ART

In quantum mechanics, the consistent histories[1] (also referred to as decoherent histories)[2] approach is intended to give a modern interpretation of quantum mechanics, generalising the conventional Copenhagen interpretation and providing a natural interpretation of quantum cosmology.[3] This interpretation of quantum mechanics is based on a consistency criterion that then allows probabilities to be assigned to various alternative histories of a system such that the probabilities for each history obey the rules of classical probability while being consistent with the Schrödinger equation. In contrast to some interpretations of quantum mechanics, particularly the Copenhagen interpretation, the framework does not include "wavefunction collapse" as a relevant description of any physical process, and emphasizes that measurement theory is not a fundamental ingredient of quantum mechanics.

Histories

A homogeneous history \( H_{i} \) (here i labels different histories) is a sequence of Propositions \( P_{i,j} \) specified at different moments of time \( t_{{i,j}} \) (here j labels the times). We write this as:

\( H_{i}=(P_{i,1},P_{i,2},\ldots ,P_{i,n_{i}}) \)

and read it as "the proposition \( P_{{i,1}} \) is true at time \( t_{{i,1}} \) and then the proposition \( P_{{i,2}} \) is true at time \( t_{{i,2}} \) and then \( \ldots \) ". The times \( t_{i,1}<t_{i,2}<\ldots <t_{i,n_{i}} \) are strictly ordered and called the temporal support of the history.

Inhomogeneous histories are multiple-time propositions which cannot be represented by a homogeneous history. An example is the logical OR of two homogeneous histories: \( {\displaystyle H_{i}\lor H_{j}} \).

These propositions can correspond to any set of questions that include all possibilities. Examples might be the three propositions meaning "the electron went through the left slit", "the electron went through the right slit" and "the electron didn't go through either slit". One of the aims of the theory is to show that classical questions such as, "where are my keys?" are consistent. In this case one might use a large number of propositions each one specifying the location of the keys in some small region of space.

Each single-time proposition \( P_{i,j} \) can be represented by a projection operator\( {\hat {P}}_{i,j} \) acting on the system's Hilbert space (we use "hats" to denote operators). It is then useful to represent homogeneous histories by the time-ordered product of their single-time projection operators. This is the history projection operator (HPO) formalism developed by Christopher Isham and naturally encodes the logical structure of the history propositions.
Consistency

An important construction in the consistent histories approach is the class operator for a homogeneous history:

\( {\displaystyle {\hat {C}}_{H_{i}}:=T\prod _{j=1}^{n_{i}}{\hat {P}}_{i,j}(t_{i,j})={\hat {P}}_{i,n_{i}}\cdots {\hat {P}}_{i,2}{\hat {P}}_{i,1}} \)

The symbol T indicates that the factors in the product are ordered chronologically according to their values of \( t_{{i,j}} \): the "past" operators with smaller values of t appear on the right side, and the "future" operators with greater values of t appear on the left side. This definition can be extended to inhomogeneous histories as well.

Central to the consistent histories is the notion of consistency. A set of histories \( \{H_{i}\} \) is consistent (or strongly consistent) if

\( \operatorname {Tr} ({\hat {C}}_{H_{i}}\rho {\hat {C}}_{H_{j}}^{\dagger })=0 \)

for all \( i\neq j \). Here \( \rho \) represents the initial density matrix, and the operators are expressed in the Heisenberg picture.

The set of histories is weakly consistent if

\( \operatorname {Tr} ({\hat {C}}_{H_{i}}\rho {\hat {C}}_{H_{j}}^{\dagger })\approx 0 \)

for all \( i\neq j . \)
Probabilities

If a set of histories is consistent then probabilities can be assigned to them in a consistent way. We postulate that the probability of history H i {\displaystyle H_{i}} H_{i} is simply

\( \operatorname {Pr} (H_{i})=\operatorname {Tr} ({\hat {C}}_{H_{i}}\rho {\hat {C}}_{H_{i}}^{\dagger }) \)

which obeys the axioms of probability if the histories \( H_{i} \) come from the same (strongly) consistent set.

As an example, this means the probability of " \( H_{i} \) OR \( H_{j} \)" equals the probability of " \( H_{i} \)" plus the probability of " \( H_{j} \)" minus the probability of " \( H_{i} \) AND \( H_{j} \)", and so forth.
Interpretation

The interpretation based on consistent histories is used in combination with the insights about quantum decoherence. Quantum decoherence implies that irreversible macroscopic phenomena (hence, all classical measurements) render histories automatically consistent, which allows one to recover classical reasoning and "common sense" when applied to the outcomes of these measurements. More precise analysis of decoherence allows (in principle) a quantitative calculation of the boundary between the classical domain and the quantum domain covariance. According to Roland Omnès,[4]

[the] history approach, although it was initially independent of the Copenhagen approach, is in some sense a more elaborate version of it. It has, of course, the advantage of being more precise, of including classical physics, and of providing an explicit logical framework for indisputable proofs. But, when the Copenhagen interpretation is completed by the modern results about correspondence and decoherence, it essentially amounts to the same physics.

[... There are] three main differences:

1. The logical equivalence between an empirical datum, which is a macroscopic phenomenon, and the result of a measurement, which is a quantum property, becomes clearer in the new approach, whereas it remained mostly tacit and questionable in the Copenhagen formulation.

2. There are two apparently distinct notions of probability in the new approach. One is abstract and directed toward logic, whereas the other is empirical and expresses the randomness of measurements. We need to understand their relation and why they coincide with the empirical notion entering into the Copenhagen rules.

3. The main difference lies in the meaning of the reduction rule for 'wave packet collapse'. In the new approach, the rule is valid but no specific effect on the measured object can be held responsible for it. Decoherence in the measuring device is enough.

In order to obtain a complete theory, the formal rules above must be supplemented with a particular Hilbert space and rules that govern dynamics, for example a Hamiltonian.

In the opinion of others[5] this still does not make a complete theory as no predictions are possible about which set of consistent histories will actually occur. That is the rules of consistent histories, the Hilbert space, and the Hamiltonian must be supplemented by a set selection rule. However, Robert B. Griffiths holds the opinion that asking the question of which set of histories will "actually occur" is a misinterpretation of the theory;[6] histories are a tool for description of reality, not separate alternate realities.

Proponents of this consistent histories interpretation—such as Murray Gell-Mann, James Hartle, Roland Omnès and Robert B. Griffiths—argue that their interpretation clarifies the fundamental disadvantages of the old Copenhagen interpretation, and can be used as a complete interpretational framework for quantum mechanics.

In Quantum Philosophy,[7] Roland Omnès provides a less mathematical way of understanding this same formalism.

The consistent histories approach can be interpreted as a way of understanding which properties of a quantum system can be treated in a single framework, and which properties must be treated in different frameworks and would produce meaningless results if combined as if they belonged to a single framework. It thus becomes possible to demonstrate formally why it is that the properties which J. S. Bell assumed could be combined together, cannot. On the other hand, it also becomes possible to demonstrate that classical, logical reasoning does apply, even to quantum experiments – but we can now be mathematically exact about how such reasoning applies.
See also

HPO formalism

References

Griffiths, Robert B. (1984). "Consistent histories and the interpretation of quantum mechanics". Journal of Statistical Physics. Springer Science and Business Media LLC. 36 (1–2): 219–272. doi:10.1007/bf01015734. ISSN 0022-4715.
Griffiths, Robert B. "The Consistent Histories Approach to Quantum Mechanics". Stanford Encyclopedia of Philosophy. Stanford University. Retrieved 2016-10-22.
Dowker, Fay; Kent, Adrian (1995-10-23). "Properties of Consistent Histories". Physical Review Letters. American Physical Society (APS). 75 (17): 3038–3041.arXiv:gr-qc/9409037. doi:10.1103/physrevlett.75.3038. ISSN 0031-9007.
Omnès, Roland (1999). Understanding Quantum Mechanics. Princeton University Press. pp. 179, 257. ISBN 978-0-691-00435-8. LCCN 98042442.
Kent, Adrian; McElwaine, Jim (1997-03-01). "Quantum prediction algorithms". Physical Review A. American Physical Society (APS). 55 (3): 1703–1720.arXiv:gr-qc/9610028. doi:10.1103/physreva.55.1703. ISSN 1050-2947.
Griffiths, R. B. (2003). Consistent Quantum Theory. Cambridge University Press.

R. Omnès, Quantum Philosophy, Princeton University Press, 1999. See part III, especially Chapter IX

External links

The Consistent Histories Approach to Quantum Mechanics – Stanford Encyclopedia of Philosophy

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