 98, 103, 103
chair 297 clinton davisson 86, 87, 474
chandrasekhar limit 92, 92, 93 clinton joseph davisson 648
chaos 409 closed graph theorem 180
chaos theory 412 closed system 267
characteristic equation 89 closing 607
characteristic polynomial 308 clyde l. cowan 474
characteristic x-ray 211 coefficients 50
charge conjugation 301, 615 coherent 56
charge decay 376 coherent state 450, 564, 565, 565, 657, 659
charge density 271 colatitude 188
chargino 122 cold cathode 145
charles galton darwin 483 cold emission 498
checkerboard nightmare 544 collapse of the wavefunction 272
chemical bond 117, 503 collision 267, 523
chemical bonding 187 color 653
chemical compound 10 color charge 119
A
chemical element 519, 520
chemical energy 411
chemical potential 25, 31, 32, 34, 131, 133,
colors of noise 409
colour confinement 120
commutation relation 259, 263
137, 138, 140, 285, 288, 383 commutative 433
chemical reaction 391 commutator 47, 47, 79, 80, 112, 173, 174,
chemical vapor deposition 504 283, 284, 422, 445, 597, 599, 620
chemistry 8, 8, 79, 312, 459, 459, 470, 485, commute 239
588, 589 compact 390
chien-shiung wu 328 compact lie group 52
chiral spin state 604 complementarity 96
DR
chladni’s figures 460 complementarity (physics) 456
chris fuchs 215 complementary 565
christiaan huygens 644, 646 completely positive 279
christian huygens 226 complete set of commuting operators 545
christopher fuchs 223 complex conjugate 42, 171, 176, 649
christopher stasheff 542 complex conjugation 613
church bells 51 complexity 488
circle 343, 512 complex number 42, 43, 147, 175, 176, 280,
ckm matrix 615 332, 350, 403, 452, 461, 631
clairvoyance 496, 540 complex numbers 215
classical electrodynamics 430 complex plane 194
classical electromagnetism 11, 459 complex pole 510
classical electron radius 73 complex projective line 404
classical field theory 425 composite field 40
classical information channel 415 composite particle 119
classical limit of quantum mechanics 272 compton effect 645
classical mechanics 15, 16, 51, 75, 129, 174, compton suppression 71
213, 260, 266, 267, 314, 330, 331, 348, compton wavelength 69, 379
348, 410, 411, 412, 425, 459, 469, 497, computational chemistry 150, 299, 459, 470
524, 640 computer 365, 471
679

computer code 129 coulomb potential 111, 364


condensed matter physics 138, 226, 421, countable 178
422, 424, 428, 434, 459, 603, 605 countably many 309
conduction electron 365, 428 counterfactual definiteness 456
configuration interaction 557 coupling constant 364

FT
configuration space 343 courant-hilbert 279
confinement 40 covalent bond 473
conical intersection 98 covariant derivative 76
conjugate quantities 619 cpt invariance 592, 614, 615
conjugate transpose 41 cpt symmetry 302, 303, 329, 589, 615
connected correlation function 511 cpt violation 615
connection 17 cp violation 303, 615, 615
consciousness 201, 495 creation and annihilation operators 80, 152,
consciousness causes collapse 219 259, 426, 450
conservation law 400, 529 creation operator 259
conservation laws 438 critical exponent 633
conservation of energy 77, 320, 322, 436 critical phenomena 633
conservation of probability 320 cross product 260
consistent histories 213, 214, 218, 222, 272, cross section 320, 322, 630
536 crystal 117, 581, 603
conspiracy theory 542 crystal lattice 568
constants of motion 261 crystalline 87
constraint 74 crystallography 555
A
constructive quantum field theory 272, 430
context 199
contextualism 204
crystals 311
crystal structure 342
cubical atom 11
continuity equation 399, 529 curie 312
continuous spectrum 292, 390, 523 curl 4
continuous symmetry 263 curvature form 17
contradiction 198 curvilinear 196
convergent series 633 cymbals 51
convex hull 125, 274
convex set 550 d
DR
cooper pair 27, 93, 255, 365 d’alembert operator 257
coordinate 322, 343, 512 d. hestenes 98
copenhagen interpretation 106, 107, 126, daniel kleppner 587
213, 218, 222, 248, 282, 294, 296, 297, dark energy 85
298, 399, 471, 534, 626, 649 dark matter 85, 439
core hamiltonian 150 darmstadt 265
corpuscular theory 226, 646 david bohm 195, 206, 219
correlation 414 david griffiths 474
correspondence limit 469 david gross 77, 434
correspondence principle 51, 54, 113, 248, david hilbert 175, 245, 272, 279
248, 411, 444, 469, 628 david hume 478
cosmic radiation 509 david politzer 77, 434
cosmic ray 588, 589, 591 davisson-germer experiment 87
cosmological constant 84 deborah s. jin 27
cotangent bundle 126 de broglie 481
coulomb 315, 517, 596 de broglie hypothesis 17, 647
coulomb’s law 188, 311 de broglie wavelength 73, 504, 601
coulomb collision 518 debye model 38
coulomb force 182 decay rate 364
coulomb operator 151 december 12 652
680

december 27 476 diode 470


decoherence 114, 218, 272, 294, 358, 595, diode laser 504
605 dipole 546, 582
deductive reasoning 478 dirac 660
deep inelastic scattering 120, 517 dirac’s constant 75, 87

FT
deflection 582 dirac delta function 41, 428
deformation quantization 47 dirac equation 11, 45, 45, 98, 157, 185, 246,
degeneracy 31, 285, 604 258, 299, 327, 422, 424, 469, 473, 528,
degeneracy (disambiguation) 89 530, 649
degeneracy pressure 141 dirac notation 636
degenerate 183, 579 dirac picture 276, 369
degenerate dwarf 94 dirac sea 439
degenerate energy level 124 dirac string 5
degenerate gases 246 direct product 433
degenerate star 94 direct sum 52, 151, 397
degree of coherence 60, 61 dirichlet problem 175
delta function 13, 292 dirk gently’s holistic detective agency 538,
democritus 9, 10, 407 540
denmark 553 discrete 519
denominator 321 discrete delta-potential method 530
density 22, 309 discrete spectrum 274, 293, 523, 631
density functional theory 114, 365, 530 discworld 542
density matrix 22, 266, 362, 393, 398, 417, dislocation 608
A
497, 550, 560, 656, 657
density of states 89, 141, 383, 504
density operator 274, 487, 489, 490
dispersion 617
dispersion relation 139
distance 174
density state 594 distant anticipation 496
de rham cohomology 17 divergence theorem 401
derivation 356 doctoral thesis 348
derivative 48, 112, 307, 432 doctor who 541
desy 121 dolbeault operator 103
detailed balance 169 domain 157, 196
detector 255 don misener 24
DR
determinant 308, 322, 515, 556, 557 doppler 74
determinism 216, 293, 477, 478, 626 dot product 175
deuterium 129, 327 double-slit experiment 96, 214, 293, 473,
dewitt notation 354 474, 489, 494, 539, 617, 645, 646, 649
dielectric constant 509 double slit 643
differential equation 147, 306, 311, 332, double slit experiment 95
523, 607, 645, 649 douglas adams 538, 540
differential equations 280 driven harmonic motion 511
differential geometry 6, 269 drop 522
differential operator 102 dual space 41, 179, 354
diffraction 87, 322, 645, 646 duncan macinnes 552
diffraction pattern 86, 644 duration 624
diffusion 200, 348 duru-kleinert transformation 351
diffusion equation 352 dye laser 193
digital devil saga 543 dynamics 273, 612
dimension 122, 303, 413, 499 dyson series 371
dimensional analysis 73
dimensional regularization 78, 79 e
dimensionless 655 earthquake 573
dinitrogen trioxide 10 eccentricity vector 264
681

echolocation 522 electromagnetic potential 6


edward condon 498 electromagnetic radiation 380, 405, 458
edward witten 431 electromagnetic spectrum 128, 406
effective action 548 electromagnetic wave 11, 15, 55, 67, 461,
effective field theory 432 466, 646

FT
effective theory 604 electromagnetic waves 229
ehrenfest’s theorem 248 electromagnetism 213, 300, 328, 421, 461,
ehrenfest theorem 266 641
eigenequation 650 electron 4, 10, 12, 14, 15, 19, 30, 40, 45,
eigenfunction 280, 344, 650 51, 67, 67, 78, 86, 89, 90, 104, 104, 115,
eigenspace 631 117, 118, 120, 128, 133, 139, 141, 142,
eigenstate 2, 89, 116, 144, 187, 245, 293, 153, 182, 187, 211, 213, 234, 240, 266,
296, 310, 345, 423, 436, 442, 463, 494, 270, 271, 287, 297, 315, 327, 331, 344,
505, 565, 581, 596 353, 360, 364, 371, 373, 375, 388, 405,
eigenstates 182, 294 417, 458, 459, 460, 464, 467, 470, 473,
eigenvalue 2, 67, 89, 116, 168, 183, 259, 473, 474, 494, 498, 519, 557, 559, 580,
274, 280, 281, 292, 300, 306, 326, 343, 588, 588, 590, 591, 616, 645, 661, 664
467, 513, 558, 565, 631, 650 electron-degenerate matter 94
eigenvalues 308, 362, 390, 424 electron affinity 663
eigenvector 280, 282, 292, 452, 467, 558, electron charge 73
624 electron configuration 124, 187
eigenvectors 168, 169, 306, 308, 310, 579 electron correlation 150
eighteenth century 451 electron cryomicroscope 555
A
einstein 472, 480
einstein’s summation convention 103
einstein (disambiguation) 30
electron diffraction 474, 647
electron hole 13
electronic configuration 89, 664
einstein-podolsky-rosen paradox 358 electronic molecular hamiltonian 115, 299
elastic 320 electronics 8, 30, 133, 140, 287, 470, 628,
electric 315 662
electrical charge 407 electronic structure 591
electrical conductivity 140 electron mass 73
electrical engineer 507 electron microscope 247, 470
electrical engineering 140, 656 electron microscopy 145, 651
DR
electrical potential 4, 429 electron neutrino 118
electrical resistance 254, 440 electrons 79, 87, 498, 582, 586, 605, 647
electric charge 11, 15, 49, 75, 120, 300, 315, electron spin 185
327, 474, 527 electron volt 661
electric current 376, 647 electroscope 376
electric dipole moment 578, 613, 613 electrostatic force 14
electric discharge 15 electrostatic levitation 377
electric field 3, 120, 124, 145, 154, 316, electrostatics 260
364, 576, 578 electroweak force 121, 473
electricity 459 electroweak interaction 121
electric potential 115, 364 electroweak theory 329
electrochemistry 552, 588 elementary charge 116, 188, 440
electrode 315, 572 elementary particle 10, 408, 619
electrodynamics 324, 522 elementary particles 88, 605
electromagnetically induced transparency 26, elitzur-vaidman bomb-testing problem 109,
27 298
electromagnetic field 48, 61, 186, 314, 338, ellipse 261
421, 429, 432, 434, 473, 611 emanuel swedenborg 206
electromagnetic force 75 emission line 15, 128
electromagnetic interaction 328 emission spectrum 405, 459
682

empiricism 478 everett interpretation 295


energy 12, 15, 23, 31, 39, 42, 67, 127, 138, everett many-worlds interpretation 97, 471
139, 142, 146, 191, 268, 275, 285, 292, evolution operator 302, 630, 630
293, 295, 310, 332, 343, 344, 360, 363, exactly solvable model 524, 608
381, 436, 443, 460, 463, 466, 486, 494, exchange operator 151

FT
513, 519, 524, 581, 621, 624, 647, 661, 664 excited state 40, 117, 581
energy density 381 exciton 392
energy eigenfunctions 182 excitons 580
energy level 11, 19, 89, 124, 193, 300 exemplar 461
energy levels 182, 312, 586 existence of god 480
energy level splitting 89 exotic atom 190
energy operator 150 exotic baryon 129
energy spectrum 39 exotic meson 129
englert 62 expectation value 112, 173, 283, 594, 631
englert-greenberger duality 109 expected value 170, 274, 366, 368
enrico fermi 130, 131, 591 experiment 104, 196, 314, 315, 361, 619
ensemble 656 exponential function 31, 286
entangled 60, 465 extension 595
entangled state 471 extensive variable 165
entanglement 114, 214, 293, 534, 537, 650 extra dimensions 122
entanglement of formation 563 extreme point 274
entanglement witness 362, 420, 551
entities 196, 197 f
A
entropy 165, 291, 391, 486, 610, 611
epicurean 407
epistemology 479
f. david peat 206
f. kurlbaum 405
f. paschen 405
eponym 253 factorial 424
epr paradox 214, 217, 247, 293, 298, 298, fano resonance 510
414, 416, 456, 471, 621 faraday isolator 612
equation 129, 647 faster-than-light 644
equations of state 92, 93 fate 626
equilibrium 254, 309 fault tolerant 605
equipartition theorem 386, 617 feminist science fiction 541
DR
equivalence class 403 fermi’s golden rule 371
equivalence principle 126 fermi-dirac 290
equivalent 594 fermi-dirac statistics 33, 90, 138, 139, 140,
eric cornell 24 141, 291, 601
ernest marsden 163, 517 fermi energy 92, 131, 132, 135, 141, 661
ernest rutherford 10, 15, 163, 312, 388, 474, fermi gas 130, 140, 142, 601
516, 517, 588, 591 fermilab 616
ernst chladni 50 fermi liquid theory 141
ernst pringsheim 405 fermion 62, 73, 80, 90, 118, 118, 118, 122,
erwin schroedinger 477 130, 138, 138, 142, 246, 327, 327, 328,
erwin schrödinger 11, 54, 57, 226, 240, 271, 424, 433, 556, 589, 592, 640, 648
271, 365, 405, 460, 462, 494, 524, 532, fermionic condensate 27, 27
540, 596, 624 fermions 27, 30, 133, 141, 286, 312, 556,
estermann 648 557, 557, 559
euclidean quantum gravity 195 fermi surface 138, 139
euclidean space 174, 175 fermi temperature 135, 141
eugene wigner 175, 422, 613, 656 feshbach-fano partioning 130
euler-lagrange equation 48, 350, 354 feshbach resonance 25, 510
ev 139 feynman 468
event horizon 84 feynman diagram 186, 353, 364, 371
683

feynman path integral 162, 272, 425, 434 freezing 486


feynman slash notation 103 frequencies 407
fiction 644 frequency 50, 86, 87, 309, 373, 381, 509,
field emission 498 621, 647, 654
field emission display 145 frequency spectrum 310

FT
field equation 257 friction 8, 609
field strength 6 fritz london 266, 473
fine-structure constant 73, 75 fubini-study metric 404, 417, 420
fine structure 20, 192 fullerenes 88, 648
fine structure constant 19, 79, 184, 191, 440 functional 560
finite potential well 338, 529 functional analysis 174, 181, 268, 271, 280,
first-order phase transition 486 393, 430, 619
flash memory 145, 498 functional calculus 169
flatland 541 functional derivative 354
flatterland 541, 542 functional derivative operator 356
fluorescence 458 functional integral 348, 351, 353, 355
flutter 511 functional integration (neurobiology) 157
flux 197 functional integration (sociology) 157
fluxon 149 functional measure 354, 355
flux pinning 150 functional space 174
focal length 65 functions 282
focal plane 64 function space 177
fock space 153, 426, 521 fundamental field 511
A
fock state 56, 57, 60, 151, 152, 658
fokker-planck equation 597
folklore 279
fundamental force 118, 470
fundamental forces 615
fundamental group 607
force 155, 257, 517, 524 fundamental interaction 311, 433
force constant 442 fundamental particle 121, 511, 522
formalism 272 fundamental particles 196
formant 511 fundamental representation 78
fortune teller 496 futurama 541
fourier analysis 176, 311
fourier series 174, 232, 240, 344, 622 g
DR
fourier transform 74, 174, 280, 355, 428, g-parity 301
642, 657 g. johnstone stoney 588
fowler-nordheim equation 145 galilean relativity 315
fractional statistics 603 gallium arsenide 441, 504
frames of reference 315 gamma function 35, 166, 602
franck-hertz experiment 20 gamma ray 67, 211, 312
frank drake 192 gamma spectroscopy 71
frankfurt 585 garrett birkhoff 219
frank jewett 552 gas 8, 15
frank wilczek 77, 434 gas giants 509
fraunhofer diffraction 64, 645 gas in a box 27, 34, 35, 38, 141, 331, 382
frederick reines 474 gas in a harmonic trap 35, 450
frederik pohl 542 gauge anomaly 432
free electron gas 139 gauge boson 117, 118, 432, 433
free electron model 338, 342 gauge field 432
free energy 663 gauge group 433
free field 80, 82 gauge invariance 440
free field theory 421 gauge invariant 49
free particle 257, 337, 463, 529 gauge symmetry 356, 432
free states 169 gauge theory 257, 432, 434
684

gauge transform 404 gravitation 63


gauge transformation 432, 433 gravitational coupling constant 79
gauss’ law 74 gravitational field 63
gaussian 56, 528 gravitational singularities 194
gaussian distribution 622 gravitational wave 568

FT
gaussian integral 160, 160 gravitino 590
gautama buddha 9 graviton 118, 118, 122, 590
geiger counter 540 gravity 14, 23, 84, 118, 122, 141, 195, 260,
generalized coordinate 16 328, 392, 433, 439, 462, 470, 590, 648
generalized laguerre polynomials 183 gravity probe b 574
generalized momentum 11, 16 gravity wave 574
general relativity 85, 118, 122, 126, 226, greatest lower bound 631
360, 433, 459, 470, 574 greebo 540
general semantics 538 greek philosophers 9
general theory of relativity 462 greg egan 541
generating functional 355 gregory breit 45
geometric algebra 98, 102 grid 646
geometrical optics 430 ground state 40, 117, 127, 138, 185, 312,
geometric quantization 17, 272 423, 424, 486, 604, 631, 632
geophysical 522 group 323
george alec effinger 542 group representation 116
george chapline 84, 84 group velocity 87
george gamow 231, 498 gurps 544
A
george mackey 475
george paget thomson 648
george sudarshan 505
gustav ludwig hertz 153
göttingen university 279

geothermal energy 573 h


gerard ’t hooft 434 haag’s theorem 80, 211, 276
germanium whisker 608 haag-kastler 430
german physical society 405 hadron 76, 117, 119, 120, 129, 329, 591,
germany 265, 585 591
gev 77, 78 hadronic showers 265
ghz 293 hagen kleinert 111, 358, 475, 633
DR
gibbs paradox 30, 133, 286 hahn-banach theorem 125, 362
gigabyte 604 half-integer 139, 557
gilbert n. lewis 11 half-iterate 102
glauber p representation 660 halides 459
global minimum 632 hall effect 440
global symmetry 432 hamel basis 178
gluon 78, 118, 119, 120, 120, 364, 433, 470, hamilton-jacobi equation 169, 271
590 hamilton-jacobi equations 127
gluonic vacuum field 511 hamiltonian 49, 89, 266, 423, 578, 581, 665
gluons 584 hamiltonian mechanics 74, 112, 171, 172,
glycerine 200 173, 283, 309
gns construction 490 hanbury-brown and twiss effect 651
gold 10, 163, 517, 522, 572 handbells 51
google 203 hanns johst 537
gradient 528 han purple 413
graffiti 628 hans bethe 321, 476
grand canonical ensemble 290 hans geiger 163, 517
grand partition function 34 harmonic analysis 181, 280, 344
grand unification theory 121 harmonic function 391
gravastar 84, 85 harmonic oscillator 153, 310, 511, 596
685

harmonics 310 hilbert-schmidt 125


harmonic series (music) 310 hilbert space 39, 41, 42, 48, 74, 81, 114,
hartle-hawking state 194 151, 168, 213, 215, 268, 271, 273, 273,
hartree-fock 11, 151, 268, 530, 557 280, 314, 323, 348, 393, 393, 394, 403,
hartree-fock method 632 415, 422, 426, 452, 467, 490, 497, 521,

FT
hartree product 556 523, 524, 549, 559, 564, 609, 630, 631,
hartree theory 556 638, 650
hawking radiation 85, 439, 470, 568 hill potential 499
healer 496 hindu 9
heat 405 history of quantum field theory 434
heat death 611 history of science 8
heat kernel 103, 437 history of thermodynamics 12
heinrich hertz 226, 373 hole 580
heinrich rubens 405 hole theory 422
heisenberg 47, 277, 348, 565, 640, 660 holism 196
heisenberg’s microscope 621 hologram 200
heisenberg’s uncertainty principle 166 holographic principle 206
heisenberg compensator 628 holomorphic functional calculus 169
heisenberg group 48, 660 holomovement 197
heisenberg picture 112, 210, 275, 423, 532, holonomy 6
599 homodyne detection 55, 659
heisenberg uncertainty principle 96, 207 homogeneous space 403
heisenbug 628 homotopy group 607
A
helium 17, 142, 376, 468, 517, 573
hellsing 538, 542
hemt 504
hopf algebra 595
howard percy robertson 624
html 537
henri becquerel 473 hugh everett 474
henri poincaré 297, 472 husimi q representation 660
hera collider 121 hydrocyanic acid 540
hermann weyl 47, 175, 271, 271, 656 hydrogen 15, 182, 186, 262, 345, 518, 519,
hermitean 302 578, 589, 596
hermite polynomials 443 hydrogen-like atom 183, 263, 529
hermitian 17, 57, 113, 173, 281, 282, 283, hydrogen atom 16, 17, 40, 67, 128, 182,
DR
292, 393, 416, 467, 494, 558, 630 184, 187, 260, 267, 270, 364, 460, 468,
hermitian adjoint 259 469, 472, 488, 523, 529, 581
hermitian conjugate 41 hydrogen molecule 473
hermitian operator 168, 259, 365, 515, 631 hypercube 166
hertz 654 hypercylinder 166
heterostructure 331, 441 hyperfine structure 11, 20, 586
hidden-variable 273 hypernucleus 312
hidden variable 199 hyperon 312
hidden variables 207, 248, 650 hyperspace theory 439
hidden variable theory 414 hypersphere 166
hideki yukawa 591 hysteresis 255
hierarchy problem 121 hückel’s rule 514
higgs boson 118, 118, 120, 121, 121, 590,
591 i
higgs mechanism 121 i. a. richards 203
high-energy physics 102, 596 ian mcewan 542
high energy physics 589 ideal gas 34, 38, 139, 141, 165, 165, 291,
high frequency approximation 548 601
high temperature superconductivity 604 identical particles 24, 43, 424
hilary putnam 215 image 197, 200
686

imaginary number 189, 194 456, 471


imaginary unit 47 interstellar medium 192
immanuel kant 204, 206, 477, 478 intertwiner 595
impact ionization 13 introduction to quantum mechanics 459
impedance 511 invariant 325, 595

FT
index of refraction 321 invariant theory 52
indicator function 390 inverse-square law 517
indium 572 inversely proportional 86, 620
inductive reasoning 478 inverse scattering problem 522
industrial revolution 8 inverse scattering transform 524, 608
inequivalent 594 inversion layer 440
inertia 439 inversion of series 37
inertial frame 185 ion 91, 128, 187, 193, 338, 568
inert pair effect 20 ion-trap quantum computing 193
infimum 560 ionization energy 661
infinite divisibility 10, 375 ionosphere 2
infinite potential well 146, 149 iron 163
infinitesimal 433 irreducible representation 52, 325
influence 200 irreducible representations 323
information 414, 467, 470 ir spectroscopy 503
information retrieval 204 isaac newton 641, 644, 646
information theory 562 isidor rabi 586
infrared 255, 405 islam 9
A
infrared imaging 256
inhomogeneous 583
initial condition 308
islamic golden age 9
is logic empirical? 215
isomorphic 176
inner product 40, 41, 42, 69, 468, 639 isospin 312, 607
inner product space 174 isotope 25, 26, 142
inside joke 541 isotropic 182
instanton 548
institute for advanced study 552 j
insulator 254, 486 j. e. evron 3
integer 5, 24, 263, 440 j. j. thomson 388, 588, 591
DR
integrable 490 j.j. thomson 10, 517, 645
integral 157, 399, 649 j. robert oppenheimer 552
integrate 321 jack steinberger 327
integration 158 jacobi identity 174
intensity 321 jahn-teller effect 124
intentionality 203 jain 9
interacting boson model 313 james chadwick 11, 589, 591
interaction 75, 129, 521 james clerk maxwell 226, 374, 375, 610
interaction energy 39 james edward zimmerman 571
interaction picture 77, 80, 174, 276, 284, james franck 153
423, 431, 532 james jeans 386, 618
interference 25, 25, 26, 104, 226, 309, 350, january 17 554
468, 489, 494, 506, 571, 616, 645, 646 jean perrin 10
interferometer 63 johannes stark 577
internal conversion (chemistry) 211 john archibald wheeler 348, 553
internal conversion coefficient 212, 212 john c. slater 556
international space station 319 john cramer 219, 650
interpretation of quantum mechanics 22, john dalton 8, 9, 12
249, 294, 357, 360, 625 johnny mnemonic 574
interpretations of quantum mechanics 361, john searle 203
687

johnson-nyquist noise 409 landau pole 77


john stewart bell 484 lande interval rule 192
john von neumann 175, 219, 271, 271, 272, landé g-factor 665
284, 452, 467, 473 laplace 262
joseph john thomson 374, 473 laplace operator 528

FT
josephson constant 253 laplacian 102, 299
josephson effect 502 large hadron collider 122, 433, 590, 592
josephson junction 571 laser 56, 65, 193, 369, 470, 568
joule 620 laser cooling 25
joule second 230 latin 404, 460
julian schwinger 226, 545, 552 lattice 603
june 15 413 law of cosines 68
june 5 24 law of multiple proportions 10
laws 196
k lax pair 285
k. k. darrow 552 lead 163, 572
kanada 9, 407 lebesgue measure 168, 390
kaon 301, 302, 303, 329, 591, 630 leiden university 23
karl popper 203 lenz 262
karl pribram 201 leo kadanoff 434
keith mayes 110 leon lederman 328
kek 615 leon rosenfeld 422
kekulé 514 lepton 117, 118, 327, 590
A
kelvin 141, 255, 285, 413, 654
kenneth wilson 434
kinematics 612
leptons 119
leptoquark 591
lester germer 86, 87, 474
kinetic energy 13, 51, 86, 115, 373, 374, lester halbert germer 648
391, 469, 508, 517, 527 leucippus 9, 407
kinetic theory 8, 12 lev davidovich landau 142
kingdom of loathing 544 levi-civita symbol 50, 262
klaus von klitzing 440 lev landau 77
klein-gordon equation 157, 424, 469, 528, lhc 121
649 lie algebra 50, 264, 595
DR
klein-nishina formula 68 lie bracket 262, 264
klystrode 145 lie derivative 50
klystron 145 lie superalgebra 595, 597
knowledge 204 lifetime 510
kochen-specker theorem 273 light 104, 376, 459, 522, 617, 644, 646, 653
korteweg-de vries equation 608 light bulb 55
kristofer straub 544 lightbulb 654
kronecker delta 48, 426 light emission 518
light quantum 404
l light wave 565
ladder operators 79, 84 limits to computing 615
lagrange’s formula 260 lindblad equation 112
lagrange multipliers 32, 137, 288 linear 81, 527
lagrangian 48, 75, 76, 122, 349, 356, 425 linear algebra 43, 178, 259, 271, 280, 306,
lagrangian mechanics 309 310, 314
laguerre polynomials 183, 488, 596 linear combination 187, 558
lake 392 linear combination of atomic orbitals molecular
lamb-rutherford experiment 186 orbital method 187
lamb shift 186 linear functional 40, 41
laminar flow 206 linear operator 42, 173, 272, 280, 283, 426,
688

494 m
linear span 177 macromolecules 88, 495
line broadening 369 macroscopic 8, 297, 410, 461
line bundle 6, 17 magnet 603
line integral 348 magnetic dipole 191

FT
linus pauling 473 magnetic evaporative cooling 25
liouville equation 266 magnetic field 3, 20, 50, 124, 149, 185, 257,
liquid 8, 9, 142, 603 266, 440, 486, 504, 571, 576, 582, 663
liquid drop model 313 magnetic flux 4, 254
lisa mason 542 magnetic flux quantum 150, 253, 408
lise meitner 13 magnetic moment 185, 241, 246, 266, 266,
list of astronomical topics 71 364, 558
list of famous experiments 164 magnetic monopole 5, 546
list of mathematical topics in quantum theory magnetic quantum number 249, 530
280 magnetic resonance imaging 587
list of noise topics 409 magneto-optic effect 612
list of particles 123, 592 magnetoencephalography 573
list of physics topics 71, 222 magnetometer 571
list of quantum field theories 434 magnetoreception 411
lithium 20 magnon 25
local hidden variable theory 218, 471 majorana 327
locality 293 manga 538, 542
local maximum 392 manhattan project 552
A
local minimum 391
local realism 414, 627, 650
local symmetry 432
manifold 523
many-body theory 79, 422
many-minds interpretation 222
locc 563 many-worlds interpretation 218, 222, 279,
logarithmic derivative 347 298, 538, 544, 644
logic 478 many-worlds interpretation of quantum me-
london force 267 chanics 536
london moment 266 many worlds 248
loop expansion 548 many worlds interpretation 107, 649
lord rayleigh 386, 618 marshall stone 452
DR
lords and ladies 539 martin gutzwiller 412
lorentz covariant 423 martinus veltman 434
lorentz factor 509 maser 587
lorentz force 611 mass 72, 120, 163, 306, 332, 424, 517, 528,
lorentz force law 49 589, 601
lorentz group 595 mass gap 521
lorentzian function 510 mass number 589
lorentz invariance 173, 283 mass renormalization 302, 432
lorentz invariant 422, 657 mass shell 77
lorentz transformation 45, 264 master equation 112, 266
loschmidt’s paradox 611, 615 master of mosquiton 543
louis-victor de broglie 237, 647 material science 8
louis de broglie 86, 219, 226, 645 mathematical 74, 196, 459
lower bound 619 mathematical analysis 181
ls-coupling 665 mathematical formulation of quantum mechan-
ls coupling 547 ics 48, 168, 175, 298, 314, 398, 415, 442,
lsz formalism 521 467, 524
ludwig boltzmann 21, 385 mathematical physics 279, 522
lyman series 20, 128, 519 mathematician 507
léon brillouin 552 mathematics 40, 79, 102, 157, 174, 403,
689

522, 523, 619, 630, 633 middle ages 9


mathworld 656 mind 495
matrix mechanics 172, 174, 213, 240, 249, mind’s eye 206
271, 279, 462, 472, 622 minimum 506
matrix population models 285 minkowski space 102

FT
matter 8, 10, 22, 86, 90, 118, 118, 201, 225, mit opencourseware 476
269, 405, 470, 644 mixed state 112, 456, 594
max born 2, 22, 213, 225, 226, 244, 267, mixing paradox 165
271, 281, 400, 421, 460, 471, 498, 585 modes of vibration 50
max planck 11, 33, 226, 229, 372, 375, 380, molecular beam epitaxy 504
381, 405, 460, 472, 591, 618 molecular geometry 98, 116, 299
max tegmark 537 molecular hamiltonian 98, 116
maxwell 485 molecular physics 459
maxwell’s demon 542, 610, 615 molecule 8, 117, 127, 182, 191, 267, 299,
maxwell’s equations 609 522, 555
maxwell-boltzmann distribution 137 molecules 27, 124, 407, 604
maxwell-boltzmann statistics 32, 33, 138, 601 moment 266
maxwell equation 49 momentum 42, 44, 67, 72, 81, 86, 87, 138,
maxwell equations 528 139, 139, 142, 163, 165, 180, 240, 242,
may 31 413 260, 268, 292, 442, 463, 528, 565, 619,
may 6 652 619, 640, 647
mcscf 557 momentum operator 113
mean 282, 624 momentum space 139, 488, 636
A
measurable space 394
measure 161
measurement 214, 277, 414, 619
monochromatic 376
moon 377
morse potential 391, 450
measurement in quantum mechanics 310, mosfet 440
465, 582 mossbauer effect 313
measurement in quantum theory 437 mott insulator 26
measurement problem 222, 294, 298, 314 multiplicative quantum number 614
measurement theory 320, 548 muon 118, 119, 590, 591
measure space 177 muon neutrino 118
medical imaging 313, 522 murray gell-mann 226
DR
meissner effect 149 music 310, 407
mellin transform 35 musical instruments 509
memories 202 mutual information 560
mems 145
mendeleev 312 n
meson 117, 120, 129, 129, 311, 591 n-body problem 267
metal 139, 141, 142, 254, 266, 315, 662 nanometer 648, 653
metallic hydrogen 90 nanotesla 571
metaphysical 201, 297, 644 nathan rosen 455
meter 652 national academy of sciences 552
michael berry 412 national bureau of standards 328
michael fisher 434 natural unit 597
michael pepper 440 nearly-free electron model 342
michel houellebecq 117 neil gaiman 540
michio kaku 439 nethack 539, 543
micro black hole 379 neutralino 122
microscope 163, 316 neutrino 119, 301, 311, 327, 474, 557, 588,
microscopic 297 590, 591
microscopy 555 neutrino experiment 474
microwave 193, 193 neutrino mass 615
690

neutrino oscillation 303 normalization factor 556


neutron 11, 40, 63, 90, 117, 141, 142, 301, normalized wavefunction 403
311, 312, 327, 327, 470, 557, 584, 588, normal mode 425
589, 591, 591, 648 norman f. ramsey 587
neutron-degenerate matter 94 noumenon 206

FT
neutron interferometer 648 november 5 374, 519
neutron interferometry 63 np-hard 417
neutronium 90 nuclear compton scattering 67
neutron matter 94 nuclear engineering 313
neutron star 8, 92, 92, 93, 94, 141 nuclear fission 313
new scientist 253, 298, 476 nuclear force 313
newton 226, 478 nuclear fusion 91, 313
newton’s laws 477 nuclear magnetic dipole 191
newton’s laws of motion 8, 330, 612 nuclear magnetic resonance 46, 313
newton’s rings 108 nuclear medicine 313
newton’s second law 113, 464 nuclear physics 129, 422, 459, 498, 516,
newtonian physics 231, 494 588, 589
new york academy of sciences 552 nuclear power 313, 313
new york times 319 nuclear reaction 313
nickel 86, 87 nuclear spin 192
niels bohr 11, 15, 15, 51, 62, 153, 164, 182, nuclear structure 313
190, 226, 234, 237, 294, 297, 422, 460, nuclear technology 313
471, 477, 480, 496, 553, 619, 625 nuclear units 312
A
niels bohr institute 553
night court 628
nikola tesla 374
nuclear weapon 313
nuclei 115
nucleon 40, 120, 589, 591, 614
niobium 572 nucleus 11, 211
nist 506 nukees 544
nitrogen dioxide 10 numb3rs 543
nitrous oxide 10 number operator 152, 327
nobel prize 315, 375, 400 numerical 404
nobel prize for physics 86, 648
nobel prize in physics 25, 67, 153, 405, 440, o
DR
647 o.r. lummer 405
nobel prizes for physics 62 observable 168, 210, 215, 245, 273, 274,
no cloning theorem 415, 416 293, 390, 436, 451, 463, 594, 619
nodal line 185 observables 200, 281
nodal surface 185 observation 534, 619, 620
noether’s theorem 263, 300, 356 observer 196, 297, 315
noise-equivalent power 409 observer effect 619, 627
noise level 409 obsolete scientific theory 15
noise power 409 occam’s razor 594
non-linear optics 568 oil 315
noncommutative 239, 494 oil-drop experiment 474
noncommutative algebra 75 old quantum theory 269
nondestructive testing 522 one-dimensional periodic case 344
nonholonomic mapping 352 one-loop feynman diagram 548
nonlinear sigma model 354, 355 on shell 354, 356
nonlocal 257 ontological 196
nonlocality 358 open ball 175
norm 432 open quantum system 548
normalisable wavefunction 640 operator 48, 67, 79, 150, 259, 355, 393,
normalization constant 189 422, 467, 564, 630, 656
691

operator algebra 181 particle detector 591


operator ordering problem 351 particle in a box 156, 295, 338, 343, 345,
operators 282, 619 364, 382, 402, 428, 466, 504, 529, 620
optical aberration 498 particle in a one-dimensional lattice (periodic
optical fiber 580 potential) 529

FT
optical field 57 particle in a ring 331, 338, 529
optical mode 398 particle in a spherically symmetric potential
optical society of america 523 331, 529
optical theorem 432, 630 particle number operator 436
optics 55, 306, 310, 321, 412, 656 particle physics 117, 123, 129, 226, 301,
orbit 261, 264 312, 329, 365, 421, 422, 434, 459, 522,
orbital 117, 236 588, 589, 592, 607, 615
orbital angular momentum 185, 665 particles 142, 151, 197, 522
orbital elements 264 particle statistics 130
orbital resonance 509 particle zoo 123
ordinary differential equation 527 partite 167
ordinary differential equations 280 partition function 286, 288
organic chemistry 514 paschen-back effect 665, 665
orthodox 62 paschen series 128
orthogonal 282, 433 pascual jordan 281, 421
orthogonal group 404 path integral formulation 107, 468, 649
orthogonality 174 patient 496
orthogonal polynomials 174 pattern 646
A
orthonormal 168, 292
orthonormal basis 176, 177, 310, 452, 490,
639
patterns 197
paul dirac 40, 83, 102, 130, 131, 213, 226,
241, 246, 271, 271, 348, 350, 368, 421,
osama bin laden 538 445, 452, 460, 462, 467, 472, 480, 488
oscillator 282, 380 paul ehrenfest 112, 617
oskar klein 256, 258 paul gordan 52
otto stern 474, 582, 582, 585, 648 pauli 192
outer product 43 pauli exclusion principle 27, 30, 90, 91, 94,
overlap matrix 515 130, 133, 136, 139, 152, 246, 286, 422,
oxygen 10 427, 556, 557, 648
DR
pauli matrices 99, 103, 302, 558, 559, 593
p pauli principle 141, 556, 557
p. a. m. dirac 474 pbs 123, 512
parable of the cave 206 penning trap 506
parachronics 544 penrose interpretation 222, 494
paradigm 196 peres-horodecki criterion 551
paradox 298, 645 pericenter 261
parastatistics 33, 138 periodic function 304, 339
paravector 102 periodic table 187, 459
paravectors 98 periodic table of elements 312
parity (physics) 300, 613 permittivity 369
partial derivative 171, 524 permittivity of free space 75
partial differential equation 43, 126, 175, permutation 424, 640
177, 188, 269, 370, 522, 526, 527 perpendicular 190
partial differential equations 157, 280, 522 perturbation 89, 191
partial trace 315, 418, 561 perturbation theory 16, 45, 423, 578
particle 39, 113, 153, 303, 314, 453, 494, peter debye 71
497, 582, 588, 608, 619, 644 petrochemical 8
particle accelerator 509, 588, 589, 591 pharmaceutical 8
particle collider 122 phase distribution 567
692

phase noise 409 planck’s constant 16, 17, 47, 51, 69, 73, 87,
phase shift 3 173, 249, 269, 283, 350, 373, 380, 407,
phases of matter 8 440, 455, 463, 466, 499, 597, 601, 613,
phase space 17, 33, 47, 58, 165, 268, 270, 618, 620, 646, 647, 654, 662
273, 611, 656 planck’s law 618

FT
phase transition 486, 603, 607 planck’s law of black body radiation 33, 654,
phase transitions 434 654
phase velocity 310 planck length 73, 88, 379, 438, 649
phenomenon 266 planck mass 73, 379, 649
philosophical interpretation of classical physics planck time 438, 439
222 planetary mechanics 16
phonon 127, 365, 428, 450, 568 plane wave 156, 175, 257, 320, 321, 401,
phosphorescence 458 463, 529
photoelectric 230 plasma physics 1
photoelectric effect 68, 71, 106, 211, 269, plastic 267
382, 386, 472, 473, 645, 647, 662 plato 206
photoelectric emission 662 plum 388
photographic 200 plum-pudding model 517
photon 13, 16, 19, 23, 31, 40, 67, 73, 118, plum pudding 389
119, 120, 121, 127, 128, 211, 230, 297, plum pudding model 10, 163, 516
353, 364, 371, 373, 375, 382, 386, 396, poet laureate 537
404, 422, 424, 432, 433, 438, 458, 461, poincare group 301
469, 472, 519, 588, 590, 591, 618, 644, poincare symmetry 589, 592
A
645, 647, 662
photon number distribution 567
photons 33, 104, 106, 229, 230, 269, 312,
poison 532
poisson bracket 47, 112, 173, 262, 263, 283,
348
508, 535, 584, 605, 641, 648 poisson brackets 262
photosynthesis 411 poissonian 57, 60
physical cosmology 194, 548 poissonian distribution 58
physical limits to computing 611 polar coordinates 343, 512
physical review letters a 409 polarization 108, 369, 396, 421, 432, 585
physical system 634 polonium 10
physicist 84, 84, 213, 519, 524 polyatomic molecule 115
DR
physicists 87 polylogarithm 35, 384
physics 8, 12, 39, 47, 52, 62, 67, 74, 75, 79, polynomial 354, 391
86, 89, 124, 126, 130, 138, 153, 155, 157, polynomially bounded 353
167, 172, 178, 179, 311, 314, 315, 320, pontryagin duality 74
320, 322, 380, 382, 388, 404, 404, 410, position 72, 463, 619, 619
411, 434, 458, 485, 486, 486, 509, 514, position manifold 126
517, 522, 524, 546, 548, 555, 564, 601, position operator 292
603, 607, 612, 630, 641, 644, 652 positive linear functional 594
physics experiment 153 positive matrix 417
physics today 535 positron 40, 119, 311, 353, 424, 473
physics world 105 positronium 40, 190
pi 283 possibility 199
picometer 648 post hartree-fock 530
pier giorgio merli 104 postulate 349, 380
pierre victor auger 12, 13 potential 13, 113, 129, 188, 295, 310, 342,
pieter zeeman 665 345, 424, 498, 506
pilot wave 273 potential difference 154, 316
pin group 327 potential energy 330, 391, 468, 517, 527,
pion 311, 327, 591 596
pioneer plaque 192 potential energy surface 98, 116, 116, 299,
693

392 proton collapse experiment 474


potential theory 391 proton conductor 267
potential well 149, 504 proton decay 121
pound-rebka falling photon experiment 648 pseudoscalar 323, 328
power series 364, 443, 499, 633 pseudoscience 651

FT
power spectrum 310 psychic healing 496
precision 619, 619 psychokinesis 496
precognition 496 psychology 628
preon 94, 122 pudding 388
preon-degenerate matter 94 pure mathematics 268
preon matter 94 pure state 215, 393, 456, 491, 492, 549
preon star 94 purification of quantum state 393
pressure 90, 309 pythagoreans 407
primitive cell 661
princeton university press 474 q
principal quantum number 16, 19, 241, 248, qcd 522
530 qed 522
principle 196 q factor 193, 510, 511
principle of complementarity 62, 237 qft 355
principle of emergence 603 quadratic form 272
principle of locality 415 quadrupole 546
private investigator 541 quant-ph 507
probability 22, 146, 157, 310, 349, 400, 488, quanta 11, 80, 230, 281, 375, 422, 460
A
529, 546, 630, 640, 646, 649
probability amplitude 41, 52, 103, 468, 471,
488, 490, 630
quantization 48, 273, 277, 408
quantization condition 16
quantization of constrained systems 277
probability current 399, 528 quantization of gauge theories 277
probability density 184, 400, 460, 528, 529, quantized 586
635 quantum 459, 486, 656
probability density function 304, 399 quantum-mechanical 266
probability distribution 213, 215, 243, 451, quantum-mechanical circuit 253
463, 471, 481, 619, 622 quantum-well intermixing 504
probability flux 399, 528 quantum annealing 438
DR
probability measure 274 quantum chaos 468
probability theory 269 quantum chemistry 12, 79, 98, 114, 115,
proca equation 528 186, 226, 365, 391, 459, 470, 473, 522,
processes 197 522, 656
process physics 439 quantum chromodynamics 78, 79, 159, 364,
product rule 113, 173, 284 430, 433, 434, 470, 473, 614
professor walter ernhart-plank 474 quantum coherence 61, 109, 294
projection 67 quantum coin-flipping 395
projection-valued measure 278, 390, 398, 452 quantum communication 550
projection operator 43, 216, 418 quantum computation 494, 605
projective hilbert space 416, 551 quantum computer 254, 256, 408, 470
projective representation 323, 403 quantum computing 40, 536, 537, 605, 611,
projective space 403 615
proof of the existence of god 644 quantum cosmology 218, 320
propagator 353, 511 quantum cryptography 395, 408, 414, 470,
proquest 554 536
prospecting 573 quantum cybernetics 650
proton 11, 40, 67, 88, 117, 142, 182, 270, quantum decoherence 293, 296, 297, 472,
311, 312, 312, 327, 470, 520, 557, 584, 535, 541
588, 589, 591, 591, 648 quantum dot 392, 504
694

quantum electrodynamics 55, 75, 77, 79, quantum mechanics, philosophy and controver-
158, 186, 326, 353, 364, 432, 433, 434, sy 249
434, 440, 469, 473 quantum memory 605
quantum electronics 408 quantum metaphysics 222
quantum entanglement 66, 247, 456, 462, quantum mind 456

FT
467, 493, 549, 560, 603, 605 quantum money 395
quantum eraser experiment 95, 95, 109 quantum monte carlo 530
quantum field theory 2, 11, 39, 48, 61, 61, quantum noise 565
72, 75, 76, 77, 77, 78, 79, 83, 186, 258, quantum number 128, 182, 187, 189, 248,
272, 276, 295, 300, 302, 348, 352, 353, 300, 301, 408, 462, 487, 530, 582, 589
353, 420, 432, 445, 462, 469, 475, 510, quantum numbers 193, 241
522, 545, 548, 568, 581, 589, 590, 596, quantum operation 214, 217, 279, 314, 398
614, 614, 615 quantum optics 60, 61, 273, 320, 404, 569,
quantum field theory in curved spacetime 272 656
quantum fluctuation 438 quantum particle 392
quantum gravity 122, 360, 444, 462, 470, quantum physics 85, 104, 144, 314, 392,
548 436, 451, 519, 544, 619
quantum gyroscope 254, 256 quantum process tomography 497
quantum hall effect 504 quantum state 22, 40, 57, 91, 112, 117, 127,
quantum harmonic oscillator 54, 79, 80, 83, 128, 139, 151, 167, 167, 186, 187, 241,
259, 266, 277, 364, 426, 468, 469, 488, 293, 302, 344, 408, 414, 421, 424, 463,
490, 503, 529, 564 495, 497, 514, 521, 558, 603, 609, 614
quantum immortality 408 quantum states 24, 196, 324, 348, 613
A
quantum indeterminacy 66, 213, 222, 396,
628
quantum information 393, 395, 507, 551
quantum stationary hamilton-jacobi equation
126
quantum statistical mechanics 489, 624
quantum information theory 124, 167, 420, quantum suicide 537, 539
420, 592 quantum superposition 471
quantum jump 16, 19 quantum system 114
quantum leap 11, 270, 331 quantum teleportation 414, 415, 471
quantum logic 219, 272, 398, 467, 490 quantum theory 11, 51, 62, 86, 89, 310,
quantum magnet 486 405, 495, 588, 591, 608
quantum measurement 272, 398, 451, 456, quantum trajectory representation theories of
DR
650 quantum mechanics 126
quantum mechanic 506, 543 quantum tunneling 145
quantum mechanical 345 quantum tunnelling 149
quantum mechanics 2, 13, 15, 22, 23, 39, quantum well 392
40, 41, 45, 51, 52, 54, 67, 72, 73, 74, 84, quantum wire 392, 504
85, 87, 97, 98, 102, 104, 109, 114, 115, quark 40, 78, 78, 94, 117, 118, 119, 120,
122, 126, 127, 146, 150, 151, 153, 153, 129, 316, 364, 470, 588, 590
157, 172, 172, 174, 174, 176, 180, 182, quark-degenerate matter 94
186, 194, 207, 210, 212, 222, 224, 248, quark matter 94
259, 260, 264, 266, 267, 281, 283, 284, quark model 129, 591, 591, 592
291, 296, 298, 302, 303, 306, 306, 310, quarks 118, 557, 584
312, 320, 324, 324, 330, 338, 343, 345, quark star 93, 94
348, 348, 352, 363, 372, 382, 386, 390, quartz 374
391, 399, 400, 403, 404, 405, 408, 409, quasar 509
410, 411, 412, 421, 423, 425, 438, 451, quasiparticle 25, 142, 422, 603
456, 459, 487, 489, 494, 507, 512, 522, qubit 193, 254, 404, 490, 537
531, 532, 540, 546, 548, 549, 556, 557, qubits 563
557, 559, 564, 581, 582, 589, 594, 596,
596, 612, 613, 614, 617, 619, 621, 630, r
631, 634, 640, 640, 644, 645, 649, 656 r. a. fairthorne 203
695

r.e. siday 3 relativistic mass 509


r. f. streater 280 relativistic plasma 509
r. neill graham 474 relativistic quantum field theory 295
radar 74 relativistic wave equations 509, 528
radar ambiguity function 74 relativity 415

FT
radian 232 relativity theory 45
radiant energy 374 relic particles 590
radiation 269 renaissance 407
radiation therapy 71 renninger negative-result experiment 298
radioactive 498 renormalization 79, 430, 437
radioactive decay 211, 313 renormalization group 348, 434
radioactive nuclei 39 rené descartes 478
radioactivity 312, 313, 473 representation of a hopf algebra 595
radiobiology 71 representation theory 52, 656
radio drama 541 reproducing kernel hilbert space 181
radio wave 522 residue theorem 387
radius 317 resolution 293
rain 522 resolution of the identity 367, 527
rainbow 522, 646 resonance 39, 40, 127, 130, 399
ralph kronig 246, 342 resonance (disambiguation) 509
raman transition 193 resonator 511
random 465 rest energy 19
random variables 560 rest mass 87, 508
A
range criterion 551
rationalism 478
ray 168
reversible computing 611, 615
richard courant 279
richard feynman 63, 107, 111, 213, 226,
rayleigh-jeans law 617 318, 348, 552
rayleigh scattering 653 richard p. feynman 474, 475, 484
ray tracing 658 riemannian manifold 102
reality 196, 469 riemann zeta function 35
real line 635 riesz representation theorem 41, 125, 179,
real number 168, 194, 213, 525 179
reciprocal 354 rigged hilbert space 44, 175, 181, 277
DR
recorded 203 right triangle 407
redshift 71 ring wave guide 529
reduced mass 189, 442 ripple tank 65, 104
reduced planck constant 189, 275 ritz method 632
reducible 594 robert a. heinlein 541
reduction criterion 551 robert andrews millikan 375
reductionism 196, 470 robert anton wilson 538, 542
redundancy 145 robert b. laughlin 441
reeh-schlieder theorem 420 robert b. leighton 474
reference frame 626 robert brown 10
refraction 646 robert h. scanlan 511
reginald cahill 439 robert millikan 315, 474
region 200 robert oppenheimer 422
relationship between string theory and quan- robertson-schrödinger relation 624
tum field theory 434 rockefeller foundation 553
relative state interpretation 279, 315, 465 roleplaying game 544
relativistic 39, 173, 283 ronald gurney 498
relativistic beaming 509 roothaan equations 150, 151
relativistic breit-wigner distribution 511 rotation 266, 322, 323
relativistic jet 509 rotational symmetry 183
696

rotation group 263, 264 schwinger-dyson equation 354, 355, 356, 434
rotation operator 170 science 8
royal society of london 106 science fiction 542, 574, 628
roy j. glauber 55, 60 scientific notation 230
rubidium 22, 25, 193 scientific rigor 644

FT
rudolf grimm 27 screen 200
rule of thumb 112 screw dislocation 608
rumor 538 sea level 392
rutherford 474, 645 sea water 522
rutherford backscattering 517 second 193, 620
rutherford model 388 second law of thermodynamics 609, 610, 615
rutherford scattering 522 second quantization 79, 422
rydberg-ritz combination principle 518, 521 secular equation 515
rydberg constant 235, 519, 520 segre embedding 417, 549, 551
rydberg formula 15, 19, 128 segre mapping 404
seismology 656
s selenium 373
s-matrix 39, 510, 630 self-adjoint 453
sackur-tetrode equation 38, 165, 167 self-adjoint operator 48, 168, 216, 274, 314,
saint-germain 543 393, 452, 524, 525, 594, 595
samsara 206 self adjoint operator 273, 292, 390
saturnine martial & lunatic 543 semantics 217
satyendra nath bose 23, 33, 34, 386 semiclassical 320
A
scalar 321, 323, 528
scalar field 78, 126, 526
scalar potential 49
semiclassical gravity 548
semiconductor 8, 13, 140, 141, 392, 498, 662
semidefinite programming 296
scale anomaly 77 seminar 498
scanning squid microscope 573 separability 563
scanning tunnelling microscope 498, 502 separable 422
scattering 467, 522, 522 separable states 420
scattering amplitude 320, 321 separation of variables 188, 280, 335, 336
scattering theory 320, 521, 651 sequence 175, 175
schottky diode 662 sequence space 176
DR
schroedinbug 539 set 176
schroedinger equation 49, 522, 523, 645, 649 shamanism 203
schrödinger 16, 348, 352 shannon entropy 418
schrödinger’s cat 297, 298, 494, 540 shell model 313
schrödinger’s cat in fiction 539 shin megami tensei 543
schrödinger’s cat trilogy 542 shot noise 60, 409, 441
schrödinger’s equation 171, 246, 249, 656 si 4, 315, 380
schrödinger equation 6, 11, 20, 43, 54, 80, signal-to-noise ratio 409
112, 127, 156, 169, 173, 174, 182, 188, signals 198
210, 211, 214, 218, 256, 260, 275, 277, simple harmonic motion 511
283, 284, 294, 295, 296, 298, 299, 303, simultaneous equation 366
306, 310, 310, 330, 331, 339, 343, 345, sinclair ql 458
350, 364, 399, 402, 423, 443, 464, 467, sine-gordon equation 608
499, 512, 521, 596, 640, 640 sine curve 232
schrödinger picture 172, 173, 210, 275, 283, singularity 84
369, 423, 530 sir roger penrose 360
schumann resonance 511 skyrmion 607
schwarzian derivative 126, 127 slac 615
schwarzschild 578 slater-type orbital 187
schwarzschild radius 73, 379 slater determinant 114, 151
697

slepton 122 spherical harmonics 54, 182, 488


sliders 538, 541 spin 45, 52, 396
slow glass 27 spin-dirac operator 103
slow light 27 spin-orbital 556
s matrix 521, 523, 523 spin-orbit coupling 185

FT
smoke detector 313 spin-orbit interaction 46, 665
so(3) 50, 52, 263 spin angular momentum 665
so(3,1) 264 spinning 266
so(4) 264 spinor 45, 102, 323, 327, 558, 559, 613
sobolev space 177 spin quantum number 246, 249, 530
social science 628 spin singlet 455
solar cell 376 spin statistics theorem 557, 592, 614
solar mass 92, 92 spiral galaxy 439
solar power 376 spontaneous parametric down conversion 96
solar system 14, 509 spontaneous symmetry breaking 121
solenoid 3, 4 spooky action at a distance 650
solid 8, 30, 90, 133, 287, 428, 450, 603 square-integrable 467
solid helium 486 square integrable 41, 399
solid state physics 450, 608 square potential 501
soliton 364, 607 squark 122
solvay conference 553 squashed entanglement 420
sommerfeld-wilson-ishiwara quantization squeezed coherent state 58
269, 270 squeezing operator 567
A
sound 410
sound wave 522, 621
source field 354
squid 254, 502
squid (disambiguation) 571
stability 506
space 467 stable 39
space-like 422 standard deviation 242, 624
space-time 196, 353, 432 standard model 117, 118, 121, 123, 158,
spacecraft 377 249, 327, 328, 329, 421, 422, 434, 522,
space quantization 586 590, 591, 592, 615, 615
spacetime 22, 122, 194, 285, 425 standing wave 231, 309
spallation neutron source 574 standing waves 407
DR
spark gap 373 stanford encyclopedia of philosophy 220, 220
sparticle 118, 122 star 92
special orthogonal group 323 stargate sg-1 538, 541, 628
special relativity 51, 72, 102, 185, 256, 315, stark broadening 577
424, 462, 467, 469, 473, 508, 509, 589, 625 stark effect 16, 20, 124, 364, 663, 666
special unitary group 323 star trek 628
spectral line 71, 182, 364, 369, 472, 576, 663 state 22
spectral measure 169, 456, 490 state vector 172, 210, 283, 494
spectral theorem 292, 416, 417, 452 static 196
spectral theory 272, 274, 280 stationary point 350
spectrometer 459 statistical 196
spectroscopy 19, 229, 270, 281, 412 statistical ensemble 314, 490
spectrum 168, 523 statistical mechanics 23, 30, 130, 165, 269,
speculative fiction 644 285, 320, 348, 352, 353, 355, 383, 391,
speed 508, 509 470, 529, 597, 617, 656
speed of light 49, 69, 72, 75, 87, 185, 192, statistics 30
414, 508, 509, 654 stefan-boltzmann law 385
sphere 139, 266 stellar mass black hole 85
spherical coordinates 183, 188 stephen donaldson 628
spherical harmonic 52, 183, 190, 345, 345 stephen hawking 194, 537
698

stephen notley 544 superconductors 253


stern-gerlach experiment 474, 558 superdense coding 420
steve martin 542 superfield 600
steven chu 25 superfluid 24, 27, 85, 142, 410, 428, 461,
steven weinberg 327, 535 603

FT
stiffness 306 superfluidity 26, 248
stimulated emission 458 supergravity 433, 590
stirling’s approximation 32, 137, 166, 288 supernova 26
stochastic 495 superposition 114, 293, 308, 309, 310, 360,
stochastic process 175 469, 505, 537
stokes’ law 317 superposition principle 494
stokes theorem 4, 6 superpotential 598
stone’s theorem 275 superselection 114
stone’s theorem on one-parameter unitary superselection sector 275
groups 275 supersolid 27
stone-von neumann theorem 48, 173, 277, supersymmetry 122, 122, 274, 356, 433, 590
283 surveyor program 377
strange matter 90, 93 switch 145
strangeness 312 symmetric 188, 640
strange quark 93 symmetry 77, 89, 345
stream 197 symmetry of second derivatives 386
string-net condensation 603, 605 symplectic form 17
string theory 61, 78, 122, 285, 433, 462, 470 symposium 410
A
strong cp problem 615
strong cp violation 614
strong force 590, 591
synchronicity 496
synchrotron radiation 15
synonym 462
strong interaction 76, 119, 120, 121, 129, system 414
329
strong interactions 328, 614 t
strong nuclear force 469 t-symmetry 301
structures 197 table of clebsch-gordan coefficients 53
stuart kauffman 205 tacoma narrows bridge 511
sturm-liouville theory 280, 311 talbot lau interferometer 648
DR
su(2) 433, 607 tangent bundle 159
su(3) 78, 433 taoism 206
sub-poissonian 567 target manifold 356
subalgebra 594, 595 target space 355
subatomic particle 21, 123, 129, 379, 408 tau lepton 119, 590
subatomic particles 10, 225 tau neutrino 118
submarine 522 taylor series 356
subspace 180 tears for fears 539, 543
substance 200 technology 470
sudarshan-glauber p representation 61 telepathy 496
sun 92, 406, 588, 653 teleportation 644
sunlight 522 temperature 22, 31, 32, 34, 132, 137, 140,
sunyaev zel’dovich effect 71 141, 142, 286, 288, 413, 440, 486, 581,
super-consciousness 644 601, 652
super-kamiokande 121 tensor 323
supercommutator 599 tensor category 603
superconducting 571 tensor product 43, 52, 151, 274, 415, 549
superconductive 266 terminal velocity 317
superconductivity 27, 365, 421 terry pratchett 539, 540, 542
superconductor 93, 266, 486, 498 tetraphenylporphyrin 648
699

the bohr model 182 tomography 659


the cat who walks through walls 541 tonks-girardeau gas 26, 27
the coming of the quantum cats 542 topological defect 607
the compass rose 541 topological entropy 603
the elegant universe 123 topological invariant 607

FT
the feynman lectures on physics 474 topological order 441, 604
the homing pigeons 542 topological quantum field theory 604
the last hero 539, 542 topological vector space 175, 176
the meaning of meaning 203 topologies on the set of operators on a hilbert
theorem 344, 619, 656 space 181
theoretical physics 74, 285, 434, 459, 596, torus 333
630 trace-class 124
theory 8, 198, 269 trace class 168, 274, 417, 550
theory of everything 196 trajectory 645
theory of relativity 226, 480 transactional interpretation 219, 222, 298,
thermal de broglie wavelength 30, 132, 286, 650
291 transduction 509
thermal equilibrium 30, 136, 285 transformation law 315
thermal light 56 transistor 470, 605
thermal noise 255, 409 transition radiation 509
thermal wavelength 35 transition radiation detector 509
thermionic emission 145, 662, 663 transition rate 11
thermodynamics 8, 140, 420 transition rule 664
A
the trick top hat 542
the universe next door 542
thin film 486
translational invariance 355
translationally invariant 354
transmission electron microscopy 555
things 197 trap 568
thinkgeek 539 tuned circuit 511
third law of thermodynamics 581 tungsten 662
thomas-fermi approximation 34 tunnel diode 502
thomas kuhn 205 tunneling time 364
thomas young 226 twentieth century 412
thomson scattering 67, 71, 73 two-photon generation 567
DR
thought 196 two lumps 544
thought experiment 269, 532, 540, 610, 625 type ii superconductor 149
three-body force 129 type i superconductor 149
tidal force 392
tidal resonance 509 u
time 169, 188, 310, 523, 531, 621 u(1) 433
time-resolved spectroscopy 510 ultraviolet 506
time evolution 214, 217, 273, 275, 279 ultraviolet catastrophe 269, 386, 641
timeline of chemical element discovery 12 ultraviolet divergence 617
timeline of cosmic microwave background as- ultraviolet light 374
tronomy 71 ultraviolet radiation 371, 374
timeline of quantum mechanics, molecular umklapp scattering 143
physics, atomic physics, nuclear physics, and unbounded operator 169
particle physics 12 uncertainty 619
timeline of thermodynamics, statistical mechan- uncertainty principle 47, 51, 63, 77, 94, 190,
ics, and random processes 12 242, 248, 275, 283, 335, 348, 396, 425,
time ordered 353 436, 438, 461, 462, 463, 466, 468, 472,
time ordering 355 480, 494, 564, 624, 640, 657
time reversal invariance 301 uncertainty relation 57, 77, 565
tolman-oppenheimer-volkoff limit 93 uncountable set 429
700

undergraduate 557 vector space 175, 461


unit 315, 594 velocity 87, 257, 497
unital 594 vernacular 458
unitarity 630 vertex renormalization 432
unitary group 404 vibrational spectroscopy 310

FT
unitary operator 48, 169, 214, 324, 391, vienna 88
419, 581 virial theorem 18
unitary representation 594 virtual particle 353, 436, 438, 438, 582, 624
unitary representation of a star lie superalgebra virtual particles 77, 438
595 viscosity 24, 317
unitary transformation 98 viscous 200
unit vector 524 vladimir fock 422
universe 21, 117, 194, 498 voigt 578
university of aberdeen 648 volt 254, 377
university of colorado at boulder 22, 27 volume 381
university of frankfurt 585 von-karman boundary condition 337
university of innsbruck 27 von klitzing constant 440
university of manchester 163 von neumann 294, 475, 620
university of tübingen 104 von neumann entropy 419
university of vienna 648 vortex 197
unobservables 206 voyager golden record 192
unruh effect 568 vulgar fraction 440
unsolved problems in physics 222, 469, 615
A
unstable 39
unstable particle 510
upper bound 13
w
walter gerlach 474
walter heitler 473
upthrust 317 walter ritz 514
ursula k. le guin 541 walther bothe 71
uv 376 walther gerlach 582, 582, 585
w and z bosons 118, 120, 120, 584
v ward-takahashi identity 356
v. a. fock 2, 151, 153 water 104
vacuum 316, 376, 519, 520 wave 67, 86, 104, 225, 310, 320, 321, 376,
DR
vacuum energy 439 463, 494, 511, 620, 644
vacuum expectation value 61, 353 wave-function renormalization 431
vacuum fluctuation 186 wave-particle duality 26, 56, 62, 66, 86, 86,
vacuum state 61, 82, 326, 426, 581, 582 96, 104, 237, 271, 348, 371, 461, 462, 463,
vacuum tube 459, 662, 662 467, 532, 608, 617, 620
vaisheshika 407 wave equation 257, 522, 641
valence shell 187 wave function 63, 71, 98, 271, 281, 284,
valentine bargmann 13 296, 303, 312, 314, 343, 399, 400, 403,
vapour pressure 316 432, 512, 524, 565, 645
variance 58, 162 wavefunction 22, 30, 44, 79, 80, 114, 128,
variational method 514 132, 152, 175, 183, 189, 195, 211, 213,
variational method (quantum mechanics) 633 214, 286, 297, 306, 310, 310, 332, 460,
variational parameters 633 487, 488, 490, 494, 505, 526, 533, 556,
variational perturbation theory 364 559, 656
variational principle 530, 631 wavefunction collapse 218, 222, 245, 293,
vector 262, 282 294, 295, 297, 298, 464, 469, 533, 620
vector boson 326 wave functions 461
vector bundle 102 wavelength 67, 86, 87, 93, 108, 170, 381,
vector mesons 584 406, 463, 495, 498, 519, 519, 520, 647, 652
vector potential 49, 429, 598 wavelet 176
701

wave mechanics 16, 263, 271, 281, 462, 472 wigner function 57
wavenumber 87 wiki 507
wave packet 56, 399, 464, 494, 565, 640 wild arms 3 543
wavepacket 565 wilhelm wien 381, 654, 654
waves 522 william chinowsky 327

FT
wave vector 156 william d. phillips 25
w boson 590 willoughby smith 373
weak decay 615 winding number 607, 608
weak force 590 wiretap 534
weak gauge boson 121 wkb approximation 365, 502, 530, 548, 597
weak interaction 121, 329, 591 wojciech h. zurek 114
weak interactions 328 wolfgang ketterle 25, 27
weak measurement 297 wolfgang pauli 11, 226, 246, 263, 422, 460,
weak nuclear force 469, 470, 609 552, 591
webcomic 544 work function 371, 372, 373
well-behaved 176 wormhole 438, 439
wendell furry 422 w state 167
werner ehrenberg 3
werner heisenberg 62, 226, 238, 271, 271, x
281, 281, 320, 405, 421, 436, 460, 462, x-ray 13, 67, 86, 255, 376
480, 619 x-rays 371
wess-zumino-witten model 608 x boson 121
weyl 475 x ray 316
A
weyl quantization 47
wheeler’s delayed choice experiment 95
white 653
y
ybco 573
white dwarf 30, 90, 91, 92, 94, 133, 141, 286 yukawa interaction 73
white dwarf material 94
white noise 409 z
whole number 10 z boson 77, 590
wick rotation 195, 350, 352, 353, 354, 356 zeeman effect 20, 50, 124, 228, 557, 576
wiener measure 158 zero-point energy 331, 438, 439, 444
wigner’s friend 220, 537 zig-zag 351
DR
wigner 3-j symbol 54 zinc sulfide 163
wigner distribution 656 zorn’s lemma 178
DR
AFT

Das könnte Ihnen auch gefallen