ART

In quantum mechanics, the interaction picture (also known as the Dirac picture after Paul Dirac) is an intermediate representation between the Schrödinger picture and the Heisenberg picture. Whereas in the other two pictures either the state vector or the operators carry time dependence, in the interaction picture both carry part of the time dependence of observables.[1] The interaction picture is useful in dealing with changes to the wave functions and observables due to interactions. Most field-theoretical calculations[2] use the interaction representation because they construct the solution to the many-body Schrödinger equation as the solution to the free-particle problem plus some unknown interaction parts.

Equations that include operators acting at different times, which hold in the interaction picture, don't necessarily hold in the Schrödinger or the Heisenberg picture. This is because time-dependent unitary transformations relate operators in one picture to the analogous operators in the others.

The interaction picture is a special case of unitary transformation applied to the Hamiltonian and state vectors.

Definition

Operators and state vectors in the interaction picture are related by a change of basis (unitary transformation) to those same operators and state vectors in the Schrödinger picture.

To switch into the interaction picture, we divide the Schrödinger picture Hamiltonian into two parts:

\( {\displaystyle H_{\text{S}}=H_{0,{\text{S}}}+H_{1,{\text{S}}}.} \)

Any possible choice of parts will yield a valid interaction picture; but in order for the interaction picture to be useful in simplifying the analysis of a problem, the parts will typically be chosen so that H0,S is well understood and exactly solvable, while H1,S contains some harder-to-analyze perturbation to this system.

If the Hamiltonian has explicit time-dependence (for example, if the quantum system interacts with an applied external electric field that varies in time), it will usually be advantageous to include the explicitly time-dependent terms with H1,S, leaving H0,S time-independent. We proceed assuming that this is the case. If there is a context in which it makes sense to have H0,S be time-dependent, then one can proceed by replacing \( {\displaystyle e^{\pm iH_{0,{\text{S}}}t/\hbar }} \) by the corresponding time-evolution operator in the definitions below.
State vectors

Let \( {\displaystyle |\psi _{\text{S}}(t)\rangle ={\text{e}}^{-iH_{\text{S}}t/\hbar }|\psi (0)\rangle } \)be the time-dependent state vector in the Schrödinger picture. A state vector in the interaction picture, \( {\displaystyle |\psi _{\text{I}}(t)\rangle }\), is defined with an additional time-dependent unitary transformation.[3]

\( {\displaystyle |\psi _{\text{I}}(t)\rangle ={\text{e}}^{iH_{0,{\text{S}}}t/\hbar }|\psi _{\text{S}}(t)\rangle .} \)

Operators

An operator in the interaction picture is defined as

\( {\displaystyle A_{\text{I}}(t)=e^{iH_{0,{\text{S}}}t/\hbar }A_{\text{S}}(t)e^{-iH_{0,{\text{S}}}t/\hbar }.} \)

Note that AS(t) will typically not depend on t and can be rewritten as just AS. It only depends on t if the operator has "explicit time dependence", for example, due to its dependence on an applied external time-varying electric field.
Hamiltonian operator

For the operator\( H_{0} \) itself, the interaction picture and Schrödinger picture coincide:

\( {\displaystyle H_{0,{\text{I}}}(t)=e^{iH_{0,{\text{S}}}t/\hbar }H_{0,{\text{S}}}e^{-iH_{0,{\text{S}}}t/\hbar }=H_{0,{\text{S}}}.} \)

This is easily seen through the fact that operators commute with differentiable functions of themselves. This particular operator then can be called H 0 {\displaystyle H_{0}} H_{0} without ambiguity.

For the perturbation Hamiltonian \( {\displaystyle H_{1,{\text{I}}}} \), however,

\( {\displaystyle H_{1,{\text{I}}}(t)=e^{iH_{0,{\text{S}}}t/\hbar }H_{1,{\text{S}}}e^{-iH_{0,{\text{S}}}t/\hbar },} \)

where the interaction-picture perturbation Hamiltonian becomes a time-dependent Hamiltonian, unless [H1,S, H0,S] = 0.

It is possible to obtain the interaction picture for a time-dependent Hamiltonian H0,S(t) as well, but the exponentials need to be replaced by the unitary propagator for the evolution generated by H0,S(t), or more explicitly with a time-ordered exponential integral.
Density matrix

The density matrix can be shown to transform to the interaction picture in the same way as any other operator. In particular, let ρI and ρS be the density matrices in the interaction picture and the Schrödinger picture respectively. If there is probability pn to be in the physical state |ψn〉, then

\( {\displaystyle \rho _{\text{I}}(t)=\sum _{n}p_{n}(t)|\psi _{n,{\text{I}}}(t)\rangle \langle \psi _{n,{\text{I}}}(t)|=\sum _{n}p_{n}(t)e^{iH_{0,{\text{S}}}t/\hbar }|\psi _{n,{\text{S}}}(t)\rangle \langle \psi _{n,{\text{S}}}(t)|e^{-iH_{0,{\text{S}}}t/\hbar }=e^{iH_{0,{\text{S}}}t/\hbar }\rho _{\text{S}}(t)e^{-iH_{0,{\text{S}}}t/\hbar }.}\)

Time-evolution
Time-evolution of states

Transforming the Schrödinger equation into the interaction picture gives

\( {\displaystyle i\hbar {\frac {d}{dt}}|\psi _{\text{I}}(t)\rangle =H_{1,{\text{I}}}(t)|\psi _{\text{I}}(t)\rangle ,}\)

which states that in the interaction picture, a quantum state is evolved by the interaction part of the Hamiltonian as expressed in the interaction picture.[4]
Time-evolution of operators

If the operator AS is time-independent (i.e., does not have "explicit time dependence"; see above), then the corresponding time evolution for AI(t) is given by

\( {\displaystyle i\hbar {\frac {d}{dt}}A_{\text{I}}(t)=[A_{\text{I}}(t),H_{0,{\text{S}}}].}\)

In the interaction picture the operators evolve in time like the operators in the Heisenberg picture with the Hamiltonian H' = H0.
Time-evolution of the density matrix

The evolution of the density matrix in the interaction picture is

\( {\displaystyle i\hbar {\frac {d}{dt}}\rho _{\text{I}}(t)=[H_{1,{\text{I}}}(t),\rho _{\text{I}}(t)],}\)

in consistency with the Schrödinger equation in the interaction picture.
Expectation values

For a general operator A {\displaystyle A} A, the expectation value in the interaction picture is given by

\( {\displaystyle \langle A_{\text{I}}(t)\rangle =\langle \psi _{\text{I}}(t)|A_{\text{I}}(t)|\psi _{\text{I}}(t)\rangle =\langle \psi _{\text{S}}(t)|e^{-iH_{0,{\text{S}}}t}e^{iH_{0,{\text{S}}}t}\,A_{\text{S}}\,e^{-iH_{0,{\text{S}}}t}e^{iH_{0,{\text{S}}}t}|\psi _{\text{S}}(t)\rangle =\langle A_{\text{S}}(t)\rangle .}\)

Using the density-matrix expression for expectation value, we will get

\( {\displaystyle \langle A_{\text{I}}(t)\rangle =\operatorname {Tr} {\big (}\rho _{\text{I}}(t)\,A_{\text{I}}(t){\big )}.}\)

Use

The purpose of the interaction picture is to shunt all the time dependence due to H0 onto the operators, thus allowing them to evolve freely, and leaving only H1,I to control the time-evolution of the state vectors.

The interaction picture is convenient when considering the effect of a small interaction term, H1,S, being added to the Hamiltonian of a solved system, H0,S. By utilizing the interaction picture, one can use time-dependent perturbation theory to find the effect of H1,I,[5]:355ff e.g., in the derivation of Fermi's golden rule,[5]:359–363 or the Dyson series[5]:355–357 in quantum field theory: in 1947, Shin'ichirō Tomonaga and Julian Schwinger appreciated that covariant perturbation theory could be formulated elegantly in the interaction picture, since field operators can evolve in time as free fields, even in the presence of interactions, now treated perturbatively in such a Dyson series.
Summary comparison of evolution in all pictures

For a time-independent Hamiltonian HS, where H0,S is Free Hamiltonian,

Evolution Picture
of: Heisenberg Interaction Schrödinger
Ket state constant \( | \psi_{I}(t) \rang = e^{i H_{0, S} ~t / \hbar} | \psi_{S}(t) \rang \) \( | \psi_{S}(t) \rang = e^{-i H_{ S} ~t / \hbar} | \psi_{S}(0) \rang \)
Observable \( A_H (t)=e^{i H_{ S}~ t / \hbar} A_S e^{-i H_{ S}~ t / \hbar} \) \( A_I (t)=e^{i H_{0, S} ~t / \hbar} A_S e^{-i H_{0, S}~ t / \hbar} \) constant
Density matrix constant \( \rho_I (t)=e^{i H_{0, S} ~t / \hbar} \rho_S (t) e^{-i H_{0, S}~ t / \hbar} \) \( \rho_S (t)= e^{-i H_{ S} ~t / \hbar} \rho_S(0) e^{i H_{ S}~ t / \hbar} \)

References

Albert Messiah (1966). Quantum Mechanics, North Holland, John Wiley & Sons. ISBN 0486409244; J. J. Sakurai (1994). Modern Quantum Mechanics (Addison-Wesley) ISBN 9780201539295.
J. W. Negele, H. Orland (1988), Quantum Many-particle Systems, ISBN 0738200522.
The Interaction Picture, lecture notes from New York University.
Quantum Field Theory for the Gifted Amateur, Chapter 18 - for those who saw this being called the Schwinger-Tomonaga equation, this is not the Schwinger-Tomonaga equation. That is a generalization of the Schrödinger equation to arbitrary space-like foliations of spacetime.

Sakurai, J. J.; Napolitano, Jim (2010), Modern Quantum Mechanics (2nd ed.), Addison-Wesley, ISBN 978-0805382914

L.D. Landau; E.M. Lifshitz (1977). Quantum Mechanics: Non-Relativistic Theory. Vol. 3 (3rd ed.). Pergamon Press. ISBN 978-0-08-020940-1. Online copy

Townsend, John S. (2000). A Modern Approach to Quantum Mechanics, 2nd ed. Sausalito, California: University Science Books. ISBN 1-891389-13-0.

See also

Bra–ket notation
Schrödinger equation
Haag's theorem

vte

Quantum mechanics
Background

Introduction History
timeline Glossary Classical mechanics Old quantum theory

Fundamentals

Bra–ket notation Casimir effect Coherence Coherent control Complementarity Density matrix Energy level
degenerate levels excited state ground state QED vacuum QCD vacuum Vacuum state Zero-point energy Hamiltonian Heisenberg uncertainty principle Pauli exclusion principle Measurement Observable Operator Probability distribution Quantum Qubit Qutrit Scattering theory Spin Spontaneous parametric down-conversion Symmetry Symmetry breaking
Spontaneous symmetry breaking No-go theorem No-cloning theorem Von Neumann entropy Wave interference Wave function
collapse Universal wavefunction Wave–particle duality
Matter wave Wave propagation Virtual particle

Quantum

quantum coherence annealing decoherence entanglement fluctuation foam levitation noise nonlocality number realm state superposition system tunnelling Quantum vacuum state

Mathematics
Equations

Dirac Klein–Gordon Pauli Rydberg Schrödinger

Formulations

Heisenberg Interaction Matrix mechanics Path integral formulation Phase space Schrödinger

Other

Quantum
algebra calculus
differential stochastic geometry group Q-analog
List

Interpretations

Bayesian Consistent histories Cosmological Copenhagen de Broglie–Bohm Ensemble Hidden variables Many worlds Objective collapse Quantum logic Relational Stochastic Transactional

Experiments

Afshar Bell's inequality Cold Atom Laboratory Davisson–Germer Delayed-choice quantum eraser Double-slit Elitzur–Vaidman Franck–Hertz experiment Leggett–Garg inequality Mach-Zehnder inter. Popper Quantum eraser Quantum suicide and immortality Schrödinger's cat Stern–Gerlach Wheeler's delayed choice

Science

Measurement problem QBism

Quantum

biology chemistry chaos cognition complexity theory computing
Timeline cosmology dynamics economics finance foundations game theory information nanoscience metrology mind optics probability social science spacetime

Technologies

Quantum technology
links Matrix isolation Phase qubit Quantum dot
cellular automaton display laser single-photon source solar cell Quantum well
laser

Extensions

Dirac sea Fractional quantum mechanics Quantum electrodynamics
links Quantum geometry Quantum field theory
links Quantum gravity
links Quantum information science
links Quantum statistical mechanics Relativistic quantum mechanics De Broglie–Bohm theory Stochastic electrodynamics

Related

Quantum mechanics of time travel Textbooks

Physics Encyclopedia

World

Index

Hellenica World - Scientific Library

Retrieved from "http://en.wikipedia.org/"
All text is available under the terms of the GNU Free Documentation License