Neutrino oscillation

Last updated

Neutrino oscillation is a quantum mechanical phenomenon in which a neutrino created with a specific lepton family number ("lepton flavor": electron, muon, or tau) can later be measured to have a different lepton family number. The probability of measuring a particular flavor for a neutrino varies between three known states, as it propagates through space. [1]

Contents

First predicted by Bruno Pontecorvo in 1957, [2] [3] neutrino oscillation has since been observed by a multitude of experiments in several different contexts. Most notably, the existence of neutrino oscillation resolved the long-standing solar neutrino problem.

Neutrino oscillation is of great theoretical and experimental interest, as the precise properties of the process can shed light on several properties of the neutrino. In particular, it implies that the neutrino has a non-zero mass outside the Einstein-Cartan torsion, [4] [5] which requires a modification to the Standard Model of particle physics. [1] The experimental discovery of neutrino oscillation, and thus neutrino mass, by the Super-Kamiokande Observatory and the Sudbury Neutrino Observatories was recognized with the 2015 Nobel Prize for Physics. [6]

Observations

A great deal of evidence for neutrino oscillation has been collected from many sources, over a wide range of neutrino energies and with many different detector technologies. [7] The 2015 Nobel Prize in Physics was shared by Takaaki Kajita and Arthur B. McDonald for their early pioneering observations of these oscillations.

Neutrino oscillation is a function of the ratio LE, where L is the distance traveled and E is the neutrino's energy. (Details in § Propagation and interference below.) All available neutrino sources produce a range of energies, and oscillation is measured at a fixed distance for neutrinos of varying energy. The limiting factor in measurements is the accuracy with which the energy of each observed neutrino can be measured. Because current detectors have energy uncertainties of a few percent, it is satisfactory to know the distance to within 1%.

Solar neutrino oscillation

The first experiment that detected the effects of neutrino oscillation was Ray Davis' Homestake experiment in the late 1960s, in which he observed a deficit in the flux of solar neutrinos with respect to the prediction of the Standard Solar Model, using a chlorine-based detector. [8] This gave rise to the solar neutrino problem. Many subsequent radiochemical and water Cherenkov detectors confirmed the deficit, but neutrino oscillation was not conclusively identified as the source of the deficit until the Sudbury Neutrino Observatory provided clear evidence of neutrino flavor change in 2001. [9]

Solar neutrinos have energies below 20  MeV. At energies above 5 MeV, solar neutrino oscillation actually takes place in the Sun through a resonance known as the MSW effect, a different process from the vacuum oscillation described later in this article. [1]

Atmospheric neutrino oscillation

Following the theories that were proposed in the 1970s suggesting unification of electromagnetic, weak, and strong forces, a few experiments on proton decay followed in the 1980s. Large detectors such as IMB, MACRO, and Kamiokande II have observed a deficit in the ratio of the flux of muon to electron flavor atmospheric neutrinos (see muon decay ). The Super-Kamiokande experiment provided a very precise measurement of neutrino oscillation in an energy range of hundreds of MeV to a few TeV, and with a baseline of the diameter of the Earth; the first experimental evidence for atmospheric neutrino oscillations was announced in 1998. [10]

Reactor neutrino oscillation

Illustration of neutrino oscillations. 41467 2015 Article BFncomms7935 Fig1 HTML-en.svg
Illustration of neutrino oscillations.

Many experiments have searched for oscillation of electron anti-neutrinos produced in nuclear reactors. No oscillations were found until a detector was installed at a distance 1–2 km. Such oscillations give the value of the parameter θ13. Neutrinos produced in nuclear reactors have energies similar to solar neutrinos, of around a few MeV. The baselines of these experiments have ranged from tens of meters to over 100 km (parameter θ12). Mikaelyan and Sinev proposed to use two identical detectors to cancel systematic uncertainties in reactor experiment to measure the parameter θ13. [11]

In December 2011, the Double Chooz experiment found that θ13  0. [12] Then, in 2012, the Daya Bay experiment found that θ13  0 with a significance of 5.2 σ; [13] These results have since been confirmed by RENO. [14]

Beam neutrino oscillation

Neutrino beams produced at a particle accelerator offer the greatest control over the neutrinos being studied. Many experiments have taken place that study the same oscillations as in atmospheric neutrino oscillation using neutrinos with a few GeV of energy and several-hundred-km baselines. The MINOS, K2K, and Super-K experiments have all independently observed muon neutrino disappearance over such long baselines. [1]

Data from the LSND experiment appear to be in conflict with the oscillation parameters measured in other experiments. Results from the MiniBooNE appeared in Spring 2007 and contradicted the results from LSND, although they could support the existence of a fourth neutrino type, the sterile neutrino. [1]

In 2010, the INFN and CERN announced the observation of a tau particle in a muon neutrino beam in the OPERA detector located at Gran Sasso, 730 km away from the source in Geneva. [15]

T2K, using a neutrino beam directed through 295 km of earth and the Super-Kamiokande detector, measured a non-zero value for the parameter θ13 in a neutrino beam. [16] NOνA, using the same beam as MINOS with a baseline of 810 km, is sensitive to the same.

Theory

Neutrino oscillation arises from mixing between the flavor and mass eigenstates of neutrinos. That is, the three neutrino states that interact with the charged leptons in weak interactions are each a different superposition of the three (propagating) neutrino states of definite mass. Neutrinos are emitted and absorbed in weak processes in flavor eigenstates [lower-alpha 1] but travel as mass eigenstates. [17]

As a neutrino superposition propagates through space, the quantum mechanical phases of the three neutrino mass states advance at slightly different rates, due to the slight differences in their respective masses. This results in a changing superposition mixture of mass eigenstates as the neutrino travels; but a different mixture of mass eigenstates corresponds to a different mixture of flavor states. For example, a neutrino born as an electron neutrino will be some mixture of electron, mu, and tau neutrino after traveling some distance. Since the quantum mechanical phase advances in a periodic fashion, after some distance the state will nearly return to the original mixture, and the neutrino will be again mostly electron neutrino. The electron flavor content of the neutrino will then continue to oscillate as long as the quantum mechanical state maintains coherence. Since mass differences between neutrino flavors are small in comparison with long coherence lengths for neutrino oscillations, this microscopic quantum effect becomes observable over macroscopic distances.

In contrast, due to their larger masses, the charged leptons (electrons, muons, and tau leptons) have never been observed to oscillate. In nuclear beta decay, muon decay, pion decay, and kaon decay, when a neutrino and a charged lepton are emitted, the charged lepton is emitted in incoherent mass eigenstates such as |
e
〉, because of its large mass. Weak-force couplings compel the simultaneously emitted neutrino to be in a "charged-lepton-centric" superposition such as |
ν
e
〉, which is an eigenstate for a "flavor" that is fixed by the electron's mass eigenstate, and not in one of the neutrino's own mass eigenstates. Because the neutrino is in a coherent superposition that is not a mass eigenstate, the mixture that makes up that superposition oscillates significantly as it travels. No analogous mechanism exists in the Standard Model that would make charged leptons detectably oscillate. In the four decays mentioned above, where the charged lepton is emitted in a unique mass eigenstate, the charged lepton will not oscillate, as single mass eigenstates propagate without oscillation.

The case of (real) W boson decay is more complicated: W boson decay is sufficiently energetic to generate a charged lepton that is not in a mass eigenstate; however, the charged lepton would lose coherence, if it had any, over interatomic distances (0.1  nm) and would thus quickly cease any meaningful oscillation. More importantly, no mechanism in the Standard Model is capable of pinning down a charged lepton into a coherent state that is not a mass eigenstate, in the first place; instead, while the charged lepton from the W boson decay is not initially in a mass eigenstate, neither is it in any "neutrino-centric" eigenstate, nor in any other coherent state. It cannot meaningfully be said that such a featureless charged lepton oscillates or that it does not oscillate, as any "oscillation" transformation would just leave it the same generic state that it was before the oscillation. Therefore, detection of a charged lepton oscillation from W boson decay is infeasible on multiple levels. [18] [19]

Pontecorvo–Maki–Nakagawa–Sakata matrix

The idea of neutrino oscillation was first put forward in 1957 by Bruno Pontecorvo, who proposed that neutrino–antineutrino transitions may occur in analogy with neutral kaon mixing. [2] Although such matter–antimatter oscillation had not been observed, this idea formed the conceptual foundation for the quantitative theory of neutrino flavor oscillation, which was first developed by Maki, Nakagawa, and Sakata in 1962 [20] and further elaborated by Pontecorvo in 1967. [3] One year later the solar neutrino deficit was first observed, [21] and that was followed by the famous article by Gribov and Pontecorvo published in 1969 titled "Neutrino astronomy and lepton charge". [22]

The concept of neutrino mixing is a natural outcome of gauge theories with massive neutrinos, and its structure can be characterized in general. [23] In its simplest form it is expressed as a unitary transformation relating the flavor and mass eigenbasis and can be written as


where

represents the Pontecorvo–Maki–Nakagawa–Sakata matrix (also called the PMNS matrix, lepton mixing matrix, or sometimes simply the MNS matrix). It is the analogue of the CKM matrix describing the analogous mixing of quarks. If this matrix were the identity matrix, then the flavor eigenstates would be the same as the mass eigenstates. However, experiment shows that it is not.

When the standard three-neutrino theory is considered, the matrix is 3×3. If only two neutrinos are considered, a 2×2 matrix is used. If one or more sterile neutrinos are added (see later), it is 4×4 or larger. In the 3×3 form, it is given by [24]

where cij cos θij, and sij sin θij. The phase factors α1 and α2 are physically meaningful only if neutrinos are Majorana particles — i.e. if the neutrino is identical to its antineutrino (whether or not they are is unknown) — and do not enter into oscillation phenomena regardless. If neutrinoless double beta decay occurs, these factors influence its rate. The phase factor δ is non-zero only if neutrino oscillation violates CP symmetry; this has not yet been observed experimentally. If experiment shows this 3×3 matrix to be not unitary, a sterile neutrino or some other new physics is required.

Propagation and interference

Since are mass eigenstates, their propagation can be described by plane wave solutions of the form

where

In the ultrarelativistic limit, we can approximate the energy as

where E is the energy of the wavepacket (particle) to be detected.

This limit applies to all practical (currently observed) neutrinos, since their masses are less than 1 eV and their energies are at least 1 MeV, so the Lorentz factor, γ, is greater than 106 in all cases. Using also tL, where L is the distance traveled and also dropping the phase factors, the wavefunction becomes

Eigenstates with different masses propagate with different frequencies. The heavier ones oscillate faster compared to the lighter ones. Since the mass eigenstates are combinations of flavor eigenstates, this difference in frequencies causes interference between the corresponding flavor components of each mass eigenstate. Constructive interference causes it to be possible to observe a neutrino created with a given flavor to change its flavor during its propagation. The probability that a neutrino originally of flavor α will later be observed as having flavor β is

This is more conveniently written as

where

The phase that is responsible for oscillation is often written as (with c and restored)

where 1.27 is unitless. In this form, it is convenient to plug in the oscillation parameters since:

If there is no CP-violation (δ is zero), then the second sum is zero. Otherwise, the CP asymmetry can be given as

In terms of Jarlskog invariant

the CP asymmetry is expressed as

Two-neutrino case

The above formula is correct for any number of neutrino generations. Writing it explicitly in terms of mixing angles is extremely cumbersome if there are more than two neutrinos that participate in mixing. Fortunately, there are several meaningful cases in which only two neutrinos participate significantly. In this case, it is sufficient to consider the mixing matrix

Then the probability of a neutrino changing its flavor is

Or, using SI units and the convention introduced above

This formula is often appropriate for discussing the transition νμντ in atmospheric mixing, since the electron neutrino plays almost no role in this case. It is also appropriate for the solar case of νeνx , where νx is a mix (superposition) of νμ and ντ . These approximations are possible because the mixing angle θ13 is very small and because two of the mass states are very close in mass compared to the third.

Classical analogue of neutrino oscillation

Coupled.svg
Spring-coupled pendulums
Schwebungsfall.svg
Time evolution of the pendulums
Gleichsinnig.svg
Lower frequency normal mode
Gegensinnig.svg
Higher frequency normal mode

The basic physics behind neutrino oscillation can be found in any system of coupled harmonic oscillators. A simple example is a system of two pendulums connected by a weak spring (a spring with a small spring constant). The first pendulum is set in motion by the experimenter while the second begins at rest. Over time, the second pendulum begins to swing under the influence of the spring, while the first pendulum's amplitude decreases as it loses energy to the second. Eventually all of the system's energy is transferred to the second pendulum and the first is at rest. The process then reverses. The energy oscillates between the two pendulums repeatedly until it is lost to friction.

The behavior of this system can be understood by looking at its normal modes of oscillation. If the two pendulums are identical then one normal mode consists of both pendulums swinging in the same direction with a constant distance between them, while the other consists of the pendulums swinging in opposite (mirror image) directions. These normal modes have (slightly) different frequencies because the second involves the (weak) spring while the first does not. The initial state of the two-pendulum system is a combination of both normal modes. Over time, these normal modes drift out of phase, and this is seen as a transfer of motion from the first pendulum to the second.

The description of the system in terms of the two pendulums is analogous to the flavor basis of neutrinos. These are the parameters that are most easily produced and detected (in the case of neutrinos, by weak interactions involving the W boson). The description in terms of normal modes is analogous to the mass basis of neutrinos. These modes do not interact with each other when the system is free of outside influence.

When the pendulums are not identical the analysis is slightly more complicated. In the small-angle approximation, the potential energy of a single pendulum system is , where g is the standard gravity, L is the length of the pendulum, m is the mass of the pendulum, and x is the horizontal displacement of the pendulum. As an isolated system the pendulum is a harmonic oscillator with a frequency of . The potential energy of a spring is where k is the spring constant and x is the displacement. With a mass attached it oscillates with a period of . With two pendulums (labeled a and b) of equal mass but possibly unequal lengths and connected by a spring, the total potential energy is

This is a quadratic form in xa and xb, which can also be written as a matrix product:

The 2×2 matrix is real symmetric and so (by the spectral theorem) it is orthogonally diagonalizable . That is, there is an angle θ such that if we define

then

where λ1 and λ2 are the eigenvalues of the matrix. The variables x1 and x2 describe normal modes which oscillate with frequencies of and . When the two pendulums are identical (La = Lb), θ is 45°.

The angle θ is analogous to the Cabibbo angle (though that angle applies to quarks rather than neutrinos).

When the number of oscillators (particles) is increased to three, the orthogonal matrix can no longer be described by a single angle; instead, three are required (Euler angles). Furthermore, in the quantum case, the matrices may be complex. This requires the introduction of complex phases in addition to the rotation angles, which are associated with CP violation but do not influence the observable effects of neutrino oscillation.

Theory, graphically

Two neutrino probabilities in vacuum

In the approximation where only two neutrinos participate in the oscillation, the probability of oscillation follows a simple pattern:

Oscillations two neutrino.svg

The blue curve shows the probability of the original neutrino retaining its identity. The red curve shows the probability of conversion to the other neutrino. The maximum probability of conversion is equal to sin22θ. The frequency of the oscillation is controlled by Δm2.

Three neutrino probabilities

If three neutrinos are considered, the probability for each neutrino to appear is somewhat complex. The graphs below show the probabilities for each flavor, with the plots in the left column showing a long range to display the slow "solar" oscillation, and the plots in the right column zoomed in, to display the fast "atmospheric" oscillation. The parameters used to create these graphs (see below) are consistent with current measurements, but since some parameters are still quite uncertain, some aspects of these plots are only qualitatively correct. [25]

Electron neutrino oscillations, long range. Here and in the following diagrams black means electron neutrino, blue means muon neutrino and red means tau neutrino. Oscillations electron long.svg
Electron neutrino oscillations, long range. Here and in the following diagrams black means electron neutrino, blue means muon neutrino and red means tau neutrino.
Electron neutrino oscillations, short range Oscillations electron short.svg
Electron neutrino oscillations, short range
Muon neutrino oscillations, long range Oscillations muon long.svg
Muon neutrino oscillations, long range
Muon neutrino oscillations, short range Oscillations muon short.svg
Muon neutrino oscillations, short range
Tau neutrino oscillations, long range Oscillations tau long.svg
Tau neutrino oscillations, long range
Tau neutrino oscillations, short range Oscillations tau short.svg
Tau neutrino oscillations, short range

The illustrations were created using the following parameter values: [25]

Observed values of oscillation parameters

Solar neutrino experiments combined with KamLAND have measured the so-called solar parameters Δm2
sol
and sin2θsol . Atmospheric neutrino experiments such as Super-Kamiokande together with the K2K and MINOS long baseline accelerator neutrino experiment have determined the so-called atmospheric parameters Δm2
atm
and sin2θatm . The last mixing angle, θ13, has been measured by the experiments Daya Bay, Double Chooz and RENO as sin2(2 θ13″) .

For atmospheric neutrinos the relevant difference of masses is about Δm2 = 24.×10−4 eV2 and the typical energies are ≈1 GeV; for these values the oscillations become visible for neutrinos traveling several hundred kilometres, which would be those neutrinos that reach the detector traveling through the earth, from below the horizon.

The mixing parameter θ13 is measured using electron anti-neutrinos from nuclear reactors. The rate of anti-neutrino interactions is measured in detectors sited near the reactors to determine the flux prior to any significant oscillations and then it is measured in far detectors (placed kilometres from the reactors). The oscillation is observed as an apparent disappearance of electron anti-neutrinos in the far detectors (i.e. the interaction rate at the far site is lower than predicted from the observed rate at the near site).

From atmospheric and solar neutrino oscillation experiments, it is known that two mixing angles of the MNS matrix are large and the third is smaller. This is in sharp contrast to the CKM matrix in which all three angles are small and hierarchically decreasing. The CP-violating phase of the MNS matrix is as of April 2020 to lie somewhere between 2 and 178 degrees, from the T2K experiment. [28]

If the neutrino mass proves to be of Majorana type (making the neutrino its own antiparticle), it is then possible that the MNS matrix has more than one phase.

Since experiments observing neutrino oscillation measure the squared mass difference and not absolute mass, one might claim that the lightest neutrino mass is exactly zero, without contradicting observations. This is however regarded as unlikely by theorists.

Origins of neutrino mass

The question of how neutrino masses arise has not been answered conclusively. In the Standard Model of particle physics, fermions only have intrinsic mass because of interactions with the Higgs field (see Higgs boson ). These interactions require both left- and right-handed versions of the fermion (see chirality ). However, only left-handed neutrinos have been observed so far.

Neutrinos may have another source of mass through the Majorana mass term. This type of mass applies for electrically neutral particles since otherwise it would allow particles to turn into anti-particles, which would violate conservation of electric charge.

The smallest modification to the Standard Model, which only has left-handed neutrinos, is to allow these left-handed neutrinos to have Majorana masses. The problem with this is that the neutrino masses are surprisingly smaller than the rest of the known particles (at least 600,000 times smaller than the mass of an electron), which, while it does not invalidate the theory, is widely regarded as unsatisfactory as this construction offers no insight into the origin of the neutrino mass scale.

The next simplest addition would be to add into the Standard Model right-handed neutrinos that interact with the left-handed neutrinos and the Higgs field in an analogous way to the rest of the fermions. These new neutrinos would interact with the other fermions solely in this way and hence would not be directly observable, so are not phenomenologically excluded. The problem of the disparity of the mass scales remains.

Seesaw mechanism

The most popular conjectured solution currently is the seesaw mechanism, where right-handed neutrinos with very large Majorana masses are added. If the right-handed neutrinos are very heavy, they induce a very small mass for the left-handed neutrinos, which is proportional to the reciprocal of the heavy mass.

If it is assumed that the neutrinos interact with the Higgs field with approximately the same strengths as the charged fermions do, the heavy mass should be close to the GUT scale. Because the Standard Model has only one fundamental mass scale, [lower-alpha 2] all particle masses [lower-alpha 3] must arise in relation to this scale.

There are other varieties of seesaw [29] and there is currently great interest in the so-called low-scale seesaw schemes, such as the inverse seesaw mechanism. [30]

The addition of right-handed neutrinos has the effect of adding new mass scales, unrelated to the mass scale of the Standard Model, hence the observation of heavy right-handed neutrinos would reveal physics beyond the Standard Model. Right-handed neutrinos would help to explain the origin of matter through a mechanism known as leptogenesis.

Other sources

There are alternative ways to modify the standard model that are similar to the addition of heavy right-handed neutrinos (e.g., the addition of new scalars or fermions in triplet states) and other modifications that are less similar (e.g., neutrino masses from loop effects and/or from suppressed couplings). One example of the last type of models is provided by certain versions supersymmetric extensions of the standard model of fundamental interactions, where R parity is not a symmetry. There, the exchange of supersymmetric particles such as squarks and sleptons can break the lepton number and lead to neutrino masses. These interactions are normally excluded from theories as they come from a class of interactions that lead to unacceptably rapid proton decay if they are all included. These models have little predictive power and are not able to provide a cold dark matter candidate.

Oscillations in the early universe

During the early universe when particle concentrations and temperatures were high, neutrino oscillations could have behaved differently. [31] Depending on neutrino mixing-angle parameters and masses, a broad spectrum of behavior may arise including vacuum-like neutrino oscillations, smooth evolution, or self-maintained coherence. The physics for this system is non-trivial and involves neutrino oscillations in a dense neutrino gas.

See also

Notes

  1. More formally, the neutrinos are emitted in an entangled state with the other bodies in the decay or reaction, and the mixed state is properly described by a density matrix. However, for all practical situations, the other particles in the decay may be well localized in time and space (e.g. to within a nuclear distance), leaving their momentum with a large spread. When these partner states are projected out, the neutrino is left in a state that for all intents and purposes behaves as the simple superposition of mass states described here. For more information, see: Cohen, Andrew G.; Glashow, Sheldon L. & Ligeti, Zoltan (13 July 2009). "Disentangling neutrino oscillations". Physics Letters B. 678 (2): 191–196. arXiv: 0810.4602 . Bibcode:2009PhLB..678..191C. doi: 10.1016/j.physletb.2009.06.020 .
  2. The Standard Model fundamental mass scale can be taken as the scale of SU(2)L × U(1)Y breaking.
  3. The mass of the electron and the mass of the Z boson are examples of particle masses set by the Standard Model fundamental mass scale.

Related Research Articles

<span class="mw-page-title-main">Pauli matrices</span> Matrices important in quantum mechanics and the study of spin

In mathematical physics and mathematics, the Pauli matrices are a set of three 2 × 2 complex matrices that are Hermitian, involutory and unitary. Usually indicated by the Greek letter sigma, they are occasionally denoted by tau when used in connection with isospin symmetries.

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.

<span class="mw-page-title-main">Gyrocompass</span> Type of non-magnetic compass based on the rotation of the Earth

A gyrocompass is a type of non-magnetic compass which is based on a fast-spinning disc and the rotation of the Earth to find geographical direction automatically. A gyrocompass makes use of one of the seven fundamental ways to determine the heading of a vehicle. A gyroscope is an essential component of a gyrocompass, but they are different devices; a gyrocompass is built to use the effect of gyroscopic precession, which is a distinctive aspect of the general gyroscopic effect. Gyrocompasses, such as the fibre optic gyrocompass are widely used to provide a heading for navigation on ships. This is because they have two significant advantages over magnetic compasses:

In mechanics and geometry, the 3D rotation group, often denoted SO(3), is the group of all rotations about the origin of three-dimensional Euclidean space under the operation of composition.

The Flipped SU(5) model is a grand unified theory (GUT) first contemplated by Stephen Barr in 1982, and by Dimitri Nanopoulos and others in 1984. Ignatios Antoniadis, John Ellis, John Hagelin, and Dimitri Nanopoulos developed the supersymmetric flipped SU(5), derived from the deeper-level superstring.

In quantum mechanics, the particle in a one-dimensional lattice is a problem that occurs in the model of a periodic crystal lattice. The potential is caused by ions in the periodic structure of the crystal creating an electromagnetic field so electrons are subject to a regular potential inside the lattice. It is a generalization of the free electron model, which assumes zero potential inside the lattice.

In physics, the S-matrix or scattering matrix relates the initial state and the final state of a physical system undergoing a scattering process. It is used in quantum mechanics, scattering theory and quantum field theory (QFT).

<span class="mw-page-title-main">Fermi's interaction</span> Mechanism of beta decay proposed in 1933

In particle physics, Fermi's interaction is an explanation of the beta decay, proposed by Enrico Fermi in 1933. The theory posits four fermions directly interacting with one another. This interaction explains beta decay of a neutron by direct coupling of a neutron with an electron, a neutrino and a proton.

In the Standard Model of particle physics, the Cabibbo–Kobayashi–Maskawa matrix, CKM matrix, quark mixing matrix, or KM matrix is a unitary matrix which contains information on the strength of the flavour-changing weak interaction. Technically, it specifies the mismatch of quantum states of quarks when they propagate freely and when they take part in the weak interactions. It is important in the understanding of CP violation. This matrix was introduced for three generations of quarks by Makoto Kobayashi and Toshihide Maskawa, adding one generation to the matrix previously introduced by Nicola Cabibbo. This matrix is also an extension of the GIM mechanism, which only includes two of the three current families of quarks.

In rotordynamics, the rigid rotor is a mechanical model of rotating systems. An arbitrary rigid rotor is a 3-dimensional rigid object, such as a top. To orient such an object in space requires three angles, known as Euler angles. A special rigid rotor is the linear rotor requiring only two angles to describe, for example of a diatomic molecule. More general molecules are 3-dimensional, such as water, ammonia, or methane.

In the theory of grand unification of particle physics, and, in particular, in theories of neutrino masses and neutrino oscillation, the seesaw mechanism is a generic model used to understand the relative sizes of observed neutrino masses, of the order of eV, compared to those of quarks and charged leptons, which are millions of times heavier. The name of the seesaw mechanism was given by Tsutomu Yanagida in a Tokyo conference in 1981.

<span class="mw-page-title-main">Two-state quantum system</span> Simple quantum mechanical system

In quantum mechanics, a two-state system is a quantum system that can exist in any quantum superposition of two independent quantum states. The Hilbert space describing such a system is two-dimensional. Therefore, a complete basis spanning the space will consist of two independent states. Any two-state system can also be seen as a qubit.

<span class="mw-page-title-main">Mathematical formulation of the Standard Model</span> Mathematics of a particle physics model

This article describes the mathematics of the Standard Model of particle physics, a gauge quantum field theory containing the internal symmetries of the unitary product group SU(3) × SU(2) × U(1). The theory is commonly viewed as describing the fundamental set of particles – the leptons, quarks, gauge bosons and the Higgs boson.

In particle physics, neutral particle oscillation is the transmutation of a particle with zero electric charge into another neutral particle due to a change of a non-zero internal quantum number, via an interaction that does not conserve that quantum number. Neutral particle oscillations were first investigated in 1954 by Murray Gell-mann and Abraham Pais.

The Mikheyev–Smirnov–Wolfenstein effect is a particle physics process which modifies neutrino oscillations in matter of varying density. The MSW effect is broadly analogous to the differential retardation of sound waves in density-variable media, however it also involves the propagation dynamics of three separate quantum fields which experience distortion.

<span class="mw-page-title-main">Covariant formulation of classical electromagnetism</span> Ways of writing certain laws of physics

The covariant formulation of classical electromagnetism refers to ways of writing the laws of classical electromagnetism in a form that is manifestly invariant under Lorentz transformations, in the formalism of special relativity using rectilinear inertial coordinate systems. These expressions both make it simple to prove that the laws of classical electromagnetism take the same form in any inertial coordinate system, and also provide a way to translate the fields and forces from one frame to another. However, this is not as general as Maxwell's equations in curved spacetime or non-rectilinear coordinate systems.

In particle physics, the Pontecorvo–Maki–Nakagawa–Sakata matrix, Maki–Nakagawa–Sakata matrix, lepton mixing matrix, or neutrino mixing matrix is a unitary mixing matrix which contains information on the mismatch of quantum states of neutrinos when they propagate freely and when they take part in weak interactions. It is a model of neutrino oscillation. This matrix was introduced in 1962 by Ziro Maki, Masami Nakagawa, and Shoichi Sakata, to explain the neutrino oscillations predicted by Bruno Pontecorvo.

Tribimaximal mixing is a specific postulated form for the Pontecorvo–Maki–Nakagawa–Sakata (PMNS) lepton mixing matrix U. Tribimaximal mixing is defined by a particular choice of the matrix of moduli-squared of the elements of the PMNS matrix as follows:

Trimaximal mixing refers to the highly symmetric, maximally CP-violating, fermion mixing configuration, characterised by a unitary matrix having all its elements equal in modulus (, ) as may be written, e.g.:

Lorentz-violating neutrino oscillation refers to the quantum phenomenon of neutrino oscillations described in a framework that allows the breakdown of Lorentz invariance. Today, neutrino oscillation or change of one type of neutrino into another is an experimentally verified fact; however, the details of the underlying theory responsible for these processes remain an open issue and an active field of study. The conventional model of neutrino oscillations assumes that neutrinos are massive, which provides a successful description of a wide variety of experiments; however, there are a few oscillation signals that cannot be accommodated within this model, which motivates the study of other descriptions. In a theory with Lorentz violation, neutrinos can oscillate with and without masses and many other novel effects described below appear. The generalization of the theory by incorporating Lorentz violation has shown to provide alternative scenarios to explain all the established experimental data through the construction of global models.

References

  1. 1 2 3 4 5 Barger, Vernon; Marfatia, Danny; Whisnant, Kerry Lewis (2012). The Physics of Neutrinos. Princeton University Press. ISBN   978-0-691-12853-5.
  2. 1 2 "Mesonium and anti-mesonium". Zh. Eksp. Teor. Fiz. 33 (2): 549–551. February 1957. reproduced and translated in B. Pontecorvo (February 1957). "Mesonium and Antimesonium". Sov. Phys. JETP . 6 (2): 429–431. Bibcode:1958JETP....6..429P.
  3. 1 2 B. Pontecorvo (May 1968). "Neutrino Experiments and the Problem of Conservation of Leptonic Charge". Zh. Eksp. Teor. Fiz. 53: 1717–1725. Bibcode:1968JETP...26..984P. reproduced and translated in B. Pontecorvo (May 1968). "Neutrino Experiments and the Problem of Conservation of Leptonic Charge". Sov. Phys. JETP . 26: 984–988. Bibcode:1968JETP...26..984P.
  4. V. De Sabbata; M. Gasperini (1981). "Neutrino oscillations in the presence of torsion". Il Nuovo Cimento A (1971-1996). 65: 479–500. doi:10.1007/BF02902051.
  5. Subhasish Chakrabarty; Amitabha Lahiri (2019). "Geometrical contribution to neutrino mass matrix". The European Physical Journal C. 79: 697. arXiv: 1904.06036 . doi:10.1140/epjc/s10052-019-7209-2.
  6. Webb, Jonathan (6 October 2015). "Neutrino 'flip' wins physics Nobel Prize". BBC News . Retrieved 6 October 2015.
  7. M. C. Gonzalez-Garcia & Michele Maltoni (April 2008). "Phenomenology with Massive Neutrinos". Physics Reports . 460 (1–3): 1–129. arXiv: 0704.1800 . Bibcode:2008PhR...460....1G. CiteSeerX   10.1.1.312.3412 . doi:10.1016/j.physrep.2007.12.004. S2CID   119651816.
  8. Davis, Raymond; Harmer, Don S.; Hoffman, Kenneth C. (1968). "Search for Neutrinos from the Sun". Physical Review Letters. 20 (21): 1205–1209. Bibcode:1968PhRvL..20.1205D. doi:10.1103/PhysRevLett.20.1205.
  9. Ahmad, Q.R.; Allen, R.C.; Andersen, T.C.; Anglin, J.D.; Bühler, G.; Barton, J.C.; et al. (SNO Collaboration) (25 July 2001). "Measurement of the rate of νe + d → p + p + e interactions produced by 8B Solar neutrinos at the Sudbury Neutrino Observatory". Physical Review Letters. 87 (7): 071301. arXiv: nucl-ex/0106015 . Bibcode:2001PhRvL..87g1301A. doi:10.1103/physrevlett.87.071301. ISSN   0031-9007. PMID   11497878.
  10. Fukuda, Y.; et al. (Super-Kamiokande Collaboration) (24 August 1998). "Evidence for Oscillation of Atmospheric Neutrinos". Physical Review Letters. 81 (8): 1562–1567. arXiv: hep-ex/9807003 . Bibcode:1998PhRvL..81.1562F. doi:10.1103/PhysRevLett.81.1562. S2CID   7102535.
  11. L, Mikaelyan and; V, Sinev (2000). "Neutrino oscillations at reactors: What is next?". Physics of Atomic Nuclei. 63 (6): 1002. arXiv: hep-ex/9908047 . Bibcode:2000PAN....63.1002M. doi:10.1134/1.855739. S2CID   15221390.
  12. Y, Abe; et al. (Double Chooz collaboration) (28 March 2012). "Indication for the disappearance of reactor electron antineutrinos in the Double Chooz experiment". Physical Review Letters. 108 (19): 131801. arXiv: 1112.6353 . Bibcode:2012PhRvL.108m1801A. doi:10.1103/PhysRevLett.108.131801. PMID   22540693. S2CID   19008791.
  13. An, F.P.; Bai, J.Z.; Balantekin, A. B.; Band, H.R.; Beavis, D.; Beriguete, W.; et al. (Daya Bay Collaboration) (23 April 2012). "Observation of Electron-Antineutrino Disappearance at Daya Bay". Physical Review Letters. 108 (17): 171803. arXiv: 1203.1669 . Bibcode:2012PhRvL.108q1803A. doi:10.1103/physrevlett.108.171803. ISSN   0031-9007. PMID   22680853. S2CID   16580300.
  14. Kim, Soo-Bong; et al. (RENO collaboration) (11 May 2012). "Observation of Reactor Electron Antineutrino Disappearance in the RENO Experiment". Physical Review Letters. 108 (19): 191802. arXiv: 1204.0626 . Bibcode:2012PhRvL.108s1802A. doi:10.1103/PhysRevLett.108.191802. PMID   23003027. S2CID   33056442.
  15. Agafonova, N.; et al. (OPERA Collaboration) (26 July 2010). "Observation of a first ντ candidate event in the OPERA experiment in the CNGS beam". Physics Letters B. 691 (3): 138–145. arXiv: 1006.1623 . Bibcode:2010PhLB..691..138A. doi:10.1016/j.physletb.2010.06.022. S2CID   119256958.
  16. Abe, K.; et al. (T2K Collaboration) (August 2013). "Evidence of electron neutrino appearance in a muon neutrino beam". Physical Review D. 88 (3): 032002. arXiv: 1304.0841 . Bibcode:2013PhRvD..88c2002A. doi:10.1103/PhysRevD.88.032002. ISSN   1550-7998. S2CID   53322828.
  17. Aartsen, M.G.; Ackermann, M.; Adams, J.; Aguilar, J.A.; Ahlers, M.; Ahrens, M.; al Samarai, I.; Altmann, D.; Andeen, K.; Anderson, T.; et al. (IceCube Collaboration) (12 April 2018). "Search for nonstandard neutrino interactions with IceCube DeepCore". Physical Review D. 97 (7): 072009. Bibcode:2018PhRvD..97g2009A. doi: 10.1103/PhysRevD.97.072009 . hdl: 2440/112089 .
  18. Akhmedov, Evgeny Kh. (26 September 2007). "Do charged leptons oscillate?". Journal of High Energy Physics. 2007 (9): 116. arXiv: 0706.1216 . Bibcode:2007JHEP...09..116A. doi:10.1088/1126-6708/2007/09/116. S2CID   13895776.
  19. Waltham, Chris (June 2004). "Teaching neutrino oscillations". American Journal of Physics. 72 (6): 742–752. arXiv: physics/0303116 . Bibcode:2004AmJPh..72..742W. doi:10.1119/1.1646132. S2CID   14205602.
  20. Z. Maki; M. Nakagawa; S. Sakata (November 1962). "Remarks on the Unified Model of Elementary Particles". Progress of Theoretical Physics . 28 (5): 870. Bibcode:1962PThPh..28..870M. doi: 10.1143/PTP.28.870 .
  21. Raymond Davis Jr.; Don S. Harmer; Kenneth C. Hoffman (May 1968). "Search for neutrinos from the Sun". Physical Review Letters . 20 (21): 1205–1209. Bibcode:1968PhRvL..20.1205D. doi:10.1103/PhysRevLett.20.1205.
  22. Gribov, V.; Pontecorvo, B. (20 January 1969). "Neutrino astronomy and lepton charge". Physics Letters B . 28 (7): 493–496. Bibcode:1969PhLB...28..493G. doi:10.1016/0370-2693(69)90525-5.
  23. Schechter, Joseph; Valle, José W. F. (1 November 1980). "Neutrino masses in SU(2) U(1) theories". Physical Review D . 22 (9): 2227–2235. Bibcode:1980PhRvD..22.2227S. doi:10.1103/PhysRevD.22.2227.
  24. Eidelman, S.; Hayes; Olive; Aguilar-Benitez; Amsler; Asner; et al. (Particle Data Group) (15 July 2004). "Chapter 15: Neutrino mass, mixing, and flavor change" (PDF). Physics Letters B . Review of Particle Physics. 592 (1–4): 1–1109. arXiv: astro-ph/0406663 . Bibcode:2004PhLB..592....1P. doi:10.1016/j.physletb.2004.06.001. S2CID   118588567. Revised September 2005
  25. 1 2 3 4 5 6 7 8 Meszéna, Balázs. "Neutrino Oscillations". Wolfram Demonstrations Project. Retrieved 8 October 2015. These images were created with Mathematica. The demonstration allows exploration of the parameters.
  26. 1 2 3 4 Olive, K.A.; et al. (Particle Data Group) (2014). "2014 Review of Particle Physics".{{cite journal}}: Cite journal requires |journal= (help)
  27. Nakamura, K.; et al. (Particle Data Group) (2010). "Review of Particle Physics". Journal of Physics G . 37 (7A): 1–708. Bibcode:2010JPhG...37g5021N. doi: 10.1088/0954-3899/37/7a/075021 . hdl: 10481/34593 . PMID   10020536.
  28. Abe, K.; Akutsu, R.; Ali, A.; Alt, C.; Andreopoulos, C.; Anthony, L.; et al. (15 April 2020). "Constraint on the matter–antimatter symmetry-violating phase in neutrino oscillations". Nature. 580 (7803): 339–344. arXiv: 1910.03887 . Bibcode:2020Natur.580..339T. doi:10.1038/s41586-020-2177-0. PMID   32296192. S2CID   203951445.
  29. Valle, J.W.F. (2006). "Neutrino physics overview". Journal of Physics . Conference Series. 53 (1): 473–505. arXiv: hep-ph/0608101 . Bibcode:2006JPhCS..53..473V. doi:10.1088/1742-6596/53/1/031. S2CID   2094005.
  30. Mohapatra, R.N. & Valle, J.W.F. (1986). "Neutrino mass and baryon number nonconservation in superstring models". Physical Review D . 34 (5): 1642–1645. Bibcode:1986PhRvD..34.1642M. doi:10.1103/PhysRevD.34.1642. hdl: 10550/47211 . PMID   9957332.
  31. Kostelecký, Alan; Samuel, Stuart (March 1994). "Nonlinear neutrino oscillations in the expanding universe" (PDF). Phys. Rev. D . 49 (4): 1740–1757. Bibcode:1994PhRvD..49.1740K. doi:10.1103/PhysRevD.49.1740. hdl: 2022/18663 . PMID   10017160.

Further reading