This document was uploaded by user and they confirmed that they have the permission to share
it. If you are author or own the copyright of this book, please report to us by using this DMCA
report form. Report DMCA
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 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 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) 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 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 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 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 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 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 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 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 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 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 ature superconductivity high-energy physics Hilary Putnam Hilbert space Hilbert-Schmidt hill potential Hindu History of nuclear weapons 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 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 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 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 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 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 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 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 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 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 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) 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 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 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) 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 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 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 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 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 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 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 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 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 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 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 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 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 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
ics Bound state Bra-ket notation Breit equation t Clebsch-Gordan coefficients Coherent state ables Compton scattering Compton wavelength ion and annihilation operators Dark energy star energy level Degenerate matter Delayed choice cal space Dirac operator Double-slit experiment ctronic density Electronic Hamiltonian Electronic ess EP Quantum Mechanics Excited state Exotic gas Fermi liquid Fermi's golden rule Fermi-Dirac ock matrix Fock space Fock state Franck-Hertz ger-Marsden experiment Gibbs paradox nics) Heisenberg picture Hilbert space Hydrogen plicate and Explicate Order according to David Internal conversion Interpretation of quantum ffect Klein-Gordon equation Ladder operators d equation London moment Many-body problem nics Matrix model Maxwell-Boltzmann statistics Multiplicative quantum number Neutral particle physics Observable Oil-drop experiment Open ox Particle in Adiabatic a one-dimensional lattice (periodic invariant ic potential Path integral formulation Penrose uantum mechanics) Photoelectric effect Planck Plum pudding model Position operator Potential Zeeman effect ility current Projective Hilbert space Pure gauge ology Quantum chaos Quantum Critical Point on Quantum foam Quantum Hall effect Quantum Quantum level Quantum mechanics Quantum antum phase transition Quantum solid Quantum Quantum Theory Parallels to Consciousness uantum well Quantum Zeno effect Quasistability guide Ritz method Rutherford model Rutherford ory Schrödinger equation Schrödinger picture riational principle Selection rule Semiclassical struction Slater determinant Spin-1/2 Spin-orbital effect Stationary state Stern-Gerlach experiment upersymmetric quantum mechanics T-symmetry antum number Transformation theory (quantum To Scilesco astrophe Uncertainty principle Unitarity Unitarity rbation theory Wave packet Wave-particle duality istribution Wigner-Eckart theorem Work function
Quantum Mechanics
FT
Quantum Mechanics
DR A
Adiabatic invariant -
Zeeman effect
compiled by To Scilesco
Quantum Mechanics Compiled by: To Scilesco BookId: dcchcaruqoqeqfro
FT
Date: 14.07.2006
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 determination of the principal authors.
DR A
Logo design by Joerg Pelka
Printed by InstaBook Corporation (instabook.net)
Published by pediapress.com a service offered by brainbot technologies AG , Mainz, Germany
An adiabatic invariant in general is a property of motion which is conserved 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 motion. In plasma physics there are three adiabatic invariants of charged particle motion.
The first adiabatic invariant, µ The magnetic moment of a gyrating particle, µ=
1 2 2 mv⊥
B
,
DR A
is a constant of the motion (as long as q/m does not change). In fact, it is invariant 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 the other hand, if the collision frequency is larger than the pump frequency, 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 condition for adiabatic invariance is violated and heating is possible. In particular, 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
FT
The longitudinal invariant of a particle trapped in a magnetic mirror, Rb J = a v|| ds, where the integral is between the two turning points, is also an adiabatic invariant. 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 invariant, 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.
DR A
External links • •
lecture notes on the second adiabatic invariant 1 lecture notes on the third adiabatic invariant 2
Principal Authors: Art Carlson, MathMartin, SimonP, Gurch, Linas
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 perturbation is acting on it slowly enough and if there is a gap between the eigenvalue and the rest of the Hamiltonian’s spectrum.
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 perturbation 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! 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
Principal Authors: Artur adib, Deco, SeventyThree, Conscious, Charles Matthews, BeteNoir
DR A
Aharonov-Bohm effect
The Aharonov-Bohm effect, sometimes called the Ehrenberg-SidayAharonov-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 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
FT
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 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
DR A
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 R φ = ~q P A · dx, 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 =
qΦ ~ .
This phase difference can be observed by placing a shielded solenoid between the slits of a double-slit experiment (or equivalent). A shielded solenoid encloses 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 superconducting loops. This quantization is due to the fact that the superconducting 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
DR A
FT
5
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 mathematical singularity in the vector potential, which can be expressed as an infinitely long Dirac string of infinitesimal diameter that contains the equivalent 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 chargequantization: 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 Osakabe 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
FT
Just as the phase of the wave function depends upon the magnetic vector potential, it also depends upon the scalar electric potential. By constructing a situation in which the electrostatic potential varies for two paths of a particle, through regions of zero electric field, an observable Aharonov-Bohm interference 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.
DR A
The initial theoretical proposal for this effect suggested an experiment where charges pass through conducting cylinders along two paths, which shield the 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.
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
•
•
• • •
DR A
• •
Aharonov, Y. and D. Bohm, "Significance of electromagnetic potentials in quantum theory," Phys. Rev. 115, 485–491 (1959). 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 nanotubes", 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 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 (SpringerVerlag: 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).
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, 2696 (1985).
Source: http://en.wikipedia.org/wiki/Aharonov-Bohm_effect Principal Authors: Stevenj, Linas, Liontooth, Reddi, CYD
Atomic theory
DR A
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 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 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 Main article: Atomism
FT
Early atomism
DR A
From the 6th century BC, Hindu, Buddhist and Jaina philosophers in ancient India developed the earliest atomic theories. The first philosopher who formulated 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. Indian 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 modern 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, presented 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, 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 experimentation. 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 further 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 single, unique type, and that although their shape and structure was immutable, Atomic theory
10
FT
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 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 oxygen (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
DR A
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 about, and developed a mathematical theory around it. This theory was validated 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 division 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 subatomic particles and are therefore divisible, and not the indivisible atomos postulated by Democritus. Physicists later invented a new term for such indivisible 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 other 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 volume, 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
FT
experimenters expected all the alpha particles to pass through without significant 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 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 experimentation 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.
DR A
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 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 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 determine 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 account for chemical phenomena such as bonding, rather than physical phenomena such as atomic spectra. Atomic theory
12
See also History of thermodynamics Kinetic theory Development of Quantum Theory Quantum Chemistry John Dalton
Related lists • •
FT
• • • • •
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
Principal Authors: Vsmith, Brighterorange, Ragesoss, Voyajer, Karol Langner, Linas, Rho, Timmy2, Dustimagic, Eric Forste
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
13 emitted photon, the energy can also be transferred to another electron, which is then ejected from the atom.
FT
Upon ejection the kinetic energy of the Auger electron corresponds to the difference between the energy of the initial electronic transition and the ionization 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 emission 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.
DR A
History
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.
Principal Authors: AjAldous, Keenan Pepper, Srleffler, Stokerm, Tristanb
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 R 1 2m ∞ Nl ≤ 2l+1 r|V (r)|V <0 dr ~2 0 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)
Principal Authors: TobinFricke, Covington, Amalas, Charles Matthews, Pjacobi
DR A
Bohr model
Figure 2
The Bohr model of the atom
In atomic physics, the Bohr model depicts the atom as a small, positively 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
FT
Introduced by Niels Bohr in 1913, the model’s key success was in explaining 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. The Bohr model is a primitive model of the hydrogen atom which cannot explain 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 simplicity, the Bohr model is still commonly taught to introduce students to quantum mechanics.
History
DR A
In the early 20th century, experiments by Ernest Rutherford and others had established that atoms consisted of a diffuse cloud of negatively charged electrons surrounding a small, dense, positively charged nucleus. Given this experimental 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 has several difficulties, the most serious of which is the loss of energy by synchrotron radiation.That is, an accelerating electric charge emits electromagnetic waves which carry energy; thus, with each orbit around the nucleus, the electron would radiate away a bit of its orbital energy, gradually spiralling inwards to the nucleus until the atom was no more. A quick calculation shows that this would happen almost instantly; thus, the naive planetary theory cannot explain why atoms are extremely long-lived. The naive planetary model also failed to explain atomic spectra, the observed discrete spectrum of light emitted by electrically excited atoms. Late 19th century experiments with electric discharges through various low-pressure gasses in evacuated glass tubes had shown that atoms will emit light (that is, electromagnetic 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 difference is carried off (or supplied) by a single quantum of light (called a Bohr model
16
FT
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 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.
DR A
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 classical mechanics treatment. The Bohr model is certainly not a full quantum mechanical description of the atom. Assumption 2) states that the laws of classical 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.
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 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 application 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. Calculations based on the Bohr-Sommerfeld model were able to accurately explain a number of more complex atomic spectral effects. For example, up to firstorder 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.
FT
The Bohr-Sommerfeld quantization condition as first formulated can be viewed 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
ˇ 2 (M, Z) → H ˇ 2 (M, R) → H 2 (M, R) , where the first map is the homoH DR
ˇ morphism of Cech cohomology groups induced by the inclusion of the integers ˇ in the reals, and the second map is the natural isomorphism between the Cech cohomology and the de Rham cohomology groups. This condition guarantees that the symplectic form arise as the curvature form of a connection of a Hermitian line bundle. This line bundle is then called a prequantization in the theory of geometric quantization.
Electron energy levels in hydrogen
DR A
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 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: λ=
h me v
where h is Planck’s Constant, and me is the mass of the electron. Bohr 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: =
me v 2 r
FT
kqe2 r2
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
DR A
Substituting, one obtains the energy of the different levels of hydrogen: 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.
Energy in terms of other constants Starting with what we found above, En =
−me qe4 1 8h2 20 n2
We can multiply top and bottom by c2 , and we’ll arrive at En =
−me c2 qe4 1 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 constants to: Bohr model
19 En =
−Er α2 2n2
where,
FT
En is the energy level
Er is the rest energy of the electron α is the fine structure constant
n is the principal quantum number.
Rydberg formula
DR A
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 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 level.
Since the energy of a photon is E=
hc λ
the wavelength of the photon given off is me e4 1 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
FT
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. The Bohr model also has difficulty with or fails to explain: •
The spectra of larger atoms. At best, it can make some approximate predictions 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.
DR A
See also • • • •
→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 •
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: 231232.
21 Niels Bohr (1921). " Atomic Structure 12". Nature. A. Einstein (1917). "Zum Quantensatz von Sommerfeld und Epstein". Verhandlungen der Deutschen Physikalischen Gesellschaft 19: 82-92. Reprinted in The Collected Papers of Albert Einstein, A. Engel translator, (1997) Princeton University Press, Princeton. 6 p.434. (Provides an elegant reformulation of the Bohr-Sommerfeld quantization conditions, as well as an important insight into the quantization of non-integrable (chaotic) dynamical systems.)
Modern •
FT
• •
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,
DR A
GoldenBoar, El C, Glenn
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. The boltzmon was proposed as a means of explaining what happens to the information 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 total 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 (positive or negative electrical charge, integral or fractional spin, etc.), a boltzmon
would have an infinite number of states and as a result, would be highly unstable. If disturbed, it might make a hole in spacetime and vanish into it, thus departing from our universe.
Ferris, Timothy. The Whole Shebang, 1997 Simon & Schuster.
Source: http://en.wikipedia.org/wiki/Boltzmon
Born probability
DR A
In quantum mechanics, the Born probability is a probability of an event calculated from a wavefunction or more generally from the density matrix. The probability (or its density) equals the squared modulus of the complex amplitude an : P = |an |2
The interpretation that the physical meaning of the wavefunction is probabilistic was proposed by Max Born, and it became a pillar of the Copenhagen interpretation of quantum mechanics.
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 produced 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
FT
23
DR A
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 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 currently not completely understood, such as spontaneously flowing out of their 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 condensation 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
FT
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 statistical 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, 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 threedimensional gas consisting of non-interacting particles with no apparent internal degrees of freedom is given by: 2/3 n h2 Tc = ζ(3/2) 2πmkB where:
Tc
is the critical temperature, the particle density,
DR A
n m
the mass per boson,
h
Planck’s constant,
kB
the Boltzmann constant, and
ζ
the Riemann zeta function; ζ(3/2) ≈ 2.6124.
Discovery
In 1938, Pyotr Kapitsa, John Allen and Don Misener discovered that helium-4 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 properties, including zero viscosity (the ability to flow without dissipating energy) and the existence of quantized vortices. It was quickly realized that the superfluidity 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 Ketterle (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 original 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 Wieman, and co-workers at JILA on June 5, 1995. They did this by cooling a dilute Bose–Einstein condensate
25
FT
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 independent effort led by Wolfgang Ketterle at MIT created a condensate made of sodium-23. Ketterle’s condensate had about a hundred times more atoms, allowing him to obtain several important results such as the observation of quantum mechanical interference between two different condensates. Cornell, Wieman and Ketterle won the 2001 Nobel Prize for their achievement.
DR A
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 provides 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 condensation of magnons was demonstrated in the antiferromagnet TlCuCl3 by Oosawa et al. The condensation was observed at temperatures as large as 14 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 unknown property of Bose–Einstein condensate. Cornell, Wieman, and their coworkers originally used rubidium-87, an isotope whose atoms naturally repel 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 resonance involving a sweep of the magnetic field causing spin flip collisions, the JILA researchers lowered the characteristic, discrete energies at which the rubidium 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 condensate suddenly reverted back to attraction, imploded and shrank beyond detection, 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
FT
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 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
DR A
Compared to more commonly-encountered states of matter, Bose–Einstein condensates are extremely fragile. The slightest interaction with the outside world can be enough to warm them past the condensation threshold, forming a normal 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 questions 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 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 also 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.
Related experiments in cooling fermions rather than bosons to extremely low temperatures have created degenerate gases, where the atoms do not congregate in a single state due to the Pauli exclusion principle. To exhibit Bose– Einstein condensate, the fermions must "pair up" to form compound particles (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 Colorado 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 Slow light Superconductivity Superfluid Supersolid Super-heavy atom Tonks-Girardeau gas
DR A
• • • • • • • • • • • •
External links
• •
•
• •
BEC Homepage 19 General introduction to Bose–Einstein condensation Nobel Prize in Physics 2001 20 - for the achievement of Bose–Einstein condensation 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
Alkali Quantum Gases at MIT 24 Atom Optics at UQ 25 Einstein’s manuscript on the Bose–Einstein condensate discovered at Leiden University 26 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 sodium 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 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). "Quantum phase transition from a superfluid to a Mott insulator in a gas of ultracold atoms". Nature 415: 39-44..
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 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 Condensation of Molecules". Physical Review Letters 91: 250401. C. A. Regal, M. Greiner, and D. S. Jin (2004). "Observation of Resonance 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 Feshbachresonant interactions in 85Rb condensates." Phys Rev Lett. 2002 Oct 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.
FT
•
• •
Source: http://en.wikipedia.org/wiki/Bose%E2%80%93Einstein_condensate Principal Authors: Voyajer, CYD, Brian Jackson, Schneelocke, Michael Hardy, PAR, Matt Gies, Hfastedge, Fangz, R. Koot
29 http://dx.doi.org/10.1103/PhysRevLett.84.5868
Bose–Einstein condensate
30
Bose-Einstein statistics
FT
For other topics related to Einstein see Einstein (disambiguation).
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.
DR A
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 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 "distinguishable" 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 indistinguishable. 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 useful for studying gases F-D statistics are most often used for the study of electrons 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. ni =
gi e(i −µ)/kT −1
with i > µ and where:
FT
The expected number of particles in an energy state i for B-E statistics is:
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
DR A
k is Boltzmann’s constant T 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
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: w(n, 2) =
(n+1)! n!1! .
Bose-Einstein statistics
32
FT
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 P P (n−k+1)! (n+2)! w(n, 3) = nk=0 w(n − k, 2) = nk=0 (n−k)!1! = n!2! 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 w(n, g) =
(n+g−1)! . 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
DR A
where the approximation assumes that gi >> 1. Following the same procedure 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 ) 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: ni =
gi 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 ni =
gi e(i −µ)/kT −1
ni =
gi ei /kT /z−1
FT
Note that the above formula is sometimes written:
where z = exp(µ/kT ) is the absolute activity.
History
DR A
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 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 understand it. Bose sent his paper to Einstein who saw the significance of it and he used his influence to get it published.
See also
• →Maxwell-Boltzmann statistics • →Fermi-Dirac statistics • Parastatistics • →Planck’s law of black body radiation
Principal Authors: PAR, Michael Hardy, Stevenj, Voyajer, Mct mht, Salix alba, Phys, Youandme
Bose-Einstein statistics
34
Bose gas
FT
An ideal Bose gas is a quantum-mechanical version of a classical ideal gas. 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 Albert 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:
Z(z, β, V ) =
Q
i
1 − ze−βi
−gi
DR A
where each term in the product corresponds to a particular energy i , g i is the 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: β≡
1 kT
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 consider 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 grand potential defined as: P
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: dE
FT
1 E α−1 Γ(α) Ecα
dg =
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 (βEc )α
=
Vf 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 (βEc )α
=
f (~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.
DR A
We can solve the equation for the grand potential by integrating the Taylor 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: α+1 (z) Ω ≈ − Li(βE α c)
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 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
N = −z ∂Ω ∂z ≈
Liα (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 N=
ζ(α) (βc Ec )α
Bose gas
36
FT
This corresponds to a critical temperature T c=1/kβ c below which the ThomasFermi approximation breaks down. The above equation can be solved for the critical temperature: 1/α Ec N Tc = ζ(α) 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 temperature, because the particle numbers using the above equation become negative. The problem here is that the Thomas-Fermi approximation has set the degeneracy 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, however, 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) (βEc )α
DR A N = N0 +
where N 0 is the number of particles in the ground state condensate: N0 =
g0 z 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 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 τ=
T 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 normalized temperature as: Bose gas
FT
37
DR A
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 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.
N=
g0 z 1−z
+N
Liα (z) ζ(α)
τα
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 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:
Ω = g0 ln(1 − z) −
Liα+1 (z) (βEc )α
All thermodynamic properties may now be computed from the grand potential. The following table lists various thermodynamic quantities calculated in
Bose gas
38
Quantity
General
z Vapor fraction 0 1− N N
Equation of state PV β Ω N = −N
=
Liα (z) ζ(α)
=
Liα+1 (z) ζ(α)
τα τα
Gibbs Free Energy = ln(z) G = ln(z)
FT
the limit of low temperature and high temperature, and in the limit of infinite 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. T Tc
T Tc
=1
≈ τ α − α 2α 2 τ =1
= τα
ζ 2 (α)
ζ(α)
ζ(α+1)
ζ(α)
= ζ(α) τ α ≈ 1 − α+1 α 2 τ =0
≈ ln
ζ(α) τα
ζ(α)
− 2α τ α
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 energy and the product of pressure and volume is the same as that for a classical ideal gas over all temperatures: ∂Ω ∂β
= αP V
DR A
U=
A similar situation holds for the specific heat at constant volume Cv =
∂U ∂T
= k(α + 1) U β
The entropy is given by: TS = U + PV − G
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 •
FT
Huang, Kerson, "Statistical Mechanics", John Wiley and Sons, New York, 1967. • 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
DR A
Bound state
In physics, a bound state is a composite of two or more building blocks (particles 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 spectrum of isolated particles. (Actually, it is possible to have unstable bound states with a positive interaction energy provided that there is a "energy barrier" 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.
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 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.
• •
Composite field →Resonance
Source: http://en.wikipedia.org/wiki/Bound_state
Principal Authors: Phys, SeventyThree, Dmr2, Conscious, Tony Sidaway
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
|ψi
FT
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
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
DR A
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, which states that a Hilbert space and its dual space are isometrically isomorphic. 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 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 "braket" or "bracket", which is written as hφ|ψi.
In quantum mechanics, this is the probability amplitude for the state ψ to collapse into the state φ. Bra-ket notation
42
Properties
FT
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:
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 properties of the inner product (with c* denoting the complex conjugate of c),
DR A
•
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 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 mechanics. For example, hermitian operators are used to represent observable physical quantities, such as energy or momentum, whereas unitary linear operators 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. Composing 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 .
hφ|A|ψi.
FT
This expression is commonly written as
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 product is to construct projection operators. Given a ket |ψi of norm 1, the orthogonal projection onto the subspace spanned by |ψi is |ψihψ|.
DR A
Composite bras and kets
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 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:
FT
ψ(x) ≡ hx|ψi. 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.
DR A
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 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 •
Feynman, Leighton and Sands (1965). The Feynman Lectures on Physics Vol. III. Addison-Wesley. ISBN 0201021153.
Principal Authors: CYD, MathKnight, AxelBoldt, Trewornan, Laurascudder, Mct mht, AugPi, Theresa knott, Ancheta Wis
Bra-ket notation
45
Breit equation
Introduction
FT
The Breit equation is a relativistic wave equation derived by Gregory Breit in 1929 based on the Dirac equation, which formally describes two or more massive 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 electrodynamic effects are negligible, this equation has been shown to give results in good agreement with experiment.
DR A
The Breit equation is not only an approximation in terms of quantum mechanics, 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 ): nP o P P 1 ˆ ˆ i HD (i) + i>j rij − i>j Bij Ψ = EΨ, where
h i 2 ˆ D (i) = qi φ(ri ) + c P H s=x,y,z αs (i)πs (I) + α0 (I)m0 c
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, thus for electrons q i = - e. The one-electron Dirac hamiltonians of the particles, along with their instantaneous Coulomb interactions 1/r ij, form the Dirac-Coulomb operator. To this, Breit added the operator (now known as the Breit operator): ij ) ˆij = 1 a(i) · a(j) + (a(i)·rij )(a(j)·r , B 2rij 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
FT
The total hamiltonian of the Breit equation, sometimes called the DiracCoulomb-Breit hamiltonian (H DCB) can be decomposed into the following practical energy operators for electrons in electric and magnetic fields (also called the Breit-Pauli hamiltonian) 1, which have well-defined meanings in the interaction of molecules with magnetic fields (for instance for nuclear magnetic resonance): ˆij = H ˆ0 + H ˆ 1 + ... + H ˆ6, B in which the consequitive partial operators are: •
4 ˆ 1 = − 12 P pˆi3 is connected to the dependence of mass on velocity: H i m 8c i 2 − m c2 2 = m2 v 2 c2 . Ekin 0
DR A
•
2 ˆ 0 = P pˆi + V is the nonrelativistic hamiltonian (m_{i} is the stationary H i 2mi mass of particle i ).
•
qi qj 2rij mi mj c2
ˆ2 = − P H i>j
ˆi · p ˆj + p
ˆ i )·ˆ rij (rij p pj 2 rij
is a correction that partly ac-
counts for retardation and can be described as the interaction between the magnetic dipole moments of the particles, which arise from the orbital motion of charges (also called orbit-orbit interaction). P 2qi ˆ 3 = µB P 1 si · F(rij ) × p ˆ ˆ • H + r × p is the classical interac3 i ij j i mi j>i r c ij
tion between the orbital magnetic moments (from the orbital motion of charge) and spin magnetic moments (also called spin-orbit interaction). The first term describes the interaction of a particles spin with its own orbital moment (F ( r i) is the electric field at the particles position), and the second term between two different particles.
•
ˆ4 = H
ih 8πc2
qi ˆi i m2 p i
P
· F(ri ) is a nonclassical term characteristic for Dirac
theory, sometimes called the Darwin term. 8π 1 ˆ 5 = 4µ2 P − (s · s )δ(r + • H si · sj − 3 i j ij i>j B 3 r ij
(si ·rij )(sj ·rij ) 2 rij
is the mag-
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 •
ˆ 6 = 2µB H
P h i
H(ri ) · si +
qi ˆ mi c A(ri ) · pi
i
is the interaction between spin and
Notes •
FT
orbital magnetic moments with an external magnetic field H.
Note 1: H.A. Bethe, E.E. Salpeter, Quantum Mechanics of One- and TwoElectron Atoms, Plenum Press, New York 1977, pg.181
Principal Authors: Karol Langner, Neilc, CambridgeBayWeather
Canonical commutation relation
In physics, the canonical commutation relation is the relation
DR A
[x, p] = i~
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
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 counterparts fˆ, gˆ of classical observables f, g should satisfy \ [fˆ, gˆ] = i~{f, g}.
In 1927, Hermann Weyl showed that a literal correspondence between a quantum operator and a classical distribution in phase space could not hold. However, he did propose a mechanism, Weyl quantization, that underlies a mathematical approach to quantization known as deformation quantization.
Canonical commutation relation
48
Representations
Generalizations The simple formula [x, p] = i~,
FT
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 operators 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 socalled 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.
DR A
valid for the quantization of the simplest classical system, can be generalized 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).
πi ≡
∂L ∂(∂xi /∂t)
This definition of the canonical momentum ensures that one of the EulerLagrange equations has the form ∂ ∂t πi
=
∂L ∂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
FT
p − eA/c 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 momentum 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 + eφ p − eA c
DR A
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 A → A0 = A + ∇Λ φ → φ0 −
1 ∂Λ c ∂t
ψ → ψ0 = U ψ
H → H 0 = U HU †
where
U = exp
ieΛ ~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 e ~c hψ|r
× ∇Λ|ψi
FT
hψ|L|ψi → hψ 0 |L0 |ψ 0 i = hψ|L|ψi +
The gauge-invariant angular momentum (or "kinetic angular momentum") is given by K = r × p − eA c which has the commutation relations [Ki , Kj ] = i~ij k Kk + e~ c xk (x · B) where B =∇×A
DR A
is the magnetic field. The inequivalence of these two formulations shows up in the →Zeeman effect and the →Aharonov-Bohm effect.
See also • • •
canonical quantization CCR algebra Lie derivative
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
External links •
FT
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.
A Study of Vibrating Plates 31 by Derek Kverno and Jim Nolen
Principal Authors: Laurascudder, Michael Hardy, Choster
Classical limit
DR A
The classical limit is the ability of a physical theory to approximate or "recover" 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 electron can never be at rest; it must always have a non-zero kinetic energy, a result 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 (relative 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.
52 Principal Authors: Cyan, Michael Hardy, Charles Matthews, Salsb
FT
Clebsch-Gordan coefficients
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 irreducible 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.
DR A
In terms of classical mathematics, the CG coefficients, or at least those associated to the group SO(3), may be defined much more directly, by means of 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 P P |j1 j2 ; jmi = jm1 1 =−j1 jm2 2 =−j2 |j1 j2 ; m1 m2 ihj1 j2 ; m1 m2 |j1 j2 ; jmi 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 m2 =−j2 |j1 j2 ; m1 m2 ihj1 j2 ; m1 m2 |j1 j2 ; jmi 1 =−j1 renaming m1 to m01 and m2 to m02 and applying the J± operator P Pj2 |j1 j2 ; m01 m02 ihj1 j2 ; m01 m02 |j1 j2 ; jmi J± |j1 j2 ; jmi = (J1± +J2± ) jm1 0 =−j m0 =−j 1
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 m02 hj1 j2 ; m01 m02 |j1 j2 ; jmi P
m01
p
P
± m01 + 1)|j1 j2 ; m01 ± 1, m02 i +
p (j2 ∓ m02 )(j2 ±
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 m02 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
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 p = (j1 ∓ m1 + 1)(j1 ± m1 )hj1 j2 ; m1 ∓ 1, m2 |j1 j2 ; j, m ∓ 1i + p (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
For more coefficients, see table of Clebsch-Gordan coefficients.
Clebsch-Gordan coefficients
54
See also Wigner 3-j symbol 6-j symbol Spherical harmonics Associated Legendre polynomials Angular momentum Angular momentum coupling
FT
• • • • • •
External links •
Java TM Clebsch-Gordan Coefficient Calculator 32
References •
DR A
A.R. Edmonds, Angular Momentum in Quantum Mechanics, (1957) Princeton University Press, ISBN 0-691-07912-9. • E. U. Condon and G. H. Shortley, The Theory of Atomic Spectra, (1970) 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)
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 coherent 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
DR A
FT
in a quadratic potential well. In the quantum theory of light (quantum electrodynamics) 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 as a continuous laser wave.
Figure 5 Figure 1: The electric field, measured by optical homodyne detection, 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 photon numbers of the three states from top to bottom are =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
DR A
FT
56
Figure 6 Figure 2: The oscillating wave packet corresponding to the second coherent state depicted in Figure 1. At each phase of the light field, the distribution is a Gaussian of constant width.
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 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
FT
57
DR A
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 experimental errors.
Since a is not hermitian, α = |α|eiθ
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 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
FT
58
DR A
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 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
DR A
FT
59
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 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 |ni. n=0 n!
Coherent state
60
P (n) = e−hni
hnin n!
FT
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:
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
FT
u(−b) = u(a − b) u0 (−b) = u0 (a − b).
These conditions yield the following matrix: 1 α i(α−k)(a−b) e i(α−k)(a−b) (α − k)e 0 0 0 0
1 −α
e−i(α+k)(a−b) (α + k)e−i(α+k)(a−b)
−1 −β
−e−i(β−k)b −(β − k)e−i(β−k)b
A A0 = −ei(β+k)b B i(β+k)b B0 (β + k)e −1 β
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 2αβ
sin(βb) sin[α(a − b)]
DR A
cos(ka) = cos(βb) cos[α(a − b)] −
In order to further simplify the expression, we will perform the following approximations: b → 0 ; V0 → ∞ ; V0 b = constant
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
DR A
~ − 2m ∇2 ψ = Eψ
Using polar coordinates on the 1 dimensional ring, the wave function depends only on the angular coordinate, and so ∇2 =
1 ∂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 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 ψ(θ) =
√1 2π
r
e±i ~
√
2mEθ
The energy eigenvalues E are quantized because of the periodic boundary conditions, and they are required to satisfy
172 http://www.falstad.com/qm1dcrystal/
Particle in a ring
344 e±i ~
2mEθ
r
e±i2π ~
√
√ r
= e±i ~
2mE
2mE(θ+2π) ,
or
= 1 = ei2πn
FT
√ r
This leads to the energy eigenvalues E=
n2 ~2 2mr2
where n = 0, 1, 2, 3, . . .
The full wave functions are, therefore ψ(θ) =
√1 2π
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.
DR A
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 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.
See also • • •
Angular momentum Harmonic analysis. One-dimensional periodic case
Principal Authors: AmarChandra, Charles Matthews, Creidieki, Michael Hardy, Idril
Particle in a ring
345
FT
Particle in a spherically symmetric potential In quantum mechanics, the particle in a spherically symmetric potential describes 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 (radial 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: •
DR A
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. • 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 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 socalled radial equation: h i 2 l(l+1) d d − 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 spherical 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
FT
Let us now consider V(r) =0 (if V0 , replace everywhere E with E − V0 ). Introducing the dimensionless variable q ρ ≡ kr, k ≡ 2m~20 E r the equation becomes a Bessel equation for J defined by J(ρ) ≡ (whence the notational choice of J ): h i 2 2− l+ 1 2 J =0 ρ2 ddρJ2 + ρ dJ + ρ 2 dρ
√
ρR(r)
which regular solutions for positive energies are given by so-called Bessel functions 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, ∞]:
DR A
ψ(r) = jl (kr)Ylm (θ, φ)
√ 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
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 • •
FT
•
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.
The first constraint comes from the fact that Neumann N and Hankel H functions 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 r , r < r0 ~2
with A a constant to be determined later. Note that for bound states, V0 < E < 0.
DR A
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 constraint, selects Hankel function of the first kind as the only converging solution at infinity (the singularity at the origin of these functions does not matter since we are now outside the sphere): q (1) 0E R(r) = Bhl i −2m r , r rel="nofollow"> r0 2 ~
Second constraint on continuity of ψ at r = r0 along with normalization allows the determination of constants A and B. Continuity of the derivative (or logarithmic derivative for convenience) requires quantization of energy.
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: El =
u2l,k ~2 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
FT
Calling s, p, d, f, g, h, etc., states with l =0, 1, 2, 3, 4, 5, etc., respectively, we obtain the following spectrum: 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
DR A
This article is about a formulation of quantum mechanics. For integrals along a path, also known as line or contour integrals, see Line integral.
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 description of quantum theory which generalizes the action principle of classical mechanics. It replaces the classical notion of a single, unique history for a system with a sum, or functional integral, over an infinity of possible histories to compute a quantum amplitude. This formulation has proved crucial to the subsequent development of theoretical physics, since it provided the basis for the grand synthesis of the 1970’s called the renormalization group which unified quantum field theory with statistical 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 mechanics. 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 integral starts from the superposition law, and exploits wave-particle duality to build a generating function for quantum amplitudes. Path integral formulation
DR A
FT
349
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:
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 histories which include that event.
3. R The amplitude a certain history contributes is proportional to R i ˙ e ~ [L(q,q,t)]dt , where [L(q, q, ˙ t)]dt is the action of that history, or time 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
FT
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 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 according to Feynman’s principles will also obey the →Schrödinger equation for the Hamiltonian corresponding to the given action.
DR A
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 proportionality 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 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. 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 integral 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 postulate 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 equation, and classical mechanics is correctly recovered. Path integral formulation
351
Time Slicing Definition
FT
Action principles can seem puzzling to the student of physics because of their seemingly teleological quality: instead of predicting the future from initial conditions, 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 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 integral 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.
DR A
For a particle in a smooth potential, the path integral is approximated by Feynman 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 +∞ R +∞ i (H(x1 ,...,xj ,t)dt) ~ lim dx dx dx . . . dx e 1 2 3 n−1 −∞ −∞ −∞ −∞ ∆t→0,n→∞,n∆t=t
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 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
352 path integral to curved space using a multivalued coordinate transformation (nonholonomic mapping explained here 175).
FT
The path integral and the partition function The path integral is just the generalization of the integral above to all quantum mechanical problems— R RT Z = DxeiS[x]/~ where S[x] = 0 dtL[x(t)]
is the action of the classical problem in which one investigates the path starting 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.
DR A
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 temperature 1/T ~.
Clearly, such a deep analogy between quantum mechanics and statistical mechanics 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 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.
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 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 singleparticle restriction has already been loosened.)
DR A
However, the path-integral formulation is also extremely important in direct application to quantum field theory, in which the "paths" or histories being considered 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 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.
Such a functional integral is extremely similar to the partition function in statistical 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 integral 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 configurations (which only depends locally on the fields), then the time ordered vacuum expectation value of polynomially bounded functional F, , is given by
hF i =
R
DφF [φ]eiS[φ] R DφeiS[φ]
Path integral formulation
354
R
FT
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 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 integral. This is often the case.
In the language of functional analysis, we can write the Euler-Lagrange equaδ 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 configuration). The quantum analogues of these equations are called the SchwingerDyson equations.
DR A
If the functional measure Dφ turns out to be translationally invariant (we’ll 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]
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: E D E D δ δ F [φ] = −i F [φ] S[φ] δφ δφ 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 configurations (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:
FT
Dφei(S[φ]+hJ,φi)
Z[J] =
Note that δn Z [J] δJ(x1 )···δJ(xn )
= in Z[J] hφ(x1 ) · · · φ(xn )iJ
or
Z ,i1 ...in [J] = in Z[J] φi1 · · · φin J where
hF iJ =
R
DφF [φ]ei(S[φ]+hJ,φi) R Dφei(S[φ]+hJ,φi)
DR A
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 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 F [φ] =
∂ k1 k ∂x11
φ(x1 ) · · ·
∂ kn ∂xknn
φ(xn )
and G is a functional of J, then h i k1 δ δ F −i δJ ··· G[J] = (−i)n ∂ k1 δJ(x ) ∂x1
1
∂ kn δ G[J]. ∂xknn δJ(xn )
Then, from the properties of the functional integrals, we get the "master" Schwinger-Dyson equation: h i δS δ −i δJ Z[J] + J(x)Z[J] = 0 δφ(x) 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 translation does not even make any sense.
FT
In that case, we would have to replace the S in this equation by another functional Sˆ = S − i ln(M )
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: R P n+1 δ n e−J/~ D[x]e−S[x]/~ = −A[x] ∞ n=0 (~)
where :S is the Wick-rotated classical action of the particle,J is the classical action with an extra term "x" and delta here is the functional derivative operator R :A[x] = exp(1/~ dtX(t)
Ward-Takahashi identities
DR A
See main article Ward-Takahashi identity
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 measure 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 transformed value of the field at any given point under an infinitesimal transformation 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
FT
R hQ[F ]i + i F ∂V f µ dsµ = 0 where the integral is over the boundary. Noether’s theorem.
This is the quantum analog of
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 depends 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 assumption R that the functional measure is locally invariant:
Dφq(x)[F ][φ] = 0
DR 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
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, R d δ d xJ(x)Q[φ(x)][−i δJ ]Z[J] = 0
The path integral in quantum-mechanical interpretation In one philosophical interpretation of quantum mechanics, the "sum over histories" interpretation, the path integral is taken to be fundamental and reality Path integral formulation
358
FT
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 interpretation explains the Einstein-Podolsky-Rosen paradox without resorting to nonlocality. Some advocates of interpretations of quantum mechanics emphasizing decoherence 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 •
DR A
•
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) 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 subject. Grosche, Christian & Steiner, Frank ; Handbook of Feynman Path Integrals, Springer Tracts in Modern Physics 145, Springer-Verlag (1998) [ISBN 3540-57135-3] 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, Oxford University Press (2002) [ISBN 0-19-851572-3]. Etingof, Pavel ; Geometry and Quantum Field Theory 177, M.I.T. OpenCourseWare (2002). This course, designed for mathematicians, is a rigorous introduction to perturbative quantum field theory, using the language of functional integrals.
Papers on-line •
FT
•
DR A
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 given 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. • DeWitt-Morette, Cécile ; Feynman’s path integral - Definition without limiting 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.
Figure 33 Diagram illustrating a less complex version of the experiment to test penrose Interpretation
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 macroscopic 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. However 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 experience 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
FT
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 on the amount of energy involved: The higher the energy required to sustain 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.
DR A
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 experiment 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 returns, 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
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
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:
··· .. . ···
ρ1n .. .
DR A
ρ11 .. 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
ρP T
ρT11 .. = .
ρTn1
··· .. . ···
ρT1n .. .
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 geometric 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
Λ = Λ1 + Λ 2 ◦ T
FT
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
,where Λ1 and Λ2 are completely positive and T is the transposition map. Combining the above two facts, we can conclude PPT is also sufficient for separability 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.
DR A
References • •
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.
Principal Authors: Mct mht, Tinissimo, Stevey7788, Matthew Mattic, Charles Matthews
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 representing 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
FT
Perturbation theory is an extremely important tool for describing real quantum systems, as it turns out to be very difficult to find exact solutions to the →Schrödinger equation for Hamiltonians of even moderate complexity. The 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 calculate 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.
DR A
The expressions produced by perturbation theory are not exact, but they can 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 largeexpansion parameters, most efficiently by variational perturbation theory. In the theory of quantum electrodynamics (QED), in which the electron-photon 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 chromodynamics, 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 describe states that are not generated adiabatically from the "free model", including 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
FT
groups of particles bound to one another. An example of this phenomenon may be found in conventional superconductivity, in which the phonon-mediated attraction between conduction electrons leads to the formation of correlated electron pairs known as Cooper pairs. When faced with such systems, one usually 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 inverse 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).
DR A
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 perturbation 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 timedependent. In this section, we discuss time-independent perturbation theory, in which the perturbation Hamiltonian is static (i.e., possesses no time dependence.) Time-independent perturbation theory was invented by Erwin Schrödinger in 1926, shortly after he produced his theories in wave mechanics.
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) superscripts 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
FT
H = H0 + λV 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 + · · ·
DR A
When = 0, these reduce to the unperturbed values, which are the first term in 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 + · · ·
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
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 following way: suppose the perturbation is applied, but we keep the system in Perturbation theory (quantum mechanics)
367
FT
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 . However, the true energy shift is slightly different, because the perturbed eigenstate is not exactly the same as |n (0)>. These further shifts are given by the second and higher order deviations. To obtain the first-order deviation in the energy eigenstate, we insert our expression 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 identity, P (0) (0) (0) V |n(0) i = k |k ihk | V |n i The result is P (0) En − H0 |n(1) i = k6=n hk (0) |V |n(0) i |k (0) i
DR A
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 P hk (0) |V |n(0) i |n(1) i = k6=n (0) (0) |k (0) i 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 element , which is a measure of how much the perturbation mixes eigenstate n with eigenstate k ; it is also inversely proportional to the energy difference between eigenstates k and n, which means that the perturbation deforms the eigenstate to a greater extent if there are more eigenstates at nearby 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 P (2) |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 calculation 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.
FT
This is actually a conceptual, rather than mathematical, problem. Imagine 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, ∀ |k (0) i ∈ D
DR A
V |k (0) i = k |k (0) i + (terms not in D)
In that case, our equation for the first-order deviation in the energy eigenstate reduces to P (0) En − H0 |n(1) i = k6∈D hk (0) |V |n(0) i |k (0) i The operator on the left hand side is not singular when applied to eigenstates outside D, so we can write P hk (0) |V |n(0) i |n(1) i = k6∈D (0) (0) |k (0) i En −Ek
Time-dependent perturbation theory
Time-dependent perturbation theory, developed by Paul Dirac, studies the effect 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 perturbation 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
FT
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 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 interested in the populations of different atomic states in a gas when a timedependent electric field is applied. These probabilities are also useful for calculating the "quantum broadening" of spectral lines (see line broadening).
DR A
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 meaningful to speak of energy levels and eigenstates for the perturbed system.)
If the unperturbed system is in eigenstate |j > at time t = 0, its state at subsequent 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 Hamiltonian of the perturbed system is H = H0 + V (t)
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 P |ψ(t)i = n cn (t)e−iEn t/~ |ni
where the c n(t) s are undetermined complex functions of t which we will refer 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) on the right hand side. This is only a matter of Perturbation theory (quantum mechanics)
370
FT
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. 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
DR A
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/~ k hn|V (t)|ki ck (t) e ∂t = ~
The matrix elements of V play a similar role as in time-independent perturbation 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 sufficiently 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 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: P Rt 0 0 0 −i(Ek −En )t0 /~ cn (t) = cn (0) + −i 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)
cn (t) =
−i ~
P Rt k 0
0
dt0 hn|V (t0 )|ki ck (0) e−i(Ek −En )t /~
FT
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 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
DR A
The photoelectric effect is the emission of electrons from matter upon the absorption of electromagnetic radiation, such as ultraviolet radiation or x-rays. 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 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 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 [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
Figure 34 The photoelectric effect. Incoming EM radiation on the left ejects electrons, depicted as flying off to the right, from a substance.
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 wavelength of the light. In the photoemission process, if an electron absorbs the 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 reemitted. 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:
Algebraically: hf = φ + Ekmax where • • •
FT
Energy of photon = Energy needed to remove an electron + Kinetic energy of the emitted electron
DR A
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, 2 is the maximum kinetic energy of ejected electrons, • Ekmax = 12 mvm • f 0 is the threshold frequency for the photoelectric effect to occur, • m is the rest mass of the ejected electron, and • 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
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 ultraviolet 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 radiation.
FT
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 attempt at explaining how the observed phenomenon was brought about.
JJ Thomson: electrons
DR A
In 1899, Joseph John Thomson investigated ultraviolet light in Crookes tubes. Influenced by the work of James Clerk Maxwell, Thomson deduced that cathode 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 "corpuscle" 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.
Tesla’s radiant energy
On November 5 1901, Nikola Tesla received the U.S. Patent 685957 186 (Apparatus 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
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 electron 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
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 minutes 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 of the light doubled the number of electrons emitted from the surface. Lenard did not know of photons.
DR A
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 interaction of light with the electrons in the substance, was contained in the paper named "On a Heuristic Viewpoint Concerning the Production and Transformation 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 explanation of the photoelectric effect won him the Nobel Prize of 1921.
The idea of light quanta was motivated by Max Planck’s published law of blackbody radiation ("On the Law of Distribution of Energy in the Normal Spectrum". Annalen der Physik 4 (1901)) by assuming that Hertzian oscillators could only exist at energies E proportional to the frequency f of the oscillator 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 photoelectric effect that fit experiments. This was an enormous theoretical leap and the reality of the light quanta was strongly resisted. The idea of light quanta 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 contradict 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 increase 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 photoelectrons 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
Uses and effects
FT
The photoelectric effect helped propel the then-emerging concept of the dual nature of light, that light exhibits characteristics of waves and particles at different times. The effect was impossible to understand in terms of the classical 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.
Solar cells (used in solar power) and light-sensitive diodes use the photoelectric effect. They absorb photons from light and put the energy into electrons, creating electric current.
DR A
Electroscopes
Electroscopes 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 (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 incident photon minus the material’s work function or binding energy, the work function of a sample can be determined by bombarding it with a monochromatic X-ray source or UV source (typically a helium discharge lamp), and measuring 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.
A typical electron energy analyzer is a concentric hemispherical analyser (CHA), which uses an electric field to divert electrons different amounts depending on their kinetic energies. For every element and core atomic orbital there will be a different binding energy. The many electrons created from each will then show up as spikes in the analyzer, and can be used to determine the elemental composition of the sample. 189
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 photoelectric effect is self-limiting, though, because a more highly-charged object gives up its electrons less easily. 190
Moon dust
DR A
Light from the sun hitting lunar dust causes it to become charged through the photoelectric 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
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:
• • • • •
• • • • • • •
Physics: Atom Corona discharge →Double-slit experiment Electron Gamma ray Nobel Prize in Physics Optical phenomenon →Planck’s law of black body radiation Photon →Quantum mechanics Radiant energy →Wave-particle duality
Lists: • • • • • •
List of electronics topics List of optical topics List of physics topics Timeline of solar cells Scientific method list Timeline of mechanics and physics
DR A
• • • • • • • • • • • •
Aleksandr Grigorievich Stoletov Albert Einstein Heinrich Hertz Ernest Lawrence Robert Millikan Max Planck Joseph John Thomson
FT
Photocurrent Photomultiplier Solar cell Solar power Transducer
External 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. • 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 Laboratory, Davidson College, Davidson.
Principal Authors: Reddi, Enochlau, William M. Connolley, Heron, Omegatron
Planck 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 wavelength and Schwarzschild radius are equal (also by definition) to the Planck length.
See also •
Micro black hole
Source: http://en.wikipedia.org/wiki/Planck_particle Principal Authors: Jaraalbe, Hidaspal
The Planck Postulate (or Planck’s Postulate) was used by Max Planck in his 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
Planck’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: I(ν, T ) =
1 2hν 3 hν c2 e kT −1
where the following table provides the definition and SI units of measure for each symbol: Symbol Meaning
SI units of measure
I
spectral radiance, energy per unit time per unit surface area per unit solid angle per unit frequency
The wavelength is related to the frequency by λ = νc .
The law is sometimes written in terms of the spectral energy density u(ν, T ) =
4π c I(ν, T )
=
1 8πhν 3 hν c3 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: u(λ, T ) =
1 8πhc 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 wavelengths. He found that the above function, Planck’s function, fit the data for Planck’s law of black body radiation
382
E = hν .
FT
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, integer multiples of a fundamental unit of energy, E, proportional to the oscillation frequency ν:
DR A
Planck made this quantization assumption five years before Albert Einstein hypothesized 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. Ultimately, Planck’s assumption of energy quantization and Einstein’s photon hypothesis became the fundamental basis for the later development of Quantum Mechanics. Both scientists would eventually receive (separate) Nobel 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.)
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 superpositions of periodic functions. The wavelength λi in the three directions i = 1 . . . 3 orthogonal to the walls can be: λi =
2L ni
where the ni are integers. For each set of integers ni there are two linear independent 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
FT
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 responsible 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: Pr =
exp(−βE(r)) Z(β)
Here β ≡ 1/ (kT ).
DR A
The denominator Z (β), is the partition function of a single mode and makes Pr properly normalized: P 1 Z (β) = ∞ r=0 exp [−βE (r)] = 1−exp[−βε] Here we have defined q hc n21 + n22 + n23 ε ≡ 2L
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: hEi = −
d log(Z) dβ
=
ε exp(βε)−1
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: ε≡
hc 2L n
FT
where n is the norm of the vector ~n = (n1 , n2 , n3 ): 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 corresponds 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: g (ε) d = 2 81 4πn2 dn =
8πL3 2 ε dε h3 c3
DR A
Inserting this in Eq. (2) gives: R∞ ε3 dε (3) U = L3 h8π 3 3 0 exp(βε)−1 c
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 ): R∞ U = 0 u(ν, T )dν L3 where:
u(ν, T ) =
1 8πhν 3 c3 ehν/kT −1
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 L3
=
R∞ 0
u(λ, T )dλ
where
u(λ, T ) =
8πhc 1 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 J (hc)3
Here J is given by: R∞ x3 J = 0 exp(x)−1 dx =
π4 15
FT
u(T ) =
We prove this result in the Appendix below. The total electromagnetic energy inside the box is thus given by: U V
=
8π 5 (kT )4 15(hc)3
DR A
where V = L3 is the volume of the box. (Note - This is not the StefanBoltzmann 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 directions, and propagates at the speed of light (c), the spectral intensity (energy/time/area/solid angle/frequency) is I(ν, T ) =
u(ν,T ) c 4π
which yields
I(ν, T ) =
1 2hν 3 c2 ehν/kT −1
Derivation (Thermodynamics)
The fact that the energy density of the box containing radiation is proportional 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: P =
u 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
V
FT
P
And V ∂S ∂P = 3 T
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 differential equation: 1 dP P dT
=
4 T
This implies that u = 3P ∝ T 4
History
DR A
Many popular science accounts of quantum theory, as well as some physics textbooks, contain some serious errors in their discussions of the history of 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 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
FT
A simple way to calculate the integral R∞ x3 J = 0 exp(x)−1 dx is as follows. After multiplying the numerator and denominator of the integrand we can expand the integrand in powers of exp(−x). R ∞ x3 exp(−x) R∞ 3 P P∞ 1 π4 J = 0 1−exp(−x) dx = ∞ n=1 0 x exp (−nx) dx = 6 n=1 n4 = 15 Here we have used that
P∞
1 n=1 n4
the argument 4, which is given by the contour integral H π cot(πz) CR
z4
is the Riemann zeta function evaluated for π4 90 .
This fact can be proven by considering
DR A
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 at 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 Planck, Max, "On the Law of Distribution of Energy in the Normal Spectrum". 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 • •
388 Principal Authors: Metacomet, PAR, Unc.hbar, Diegueins, Rparson
DR A
FT
Plum pudding model
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. Thomson, 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). Instead 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 interpreted 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.
FT
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 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.
DR A
A little-known (or now forgotten) fact about the original Thomson "plum pudding" 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 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 attempted 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.
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.
Position operator In quantum mechanics, the position operator corresponds to the position observable 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 }.
DR A
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 real line and that Q has purely continuous spectrum, therefore no eigenvalues. The three dimensional case is defined analogously. We shall keep the onedimensional 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).
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 R |ΩQ (B)ψ|2 = |C B · ψ|2 = B |ψ|2 dµ
,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
FT
∂ P ψ = −i~ ∂x ψ
,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).
A potential energy surface is generally used within the adiabatic or BornOppenheimer approximation in quantum mechanics and statistical mechanics 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.
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
FT
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 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.
DR A
A potential hill is the opposite of a potential well, the region surrounding a local maximum.
Quantum confinement
Quantum confinement is when electrons and holes in a semiconductor are confined 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 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: matters 215, Nat. Mater., 2003, 2, 138 139.
Principal Authors: Patrick, Laurascudder, Linas, Fasten, Bantman
In functional analysis and quantum measurement theory, a POVM (Positive Operator 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 measurement 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.
DR A
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 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 operators P {Fi } on a Hilbert space H that sum to unity, n i=1
Fi = IH .
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
E 7→ hF (E)ξ | ξi
FT
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 nonnegative self-adjoint operators on a Hilbert space H such that F(X ) = I H and for every ξ ∈ H,
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 associated 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,
DR A
Fi = Mi† Mi ,
for some operator Mi , known as a Kraus Operator. The state of the system after the measurement ρ0 is transformed according to, ρ0 =
Mi ρMi† . tr(Mi ρMi† )
Neumark’s dilation theorem
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 dimension 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. Consider 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 π ˆi succeeds is, P (i) = T rA (T rB (ˆ πi 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.
FT
P (i) = T rA (Fi ρA )).
An example: Unambiguous quantum state discrimination
DR A
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 performing UQSD than any possible projective measurement.
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 Aˆ = a|ψ T ihψ T | + b|ψihψ|,
|hφ|ψi| = cos(θ),
FT
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 discriminates between the two states (i.e. with 100% probability). This latter ability is unnecessary for UQSD and, in fact, is impossible for anything but orthogonal states. Now consider a set that consists of two states |ψi and |φi in two-dimensional Hilbert space that are not orthogonal. i.e.,
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,
DR A
π ˆψT = |ψ T ihψ T |, π ˆφT = |φT ihφT |,
50% of the time. If π ˆφT 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 π ˆψT indicates conclusively that the system is in |φi and 0 is inconclusive. The probability that this strategy returns a conclusive result is, Pproj =
1−|hφ|ψi|2 . 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| Fˆψ = 1+|hφ|ψi|
Fˆφ =
1−|ψihψ| 1+|hφ|ψi|
POVM
397 Fˆinconcl. = 1 − Fˆψ − Fˆφ ,
DR A
FT
where the result associated with Fˆi indicates the system is in state i with certainty.
Figure 38 The POVM strategy for unambiguously discriminating between nonorthogonal states.
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, respectively. 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 strategy corresponds to passing the photons through a polarizer in either the vertical
POVM
398
See also • • • • • •
Quantum measurement →Mathematical formulation of quantum mechanics Quantum logic Density matrix Quantum operation Projection-valued measure
References •
FT
direction or -45 degree direction. If the photon passes through the vertical polarizer 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 complicated and requires another optical mode (known as an ancilla). It has a success √ probability of 1 − 1/ 2 = 29.3%.
DR A
POVMs • J.Preskill, Lecture Note for Physics: Quantum Information and Computation, 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. • 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
FT
In quantum mechanics, a probability amplitude is a complex-valued function that describes an uncertain or unknown quantity. For example, each particle has a probability amplitude describing its position. This amplitude is then called wave function. This is a complex-valued function of the position coordinates.
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.
DR A
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 result of the measurement will lie within with a probability P() given by R P () = |ψ(x)|2 dx 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 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 acceptable 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
FT
where P(x,t) is the probability density and measured in units of (probability)/(volume) = r -3. This equation is the mathematical equivalent of probability 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.
Principal Authors: Charles Matthews, P3d0, That Guy, From That Show!, Onebyone, Michael Hardy, RJFJR, NymphadoraTonks, Conscious, Paul A
Probability current
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 ~ − Ψ∇Ψ ~ ∗ ~j = ~ Ψ∗ ∇Ψ 2mi
in the position basis and satisfies the quantum mechanical continuity equation
Probability current
401 ∂ρ ∂t
~ · ~j = 0 +∇
ρ = |Ψ|2 .
FT
with the probability density ρ defined as
The divergence theorem implies the continuity equation is equivalent to the integral equation R R ∂ ~ =0 |Ψ|2 dV + ~j · dA ∂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.
DR A
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 rate at which probability flows into V .
Examples Plane wave
The probability current associated with the (three dimensional) plane wave ~
Ψ = eik·~r
is
~j =
~ 2mi
~ ~ i~k·~r ~ ~ −i~k·~r = e−ik·~r ∇e − eik·~r ∇e
~~k 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.
FT
Particle in a box 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 ~ n jn = 2mi Ψ∗n ∂Ψ =0 ∂x − Ψn ∂x since Ψn = Ψ∗n .
Derivation of continuity equation
In this section the continuity equation is derived from the definition of probability current and the basic principles of quantum mechanics.
DR A
Suppose Ψ is the wavefunction for a single particle in the position basis (i.e. Ψ is a function of x, y, and z). Then R P = V |Ψ|2 dV
is the probability that a measurement of the particle’s position will yield a value within V. The time derivative of this is R R ∂Ψ ∗ ∂ dP ∂Ψ∗ 2 dV dt = ∂t V |Ψ| dV = V ∂t Ψ + Ψ ∂t
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 i~ ∂Ψ ∂t =
−~2 2 2m ∇ Ψ + V Ψ
and use it to solve for the time derivative of Ψ : ∂Ψ ∂t
=
i~ 2 i 2m ∇ Ψ − ~ V Ψ
When substituted back into the preceding equation for R ~ dP ∗ 2 2 ∗ dt = − V 2mi Ψ ∇ Ψ − Ψ∇ Ψ dV .
Now from the product rule for the divergence operator Probability current
and since the first and third terms cancel: R dP ~ · ~ Ψ∗ ∇Ψ ~ − Ψ∇Ψ ~ ∗ dV =− ∇
If we now recall the expression for P and note that the argument of the divergence 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 ∂t
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 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
FT
P (Hn ) = CP n−1 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 FubiniStudy 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
DR A
Pure gauge
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
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.
FT
An entirely new conceptual framework was developed around this idea, during the first half of the 1900s. Usually referred to as quantum "mechanics", it is regarded by virtually every professional physicist as the most fundamental 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.
DR A
The consequences of the differences between classical and quantum mechanics quickly became obvious. But it was not until 1926, by the work of Werner Heisenberg, Erwin Schrödinger, and others, that quantum mechanics became correctly formulated and understood mathematically. Despite tremendous experimental 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 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 Society, in Berlin, where the black body work was being intensely discussed. In this way, careful experimentalists (including F. Paschen, O.R. Lummer, Ernst Pringsheim, 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 radiation 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
FT
red-hot, the red wavelength parts start to become visible. This had been studied over the previous years, as the instruments were being developed. However, 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 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 electromagnetic 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 quantum mechanics, appeared Sunday evening October 7, 1900, in a so-called backof-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.
DR A
At first, it was just a fit to the data. Only weeks later did it turn out to enforce quantization.
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 formula"). 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 simplifying the mathematics, so that this could happen. And here Planck used hard won experience from the previous years. Briefly stated, he had two mathematical expressions:
• •
(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 approximately, 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 E=
hf
hf e kT
−1
This is (essentially) what is being compared with the experimental measurements. There are two parameters to determine from the data, written in the Quantum
407
FT
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 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
DR A
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. 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. 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 •
See also →Quantum mechanics →Quantum state Quantum number Quantum cryptography Quantum electronics Quantum computer Quantum immortality Magnetic flux quantum Quantization →Subatomic particle →Elementary particle
DR A
• • • • • • • • • • •
FT
J. Mehra and H. Rechenberg, The Historical Development of Quantum Theory, Vol.1, Part 1, Springer-Verlag New York Inc., New York 1982. • 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.
Source: http://en.wikipedia.org/wiki/Quantum Principal Authors:
Dennis Estenson II, Finn-Zoltan, Deathphoenix, Bensaccount, Stevertigo,
Quantum 1/f noise is an intrinsic part of quantum mechanics. It comes from scattering of different particles of one another in solid state physics. Quantum 1/f noise is a source of Chaos in such systems.
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 University of Missouri - St. Louis, and published in Physical Review Letters A, in August 1980.
DR A
For more on Quantum 1/f, see: P.H. Handel: "1/f Macroscopic Quantum Fluctuations of Electric Currents Due to Bremsstrahlung with Infrared Radiative Corrections", Zeitschrift fuer Naturforschung 30a, p.1201 (1975)
See also • • • • • • • • • • • • •
shot noise 1/f noise white noise Johnson-Nyquist noise signal-to-noise ratio noise level noise power noise-equivalent power phase noise thermal noise list of noise topics audio system measurements Colors of noise
Principal Authors: Gordon Stangler, Charles Matthews, Wendell, Ardric47, Conscious
Quantum 1/f noise
410
Quantum acoustics
FT
In physics, quantum acoustics is the study of sound under conditions such 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. A symposium on quantum acoustics is held in Poland each year
See also •
Superfluid
References •
216
DR A
Quantum acoustics by Humphrey J. Maris in the McGraw-Hill Encyclopedia of Science & Technology Online 217 • Handbook of Acoustics by Malcolm Crocker has a chapter on quantum acoustics.
Quantum biology is the science of studying biological processes in terms of quantum mechanics. In exploring quantum mechanical explanations for biological 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 macroscopic phenenoma would be the result of quantum processes):
the absorbance of frequency-specific radiation (i.e., photosynthesis and vision); • the conversion of chemical energy into motion; • magnetoreception in animals. Quantum biological research is extremely limited by computer processing power; the analytical power required to model quantum effects increases exponentially with the number of particles involved.
References • •
nanoword.net 218 Theoretical and Computational Biophysics Group, University of Illinois at Urbana-Champaign 219
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 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.
During the first half of the twentieth century, chaotic behavior in mechanics was recognized (in celestial mechanics), but not well-understood. The foundations of modern quantum mechanics were laid in that period, essentially leaving aside the issue of the quantum-classical correspondence in systems whose 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.
DR A
Similar questions arise in many different branches of physics, ranging from nuclear to atomic, molecular and solid-state physics, and even to acoustics, microwaves 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 analyzing 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 of 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 systems are level repulsion in the spectrum, dynamical localization in the time evolution (e.g. ionization rates of atoms), and enhanced stationary wave intensities in regions of space where classical dynamics exhibits only unstable trajectories (wave function scarring). An alternative name for quantum chaos, proposed by Sir Michael Berry, is quantum chaology.
History
Important methods applied in the theoretical study of quantum chaos include random-matrix theory (significant contributions by Oriol Bohigas, see also American Scientist 220) and periodic-orbit theory (pioneered by Martin Gutzwiller).
A. Einstein (1917). "Zum Quantensatz von Sommerfeld und Epstein". Verhandlungen der Deutschen Physikalischen Gesellschaft 19: 82-92. Reprinted in The Collected Papers of Albert Einstein, A. Engel translator, (1997) Princeton University Press, Princeton. 6 p.434. (Provides an elegant reformulation of the Bohr-Sommerfeld quantization conditions, as well as an important insight into the quantization of non-integrable (chaotic) dynamical systems.) • Joeseph B. Keller (1958). "". Annals of Physics (NY) 4: 180. (An independent 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.
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 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.
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
FT
Quantum entanglement is a quantum mechanical phenomenon in which the 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 impossible 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).
DR A
Quantum entanglement is closely concerned with the emerging technologies of quantum computing and quantum cryptography, and has been used to experimentally realize quantum teleportation. At the same time, it prompts some of the more philosophically oriented discussions concerning quantum theory. The correlations predicted by quantum mechanics, and observed in experiment, reject 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 entanglement can be related to different interpretations of quantum mechanics.
Background
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 producing 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 parameters would cause the correlations. However, in 1964 Bell derived an upper limit, known as Bell’s inequality, on the strength of correlations for any
Quantum entanglement
415
FT
theory obeying "local realism" (see principle of locality). Quantum entanglement can lead to stronger correlations that violate this limit, so that quantum entanglement is experimentally distinguishable from a broad class of local hidden-variable theories. Results of subsequent experiments have overwhelmingly 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 subtle reasons: (i) quantum mechanical measurements yield probabilistic results, and (ii) the no cloning theorem forbids the statistical inspection of entangled quantum states.
DR A
Although no information can be transmitted through entanglement alone, it is possible to transmit information using a set of entangled states used in conjunction with a classical information channel. This process is known as quantum 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. 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
FT
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 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
DR A
If the composite system is in this state, it is impossible to attribute to either 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 outcomes, 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 .
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 information 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
Ensembles
FT
the Fubini-Study metric. The product of two pure states is then given by the Segre embedding.
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 |,
DR A
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 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 P B ρ = i p i ρA i ⊗ ρi B ,where ρA i ’s and ρi ’s are they themselves states on the subsystems A and B 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 genB erality that ρA 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.
FT
Reduced Density Matrices
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|.
DR A
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 .
ρA is sometimes called the reduced density matrix of ρ on subsystem A. Colloquially, 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 entangled 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:
FT
S(ρ) = −Tr (ρ log ρ) , 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 .
DR A
Entropy provides one tool which can be used to quantify entanglement (although 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. 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
FT
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 system, i.e. U (t) = exp −iHt ~
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
DR A
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 superdense coding and quantum state teleportation. Efforts to quantify this resource are often termed entanglement theory. See for example Entanglement Theory 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 →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 symposium on Theory of computing, 2003.
222 http://www.imperial.ac.uk/quantuminformation
Quantum entanglement
421
External links •
FT
An Interview With Brian Clegg, Author of "The God Effect : Quantum Entanglement, Science’s Strangest Phenomenon" 223 California Literary Review • Multiple entanglement and quantum repeating 224 • How to entangle photons experimentally 225
Principal Authors: Mct mht, CYD, Roadrunner, CSTAR, Caroline Thompson, Cortonin, Charles Matthews
Quantum field theory
DR A
Quantum field theory (QFT) is the application of quantum mechanics to fields. It provides a theoretical framework, widely used in particle physics and condensed matter physics, in which to formulate consistent quantum theories 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 superconductivity. 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.
Origin
Quantum field theory originated in the problem of computing the energy radiated 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
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.
FT
It was obvious from the beginning that the quantum treatment of the electromagnetic field required a proper treatment of relativity. Jordan and Wolfgang Pauli showed in 1928 that commutators of the field were actually Lorentz invariant. By 1933, Niels Bohr and Leon Rosenfeld had related these commutation 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 relativity, work that was completed by Wendell Furry and Robert Oppenheimer using methods developed for this purpose by Vladimir Fock. This need to put together 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.
DR A
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 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
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 spacetime point. Intuitively, this means that operators can be localized. This definition 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 function 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
ˆ(t) and p ˆ (t), x
FT
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
we now look at operators generated by ˆ t) φ(x,
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.
DR A
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 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 eigenstate is called the vacuum in particle physics and the ground state in condensed matter physics. (QFT appears in the continuum limit of condensed matter systems) This simple explanation of what QFT really is, is often obscured in treatments which jump straight to the path integral approach, which is a good computational 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 mechanics, 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 .
FT
The first problem occurs when we seek to extend the equation to large numbers of particles. As described in the article on identical particles, quantum mechanical particles of the same species are indistinguishable, in the sense that the state of the entire system must be symmetric (bosons) or antisymmetric (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 Nj ! P j |φ1 · · · φN i = p∈Sn |φp(1) i · · · |φp(N ) i N!
DR A
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. The second problem arises when trying to reconcile the Schrödinger equation with special relativity. It is possible to modify the Schrödinger equation to include the rest energy of a particle, resulting in the →Klein-Gordon equation or the Dirac equation. However, these equations have many unsatisfactory qualities; 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 mass-energy relation predicts that sufficiently massive particles can decay into several lighter particles, and sufficiently energetic particles can combine to form massive particles. For example, an electron and a positron can annihilate each other to create photons. 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 create 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.
FT
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. 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 harmonic 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: •
DR A
•
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 (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.
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 exchange 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 understanding 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 3
[|φ1 i|φ2 i|φ2 i + |φ2 i|φ1 i|φ2 i + |φ2 i|φ2 i|φ1 i]
In second quantized form, we write this as
Quantum field theory
426 |1, 2, 0, 0, 0, · · ·i
FT
which means "one particle in state 1, two particles in state 2, and zero particles in all the other states."
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 annihilate 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
DR A
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 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 annihilation operators are indeed Hermitian conjugates, which justifies the way we have written them. The bosonic creation and annihilation operators obey the commutation relation h h i i ai , a†j = δij ai , aj = 0 , a†i , a†j = 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 quantum of energy in a harmonic oscillator. The final step toward obtaining a quantum field theory is to re-write our original N -particle Hamiltonian in terms of creation and annihilation operators acting on a Fock space. For instance, the Hamiltonian of a field of free (noninteracting) bosons is P H = k Ek a†k ak
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.
FT
Canonical quantization for fermions 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 annihilation operators c and creation operators c† by cj |N1 , N2 , · · · , Nj = 0, · · ·i = 0
The fermionic creation and annihilation operators obey an anticommutation relation, n n o o ci , cj = 0 , c†i , c†j = 0 , ci , c†j = δij
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.
Significance of creation and annihilation operators When we re-write a Hamiltonian using a Fock space and creation and annihilation 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 common 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 evolution, 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 instantly 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
FT
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 and ck , we could add a "potential energy" term to our Hamiltonian such as: P V = k,q Vq (aq + a†−q )c†k+q ck
DR A
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. 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 particular point in space. In particle physics, these are often more convenient to work 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 P φ(r) ≡ i eiki ·r ai
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 relations are the same, with the commutators replaced by anticommutators.
Quantum field theory
429
FT
It should be emphasized that the field operator is not the same thing as a singleparticle 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 space representation, say P ~2 P 2 H = − 2m i ∇i + i<j U (|ri − rj |)
where the indices i and j run over all particles, then the field theory Hamiltonian is R 3 R R ~2 H = − 2m d r φ(r)† ∇2 φ(r) + d3r d3r0 φ(r)† φ(r0 )† U (|r − r0 |)φ(r0 )φ(r) This looks remarkably like an expression for the expectation value of the energy, 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.
DR A
Quantization of classical fields
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 quantum 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 electromagnetic field is the number of dynamical degrees of freedom. For a system of N particles, there are 3N coordinate variables corresponding to the position 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
Path integral methods The axiomatic approach
FT
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.
There have been many attempts to put quantum field theory on a firm mathematical footing by formulating a set of axioms for it. These attempts fall into two broad classes.
DR A
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 axioms enjoyed limited success. It was possible to prove that any QFT satisfying these axioms satisfied certain general theorems, such as the spin-statistics theorem 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 analytic 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 (associated 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 useful 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 difficult 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 structures which permit the Feynman path integrals to exist.
Renormalization
Some of the problems and phenomena eventually addressed by renormalization 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
DR A
FT
431
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 to the charge. A single particle state in quantum field theory incorporates within it multiparticle 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 propagation, and are therefore quintessentially a product of quantum field theory. Corrections such as these are incorporated into wave-function renormalization
Quantum field theory
432
Gauge theories
FT
and mass renormalization. Similar corrections to the interaction Hamiltonian, H I, include vertex renormalization, or, in modern language, effective field theory.
DR A
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 electrodynamics, the theory is also invariant under a local change of phase, that is - one may shift the phase of all wave functions so that 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. In quantum field theory the excitations of fields represent particles. The particle associated with excitations of the gauge field is the gauge boson, which is the photon in the case of quantum electrodynamics. The degrees of freedom in quantum field theory are local fluctuations of the fields. The existence of a gauge symmetry reduces the number of degrees of freedom, simply because some fluctuations of the fields can be transformed to zero by gauge transformations, so they are equivalent to having no fluctuations at all, and they therefore have no physical meaning. Such fluctuations are usually called "non-physical degrees of freedom" or gauge artifacts; Usually some of them have a negative norm, making them inadequate for a consistent theory. Therefore, if a classical field theory has a gauge symmetry, then its quantized version (i.e. the corresponding quantum field theory) will have this symmetry as well. In other words, a gauge symmetry cannot have a quantum anomaly. If a gauge symmetry is anomalous (i.e. not kept in the quantum theory) then the theory is non-consistent: for example, in quantum electrodynamics, had there been a gauge anomaly, this would require the appearance of photons with longitudinal polarization and polarization in the time direction, the latter having a negative norm, rendering the theory inconsistent; another possibility would be for these photons to appear only in intermediate processes but not in the final products of any intercation, making the theory non unitary and again inconsistent (see optical theorem).
Quantum field theory
433
FT
In general, the gauge transformations of a theory consist several different transformations, which may not be commutative. These transformations are together described by a mathematical object known as a gauge group. Infinitesimal gauge transformations are the gauge group generators. Therefore the number of gauge bosons is the group rank (i.e. number of generators forming an orthogonal basis). All the fundamental interactions in nature are described by gauge theories. These are: •
DR A
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 product of U(1) and SU(2)). • Gravity, whose classical theory is general relativity, admits the equivalence principle which is a form of gauge symmetry.
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 socalled Hierarchy Problem, that is, to explain why particles not protected by any symmetry (like the Higgs boson) do not receive radiative corrections to its mass driving it to the larger scales (GUT, Planck...). It was soon realized that supersymmetry has other interesting properties: its gauged version is an extension of general relativity (Supergravity), and it is a key ingredient for the consistency of string theory. The way supersymmetry protects the hierarchies is the following: since for every particle there is a superpartner with the same mass, any loop in a radiative correction is cancelled by the loop corresponding to its superpartner, rendering the theory UV finite. Since no superpartners have yet been observed, if supersymmetry exists it must be broken (through a so-called soft term, which breaks supersymmetry without 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
FT
History 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, Feynman, 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.
DR A
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 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.
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 laureate 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.
FT
•
•
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.
DR 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.
External links • •
Siegel, Warren ; Fields 229 (also available from arXiv:hep-th/9912205) ’t Hooft, Gerard ; The Conceptual Basis of Quantum Field Theory, Handbook of the Philosophy of Science, Elsevier (to be published). Review article written by a master of gauge theories, [http://nobelprize.org/physics/laureates/1999/thooft-autobio.html’’Nobel laureate 1999]. Full text available in 230.
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 uncertainty principle.
DR A
According to one formulation of the principle, energy and time can be related by the relation ∆E∆t ≈
h 2π
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 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 structure 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 observed 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
fields, renormalization complicates matters a lot). For the quantized KleinGordon 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 ϕ˜t (k) to be h i R d3 k ∗ p ρ0 [ϕt ] = exp − ~1 (2π) ˜t (k) |k|2 + m2 ϕ˜t (k) . 3ϕ
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 i R d3 k ∗ 1 2 + m2 ) ϕ 1 ˜ (k) (|k| ˜ (k) . ρE [ϕt ] = exp [−H[ϕt ]/kT ] = exp − kT t 3ϕ 2 (2π) t
DR A
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), (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 probability density as the quantum vacuum state, so that the principal difference from quantum field theory is the measurement theory (measurement in quantum theory is different from measurement for a classical continuous random 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 quantum fluctuations Quantum fluctuations are recently being used to anneal glassy systems (physical glass, or equivalently, hard combinatorial optimization problem with ragged energy/cost landscape) to their minimal/ground states. Thus it provides a general algorithmic scheme for classical/quantum computers.
Principal Authors: Arnab das, AstroNomer, Eequor, Michael Hardy, RoboDick, Phys, Nowhither, Andre Engels
Quantum foam
DR A
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.
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.
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 period 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 arising 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.
FT
Reginald Cahill has developed a theory called Process physics, which describes space as a quantum foam system in which gravity is an inhomogeneous flow 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 centimeter 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.
DR A
See also • • • • • • •
Dirac sea Hawking radiation Hyperspace theory Planck time "Rolling ball" topology Vacuum energy Wormhole
References •
• •
John Archibald Wheeler with Kenneth Ford. Geons, Black Holes, and Quantum 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
The quantum Hall effect is a quantum-mechanical version of the Hall effect, observed in two-dimensional electron systems subjected to low temperatures and strong magnetic fields, in which the Hall conductance σ takes on the quantized values
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.)
DR A
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 billion. This phenomenon, referred to as "exact quantization", has been shown to be a subtle manifestation of the principle of gauge invariance. It has allowed 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 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 calculation. 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 quantized. For this finding, von Klitzing was awarded the 1985 Nobel Prize in Physics. The link between exact quantization and gauge invariance was subsequently found by Robert Laughlin. The fractional effect is due to completely different physics, and was experimentally discovered in 1982 by Daniel Tsui and Horst Störmer, in experiments
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 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 A
•
FT
performed on gallium arsenide heterostructures developed by Arthur Gossard. The effect was explained by Robert B. Laughlin in 1983, using a novel quantum liquid phase that accounts for the effects of interactions between electrons. Tsui, Störmer, and Laughlin were awarded the 1998 Nobel Prize for their work. 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.
The quantum harmonic oscillator is the quantum mechanical analogue of 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 article mathematical formulation of quantum mechanics.
In diatomic molecules, the natural frequency can be found by: q ω = mkr
242
where
ω = 2πf is the angular frequency, k is the bond force constant, and 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: H=
p2 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
Figure 40 Wavefunction representations for the first six bound eigenstates, n = 0 to 5. The horizontal axis shows the position x. The graphs are not normalised
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 p mω mωx2 · exp − · Hn hx|ψn i = √ 1n · mω π~ 2~ ~ 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
FT
444
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 dxn e
DR A
Hn (x) = (−1)n ex
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 following section on ladder operators, we will engage in a more detailed examination 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. Nevertheless, 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
a a†
p mω 2~ x + p mω 2~ x −
= =
i mω p i mω p
FT
The power series solution, though straightforward, is rather tedious. The "ladder operator" method, due to Paul Dirac, allows us to extract the energy eigenvalues without directly solving the differential equation. Furthermore, it is readily generalizable to more complicated problems, notably in quantum field theory. Following this approach, we define the operators a and its adjoint a †
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
x =
q
p
i
~mω 2
a† + a
a† − a
DR A
=
~ q2mω
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 [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 hψE | ~ω −
1 2
|ψE i =
E ~ω
−
1 2
≥ 0,
FT
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 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]† = H, a =
−~ωa ~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)
DR 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 multiplicative 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 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
H |0i = (~ω/2) |0i
FT
In this case, subsequent applications of the lowering operator will just produce zero kets, instead of additional energy eigenstate. Furthermore, we have shown above that
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
DR A
The quantum harmonic oscillator possesses natural scales for length and energy, which can be used to simplify the problem. These can be found by nondimensionalization. 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: 2
∂ 1 2 H = − 21 ∂u 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 .
To avoid confusion, we will not adopt these natural units in this article. However, they frequently come in handy when performing calculations.
N -dimensional harmonic oscillator
The one-dimensional harmonic oscillator is readily generalizable to N dimensions, 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 position coordinate is a momentum; we label these p 1, ..., p N . The canonical commutation relations between these operators are
[xi , pj ] = [xi , xj ] = [pi , pj ] =
i~δi,j 0 0
Quantum harmonic oscillator
448
The Hamiltonian for this system is P p2i 1 2 2 H= N i=1 2m + 2 mω xi .
FT
.
As the form of this Hamiltonian makes clear, the N -dimensional harmonic oscillator is exactly analogous to N independent one-dimensional harmonic oscillators 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.
DR A
This observation makes the solution straightforward. For a particular set of quantum numbers {n} the energy eigenfunctions for the N -dimensional oscillator are expressed in terms of the 1-dimensional eigenfunctions as: Q hx|ψ{n} i = N i=1 hxi |ψni i In the ladder operator method, we define N sets of ladder operators,
= =
ai a†i
p mω 2~ xi + p mω 2~ xi −
i mω pi i 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 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 energy 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
(n+1)(n+2) 2
Related problems
FT
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: P P P n(n+1) gn = nn1 =0 n−n1 +1 = nn1 =0 n+1− nn1 =0 n1 = (n+1)(n+1)− 2 =
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
DR A
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 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 thirdorder term.
The anharmonic oscillator Hamiltonian is the harmonic oscillator Hamiltonian with an additional x 3 potential: H=
p2 2m
+ 12 mω 2 x2 + λx3
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 x3 |ψE i . 0
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.
FT
Coupled harmonic oscillators
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 P p2i 1 2P 2 H= N 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.
DR A
Remarkably, there exists a coordinate transformation to turn this problem into 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 • • • •
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. AddisonWesley. 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
DR A
Quantum indeterminacy is the apparent necessary incompleteness in the description 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 determined all the values of its measurable properties, and conversely (b) the values of its measurable properties uniquely determined the state. 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. 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 mechanics, since it had established early on by experimentalists that errors in measurement 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 measurement. Many theories have been proposed since the beginning of quantum mechanics and quantum measurement continues to be an active research area
in both theoretical and experimental physics (Braginski and Khalili 1992.) Possibly the first systematic attempt at a mathematical theory was developed by John von Neumann. The kind of measurements he investigated in (von Neumann, 1955) are now called projective measurements. That theory was based 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 (attributed by von Neumann to Paul Dirac).
DR A
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 corresponding eigenvalue of the equation A e = e. It is immediate from this that measurement in general will be non-deterministic. Quantum mechanics, moreover, gives a recipe for computing a probability distribution Pr on the possible outcomes given the initial system state is ψ. The probability is
Pr(λ) = hE(λ)ψ | ψi
where E() is the projection onto the space of eigenvectors of A with eigenvalue .
Quantum indeterminacy
453
DR A
FT
Example
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 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
σ1 =
0 1
1 , 0
σ2 =
0 i
−i , 0
σ3 =
1 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) 2 2
•
FT
•
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 measurement of both σ 1 and σ 3 have determinate values.
DR A
There are various questions that can be asked about the above indeterminacy assertion. •
Can the indeterminacy be understood as similar to an error in measurement explainable by an error parameter? More precisely, is there a hidden parameter 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?
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, although 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 understood (particularly since measurement is disturbance), but in the most natural 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
FT
to how precisely its location can be specified. This quantum uncertainty principle 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 uncertainty 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. Indeed in the quantum mechanical formalism, for a given quantum state, each one of these measurable values will be obtained non-deterministically in accordance 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.
DR A
This indeterminacy might be regarded as a kind of essential incompleteness in our description of a physical system. Notice however, that the indeterminacy 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 description 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
FT
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 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 densely defined self-adjoint operator on H. The spectral measure of A is a projectionvalued measure defined by the condition R
EA (U ) =
U
λd E(λ),
DR A
for every Borel subset U of R. Given a mixed state S, we introduce the distribution 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 • • • • • • • •
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
• •
• •
• •
DR A
•
A. Aspect, Bell’s inequality test: more ideal than ever, Nature 398 189 (1999). 244 V. Braginski and F. Khalili, Quantum Measurements, Cambridge University Press, 1992. 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, accuracy 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 description of physical reality be considered complete? 245 Phys. Rev. 47 777 (1935). 246 G. Mackey, Mathematical Foundations of Quantum Mechanics, W. A. Benjamin, 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, 1999.
FT
•
External links •
Common Misconceptions Regarding Quantum Mechanics 247 See especially part III "Misconceptions regarding measurement".
In physics, a quantum leap or quantum jump is a change of an electron within 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 continuously. 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 intermediate between those.
Vernacular usage
DR A
In the vernacular, the term quantum leap has come to mean an abrupt change, especially an advance or augmentation. The term dates back to early-to-mid20th 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 revolutionary advance, rather than an evolutionary one.
Principal Authors: Michael Hardy, Srleffler, Maveric149, Hinakana, Djinn112
Quantum levels are fixed levels with a logarithmic, descending quantum pattern in the visible spectrum of light that can be observed through a spectrometer 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.
Principal Authors: Enochlau, JYOuyang, Michael Hardy, Borofkin, Charles Matthews
Quantum mechanics
For a non-technical introduction to the topic, please see Introduction to Quantum mechanics.
Quantum mechanics is a fundamental branch of theoretical physics that replaces 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, particle physics, and nuclear physics. Along with general relativity, quantum mechanics is one of the pillars of modern physics.
Quantum mechanics
DR A
FT
460
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 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 Heisenberg, 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
FT
Quantum mechanics is a more fundamental theory than Newtonian mechanics 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 quantum 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.
DR A
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 calculation 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 quantifies the inability to precisely locate the particle. The other exemplar that led to quantum mechanics was the study of electromagnetic waves such as light. When it was found in 1900 by Max Planck that the energy of waves could be described as consisting of small packets or quanta, Albert Einstein exploited this idea to show that an electromagnetic wave such as light could be described by a particle called the photon with a discrete 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.
FT
Since the early days of quantum theory, physicists have made many attempts to combine it with the other highly successful theory of the twentieth century, Albert Einstein’s General Theory of Relativity. While quantum mechanics is entirely consistent with special relativity, serious problems emerge when one tries to join the quantum laws with general relativity, the more elaborate description of nature which includes gravity. Resolving these inconsistencies has been a major goal of twentieth- and twenty-first-century physics. Despite the proposal 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.)
DR A
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 quantum numbers quantum mechanics "reduces" to classical mechanics and classical 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, nonrelativistic quantum mechanics is also used due to its simplicity and when relativistic effects are negligible. We will use the terms quantum mechanics, quantum physics, and quantum theory synonymously, to refer to both relativistic 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 mechanics, matrix mechanics (invented by Werner Heisenberg) and wave mechanics (invented by Erwin Schrödinger).
Quantum mechanics
463
FT
In this formulation, the instantaneous state of a quantum system encodes the probabilities of its measurable properties, or "observables". Examples of observables include energy, position, momentum, and angular momentum. Observables can be either continuous (e.g., the position of a particle) or discrete (e.g., the energy of an electron bound to a hydrogen atom).
DR A
Generally, quantum mechanics does not assign definite values to observables. Instead, it makes predictions about probability distributions; that is, the probability 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 "eigenstates" 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 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 quantum 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 wavefunction. The position and momentum of the particle are observables. The 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 wavelength 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
FT
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 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 previous 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 momentum. 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 certain, 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.
DR A
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 equation, 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 forces acting on it. However, the wave packet will also spread out as time progresses, which means that the position becomes more uncertain. This also has the effect of turning position eigenstates (which can be thought of as infinitely sharp wave packets) into broadened wave packets that are no longer position eigenstates.
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 an unexcited atom is pictured classically as a particle moving in a circular trajectory around the atomic nucleus, whereas in quantum mechanics it is described by a static, spherically symmetric wavefunction surrounding the nucle-
Quantum mechanics
DR A
FT
465
us ( ). (Note that only the lowest angular momentum states, labeled s, are spherically symmetric). The time evolution of wave functions is deterministic in the sense that, given a wavefunction at an initial time, it makes a definite prediction of what the wavefunction will be at any later time. During a measurement, the change of the 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 measurement. This is one of the most difficult aspects of quantum systems to understand. It was the central topic in the famous Bohr-Einstein debates, in which the two scientists attempted to clarify these fundamental principles by way of thought experiments. In the decades after the formulation of quantum mechanics, the question of what constitutes a "measurement" has been extensively studied. 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 independent entity. For details, see the article on measurement in quantum mechanics. Quantum mechanics
466
Quantum mechanical effects
FT
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".
DR A
The first type of quantum effect is the quantization of certain physical quantities. Quantization first arose in the mathematical formulae of Max Planck in 1900 as discussed in the introduction. Max Planck was analyzing how the radiation 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 infinite, 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.
In the example we have given, of a free particle in empty space, both the position 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 valh 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 quantized, and they play an important role in many physical systems. Examples of quantized observables include angular momentum, the total energy of a bound system, and the energy contained in an electromagnetic wave of a given frequency. Another quantum effect is the uncertainty principle, which is the phenomenon that consecutive measurements of two or more observables may possess a fundamental 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 simultaneously measured with arbitrary precision, even in principle: as the precision of the position measurement improves, the maximum precision of the momentum 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
FT
Another quantum effect is the wave-particle duality. It has been shown that, under certain experimental conditions, microscopic objects like atoms or electrons 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 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 independent 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 measurement made on one particle can produce, through the collapse of the total wavefunction, an instantaneous effect on other particles with which it is entangled, even if they are far apart. (This does not conflict with special relativity because information cannot be transmitted in this way.)
DR A
Mathematical formulation
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 mechanical 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 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 equation, in which the Hamiltonian, the operator corresponding to the total energy of the system, generates time evolution.
Quantum mechanics
468
FT
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 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 encodes information about probabilities, its phase encodes information about the interference between quantum states. This gives rise to the wave-like behavior of quantum states.
DR A
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 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 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
FT
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 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 mechanics: Is there a preferred interpretation of quantum mechanics? How does the quantum description of reality, which includes elements such as the superposition of states and wavefunction collapse, give rise to the reality we perceive?
DR A
When quantum mechanics was originally formulated, it was applied to models whose correspondence limit was non-relativistic classical mechanics. For instance, the well-known model of the quantum harmonic oscillator uses an explicitly non-relativistic expression for the kinetic energy of the oscillator, and is thus a quantum version of the classical harmonic oscillator.
Early attempts to merge quantum mechanics with special relativity involved the replacement of the 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 unsatisfactory qualities stemming from their neglect of the relativistic creation and annihilation of particles. A fully relativistic quantum theory required the development of quantum field theory, which applies quantization to a field rather than a fixed set of particles. The first complete quantum field theory, quantum electrodynamics, provides a fully quantum description of the electromagnetic interaction. The full apparatus of quantum field theory is often unnecessary for describing electrodynamic systems. A simpler approach, one employed since the inception of quantum mechanics, is to treat charged particles as quantum mechanical objects being acted on by a classical electromagnetic field. For example, the elementary quantum model of the hydrogen atom describes the electric field of the hydrogen atom using a classical 1/r Coulomb potential. This "semiclassical" 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
FT
called quantum chromodynamics, and describes the interactions of the subnuclear particles: quarks and gluons. The weak nuclear force and the electromagnetic force were unified, in their quantized forms, into a single quantum field theory known as electroweak theory. It has proven difficult to construct quantum models of gravity, the remaining fundamental force. Semi-classical approximations are workable, and have led to predictions such as Hawking radiation. However, the formulation of a complete theory of quantum gravity is hindered by apparent incompatibilities between general relativity, the most accurate theory of gravity currently known, and some of the fundamental assumptions of quantum theory. The resolution of these incompatibilities is an area of active research, and theories such as string theory are among the possible candidates for a future theory of quantum gravity.
Applications of quantum theory
DR A
Quantum mechanics has had enormous success in explaining many of the features of our world. The individual behavior of the subatomic particles that make up all forms of matter - electrons, protons, neutrons, and so forth - can often only be satisfactorily described using quantum mechanics.Quantum mechanics has strongly influenced string theory, a candidate for a theory of everything (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 quantum mechanics to chemistry is known as quantum chemistry. (Relativistic) quantum mechanics can in principle mathematically describe most of chemistry. Quantum mechanics can provide quantitative insight into ionic and covalent bonding processes by explicitly showing which molecules are energetically favorable to which others, and by approximately how much. Most of the calculations performed in computational chemistry rely on quantum mechanics. Much of modern technology operates at a scale where quantum effects are significant. Examples include the laser, the transistor, the electron microscope, and magnetic resonance imaging. The study of semiconductors led to the invention of the diode and the transistor, which are indispensable for modern electronics. Researchers are currently seeking robust methods of directly manipulating quantum states. Efforts are being made to develop quantum cryptography, which will allow guaranteed secure transmission of information. A more distant goal is the development of quantum computers, which are expected to
Quantum mechanics
471
FT
perform certain computational tasks exponentially faster than classical computers. Another active research topic is quantum teleportation, which deals with techniques to transmit quantum states over arbitrary distances.
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 amplitudes and probability distributions took decades to be appreciated.
DR A
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 mechanics 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.
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 theory, 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 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 "multiverse" composed of mostly independent parallel universes. This is not accomplished by introducing some new axiom to quantum mechanics, but on the contrary by removing the axiom of the collapse of the wave packet: All the possible consistent states of the measured system and the measuring apparatus (including the observer) are present in a real physical (not just formally mathematical, as in other interpretations) quantum superposition. (Such a superposition of consistent state combinations of different systems is called an entangled state.) While the multiverse is deterministic, we perceive non-deterministic behavior governed by probabilities, because we can observe only the universe, i.e. the Quantum mechanics
472
History
FT
consistent state contribution to the mentioned superposition, we inhabit. Everett’s interpretation is perfectly consistent with John Bell’s experiments and makes them intuitively understandable. However, according to the theory of quantum decoherence, the parallel universes will never be accessible for us, 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 system 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).
DR A
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 photoelectric 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 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 physicist 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
FT
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 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.
DR A
Beginning in 1927, attempts were made to apply quantum mechanics to fields rather than single particles, resulting in what are known as quantum field theories. Early workers in this area included Dirac, Pauli, Weisskopf, and Jordan. This area of research culminated in the formulation of quantum electrodynamics by Feynman, Dyson, Schwinger, and Tomonaga during the 1940s. Quantum electrodynamics is a quantum theory of electrons, positrons, and the electromagnetic field, and served as a role model for subsequent quantum field theories. The theory of quantum chromodynamics was formulated beginning in the early 1960s. The theory as we know it today was formulated by Politzer, Gross and Wilzcek in 1975. Building on pioneering work by Schwinger, Higgs, Goldstone, Glashow, Weinberg and Salam independently showed how the weak nuclear 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 electron and its negative charge) (1897) • The study of black body radiation between 1850 and 1900, which could not be explained without quantum concepts. • The photoelectric effect: Einstein explained this in 1905 (and later received a Nobel prize for it) using the concept of photons, particles of light with quantized energy 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 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 →Measurement in quantum mechanics Quantum electrochemistry Quantum chemistry Quantum computers
• • • •
Quantum information →Quantum field theory Quantum thermodynamics Theoretical chemistry
DR A
• • • • •
FT
•
References •
P. A. M. Dirac, The Principles of Quantum Mechanics (1930) – the beginning chapters provide a very clear and comprehensible introduction • 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
•
• •
• •
DR A
•
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 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. ISBN 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 mechanics. 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. J. von Neumann, Mathematical Foundations of Quantum Mechanics, Princeton University Press, 1955. H. Weyl, The Theory of Groups and Quantum Mechanics, Dover Publications 1950.
FT
•
•
•
Notes •
Note 1: The Davisson-Germer experiment, which demonstrates the wave 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
Quantum Physics Made Relatively Simple 256: three video lectures by Hans Bethe Decoherence 257 by Erich Joos
FT
Course material:
• 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
DR A
Media:
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. •
Principal Authors: CYD, Lethe, David R. Ingham, Ancheta Wis, Voyajer, Laurascudder, Andris, Bensaccount, El C, Anville
DR A
Quantum mechanics, philosophy and controversy Quantum mechanics has had many detractors including Albert Einstein and Erwin Schroedinger. Quantum mechanics has had a profound affect on philosophy. 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 determinism 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 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 matters, to explain the position of many physicists on quantum mechanics and the profound effect that quantum mechanics has had on philosophy.
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 observation and experiment. This position is known as determinism. However, 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 entirely 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 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, attempted to reconcile the seemingly incompatible schools of thought known as empiricism (e.g., David Hume) and rationalism (e.g., René Descartes). According to the empiricists, the only possible knowledge of the world is the knowledge that can be obtained by means of perception (inductive reasoning). 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 perceive 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 object, 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
FT
rationalists, we can never be certain of any knowledge derived from perception, 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 perceived while we are awake and the same object perceived during a dream, we 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 epistemology, 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.
DR A
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, according 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 determinism. 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 thingin-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 freedom 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 possible 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 separate aspects. This method made possible tremendous advances in philosophical thought throughout the late eighteenth and early nineteenth centuries. However, it also set strict limits on human knowledge. For Kant, we cannot ’know’
Quantum mechanics, philosophy and controversy
480
FT
freedom or any other thing-in-itself in a rigorously scientific way. Rather, freedom 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 illegitimate 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 historically 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.
DR A
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 objects 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.
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 formulated 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 analysis 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 measurements 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 features 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
FT
was not accepted by Einstein or other physicists such as Alfred Lande), the uncertainty 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 calculated 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).
DR A
It is this interpretation that Einstein was questioning when he said "I cannot believe 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 Uncertainty 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 deterministic 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 distribution (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 character 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 differential equations applying a kind of resonance argument." (Albert Einstein, On Quantum Physics, 1954)
FT
Niels Bohr himself appears to have taken Kant’s view that there are two aspects 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."
DR A
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 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 debating 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." • 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 understanding the world view that quantum mechanics represents. At least I do, Quantum mechanics, philosophy and controversy
483
FT
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 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 description 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.
DR A
Sometimes Schrödinger’s wave equation is erroneously said to give the exact 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 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
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. 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 special 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
DR A
"This is the third of four lectures on a rather difficult subject – the theory of quantum electrodynamics – and since there are obviously more people here tonight than there were before, some of you haven’t heard the other two lectures 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]
"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 carry 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, although history has justified them, were burying their heads in the sand. I feel that Einstein’s intellectual superiority over Bohr, in this instance, was enormous; 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]
"Thus the last and most successful creation of theoretical physics, namely quantum 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 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 occurrence. 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)
Quantum mineralogy is the branch of physics and chemistry that uses fundamental (quantum-level) properties of particular elements to describe the macroscopic properties of minerals containing those elements.
References •
Quantum mineralogy by Bryan C. Chakoumakos in McGraw-Hill Encyclopedia of Science & Technology Online 282
In physics, a quantum phase transition (QPT) is a phase transition between 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 absolute 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.
DR A
To understand quantum phase transitions, it is useful to contrast them to classical phase transitions (CPT) (also called thermal phase transitions). A CPT describes 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. In contrast, even at zero temperature a quantum-mechanical system has quantum 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 wellstudied superconductor/insulator transition in disordered thin films which separates two quantum phases having different symmetries. Quantum magnets provide another example of QPT.
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).
In quantum mechanics, a quantum state is any possible state in which a quantum mechanical system can be. A fully specified quantum state can be described by a state vector, a wavefunction, or a complete set of quantum numbers for a specific system. A partially known quantum state, such as an ensemble with some quantum numbers fixed, can be described by a density operator.
Quantum state
488
Bra-ket notation
Basis states
FT
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 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 basis of position vectors |r>. The resulting expression Ψ(r)=, which is known as the wavefunction, is a special representation of the quantum state, namely, its projection into position space. Other representations, such as projection into momentum space, are possible. The various representations are simply different expressions of a single physical quantum state.
DR A
Any quantum state |ψi can be expressed in terms of a sum of basis states (also called basis kets) |ki i in the form 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 normalization condition mandates that the total sum of probabilities is equal to one, P 2 i |ci | = 1. The simplest understanding of basis states is obtained by examining the quantum 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 construction operator a ˆ† and a destruction operator a ˆ in what is called the ladder operator 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 2
+
√1 |1i, 2
√1 |0i 2
−
√1 |1i, 2
and
|βi =
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 √1 ( √1 |0i 2 2
Note that in the states |αi and |βi, the two states |0i and |1i each have a prob1 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
DR A
A pure quantum state is a state which can be described by a single ket vector, 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 P P P hai = hψ|A|ψi = i ai hψ|αi ihαi |ψi = i ai |hαi |ψi|2 = i ai P (αi )
where |αi i are basis kets for the operator A, and P (αi ) is the probability of |ψi being measured in state |αi i. 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 P P P [A] = hAi = s ps hψs |A|ψs i = s i ps ai |hαi |ψs i|2 = tr(ρA) 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
FT
For a mathematical discussion on states as functionals, see GNS construction. There, the same objects are described in a C*-algebraic context.
Principal Authors: Cortonin, Fresheneesz, Laussy, Bkalafut, CSTAR
Quantum statistical mechanics
Quantum statistical mechanics is the study of statistical ensembles of quantum mechanical systems. A statistical ensemble is described by a density operator S, which is a non-negative, self-adjoint, trace-class operator of trace 1 on the →Hilbert space H describing the quantum system. This can be shown under various mathematical formalisms for quantum mechanics. One such formalism is provided by quantum logic.
Expectation
From classical probability theory we know that the expectation of a random variable X is completely determined by its distribution D X by
Exp(X) =
R
R
λ d DX (λ)
assuming, of course that the random variable is integrable or the random variable is non-negative. Similarly, let A be an observable of a quantum mechanical system. A is given by a densely defined self-adjoint operator on H. The spectral measure ofRA defined by
EA (U ) =
U
λd E(λ),
Quantum statistical mechanics
491
DA (U ) = Tr(EA (U )S).
FT
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 selfadjoint projections of H. In analogy with probability theory, given a state S, we introduce the distribution of A under S which is the probability measure defined on the Borel subsets of R by
Similarly, the expected value of A is defined in terms of the probability distribution D A by
Exp(A) =
R
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 .
DR 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 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.
Von Neumann entropy
Of particular significance for describing randomness of a state is the von Neumann 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
0 λ2
0 0
··· ··· ··· ··· ···
0
and we define P
H(S) = −
i
λn ···
··· · · · · · ·
λi log2 λi .
FT
492
λ1 0 0
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 Tbe the diagonal matrix 1 2(log2 2)2
0
1 3(log2 3)2
0
··· ··· ··· ··· ···
0 0
··· · · · · · ·
DR A
T =
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 ) 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
0
1 n
··· ... ··· ···
0 0 1 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.
FT
Entropy can be used as a measure of quantum entanglement.
Gibbs canonical ensemble
Consider an ensemble of systems described by a Hamiltonian H with average energy E. If H has pure-point spectrum and the eigenvalues 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
S=
e−βH Tr(e−βH )
where β is such that the ensemble average of energy satisfies
DR A
Tr(SH) = E ,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 m Pe−βE−βE . n n
e
Under certain conditions the Gibbs canonical ensemble maximizes the von Neumann entropy of the state subject to the energy conservation requirement.
References •
J. von Neumann, Mathematical Foundations of Quantum Mechanics, Princeton University Press, 1955.
•
F. Reif, Statistical and Thermal Physics, McGraw-Hill, 1985.
Quantum superposition is the application of the superposition principle to quantum mechanics. The superposition principle is the addition of the amplitudes of waves from interference. In quantum mechanics it is the amplitudes of wavefunctions, or state vectors, that add. It occurs when an object simultaneously "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 superposition (with a probability proportional to the square of the amplitude of that eigenstate in the linear combination).
DR A
The question naturally arose as to why "real" (macroscopic, Newtonian) objects and events do not seem to display quantum mechanical features such as superposition. In 1935, Erwin Schrödinger devised a well-known thought experiment, 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 uncertainty principle and a distinct state of one observable corresponds to a superposition of many states for the other observable.
Principal Authors: Reddi, Stevenj, Stevertigo, PopUpPirate, D.328, El C, John187, Retodon8
Quantum superposition
495
FT
Quantum Theory Parallels to Consciousness 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 prediction and provide guidelines for experimental design?
DR A
Ordering in random physical processes may be attached to "mindful" meditation 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).
During so-called "unusual", "expanded" or "transcendental" states of consciousness, certain macromolecules or entire cell ensembles in the brain may be mentally 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 matter are inherent to the stochastic behavior of inanimate matter, the psychomotorical lability of animate beings, the neuropsychoemotional ambivalence of cognitive beings as well as the ill/pathological/disturbed states of living 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 localized. 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
FT
molecular couplings to their biophysical surroundings and leading to the mental quality of "clarity". Subjective "clarity" (lucid empathy, participatory interest and vivid mental imagery void of expectation) may enable such thermodynamically decoupled states to selectively entangle with others. The quantum physicist Niels Bohr once said that mental clarity is complementary to mental precision. How these ideas allow for such phenomena as precognition, telepathy - encompassing clairvoyance, distant anticipation, remote perception, synchronicity, and the like - psychic healing, and psychokinesis between the members of a fortune teller/event-, percipient/agent-, healer/patient-, or influencer/object-pair is discussed.
References
Braud, W. G., & Schlitz, M. J. (1991). Consciousness Interactions with remote biological systems: Anomalous intentionality effects. Subtle Energies, 2(1), 1-46.
•
Jahn, R. G., & Dunne, B. J. (1986). On the quantum mechanics of consciousness, 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.
•
Schmid, G. B. (2000b). Tod durch Vorstellungskraft: Das Geheimnis psychogener 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.
DR A
•
Source: http://en.wikipedia.org/wiki/Quantum_Theory_Parallels_to_Consciousness Principal Authors: Cholmes75, Omphaloscope, Harlanpaine
Quantum Theory Parallels to Consciousness
497
Quantum tomography
FT
Quantum tomography or quantum state tomography is the process of reconstructing the quantum state (density matrix) for a source of quantum systems by measurements on the systems coming from the source. To be able to uniquely 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
DR A
Quantum tunnelling
Quantum tunnelling (or tunneling) is the quantum-mechanical effect of transitioning 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 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 pronounced. 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
FT
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 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.
DR A
Alpha decay via tunnelling was also solved concurrently by Ronald Gurney and 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 mechanics that applies to many different systems. Today the theory of tunnelling is even applied to the early cosmology of the universe. Quantum tunnelling was later applied to other situations, such as the cold emission of electrons, and perhaps most importantly semiconductor and superconductor 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 using 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
d2 Ψ(x) dx2
=
2m ~2
FT
~ d − 2m Ψ(x) + V (x)Ψ(x) = EΨ(x) dx2
(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)
2m ~2
(V (x) − E)
DR A
A0 (x) + A(x)2 − B(x)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 also want to start with as high a power of Planck’s constant as possible. P i A(x) = ~1 ∞ i=0 ~ Ai (x) B(x) =
1 ~
P∞
i=0 ~
i B (x) 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))
ı
Ψ(x) ≈ C e √
dx
(E−V (x))+θ ~2 2m [4] 2 (E−V (x)) ~
FT
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 we get R p 2m
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 +
C+ e
dx
~2
(V (x)−E)
√
−
dx
+C− e (x)−E)
(V [4] 2m ~2
~2
(V (x)−E)
DR A
Ψ(x) ≈
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) ~2 amplitude be proportional to e and thus the tunnelling be exponentially dampened by large deviations from classically permitable motion. 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 (V (x) − E) in a power series. ~2 2m ~2
(V (x) − E) = U1 (x − x1 ) + U2 (x − x1 )2 + · · ·
Let us only approximate to linear order 2
d Ψ(x) dx2
2m ~2
(V (x) − E) = U1 (x − x1 )
= 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
FT
3
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 4
DR A
C− = −C sin θ −
Now we can easily construct global solutions and solve tunnelling problems. C outgoing 2 , for a particle tunnelling through a The transmission coefficient, C incoming single potential barrier is found to be R x2 p 2m −2
T =
e
x1
−2
1+ 14 e
dx
~2
(V (x)−E)
R x2 p 2m x1
dx
~2
(V (x)−E)
2
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 coefficient correctly goes to zero. This classical limit would have failed in the unphysical, but much simpler to solve, situation of a square potential.
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. AddisonWesley. ISBN 0805387145.
Quantum tunnelling
503
External links
FT
Source: http://en.wikipedia.org/wiki/Quantum_tunnelling 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.
DR A
A molecule can vibrate in many ways and each of them we can call a vibrational 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 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 rotational 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.
Principal Authors: Wolf530, Salsb, Cypa, Grendelkhan, Conscious
Quantum vibration
504
Quantum well
Fabrication
FT
A quantum well is a potential well that confines particles, which were originally 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.
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.
DR A
Applications
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 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
FT
The quantum Zeno effect is a quantum mechanical phenomenon first described by George Sudarshan and Baidyanaith Misra of the University of Texas in 1977. It describes the situation that an unstable particle, if observed continuously, 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 transition 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 “measurement” 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.
DR A
Given a system in a state A, which is the eigenstate of some measurement operator. Say the system under free time evolution will decay with a certain probability 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 system 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 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.
Experimentally, strong suppression of the evolution of a quantum system due to environmental coupling has been observed in a number of microscopic systems. One such experiment was performed in October 1989 by Itano, Heinzen, Bollinger and Wineland at NIST ( PDF 284). Approximately 5000 9Be + ions were stored in a cylindrical Penning trap and laser cooled to below 250mK. A resonant 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.
Principal Authors: Tim Starling, Seth Ilys, Domitori, Pjacobi, Gerd Breitenbach
Quasistability
Quasistability is the local stability of a system at a local minimum of a potential. 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 mechanic effect.
Qwiki 285 is a quantum physics wiki devoted to the collective creation of content 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, electrical 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
DR A
Range criterion
In quantum mechanics, in particular quantum information, the Range criterion 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 . 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
P In general, if a matrix M is of the form M = i vi vi∗ , it is obvious that the 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
FT
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 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
DR A
∗ 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
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 Positive Partial Transposition", Physics Letters A 232, (1997).
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
FT
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 radiation. In astrophysics, jets of relativistic plasma are produced by the centers of active galaxies and quasars. 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
This article is about resonance in physics. For other senses of this term, see resonance (disambiguation).
In physics, resonance is the tendency of a system to oscillate with higher amplitude when the frequency of its oscillations matches the system’s natural frequency of vibration (its resonant frequency) than it does at other frequencies.
Examples
Examples are the acoustic resonances of musical instruments, the tidal resonance 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.
Theory
FT
See also: center frequency
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: I(ω) ∝
Γ 2 2
2
(ω−Ω) +( Γ2 )
.
DR A
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 oscillators tend to have broad linewidths, and respond to a wider range of driving frequencies around the resonant frequency. The linewidth is inversely proportional to the Q factor, which is a measure of the sharpness of the resonance.
Quantum mechanics
A resonance is a quantum state whose mean energy lies above the fragmentation 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.
•
an exponential decay of the system when the system has a mean energy close to the resonance energy (time-dependent definition, i.e. in timeresolved spectroscopy) - The lifetime (or inverse of the exponent of the exponential signal) of the resonance is proportional to the inverse of its width. Resonances are usually classified into shape and Feshbach resonances or into 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 happens 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
FT
If the resonance happens to be a "fundamental particle" (i.e. described by a "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 Driven harmonic motion Formant Harmonic oscillator Impedance Q factor Resonator Schumann resonance Simple harmonic motion Tuned circuit Wave Gluonic vacuum field
DR A
• • • • • • • • • • • •
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 • Resonance versus resonant 292 • YouTube 293 - Video of the effects of resonance on rice
FT
•
Source: http://en.wikipedia.org/wiki/Resonance
Principal Authors: Omegatron, Heron, Jitse Niesen, Michael Hardy, Hyacinth, DrBob, N.MacInnes, Phys
Ring wave guide
DR A
In quantum mechanics, the ring wave guide starts from the one dimensional, time independent →Schrödinger equation: 2
~ − 2m ∇2 ψ = Eψ
Using polar coordinates on the 1 dimensional ring, the wave function depends only on the angular coordinate, and so ∇2 =
1 ∂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 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
The energy eigenvalues E are quantized because of the periodic boundary conditions, and they are required to satisfy 2mE(θ+2π) ,
= 1 = ei2πn
or
This leads to the energy eigenvalues E=
n2 ~2 2mr2
where n = 0, 1, 2, 3, . . .
The full wave functions are, therefore ψ(θ) =
√1 2π
e±inθ
Quantum states found:
DR A
n = 0:
ψ is a constant function, and E = 0. This represents a stationary particle (no angular momentum spinning around the ring).
n = 1:
E=
~2 2mr2
and
ψ(θ) =
√1 2π
e±iθ
This produces two independent states that have the same energy level (degeneracy) 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 directions. 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
Application
FT
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.
In organic chemistry, aromatic compounds contain atomic rings, such as benzene 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.
DR A
The rule that 4n + 2 excess electrons in the ring produces an exceptionally stable ("aromatic") compound, is known as the Hückel’s rule.
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 on the ground state energy. As with other variational methods, a trial wave function is tested on the system. 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: R ˆ dτ E0 < Ψ∗ HΨ
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
FT
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: P Φ= N i=1 ci Ψi . With a known hamiltonian, we can write its expected value as
PN PN PN ∗ PN ˆ c c H ci Ψi ci Ψi |H| j=1 i j ij A PNi=1 = Pi=1 ≡ B ε = Pi=1N . N PN ∗ cΨ| i=1 i i
cΨ 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 c (Hkj −εSkj ) j=1 j ∂ε = 0, ∗ = B ∂c k
DR A
which leads to a set of N secular equations: PN for k = 1, 2, ..., N . j=1 cj Hkj − εSkj = 0
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,
which in turn is true only for N values of ε. Furthermore, since the hamiltonian 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 approximation 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
FT
The Rutherford model of the atom was devised by Ernest Rutherford around 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.
DR A
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 electrons. In the Rutherford model the nucleus consisted of protons and embedded electrons, this was however proven false later.
See also •
Atomic nucleus
External Links • • •
World Of Atoms - "Rutherford’s Model" 294 Rutherford’s Model 295 Rutherford’s Model 296
In physics, Rutherford scattering is a phenomenon that was explained by 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 process probed the insides of nuclei in the 1960s, called deep inelastic scattering.
DR A
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 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 concluded 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 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 nucleus, one can write: 1 2 2 mv
=
1 4π0
·
q1 q2 b
Rearranging:
Rutherford scattering
518 b=
1 4π0
·
2q1 q2 mv 2
For an alpha particle: m (mass) = 6.7×10 -27 kg 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
FT
• • • •
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 •
DR A
E. Rutherford, The Scattering of α and β Particles by Matter and the Structure 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
Principal Authors: Sodium, Linas, Michael Hardy, Jll, HenkvD, Art Carlson, Awolf002, Andre Engels
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.
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.
DR A
The formula was invented by the Swedish physicist Johannes Rydberg and presented on November 5, 1888.
Rydberg formula for hydrogen 1 λvac
= RH Z 2
1 n21
−
1 n22
Where
λvac is the wavelength of the light emitted in vacuum, 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 1 2 → ∞ Lyman series
Converge toward
2 3 → ∞ Balmer series
365nm
3 4 → ∞ Paschen series
821nm
4 5 → ∞ Brackett series
1459nm
5 6 → ∞ Pfund series
2280nm
91nm
Rydberg formula
520 6 7 → ∞ Humphreys series 3283nm
FT
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.
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 = RZ 2 2 λvac n n 1
where
2
λvac is the wavelength of the light emitted in vacuum; R is the Rydberg constant for this element;
DR A
Z is the atomic number, i.e. the number of protons in the atomic nucleus of this element; 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.
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
References
FT
the resulting curves were similarly shaped, he sought a single function which could generate all of them when appropriate constants were inserted.
Mike Sutton, “Getting the numbers right – the lonely struggle of Rydberg” Chemistry World, Vol. 1, No. 7, July 2004.
Principal Authors: Nixdorf, Voyajer, Laurascudder, Chris Roy, Xerxes314, Tantalate, W.marsh, Wibblywobbly
DR A
Scattering channel
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 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
FT
In mathematics and physics, scattering theory is a framework for studying 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 rainbow. Examples of particle scattering includes the motion of billiard balls or the →Rutherford scattering of alpha particles by gold nucleii. More precisely, scattering 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.
DR A
The inverse scattering problem is the problem of determining the characteristics of an object (its shape, internal constitution, etc.) from measurement data of radiation or particles scattered from the object.
Since its early statement for radiolocation, the problem has found vast number 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 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 differential 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
Topics in physics
FT
to build an intuitive understanding. When two atoms are scattered off one another, one can understand them as being the bound state solutions of some differential equation. Thus, for example, the hydrogen atom corresponds to a solution to the Schroedinger equation with an inverse-square law central potential. 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.
According to the optics classification of the Optical Society of America this field consists of the following topics: Aerosol and cloud effects Atmospheric scattering Backscattering Diffusion Extinction Index measurements Inverse scattering Linewidth
Rayleigh Scattering Scattering from rough surfaces Stimulated scattering Stimulated Brillouin scattering Stimulated Raman scattering Scintillation Turbid media
DR A
• • • • • • • •
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 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
References •
FT
An important, notable development is the inverse scattering transform, central to the solution of many exactly solvable models.
Lectures of the European school on theoretical methods for electron and positron induced chemistry, Prague, Feb. 2005 299
Principal Authors: Vb, Pflatau, Charles Matthews, David R. Ingham, Fuhghettaboutit
Schrödinger equation
DR A
In physics, the Schrödinger equation, proposed by the Austrian physicist Erwin Schrödinger in 1925, describes the space- and time-dependence of quantum mechanical systems. It is of central importance to the theory of quantum mechanics, playing a role analogous to Newton’s second law in classical mechanics.
In the mathematical formulation of quantum mechanics, each system is associated 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 system. As the state of a system generally changes over time, the state vector is a 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
525 Schrödinger equation, and must be independently determined based on the physical properties of the system.
FT
Time-independent Schrödinger equation
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: One dimensional time-independent 2 ~ d ψ(x) − 2m (dx)2 2
300
:
+ U (x)ψ(x) = Eψ(x)
3-dimensional time-independent ~2 − 2m ∇2 ψ(r) + U (r)ψ(r)
301
:
= Eψ(r)
DR A
For every time-independent Hamiltonian, H, there exists a set of quantum states, |ψn i, known as energy eigenstates, and corresponding real numbers En 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 eigenvalue of the state vector with the Hamiltonian. This eigenvalue equation is referred to as the time-independent Schrödinger equation. Self-adjoint operators such as the Hamiltonian have the property that their eigenvalues are always real numbers, as we would expect since the energy is a physically observable 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 .
Energy eigenstates are convenient to work with because their time-dependence is so simple; that is why the time-independent Schrödinger equation is so useful. 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 |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 n i~ ∂c ∂t = En cn (t) .
DR A
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 P |ψ (t)i = n e−iEn t/~ cn (0) |ni .
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 reformulated 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 only a single-particle system, for which each position eigenstate may be denoted 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: R |ri hr| d3 r = I.
Schrödinger equation
527
FT
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: R |ψ (r, t)|2 d3 r = 1.
We can thus interpret the absolute square of the wavefunction as the probability 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 wavefunctions do not change with time. Energy eigenstates thus correspond to static probability distributions.
Operators in the position basis
DR A
Any operator A acting on the wavefunction is defined in the position basis by Aψ (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 ∂ 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 Hamiltonian operator typically includes partial derivatives with respect to the position 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)
T =
p2 2m
FT
where 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,
DR A
∇ is the gradient operator, and ∇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 equation is i h 2 ~ ∇2 + V (r) ψ (r) = Eψ (r) . − 2m 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 , imagine 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:
j=
~ m
·
1 2i
(ψ ∗ ∇ψ − ψ∇ψ ∗ ) =
~ m
Im (ψ ∗ ∇ψ)
Schrödinger equation
529
and measured in units of (probability)/(area × time) = r -2t -1. ∂ ∂t P
(x, t) + ∇ · j = 0
FT
The probability flux satisfies a quantum continuity equation, i.e.:
where P (x, t) is the probability density and measured in units of (probability)/(volume) = r -3. This equation is the mathematical equivalent of probability 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.
DR A
Solutions of the Schrödinger equation
Analytical solutions of the time-independent Schrödinger equation can be obtained for a variety of relatively simple conditions. These solutions provide insight into the nature of quantum phenomena and sometimes provide a reasonable approximation of the behavior of more complex systems (e.g., in statistical mechanics, molecular vibrations are often approximated as harmonic oscillators). Several of the more common analytical solutions include:
• • • • • • • • •
The free particle 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) 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 • Dirac equation
DR A
• • •
FT
• •
References
• •
E. Schrödinger, Ann. Phys. (Leipzig) 489 (1926) p.79 E. Schrödinger, Phys. Rev. 28 (1926) p. 1049
Modern reviews •
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.
Principal Authors: CYD, MathKnight, Passw0rd, Michael Hardy, Fresheneesz, Alex valavanis, Nommonomanac, Camembert, Voyajer, Linas
Schrödinger picture
In quantum mechanics, a state function is a linear combination (a superposition) of eigenstates. In the Schrödinger picture, all of these eigenstates are constantly rotating through time.
DR A
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 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 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 Hamiltonian. Taking the time derivative of |ψe (t)i yields
which is the Schrödinger equation for time evolution.
FT
Thus, in the Schrödinger formulation of quantum mechanics, all unperturbed state functions are time-harmonic. State functions in the Schrödinger picture are never entirely static, they are always undulating. This is why state functions in the Schrödinger formulation are called wavefunctions. It reveals the undulatory nature of matter: the wave-particle duality. (Actually, wavefunctions 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.
DR A
Further reading •
Principles 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
Schrödinger’s cat is a seemingly paradoxical thought experiment devised by Erwin Schrödinger that attempts to illustrate the incompleteness of an early interpretation 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 containing 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
DR A
FT
533
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."
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.
FT
The original article appeared in the German magazine Naturwissenschaften ("Natural Sciences") in 1935: E. Schrödinger: "Die gegenwärtige Situation in der Quantenmechanik" ("The present situation in quantum mechanics"), Naturwissenschaften, 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:
DR A
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 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 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 experiment to mean that while the box is closed, the system simultaneously exists in a superposition of the states "decayed nucleus/dead cat" and "undecayed nucleus/living cat", and that only when the box is opened and an observation performed does the wave function collapse into one of the two states. 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
FT
A precise rule is that probability enters at the point where the classical approximation 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 as probability. Note, however, that this only applies to descriptions of the system, not the system itself.
DR A
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 other 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 uncertainty 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 improbability 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 fundamentally 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, 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 "observe"). However, since "observation" has been shown by experiment to have nothing to do with consciousness - or at the very least, any traditional definition 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 measurement 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
FT
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? Considerable progress has been made in recent years toward the resolution of the problem, which I cannot go into here. It is enough to say that neither Bohr nor Einstein had focused on the real problem with quantum mechanics. The Copenhagen rules clearly work, so they have to be accepted. But this leaves the task of explaining them by applying the deterministic equation for the evolution of the wavefunction, the Schrödinger equation, to observers and their apparatus.
In the many-worlds interpretation of quantum mechanics, which does not single 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 mechanism of quantum decoherence is also important for the interpretation in terms of Consistent Histories. Only the "dead cat" or "alive cat" can be a part of a consistent history in this interpretation.
In other words, when the box is opened, the universe (or at least the part of the 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 retransmits 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
FT
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 performed, although a workable quantum cryptographic communications system 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 entanglement 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.
DR A
Physicist Stephen Hawking once exclaimed, "When I hear of Schrödinger’s cat, I 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 Browning!") In fact, Hawking and many other physicists are of the opinion that the "Copenhagen 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.
Related humor
Figure 44 Another joke about Schrödinger’s cat and the