ART

In particle physics, Fermi's interaction (also the Fermi theory of beta decay or the Fermi four-fermion interaction) is an explanation of the beta decay, proposed by Enrico Fermi in 1933.[1] The theory posits four fermions directly interacting with one another (at one vertex of the associated Feynman diagram). This interaction explains beta decay of a neutron by direct coupling of a neutron with an electron, a neutrino (later determined to be an antineutrino) and a proton.[2]

Fermi first introduced this coupling in his description of beta decay in 1933.[3] The Fermi interaction was the precursor to the theory for the weak interaction where the interaction between the proton–neutron and electron–antineutrino is mediated by a virtual W− boson, of which the Fermi theory is the low-energy effective field theory.

History of initial rejection and later publication

Fermi first submitted his "tentative" theory of beta decay to the prestigious science journal Nature, which rejected it "because it contained speculations too remote from reality to be of interest to the reader.[4]" Nature later admitted the rejection to be one of the great editorial blunders in its history.[5] Fermi then submitted revised versions of the paper to Italian and German publications, which accepted and published them in those languages in 1933 and 1934.[6][7][8][9] The paper did not appear at the time in a primary publication in English.[5] An English translation of the seminal paper was published in the American Journal of Physics in 1968.[9]

Fermi found the initial rejection of the paper so troubling that he decided to take some time off from theoretical physics, and do only experimental physics. This would lead shortly to his famous work with activation of nuclei with slow neutrons.

The "tentativo"
Definitions

The theory deals with three types of particles presumed to be in direct interaction: initially a “heavy particle” in the “neutron state” ( \( {\displaystyle \rho =+1}) \), which then transitions into its “proton state” ( \( {\displaystyle \rho =-1}) \) with the emission of an electron and a neutrino.

Electron state

\( {\displaystyle \psi =\sum _{s}\psi _{s}a_{s},} \)

where \( \psi \) is the single-electron wavefunction, \( \psi _{s} \) are its stationary states.

a s {\displaystyle a_{s}} a_{s} is the operator which annihilates an electron in state s which acts on the Fock space as

\( {\displaystyle a_{s}\Psi (N_{1},N_{2},\ldots ,N_{s},\ldots )=(-1)^{N_{1}+N_{2}+\cdots +N_{s}-1}(1-N_{s})\Psi (N_{1},N_{2},\ldots ,1-N_{s},\ldots ).} \)

\( {\displaystyle a_{s}^{*}} \) is the creation operator for electron state s:

\( {\displaystyle a_{s}^{*}\Psi (N_{1},N_{2},\ldots ,N_{s},\ldots )=(-1)^{N_{1}+N_{2}+\cdots +N_{s}-1}N_{s}\Psi (N_{1},N_{2},\ldots ,1-N_{s},\ldots ).} \)

Neutrino state

Similarly,

\( {\displaystyle \phi =\sum _{\sigma }\phi _{\sigma }b_{\sigma },} \)

where \( \phi \) is the single-neutrino wavefunction, and \( {\displaystyle \phi _{\sigma }} \) are its stationary states.

\( {\displaystyle b_{\sigma }} \) is the operator which annihilates a neutrino in state \( \sigma \) which acts on the Fock space as

\( {\displaystyle b_{\sigma }\Phi (M_{1},M_{2},\ldots ,M_{\sigma },\ldots )=(-1)^{M_{1}+M_{2}+\cdots +M_{\sigma }-1}(1-M_{\sigma })\Phi (M_{1},M_{2},\ldots ,1-M_{\sigma },\ldots ).} \)

\( {\displaystyle b_{\sigma }^{*}} \) is the creation operator for neutrino state \( \sigma . \)
Heavy particle state

\( \rho \) is the operator introduced by Heisenberg (later generalized into isospin) that acts on a heavy particle state, which has eigenvalue +1 when the particle is a neutron, and −1 if the particle is a proton. Therefore, heavy particle states will be represented by two-row column vectors, where

\( {\displaystyle {\begin{pmatrix}1\\0\end{pmatrix}}} \)

represents a neutron, and

\( {\displaystyle {\begin{pmatrix}0\\1\end{pmatrix}}} \)

represents a proton (in the representation where \( \rho \) is the usual \( \sigma _{z} \) spin matrix).

The operators that change a heavy particle from a proton into a neutron and vice versa are respectively represented by

\( {\displaystyle Q=\sigma _{x}-i\sigma _{y}={\begin{pmatrix}0&1\\0&0\end{pmatrix}}} \)

and

\( {\displaystyle Q^{*}=\sigma _{x}+i\sigma _{y}={\begin{pmatrix}0&0\\1&0\end{pmatrix}}.} \)

\( u_{n} \) resp. \( v_{n} \) is an eigenfunction for a neutron resp. proton in the state n.

Hamiltonian

The Hamiltonian is composed of three parts: \( {\displaystyle H_{\text{h.p.}}} \), representing the energy of the free heavy particles, \( {\displaystyle H_{\text{l.p.}}} \), representing the energy of the free light particles, and a part giving the interaction \( {\displaystyle H_{\text{int.}}}. \)

\( {\displaystyle H_{\text{h.p.}}={\frac {1}{2}}(1+\rho )N+{\frac {1}{2}}(1-\rho )P,} \)

where N N and P are the energy operators of the neutron and proton respectively, so that if \( \rho =1 \), \( {\displaystyle H_{\text{h.p.}}=N} \), and if \({\displaystyle \rho =-1} \), \({\displaystyle H_{\text{h.p.}}=P} \).

\( {\displaystyle H_{\text{l.p.}}=\sum _{s}H_{s}N_{s}+\sum _{\sigma }K_{\sigma }M_{\sigma },}

where \( H_{s} \) is the energy of the electron in the \( {\displaystyle s^{\text{th}}} \) state in the nucleus's Coulomb field, and \( N_s \) is the number of electrons in that state; \( M_\sigma \) is the number of neutrinos in the \( {\displaystyle \sigma ^{\text{th}}} \) state, and \( {\displaystyle K_{\sigma }} \) energy of each such neutrino (assumed to be in a free, plane wave state).

The interaction part must contain a term representing the transformation of a proton into a neutron along with the emission of an electron and a neutrino (now known to be an antineutrino), as well as a term for the inverse process; the Coulomb force between the electron and proton is ignored as irrelevant to the β {\displaystyle \beta } \beta -decay process.

Fermi proposes two possible values for \( {\displaystyle H_{\text{int.}}} \) : first, a non-relativistic version which ignores spin:

H int. = g [ Q ψ ( x ) ϕ ( x ) + Q ∗ ψ ∗ ( x ) ϕ ∗ ( x ) ] , {\displaystyle H_{\text{int.}}=g\left[Q\psi (x)\phi (x)+Q^{*}\psi ^{*}(x)\phi ^{*}(x)\right],} {\displaystyle H_{\text{int.}}=g\left[Q\psi (x)\phi (x)+Q^{*}\psi ^{*}(x)\phi ^{*}(x)\right],} \)

and subsequently a version assuming that the light particles are four-component Dirac spinors, but that speed of the heavy particles is small relative to c and that the interaction terms analogous to the electromagnetic vector potential can be ignored:

\( {\displaystyle H_{\text{int.}}=g\left[Q{\tilde {\psi }}^{*}\delta \psi +Q^{*}{\tilde {\psi }}\delta \psi ^{*}\right],} \)

where \( \psi \) and \( \phi \) are now four-component Dirac spinors, \( \tilde{\psi} \)represents the Hermitian conjugate of \( \psi \), and \( \delta \) is a matrix

\( {\displaystyle {\begin{pmatrix}0&-1&0&0\\1&0&0&0\\0&0&0&1\\0&0&-1&0\end{pmatrix}}.} \)

Matrix elements

The state of the system is taken to be given by the tuple \( {\displaystyle \rho ,n,N_{1},N_{2},\ldots ,M_{1},M_{2},\ldots ,} \) where \( {\displaystyle \rho =\pm 1} \) specifies whether the heavy particle is a neutron or proton, n is the quantum state of the heavy particle, \( N_s \) is the number of electrons in state s and \( M_\sigma \) is the number of neutrinos in state \( \sigma \) .

Using the relativistic version of \( {\displaystyle H_{\text{int.}}} \) , Fermi gives the matrix element between the state with a neutron in state n and no electrons resp. neutrinos present in state s resp. \( \sigma \) , and the state with a proton in state m and an electron and a neutrino present in states s and \( \sigma \) as

\( {\displaystyle H_{\rho =-1,m,N_{s}=1,M_{\sigma }=1}^{\rho =1,n,N_{s}=0,M_{\sigma }=0}=\pm g\int v_{m}^{*}u_{n}{\tilde {\psi }}_{s}\delta \phi _{\sigma }^{*}d\tau ,} \)

where the integral is taken over the entire configuration space of the heavy particles (except for \( \rho \) ). The \( \pm \) is determined by whether the total number of light particles is odd (−) or even (+).

Transition probability
See also: Transition of state

To calculate the lifetime of a neutron in a state n {\displaystyle n} n according to the usual Quantum perturbation theory, the above matrix elements must be summed over all unoccupied electron and neutrino states. This is simplified by assuming that the electron and neutrino eigenfunctions \( \psi _{s} \) and \( {\displaystyle \phi _{\sigma }} \) are constant within the nucleus (i.e., their Compton wavelength is much smaller than the size of the nucleus). This leads to

\( {\displaystyle H_{\rho =-1,m,N_{s}=1,M_{\sigma }=1}^{\rho =1,n,N_{s}=0,M_{\sigma }=0}=\pm g{\tilde {\psi }}_{s}\delta \phi _{\sigma }^{*}\int v_{m}^{*}u_{n}d\tau ,}

where \( \psi _{s} \) and \( {\displaystyle \phi _{\sigma }} \) are now evaluated at the position of the nucleus.

According to Fermi's golden rule , the probability of this transition is

\( {\displaystyle {\begin{aligned}\left|a_{\rho =-1,m,N_{s}=1,M_{\sigma }=1}^{\rho =1,n,N_{s}=0,M_{\sigma }=0}\right|^{2}&=\left|H_{\rho =-1,m,N_{s}=1,M_{\sigma }=1}^{\rho =1,n,N_{s}=0,M_{\sigma }=0}\times {\frac {\exp {{\frac {2\pi i}{h}}(-W+H_{s}+K_{\sigma })t}-1}{-W+H_{s}+K_{\sigma }}}\right|^{2}\\&=4\left|H_{\rho =-1,m,N_{s}=1,M_{\sigma }=1}^{\rho =1,n,N_{s}=0,M_{\sigma }=0}\right|^{2}\times {\frac {\sin ^{2}\left({\frac {\pi t}{h}}(-W+H_{s}+K_{\sigma })\right)}{(-W+H_{s}+K_{\sigma })^{2}}},\end{aligned}}} \)

where W is the difference in the energy of the proton and neutron states.

Averaging over all positive-energy neutrino spin / momentum directions (where \( \Omega ^{{-1}} \) is the density of neutrino states, eventually taken to infinity), we obtain

\( {\displaystyle \left\langle \left|H_{\rho =-1,m,N_{s}=1,M_{\sigma }=1}^{\rho =1,n,N_{s}=0,M_{\sigma }=0}\right|^{2}\right\rangle _{\text{avg}}={\frac {g^{2}}{4\Omega }}\left|\int v_{m}^{*}u_{n}d\tau \right|^{2}\left({\tilde {\psi }}_{s}\psi _{s}-{\frac {\mu c^{2}}{K_{\sigma }}}{\tilde {\psi }}_{s}\beta \psi _{s}\right),}

where μ {\displaystyle \mu } \mu is the rest mass of the neutrino and β {\displaystyle \beta } \beta is the Dirac matrix.

Noting that the transition probability has a sharp maximum for values of \( p_{\sigma } \) for which \( {\displaystyle -W+H_{s}+K_{\sigma }=0} \), this simplifies to

\( {\displaystyle t{\frac {8\pi ^{3}g^{2}}{h^{4}}}\times \left|\int v_{m}^{*}u_{n}d\tau \right|^{2}{\frac {p_{\sigma }^{2}}{v_{\sigma }}}\left({\tilde {\psi }}_{s}\psi _{s}-{\frac {\mu c^{2}}{K_{\sigma }}}{\tilde {\psi }}_{s}\beta \psi _{s}\right),} \)

where \( p_{\sigma } \) and \( {\displaystyle K_{\sigma }} \) is the values for which \( {\displaystyle -W+H_{s}+K_{\sigma }=0}. \)

Fermi makes three remarks about this function:

Since the neutrino states are considered to be free, \( {\displaystyle K_{\sigma }>\mu c^{2}} \) and thus the upper limit on the continuous \( \beta \)-spectrum is \( {\displaystyle H_{s}\leq W-\mu c^{2}}. \)
Since for the electrons \( {\displaystyle H_{s}>mc^{2}} \) , in order for \( \beta \) -decay to occur, the proton–neutron energy difference must be \( {\displaystyle W\geq (m+\mu )c^{2}} \)
The factor

\( {\displaystyle Q_{mn}^{*}=\int v_{m}^{*}u_{n}d\tau } \)

in the transition probability is normally of magnitude 1, but in special circumstances it vanishes; this leads to (approximate) selection rules for β {\displaystyle \beta } \beta -decay.

Forbidden transitions
Main article: Forbidden transition

As noted above, when the inner product \( {\displaystyle Q_{mn}^{*}} \) between the heavy particle states \( u_{n} \) and \( v_{m} \) vanishes, the associated transition is "forbidden" (or, rather, much less likely than in cases where it is closer to 1).

If the description of the nucleus in terms of the individual quantum states of the protons and neutrons is good, \( {\displaystyle Q_{mn}^{*}} \) vanishes unless the neutron state \( u_{n} \) and the proton state \( v_{m} \) have the same angular momentum; otherwise, the angular momentum of the whole nucleus before and after the decay must be used.

Influence

Shortly after Fermi's paper appeared, Werner Heisenberg noted in a letter to Wolfgang Pauli[10] that the emission and absorption of neutrinos and electrons in the nucleus should, at the second order of perturbation theory, lead to an attraction between protons and neutrons, analogously to how the emission and absorption of photons leads to the electromagnetic force. He found that the force would be of the form \( {\displaystyle {\frac {\text{Const.}}{r^{5}}}} \), but that contemporary experimental data led to a value that was too small by a factor of a million.[11]

The following year, Hideki Yukawa picked up on this idea,[12] but in his theory the neutrinos and electrons were replaced by a new hypothetical particle with a rest mass approximately 200 times heavier than the electron.[13]
Later developments

Fermi's four-fermion theory describes the weak interaction remarkably well. Unfortunately, the calculated cross-section, or probability of interaction, grows as the square of the energy \( \sigma \approx G_{{{\rm {F}}}}^{2}E^{2} \). Since this cross section grows without bound, the theory is not valid at energies much higher than about 100 GeV. Here GF is the Fermi constant, which denotes the strength of the interaction. This eventually led to the replacement of the four-fermion contact interaction by a more complete theory (UV completion)—an exchange of a W or Z boson as explained in the electroweak theory.
MuonFermiDecay.gif
Fermi's interaction showing the 4-point fermion vector current, coupled under Fermi's Coupling Constant GF. Fermi's Theory was the first theoretical effort in describing nuclear decay rates for β decay.

The interaction could also explain muon decay via a coupling of a muon, electron-antineutrino, muon-neutrino and electron, with the same fundamental strength of the interaction. This hypothesis was put forward by Gershtein and Zeldovich and is known as the Vector Current Conservation hypothesis.[14]

In the original theory, Fermi assumed that the form of interaction is a contact coupling of two vector currents. Subsequently, it was pointed out by Lee and Yang that nothing prevented the appearance of an axial, parity violating current, and this was confirmed by experiments carried out by Chien-Shiung Wu.[15][16]

The inclusion of parity violation in Fermi's interaction was done by George Gamow and Edward Teller in the so-called Gamow–Teller transitions which described Fermi's interaction in terms of parity-violating "allowed" decays and parity-conserving "superallowed" decays in terms of anti-parallel and parallel electron and neutrino spin states respectively. Before the advent of the electroweak theory and the Standard Model, George Sudarshan and Robert Marshak, and also independently Richard Feynman and Murray Gell-Mann, were able to determine the correct tensor structure (vector minus axial vector, V − A) of the four-fermion interaction.[17][18]
Fermi constant

The most precise experimental determination of the Fermi constant comes from measurements of the muon lifetime, which is inversely proportional to the square of GF (when neglecting the muon mass against the mass of the W boson).[19] In modern terms:[3][20]

\( {\displaystyle G_{\rm {F}}^{0}={\frac {G_{\rm {F}}}{(\hbar c)^{3}}}={\frac {\sqrt {2}}{8}}{\frac {g^{2}}{M_{\rm {W}}^{2}c^{4}}}=1.1663787(6)\times 10^{-5}\;{\textrm {GeV}}^{-2}\approx 4.5437957\times 10^{14}\;{\textrm {J}}^{-2}\ .} \)

Here g is the coupling constant of the weak interaction, and MW is the mass of the W boson, which mediates the decay in question.

In the Standard Model, the Fermi constant is related to the Higgs vacuum expectation value

\( {\displaystyle v=\left({\sqrt {2}}\,G_{\rm {F}}^{0}\right)^{-1/2}\simeq 246.22\;{\textrm {GeV}}} \).[21]

More directly, approximately (tree level for the standard model),

\( {\displaystyle G_{\rm {F}}^{0}\simeq {\frac {\pi \alpha }{{\sqrt {2}}~M_{\rm {W}}^{2}(1-M_{\rm {W}}^{2}/M_{\rm {Z}}^{2})}}.} \)

This can be further simplified in terms of the Weinberg angle using the relation between the W and Z Bosons with \( {\displaystyle M_{\text{Z}}={\frac {M_{\text{W}}}{\cos \theta _{\text{W}}}}} \), so that

\( {\displaystyle G_{\rm {F}}^{0}\simeq {\frac {\pi \alpha }{{\sqrt {2}}~M_{\rm {Z}}^{2}\cos ^{2}\theta _{\rm {W}}\sin ^{2}\theta _{\rm {W}}}}.} \)

References

Yang, C. N. (2012). "Fermi's β-decay Theory". Asia Pacific Physics Newsletter. 1 (1): 27–30. doi:10.1142/s2251158x12000045.
Feynman, R.P. (1962). Theory of Fundamental Processes. W. A. Benjamin. Chapters 6 & 7.
Griffiths, D. (2009). Introduction to Elementary Particles (2nd ed.). pp. 314–315. ISBN 978-3-527-40601-2.
Pais, Abraham (1986). Inward Bound. Oxford: Oxford University Press. p. 418. ISBN 0-19-851997-4.
Close, Frank (February 23, 2012). Neutrino. Oxford University Press. Retrieved May 5, 2017.
Fermi, E. (1933). "Tentativo di una teoria dei raggi β". La Ricerca Scientifica (in Italian). 2 (12).
Fermi, E. (1934). "Tentativo di una teoria dei raggi β". Il Nuovo Cimento (in Italian). 11 (1): 1–19. Bibcode:1934NCim...11....1F. doi:10.1007/BF02959820. S2CID 123342095.
Fermi, E. (1934). "Versuch einer Theorie der beta-Strahlen. I". Zeitschrift für Physik (in German). 88: 161. Bibcode:1934ZPhy...88..161F. doi:10.1007/BF01351864. S2CID 125763380.
Wilson, F. L. (1968). "Fermi's Theory of Beta Decay". American Journal of Physics. 36 (12): 1150–1160. Bibcode:1968AmJPh..36.1150W. doi:10.1119/1.1974382. Includes complete English translation of Fermi's 1934 paper in German
Pauli, Wolfgang (1985). Scientific Correspondence with Bohr, Einstein, Heisenberg a.o. Volume II:1930–1939. Springer-Verlag Berlin Heidelberg GmbH. p. 250, letter #341, Heisenberg to Pauli, January 18th 1934.
Brown, Laurie M (1996). The Origin of the Concept of Nuclear Forces. Institute of Physics Publishing. Section 3.3.
Yukawa, H. (1935). "On the interaction of elementary particles. I.". Proceedings of the Physico-Mathematical Society of Japan. 17: 1.
Mehra, Jagdish (2001). The Historical Development of Quantum Theory, Volume 6 Part 2 (1932–1941). Springer. p. 832.
Gerstein, S. S.; Zeldovich, Ya. B. (1955). "Meson corrections in the theory of beta decay". Zh. Eksp. Teor. Fiz.: 698–699.
Lee, T. D.; Yang, C. N. (1956). "Question of Parity Conservation in Weak Interactions". Physical Review. 104 (1): 254–258. Bibcode:1956PhRv..104..254L. doi:10.1103/PhysRev.104.254.
Wu, C. S.; Ambler, E; Hayward, R. W.; Hoppes, D. D.; Hudson, R. P. (1957). "Experimental Test of Parity Conservation in Beta Decay". Physical Review. 105 (4): 1413–1415. Bibcode:1957PhRv..105.1413W. doi:10.1103/PhysRev.105.1413.
Feynman, R. P.; Gell-Mann, M. (1958). "Theory of the Fermi interaction" (PDF). Physical Review. 109 (1): 193. Bibcode:1958PhRv..109..193F. doi:10.1103/physrev.109.193.
Sudarshan, E. C.; Marshak, R. E. (1958). "Chirality invariance and the universal Fermi interaction". Physical Review. 109 (5): 1860. Bibcode:1958PhRv..109.1860S. doi:10.1103/physrev.109.1860.2.
Chitwood, D. B.; MuLan Collaboration; et al. (2007). "Improved Measurement of the Positive-Muon Lifetime and Determination of the Fermi Constant". Physical Review Letters. 99 (3): 032001. arXiv:0704.1981. Bibcode:2007PhRvL..99c2001C. doi:10.1103/PhysRevLett.99.032001. PMID 17678280. S2CID 3255120.
"CODATA Value: Fermi coupling constant". The NIST Reference on Constants, Units, and Uncertainty. US National Institute of Standards and Technology. June 2015. Retrieved October 31, 2016.

Plehn, T.; Rauch, M. (2005). "Quartic Higgs coupling at hadron colliders". Physical Review D. 72 (5): 053008. arXiv:hep-ph/0507321. Bibcode:2005PhRvD..72e3008P. doi:10.1103/PhysRevD.72.053008. S2CID 10737764.

vte

Quantum field theories
Standard
Theories

Chern–Simons Conformal field theory Ginzburg–Landau Kondo effect Local QFT Noncommutative QFT Quantum Yang–Mills Quartic interaction sine-Gordon String theory Toda field Topological QFT Yang–Mills Yang–Mills–Higgs

Models

Chiral Non-linear sigma Schwinger Standard Model Thirring–Wess Wess–Zumino Wess–Zumino–Witten Yukawa

Four-fermion interactions

Theories

BCS theory Fermi's interaction Luttinger liquid Top quark condensate

Models

Gross–Neveu Hubbard Nambu–Jona-Lasinio Thirring Thirring–Wess

Related

History Axiomatic QFT Loop quantum gravity Loop quantum cosmology QFT in curved spacetime Quantum chaos Quantum chromodynamics Quantum dynamics Quantum electrodynamics
links Quantum gravity
links Quantum hadrodynamics Quantum hydrodynamics Quantum information Quantum information science
links Quantum logic Quantum thermodynamics

See also: Template Template:Quantum mechanics topicsWikipedia book Book:Four-fermion interactions

vte

Standard Model
Background

Particle physics
Fermions Gauge boson Higgs boson Quantum field theory Gauge theory Strong interaction
Color charge Quantum chromodynamics Quark model Electroweak interaction
Weak interaction Quantum electrodynamics Fermi's interaction Weak hypercharge Weak isospin


Constituents

CKM matrix Spontaneous symmetry breaking Higgs mechanism Mathematical formulation of the Standard Model

Beyond the
Standard Model
Evidence

Hierarchy problem Dark matter Cosmological constant problem Strong CP problem Neutrino oscillation

Theories

Technicolor Kaluza–Klein theory Grand Unified Theory Theory of everything

Supersymmetry

MSSM Superstring theory Supergravity

Quantum gravity

String theory Loop quantum gravity Causal dynamical triangulation Canonical quantum gravity Superfluid vacuum theory Twistor theory

Experiments

Gran Sasso INO LHC SNO Super-K Tevatron

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