Fermi's interaction

Last updated

b
decay in an atomic nucleus (the accompanying antineutrino is omitted). The inset shows beta decay of a free neutron. In both processes, the intermediate emission of a virtual
W
boson (which then decays to electron and antineutrino) is not shown. Beta-minus Decay.svg

β
decay in an atomic nucleus (the accompanying antineutrino is omitted). The inset shows beta decay of a free neutron. In both processes, the intermediate emission of a virtual
W
boson
(which then decays to electron and antineutrino) is not shown.

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]

Contents

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.

According to Eugene Wigner, who together with Jordan introduced the Jordan–Wigner transformation, Fermi's paper on beta decay was his main contribution to the history of physics. [4]

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." [5] [6] It has been argued that Nature later admitted the rejection to be one of the great editorial blunders in its history, but Fermi's biographer David N. Schwartz has objected that this is both unproven and unlikely. [7] 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. [8] [9] [10] [11] 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. [11]

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” (), which then transitions into its “proton state” () with the emission of an electron and a neutrino.

Electron state

where is the single-electron wavefunction, are its stationary states.

is the operator which annihilates an electron in state which acts on the Fock space as

is the creation operator for electron state

Neutrino state

Similarly,

where is the single-neutrino wavefunction, and are its stationary states.

is the operator which annihilates a neutrino in state which acts on the Fock space as

is the creation operator for neutrino state .

Heavy particle state

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

represents a neutron, and

represents a proton (in the representation where is the usual spin matrix).

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

and

resp. is an eigenfunction for a neutron resp. proton in the state .

Hamiltonian

The Hamiltonian is composed of three parts: , representing the energy of the free heavy particles, , representing the energy of the free light particles, and a part giving the interaction .

where and are the energy operators of the neutron and proton respectively, so that if , , and if , .

where is the energy of the electron in the state in the nucleus's Coulomb field, and is the number of electrons in that state; is the number of neutrinos in the state, and 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 -decay process.

Fermi proposes two possible values for : first, a non-relativistic version which ignores spin:

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 and that the interaction terms analogous to the electromagnetic vector potential can be ignored:

where and are now four-component Dirac spinors, represents the Hermitian conjugate of , and is a matrix

Matrix elements

The state of the system is taken to be given by the tuple where specifies whether the heavy particle is a neutron or proton, is the quantum state of the heavy particle, is the number of electrons in state and is the number of neutrinos in state .

Using the relativistic version of , Fermi gives the matrix element between the state with a neutron in state and no electrons resp. neutrinos present in state resp. , and the state with a proton in state and an electron and a neutrino present in states and as

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

Transition probability

To calculate the lifetime of a neutron in a state 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 and are constant within the nucleus (i.e., their Compton wavelength is much smaller than the size of the nucleus). This leads to

where and are now evaluated at the position of the nucleus.

According to Fermi's golden rule [ further explanation needed ], the probability of this transition is

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

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

where is the rest mass of the neutrino and is the Dirac matrix.

Noting that the transition probability has a sharp maximum for values of for which , this simplifies to [ further explanation needed ]

where and is the values for which .

Fermi makes three remarks about this function:

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

Forbidden transitions

As noted above, when the inner product between the heavy particle states and 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 accurate to a good approximation, vanishes unless the neutron state and the proton state have the same angular momentum; otherwise, the total angular momentum of the entire 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 [12] 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 , but that contemporary experimental data led to a value that was too small by a factor of a million. [13]

The following year, Hideki Yukawa picked up on this idea, [14] 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. [15]

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 . 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.

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 b decay. 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. [16]

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. [17] [18]

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, VA) of the four-fermion interaction. [19] [20]

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). [21] In modern terms, the "reduced Fermi constant", that is, the constant in natural units is [3] [22]

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

. [23]

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

This can be further simplified in terms of the Weinberg angle using the relation between the W and Z bosons with , so that

Related Research Articles

<span class="mw-page-title-main">Beta decay</span> Type of radioactive decay

In nuclear physics, beta decay (β-decay) is a type of radioactive decay in which an atomic nucleus emits a beta particle, transforming into an isobar of that nuclide. For example, beta decay of a neutron transforms it into a proton by the emission of an electron accompanied by an antineutrino; or, conversely a proton is converted into a neutron by the emission of a positron with a neutrino in so-called positron emission. Neither the beta particle nor its associated (anti-)neutrino exist within the nucleus prior to beta decay, but are created in the decay process. By this process, unstable atoms obtain a more stable ratio of protons to neutrons. The probability of a nuclide decaying due to beta and other forms of decay is determined by its nuclear binding energy. The binding energies of all existing nuclides form what is called the nuclear band or valley of stability. For either electron or positron emission to be energetically possible, the energy release or Q value must be positive.

<span class="mw-page-title-main">Uncertainty principle</span> Foundational principle in quantum physics

The uncertainty principle, also known as Heisenberg's indeterminacy principle, is a fundamental concept in quantum mechanics. It states that there is a limit to the precision with which certain pairs of physical properties, such as position and momentum, can be simultaneously known. In other words, the more accurately one property is measured, the less accurately the other property can be known.

In particle physics, the Dirac equation is a relativistic wave equation derived by British physicist Paul Dirac in 1928. In its free form, or including electromagnetic interactions, it describes all spin-12 massive particles, called "Dirac particles", such as electrons and quarks for which parity is a symmetry. It is consistent with both the principles of quantum mechanics and the theory of special relativity, and was the first theory to account fully for special relativity in the context of quantum mechanics. It was validated by accounting for the fine structure of the hydrogen spectrum in a completely rigorous way.

In quantum mechanics, a density matrix is a matrix that describes the quantum state of a physical system. It allows for the calculation of the probabilities of the outcomes of any measurement performed upon this system, using the Born rule. It is a generalization of the more usual state vectors or wavefunctions: while those can only represent pure states, density matrices can also represent mixed states. Mixed states arise in quantum mechanics in two different situations:

  1. when the preparation of the system is not fully known, and thus one must deal with a statistical ensemble of possible preparations, and
  2. when one wants to describe a physical system that is entangled with another, without describing their combined state; this case is typical for a system interacting with some environment.

Electron density or electronic density is the measure of the probability of an electron being present at an infinitesimal element of space surrounding any given point. It is a scalar quantity depending upon three spatial variables and is typically denoted as either or . The density is determined, through definition, by the normalised -electron wavefunction which itself depends upon variables. Conversely, the density determines the wave function modulo up to a phase factor, providing the formal foundation of density functional theory.

A continuity equation or transport equation is an equation that describes the transport of some quantity. It is particularly simple and powerful when applied to a conserved quantity, but it can be generalized to apply to any extensive quantity. Since mass, energy, momentum, electric charge and other natural quantities are conserved under their respective appropriate conditions, a variety of physical phenomena may be described using continuity equations.

In particle, atomic and condensed matter physics, a Yukawa potential is a potential named after the Japanese physicist Hideki Yukawa. The potential is of the form:

A bound state is a composite of two or more fundamental building blocks, such as particles, atoms, or bodies, that behaves as a single object and in which energy is required to split them.

In physics, a parity transformation is the flip in the sign of one spatial coordinate. In three dimensions, it can also refer to the simultaneous flip in the sign of all three spatial coordinates :

In atomic physics, the electron magnetic moment, or more specifically the electron magnetic dipole moment, is the magnetic moment of an electron resulting from its intrinsic properties of spin and electric charge. The value of the electron magnetic moment is −9.2847647043(28)×10−24 J⋅T−1. In units of the Bohr magneton (μB), it is −1.00115965218059(13) μB, a value that was measured with a relative accuracy of 1.3×10−13.

<span class="mw-page-title-main">Charge density</span> Electric charge per unit length, area or volume

In electromagnetism, charge density is the amount of electric charge per unit length, surface area, or volume. Volume charge density is the quantity of charge per unit volume, measured in the SI system in coulombs per cubic meter (C⋅m−3), at any point in a volume. Surface charge density (σ) is the quantity of charge per unit area, measured in coulombs per square meter (C⋅m−2), at any point on a surface charge distribution on a two dimensional surface. Linear charge density (λ) is the quantity of charge per unit length, measured in coulombs per meter (C⋅m−1), at any point on a line charge distribution. Charge density can be either positive or negative, since electric charge can be either positive or negative.

The derivation of the Navier–Stokes equations as well as its application and formulation for different families of fluids, is an important exercise in fluid dynamics with applications in mechanical engineering, physics, chemistry, heat transfer, and electrical engineering. A proof explaining the properties and bounds of the equations, such as Navier–Stokes existence and smoothness, is one of the important unsolved problems in mathematics.

In physics, relativistic quantum mechanics (RQM) is any Poincaré covariant formulation of quantum mechanics (QM). This theory is applicable to massive particles propagating at all velocities up to those comparable to the speed of light c, and can accommodate massless particles. The theory has application in high energy physics, particle physics and accelerator physics, as well as atomic physics, chemistry and condensed matter physics. Non-relativistic quantum mechanics refers to the mathematical formulation of quantum mechanics applied in the context of Galilean relativity, more specifically quantizing the equations of classical mechanics by replacing dynamical variables by operators. Relativistic quantum mechanics (RQM) is quantum mechanics applied with special relativity. Although the earlier formulations, like the Schrödinger picture and Heisenberg picture were originally formulated in a non-relativistic background, a few of them also work with special relativity.

Spin is an intrinsic form of angular momentum carried by elementary particles, and thus by composite particles such as hadrons, atomic nuclei, and atoms. Spin is quantized, and accurate models for the interaction with spin require relativistic quantum mechanics or quantum field theory.

In quantum field theory, a non-topological soliton (NTS) is a soliton field configuration possessing, contrary to a topological one, a conserved Noether charge and stable against transformation into usual particles of this field for the following reason. For fixed charge Q, the mass sum of Q free particles exceeds the energy (mass) of the NTS so that the latter is energetically favorable to exist.

In Big Bang cosmology, neutrino decoupling was the epoch at which neutrinos ceased interacting with other types of matter, and thereby ceased influencing the dynamics of the universe at early times. Prior to decoupling, neutrinos were in thermal equilibrium with protons, neutrons and electrons, which was maintained through the weak interaction. Decoupling occurred approximately at the time when the rate of those weak interactions was slower than the rate of expansion of the universe. Alternatively, it was the time when the time scale for weak interactions became greater than the age of the universe at that time. Neutrino decoupling took place approximately one second after the Big Bang, when the temperature of the universe was approximately 10 billion kelvin, or 1 MeV.

In nuclear physics and atomic physics, weak charge refers to the Standard Model weak interaction coupling of a particle to the Z boson. For example, for any given nuclear isotope, the total weak charge is approximately −0.99 per neutron, and +0.07 per proton. It also shows an effect of parity violation during electron scattering.

<span class="mw-page-title-main">Symmetry in quantum mechanics</span> Properties underlying modern physics

Symmetries in quantum mechanics describe features of spacetime and particles which are unchanged under some transformation, in the context of quantum mechanics, relativistic quantum mechanics and quantum field theory, and with applications in the mathematical formulation of the standard model and condensed matter physics. In general, symmetry in physics, invariance, and conservation laws, are fundamentally important constraints for formulating physical theories and models. In practice, they are powerful methods for solving problems and predicting what can happen. While conservation laws do not always give the answer to the problem directly, they form the correct constraints and the first steps to solving a multitude of problems. In application, understanding symmetries can also provide insights on the eigenstates that can be expected. For example, the existence of degenerate states can be inferred by the presence of non commuting symmetry operators or that the non degenerate states are also eigenvectors of symmetry operators.

Lagrangian field theory is a formalism in classical field theory. It is the field-theoretic analogue of Lagrangian mechanics. Lagrangian mechanics is used to analyze the motion of a system of discrete particles each with a finite number of degrees of freedom. Lagrangian field theory applies to continua and fields, which have an infinite number of degrees of freedom.

In nuclear physics, a beta decay transition is the change in state of an atomic nucleus undergoing beta decay. (β-decay) When undergoing beta decay, a nucleus emits a beta particle and a corresponding neutrino, transforming the original nuclide into one with the same mass, but differing charge.

References

  1. Yang, C. N. (2012). "Fermi's β-decay Theory". Asia Pacific Physics Newsletter. 1 (1): 27–30. doi:10.1142/s2251158x12000045.
  2. Feynman, R.P. (1962). Theory of Fundamental Processes. W. A. Benjamin. Chapters 6 & 7.
  3. 1 2 Griffiths, D. (2009). Introduction to Elementary Particles (2nd ed.). pp. 314–315. ISBN   978-3-527-40601-2.
  4. Fermi, Enrico (2004). Fermi Remembered. University of Chicago Press. p. 241-244. ISBN   0226121119. Edited by James W. Cronin.
  5. 1 2 Close, Frank (February 23, 2012). Neutrino. Oxford University Press. p. 24. ISBN   0199695997.
  6. Pais, Abraham (1986). Inward Bound . Oxford: Oxford University Press. p.  418. ISBN   0-19-851997-4.
  7. Schwartz, David N. (2017). The Last Man Who Knew Everything. The Life and Times of Enrico Fermi, Father of the Nuclear Age. Basic Books. ISBN   0465093124. Part II, Section 8, notes 60, 61, 63. According to Schwartz, it is not proven that there was a retraction by the magazine, since the archives relating to those years were lost during a move. He argues that it is even unlikely that Fermi seriously requested publication from the journal, since at that time Nature only published short notes on such articles, and was not suitable for the publication of even a new physical theory. More suitable, if anything, would have been the Proceedings of the Royal Society .
  8. Fermi, E. (1933). "Tentativo di una teoria dei raggi β". La Ricerca Scientifica (in Italian). 2 (12).
  9. 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.
  10. Fermi, E. (1934). "Versuch einer Theorie der beta-Strahlen. I". Zeitschrift für Physik (in German). 88 (3–4): 161. Bibcode:1934ZPhy...88..161F. doi:10.1007/BF01351864. S2CID   125763380.
  11. 1 2 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
  12. 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.
  13. Brown, Laurie M (1996). The Origin of the Concept of Nuclear Forces . Institute of Physics Publishing. Section 3.3.
  14. Yukawa, H. (1935). "On the interaction of elementary particles. I.". Proceedings of the Physico-Mathematical Society of Japan. 17: 1.
  15. Mehra, Jagdish (2001). The Historical Development of Quantum Theory, Volume 6 Part 2 (1932–1941). Springer. p. 832.
  16. Gerstein, S. S.; Zeldovich, Ya. B. (1955). "Meson corrections in the theory of beta decay". Zh. Eksp. Teor. Fiz. : 698–699.
  17. 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 .
  18. 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 .
  19. 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.
  20. 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.
  21. 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.
  22. "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.
  23. 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.