Wave function collapse

Source: Wikipedia, the free encyclopedia.
(Redirected from Wavefunction collapse)

In quantum mechanics, wave function collapse, also called reduction of the state vector,[1] occurs when a wave function—initially in a superposition of several eigenstates—reduces to a single eigenstate due to interaction with the external world. This interaction is called an observation, and is the essence of a measurement in quantum mechanics, which connects the wave function with classical observables such as position and momentum. Collapse is one of the two processes by which quantum systems evolve in time; the other is the continuous evolution governed by the Schrödinger equation.[2]

Calculations of quantum decoherence show that when a quantum system interacts with the environment, the superpositions apparently reduce to mixtures of classical alternatives. Significantly, the combined wave function of the system and environment continue to obey the Schrödinger equation throughout this apparent collapse.[3] More importantly, this is not enough to explain actual wave function collapse, as decoherence does not reduce it to a single eigenstate.[4][5]

Historically, Werner Heisenberg was the first to use the idea of wave function reduction to explain quantum measurement.[6][citation needed]

Mathematical description

Before collapsing, the wave function may be any square-integrable function, and is therefore associated with the probability density of a quantum-mechanical system. This function is expressible as a linear combination of the eigenstates of any observable. Observables represent classical dynamical variables, and when one is measured by a classical observer, the wave function is projected onto a random eigenstate of that observable. The observer simultaneously measures the classical value of that observable to be the eigenvalue of the final state.[7]

Mathematical background

The quantum state of a physical system is described by a wave function (in turn—an element of a ray in projective Hilbert space). This can be expressed as a vector using Dirac or bra–ket notation :

The kets specify the different quantum "alternatives" available—a particular quantum state. They form an orthonormal eigenbasis, formally

where represents the Kronecker delta.

An observable (i.e. measurable parameter of the system) is associated with each eigenbasis, with each quantum alternative having a specific eigenvalue of the observable. A "measurable parameter of the system" could be the usual position and the momentum of, say, a particle, but also its energy , components of spin (), orbital () and total angular () momenta, etc. In the basis representation these are respectively .

The coefficients are the probability amplitudes corresponding to each basis . These are complex numbers. The moduli square of , that is (where denotes complex conjugate), is the probability of measuring the system to be in the state .

For simplicity in the following, all wave functions are assumed to be normalized; the total probability of measuring all possible states is one:

Collapse

With these definitions it is easy to describe collapse. For any observable, the wave function is initially some linear combination of the eigenbasis of that observable. When an external agency (an observer, experimenter) measures the observable associated with the eigenbasis , the wave function collapses from the full to just one of the basis eigenstates, , that is:

The probability of collapsing to a given eigenstate is the Born probability, . Immediately post-measurement, other elements of the wave function vector, , have "collapsed" to zero, and . Unless the observable being measured commutes with the Hamiltonian, the post-measurement state will in general evolve as time progresses into a superposition of different energy eigenstates as governed by the Schrödinger equation. Unless the state projected onto upon measurement has a definite energy value, the probability of having the same measurement outcome a non-zero time later will in general be less than one.

More generally, collapse is defined for an operator with eigenbasis . If the system is in state , and is measured, the probability of collapsing the system to eigenstate and measuring the eigenvalue of with respect to would be . Note that this is not the probability that the particle is in state ; it is in state until cast to an eigenstate of .

Quantum decoherence

Quantum decoherence explains why a system interacting with an environment transitions from being a pure state, exhibiting superpositions, to a mixed state, an incoherent combination of classical alternatives.[5] This transition is fundamentally reversible, as the combined state of system and environment is still pure, but for all practical purposes irreversible, as the environment is a very large and complex quantum system, and it is not feasible to reverse their interaction. Decoherence is thus very important for explaining the classical limit of quantum mechanics, but cannot explain wave function collapse, as all classical alternatives are still present in the mixed state, and wave function collapse selects only one of them.[4][8][5]

History and context

The concept of wavefunction collapse was introduced by Werner Heisenberg in his 1927 paper on the uncertainty principle, "Über den anschaulichen Inhalt der quantentheoretischen Kinematik und Mechanik", and incorporated into the mathematical formulation of quantum mechanics by John von Neumann, in his 1932 treatise Mathematische Grundlagen der Quantenmechanik.[9] Heisenberg did not try to specify exactly what the collapse of the wavefunction meant. However, he emphasized that it should not be understood as a physical process.[10] Niels Bohr also repeatedly cautioned that we must give up a "pictorial representation", and perhaps also interpreted collapse as a formal, not physical, process.[11]

Consistent with Heisenberg, von Neumann postulated that there were two processes of wave function change:

  1. The probabilistic, non-unitary, non-local, discontinuous change brought about by observation and measurement, as outlined above.
  2. The deterministic, unitary, continuous time evolution of an isolated system that obeys the Schrödinger equation (or a relativistic equivalent, i.e. the Dirac equation).

In general, quantum systems exist in superpositions of those basis states that most closely correspond to classical descriptions, and, in the absence of measurement, evolve according to the Schrödinger equation. However, when a measurement is made, the wave function collapses—from an observer's perspective—to just one of the basis states, and the property being measured uniquely acquires the eigenvalue of that particular state, . After the collapse, the system again evolves according to the Schrödinger equation.

By explicitly dealing with the interaction of object and measuring instrument, von Neumann[2] has attempted to create consistency of the two processes of wave function change.

He was able to prove the possibility of a quantum mechanical measurement scheme consistent with wave function collapse. However, he did not prove the necessity of such a collapse. Although von Neumann's projection postulate is often presented as a normative description of quantum measurement, it was conceived by taking into account experimental evidence available during the 1930s (in particular the Compton–Simon experiment was paradigmatic), but many important present-day measurement procedures do not satisfy it (so-called measurements of the second kind).[12][13][14]

The existence of the wave function collapse is required in:

On the other hand, the collapse is considered a redundant or optional approximation in:

The cluster of phenomena described by the expression wave function collapse is a fundamental problem in the interpretation of quantum mechanics, and is known as the measurement problem.

In the Copenhagen Interpretation collapse is postulated to be a special characteristic of interaction with classical systems (of which measurements are a special case). Mathematically it can be shown that collapse is equivalent to interaction with a classical system modeled within quantum theory as systems with Boolean algebras of observables[15] and equivalent to a conditional expectation value.[16]

Everett's many-worlds interpretation deals with it by discarding the collapse-process, thus reformulating the relation between measurement apparatus and system in such a way that the linear laws of quantum mechanics are universally valid; that is, the only process according to which a quantum system evolves is governed by the Schrödinger equation or some relativistic equivalent.

A general description of the evolution of quantum mechanical systems is possible by using density operators and quantum operations. In this formalism (which is closely related to the C*-algebraic formalism) the collapse of the wave function corresponds to a non-unitary quantum operation. Within the C* formalism this non-unitary process is equivalent to the algebra gaining a non-trivial centre[17] or centre of its centralizer corresponding to classical observables.[18]

The significance ascribed to the wave function varies from interpretation to interpretation, and varies even within an interpretation (such as the Copenhagen Interpretation). If the wave function merely encodes an observer's knowledge of the universe then the wave function collapse corresponds to the receipt of new information. This is somewhat analogous to the situation in classical physics, except that the classical "wave function" does not necessarily obey a wave equation. If the wave function is physically real, in some sense and to some extent, then the collapse of the wave function is also seen as a real process, to the same extent.

See also

Notes

References

  1. ^ Penrose, Roger (May 1996). "On Gravity's role in Quantum State Reduction". General Relativity and Gravitation. 28 (5): 581–600. doi:10.1007/BF02105068. ISSN 0001-7701.
  2. ^ a b J. von Neumann (1932). Mathematische Grundlagen der Quantenmechanik (in German). Berlin: Springer.
    J. von Neumann (1955). Mathematical Foundations of Quantum Mechanics. Princeton University Press.
  3. ^ Zurek, Wojciech Hubert (2009). "Quantum Darwinism". Nature Physics. 5 (3): 181–188. arXiv:0903.5082. Bibcode:2009NatPh...5..181Z. doi:10.1038/nphys1202. S2CID 119205282.
  4. ^ a b Schlosshauer, Maximilian (2005). "Decoherence, the measurement problem, and interpretations of quantum mechanics". Rev. Mod. Phys. 76 (4): 1267–1305. arXiv:quant-ph/0312059. Bibcode:2004RvMP...76.1267S. doi:10.1103/RevModPhys.76.1267. S2CID 7295619.
  5. ^ a b c Fine, Arthur (2020). "The Role of Decoherence in Quantum Mechanics". Stanford Encyclopedia of Philosophy. Center for the Study of Language and Information, Stanford University website. Retrieved 11 April 2021.
  6. ^ Heisenberg, W. (1927). Über den anschaulichen Inhalt der quantentheoretischen Kinematik und Mechanik, Z. Phys. 43: 172–198. Translation as 'The actual content of quantum theoretical kinematics and mechanics' here
  7. ^ Griffiths, David J. (2005). Introduction to Quantum Mechanics, 2e. Upper Saddle River, New Jersey: Pearson Prentice Hall. pp. 106–109. ISBN 0131118927.
  8. ^ Wojciech H. Zurek (2003). "Decoherence, einselection, and the quantum origins of the classical". Reviews of Modern Physics. 75 (3): 715. arXiv:quant-ph/0105127. Bibcode:2003RvMP...75..715Z. doi:10.1103/RevModPhys.75.715. S2CID 14759237.
  9. ^ C. Kiefer (2002). "On the interpretation of quantum theory—from Copenhagen to the present day". arXiv:quant-ph/0210152.
  10. ^ G. Jaeger (2017). ""Wave-Packet Reduction" and the Quantum Character of the Actualization of Potentia". Entropy. 19 (10): 13. Bibcode:2017Entrp..19..513J. doi:10.3390/e19100513.
  11. ^ Henrik Zinkernagel (2016). "Niels Bohr on the wave function and the classical/quantum divide". Studies in History and Philosophy of Modern Physics. 53: 9–19. arXiv:1603.00353. Bibcode:2016SHPMP..53....9Z. doi:10.1016/j.shpsb.2015.11.001. S2CID 18890207. We can thus say that, for Bohr, the collapse is not physical in the sense of a physical wave (or something else) collapsing at a point. But it is a description – in fact the best, or most complete, description – of something happening, namely the formation of a measurement record (e.g. a dot on a photographic plate).
  12. ^ W. Pauli (1958). "Die allgemeinen Prinzipien der Wellenmechanik". In S. Flügge (ed.). Handbuch der Physik (in German). Vol. V. Berlin: Springer-Verlag. p. 73.
  13. ^ L. Landau & R. Peierls (1931). "Erweiterung des Unbestimmtheitsprinzips für die relativistische Quantentheorie". Zeitschrift für Physik (in German). 69 (1–2): 56–69. Bibcode:1931ZPhy...69...56L. doi:10.1007/BF01391513. S2CID 123160388.)
  14. ^ Discussions of measurements of the second kind can be found in most treatments on the foundations of quantum mechanics, for instance, J. M. Jauch (1968). Foundations of Quantum Mechanics. Addison-Wesley. p. 165.; B. d'Espagnat (1976). Conceptual Foundations of Quantum Mechanics. W. A. Benjamin. pp. 18, 159.; and W. M. de Muynck (2002). Foundations of Quantum Mechanics: An Empiricist Approach. Kluwer Academic Publishers. section 3.2.4.
  15. ^ Belavkin, V. P. (May 1994). "Nondemolition Principle of Quantum Measurement Theory". Foundations of Physics. 24 (5): 685–714. arXiv:quant-ph/0512188. Bibcode:1994FoPh...24..685B. doi:10.1007/BF02054669. ISSN 0015-9018. S2CID 2278990.
  16. ^ Redei, Miklos; Summers, Stephen J. (2006-08-07). "Quantum Probability Theory". arXiv:quant-ph/0601158.
  17. ^ Primas, Hans (2017). Atmanspacher, Harald (ed.). Knowledge and Time. Springer International Publishing. ISBN 978-3-319-47369-7.
  18. ^ Fröhlich, J.; Schubnel, B. (2013-10-05). "Quantum Probability Theory and the Foundations of Quantum Mechanics". arXiv:1310.1484 [quant-ph].

External links