Spin–statistics_theorem

Spin–statistics theorem

Spin–statistics theorem

Theorem in quantum mechanics


The spin–statistics theorem proves that the observed relationship between the intrinsic spin of a particle (angular momentum not due to the orbital motion) and the quantum particle statistics of collections of such particles is a consequence of the mathematics of quantum mechanics. In units of the reduced Planck constant ħ, all particles that move in 3 dimensions have either integer spin and obey Bose-Einstein statistics or half-integer spin and obey Fermi-Dirac statistics.[1][2]

Spin-statistics connection

All known particles obey either Fermi-Dirac statistics or Bose-Einstein statistics. A particle's intrinsic spin always predicts the statistics of a collection of such particles and vice versa:[3]

  • Integral spin particles are bosons with Bose-Einstein statistics
  • Half-integral spin particle are fermions with Fermi-Dirac statistics.

A spin-statistics theorem shows that the mathematical logic of quantum mechanics predicts or explains this physical result.[4]

The statistics of indistinguishable particles is among the most fundamental of physical effects. The Pauli exclusion principle -- that every occupied quantum state contains at most one fermion -- controls the formation of matter. The basic building blocks of matter such as protons, neutrons, and electrons are all fermions. Conversely, particles such as the photon, which mediate forces between matter particles, are all bosons.[citation needed] A spin-statistics theorem attempts explain the origin of this fundamental dichotomy.[5]:4

Background

Naively, spin, an angular momentum property intrinsic to a particle, would be unrelated to fundamental properties of a collection of such particles. However, these are indistinguishable particles: any physical prediction relating multiple indistinguishable particles must not change when the particles are exchanged.

Quantum states and indistinguishable particles

In a quantum system, a physical state is described by a state vector. A pair of distinct state vectors are physically equivalent if they differ only by an overall phase factor, ignoring other interactions. A pair of indistinguishable particles such as this have only one state. This means that if the positions of the particles are exchanged (i.e., they undergo a permutation), this does not identify a new physical state, but rather one matching the original physical state. In fact, one cannot tell which particle is in which position.

While the physical state does not change under the exchange of the particles' positions, it is possible for the state vector to change sign as a result of an exchange. Since this sign change is just an overall phase, this does not affect the physical state.

The essential ingredient in proving the spin-statistics relation is relativity, that the physical laws do not change under Lorentz transformations. The field operators transform under Lorentz transformations according to the spin of the particle that they create, by definition.

Additionally, the assumption (known as microcausality) that spacelike-separated fields either commute or anticommute can be made only for relativistic theories with a time direction. Otherwise, the notion of being spacelike is meaningless. However, the proof involves looking at a Euclidean version of spacetime, in which the time direction is treated as a spatial one, as will be now explained.

Lorentz transformations include 3-dimensional rotations and boosts. A boost transfers to a frame of reference with a different velocity and is mathematically like a rotation into time. By analytic continuation of the correlation functions of a quantum field theory, the time coordinate may become imaginary, and then boosts become rotations. The new "spacetime" has only spatial directions and is termed Euclidean.

Exchange symmetry or permutation symmetry

Bosons are particles whose wavefunction is symmetric under such an exchange or permutation, so if we swap the particles, the wavefunction does not change. Fermions are particles whose wavefunction is antisymmetric, so under such a swap the wavefunction gets a minus sign, meaning that the amplitude for two identical fermions to occupy the same state must be zero. This is the Pauli exclusion principle: two identical fermions cannot occupy the same state. This rule does not hold for bosons.

In quantum field theory, a state or a wavefunction is described by field operators operating on some basic state called the vacuum. In order for the operators to project out the symmetric or antisymmetric component of the creating wavefunction, they must have the appropriate commutation law. The operator

(with an operator and a numerical function with complex values) creates a two-particle state with wavefunction , and depending on the commutation properties of the fields, either only the antisymmetric parts or the symmetric parts matter.

Let us assume that and the two operators take place at the same time; more generally, they may have spacelike separation, as is explained hereafter.

If the fields commute, meaning that the following holds:

then only the symmetric part of contributes, so that , and the field will create bosonic particles.

On the other hand, if the fields anti-commute, meaning that has the property that

then only the antisymmetric part of contributes, so that , and the particles will be fermionic.

Proofs

An elementary explanation for the spin-statistics theorem cannot be given despite the fact that the theorem is so simple to state. In the Feynman Lectures on Physics, Richard Feynman said that this probably means that we do not have a complete understanding of the fundamental principle involved.[3]

Numerous notable proofs have been published, with different kinds of limitations and assumptions. They are all "negative proofs", meaning that they establish that integral spin fields cannot result in fermion statistics while half-integral spin fields cannot result in boson statistics.[5]:487

Proofs that avoid using any relativistic quantum field theory mechanism have defects. Many such proofs rely on a claim that

where the operator permutes the coordinates. However, the value on the left-hand-side represents the probability of particle 1 at , particle 2 at , and so on, and is thus quantum mechanically invalid for indistinguishable particles.[6]:567

The first proof was formulated[7] in 1939 Markus Fierz, a student of Wolfgang Pauli, and was rederived in a more systematic way by Pauli the following year.[8] In a later summary, Pauli listed three postulates within relativistic quantum field theory as required for these versions of the theorem:

  1. Any state with particle occupation has higher energy than the vacuum state,
  2. Spatially separated measurements do not disturb each other (they commute),
  3. Physical probabilities are positive (the metric of the Hilbert space is postive definite).

Their analysis neglected particle interactions other than commutation/anti-commutation of the state.[9][5]:374

In 1949 Richard Feynman gave a completely different type of proof[10] based on vacuum polarization which was later critiqued by Pauli.[9][5]:368 Pauli showed that Feynman's proof explicitly relied on the first two postulates he used and implicitly used the third one by first allowing negative probabilities but then rejecting field theory results with probabilities greater than one.

A proof by Julian Schwinger in 1950 based on time-reversal invariance[11] followed a proof by Frederik Belinfante in 1940 based on charge-conjugation invariance, leading to a connection to the CPT theorem more fully developed by Pauli in 1955.[12] These proofs where notably difficult to follow.[5]:393

Work on the mathematical foundations of quantum mechanics by Arthur Wightman lead to a theorem that stated that the expectation value of the product of two fields, , could be analytically continued to all separations .[5]:425 (The first two postulates of the Pauli era proofs involve the vacuum state and fields at separate locations.) The new result allowed more rigorous proofs of the spin-statistics theorems by Gerhart Luders and Bruno Zumino[13] and by Burgoyne.[5]:393 In 1957 Res Jost derived the CPT theorem using the spin-statistics theorem and Burgoyne's proof the spin-statistics theorem in 1958 required no constraints on the interactions nor on the form of the field theories. These results are among the most rigorous practical theorems.[14]:529

In spite of these successes, Feynman, in his 1963 undergraduate lecture that discussed the spin-statistics connection, says: "We apologize for the fact that we cannot give you an elementary explanation."[3] Neuenschwander echoed this in 1994, asking if there was any progress[15] spurring additional proofs and books.[5] Neuenschwander's 2013 popularization of the spin-statistics connection suggested that simple explanations remain elusive.[16]

Experimental tests

In 1987 Greenberg and Mohaparra proposed that the spin statistics theorem could have small violations.[17][18] With the help of very precise calculations for states of the He atom that violate the Pauli exclusion principle,[19] Deilamian, Gillaspy and Kelleher[20] looked for the 1s2s 1S0 of He using an atomic beam spectrometer. The search was unsuccessful with an upper limit of 5×10−6.

Relation to representation theory of the Lorentz group

The Lorentz group has no non-trivial unitary representations of finite dimension. Thus it seems impossible to construct a Hilbert space in which all states have finite, non-zero spin and positive, Lorentz-invariant norm. This problem is overcome in different ways depending on particle spin–statistics.

For a state of integer spin the negative norm states (known as "unphysical polarization") are set to zero, which makes the use of gauge symmetry necessary.

For a state of half-integer spin the argument can be circumvented by having fermionic statistics.[21]

Quasiparticle anyons in 2 dimensions

In 1982, physicist Frank Wilczek published a research paper on the possibilities of possible fractional-spin particles, which he termed anyons from their ability to take on "any" spin.[22] He wrote that they were theoretically predicted to arise in low-dimensional systems where motion is restricted to fewer than three spatial dimensions. Wilczek described their spin statistics as "interpolating continuously between the usual boson and fermion cases".[22] The effect has become the basis for understanding the fractional quantum hall effect.[23][24]

See also


References

  1. Dirac, Paul Adrien Maurice (1981-01-01). The Principles of Quantum Mechanics. Clarendon Press. p. 149. ISBN 9780198520115.
  2. Pauli, Wolfgang (1980-01-01). General principles of quantum mechanics. Springer-Verlag. ISBN 9783540098423.
  3. Feynman, Richard P.; Robert B. Leighton; Matthew Sands (1965). The Feynman Lectures on Physics, Vol. 3. Addison-Wesley. p. 4.1. ISBN 978-0-201-02118-9.
  4. Sudarshan, E. C. G. (May 1968). "The fundamental theorem on the relation between spin and statistics". Proceedings of the Indian Academy of Sciences - Section A. 67 (5): 284–293. doi:10.1007/BF03049366. ISSN 0370-0089.
  5. Duck, Ian; Sudarshan, Ennackel Chandy George; Sudarshan, E. C. G. (1998). Pauli and the spin-statistics theorem (1. reprint ed.). Singapore: World Scientific. ISBN 978-981-02-3114-9.
  6. Curceanu, Catalina; Gillaspy, J. D.; Hilborn, Robert C. (2012-07-01). "Resource Letter SS–1: The Spin-Statistics Connection". American Journal of Physics. 80 (7): 561–577. doi:10.1119/1.4704899. ISSN 0002-9505.
  7. Markus Fierz (1939). "Über die relativistische Theorie kräftefreier Teilchen mit beliebigem Spin". Helvetica Physica Acta. 12 (1): 3–37. Bibcode:1939AcHPh..12....3F. doi:10.5169/seals-110930.
  8. Richard Feynman (1961). "The theory of positrons". Quantum Electrodynamics. Basic Books. ISBN 978-0-201-36075-2. A reprint of Feynman's 1949 paper in Physical Review
  9. Julian Schwinger (June 15, 1951). "The Quantum Theory of Fields I". Physical Review. 82 (6): 914–917. Bibcode:1951PhRv...82..914S. doi:10.1103/PhysRev.82.914. S2CID 121971249.
  10. Pauli, Wolfgang (1988). "Exclusion Principle, Lorentz Group and Reflection of Space-Time and Charge". In Enz, Charles P.; v. Meyenn, Karl (eds.). Wolfgang Pauli (in German). Wiesbaden: Vieweg+Teubner Verlag. pp. 459–479. doi:10.1007/978-3-322-90270-2_41. ISBN 978-3-322-90271-9.
  11. Lüders, Gerhart; Zumino, Bruno (1958-06-15). "Connection between Spin and Statistics". Physical Review. 110 (6): 1450–1453. doi:10.1103/PhysRev.110.1450. ISSN 0031-899X.
  12. Pais, Abraham (2002). Inward bound: of matter and forces in the physical world (Reprint ed.). Oxford: Clarendon Press [u.a.] ISBN 978-0-19-851997-3.
  13. Neuenschwander, Dwight E. (1994-11-01). "Question ♯7. The spin-statistics theorem". American Journal of Physics. 62 (11): 972–972. doi:10.1119/1.17652. ISSN 0002-9505.
  14. Neuenschwander, Dwight E. (2015-07-28). "The Spin-Statistics Theorem and Identical Particle Distribution Functions". Radiations. p. 27.
  15. Greenberg, O. W.; Mohapatra, R. N. (1987-11-30). "Local Quantum Field Theory of Possible Violation of the Pauli Principle". Physical Review Letters. 59 (22): 2507–2510. doi:10.1103/PhysRevLett.59.2507. ISSN 0031-9007.
  16. Deilamian, K.; et al. (1995). "Search for small violations of the symmetrization postulate in an excited state of Helium". Phys. Rev. Lett. 74 (24): 4787–4790. Bibcode:1995PhRvL..74.4787D. doi:10.1103/PhysRevLett.74.4787. PMID 10058599.
  17. Peskin, Michael E.; Schroeder, Daniel V. (1995). An Introduction to Quantum Field Theory. Addison-Wesley. ISBN 0-201-50397-2.
  18. Laughlin, R. B. (1999-07-01). "Nobel Lecture: Fractional quantization". Reviews of Modern Physics. 71 (4): 863–874. doi:10.1103/RevModPhys.71.863. ISSN 0034-6861.
  19. Murthy, Ganpathy; Shankar, R. (2003-10-03). "Hamiltonian theories of the fractional quantum Hall effect". Reviews of Modern Physics. 75 (4): 1101–1158. doi:10.1103/RevModPhys.75.1101. ISSN 0034-6861.

Further reading


Share this article:

This article uses material from the Wikipedia article Spin–statistics_theorem, and is written by contributors. Text is available under a CC BY-SA 4.0 International License; additional terms may apply. Images, videos and audio are available under their respective licenses.