Majorana equation

Last updated

In physics, the Majorana equation is a relativistic wave equation. It is named after the Italian physicist Ettore Majorana, who proposed it in 1937 as a means of describing fermions that are their own antiparticle. [1] Particles corresponding to this equation are termed Majorana particles, although that term now has a more expansive meaning, referring to any (possibly non-relativistic) fermionic particle that is its own anti-particle (and is therefore electrically neutral).

Contents

There have been proposals that massive neutrinos are described by Majorana particles; there are various extensions to the Standard Model that enable this. The article on Majorana particles presents status for the experimental searches, including details about neutrinos. This article focuses primarily on the mathematical development of the theory, with attention to its discrete and continuous symmetries. The discrete symmetries are charge conjugation, parity transformation and time reversal; the continuous symmetry is Lorentz invariance.

Charge conjugation plays an outsize role, as it is the key symmetry that allows the Majorana particles to be described as electrically neutral. A particularly remarkable aspect is that electrical neutrality allows several global phases to be freely chosen, one each for the left and right chiral fields. This implies that, without explicit constraints on these phases, the Majorana fields are naturally CP violating. Another aspect of electric neutrality is that the left and right chiral fields can be given distinct masses. That is, electric charge is a Lorentz invariant, and also a constant of motion; whereas chirality is a Lorentz invariant, but is not a constant of motion for massive fields. Electrically neutral fields are thus less constrained than charged fields. Under charge conjugation, the two free global phases appear in the mass terms (as they are Lorentz invariant), and so the Majorana mass is described by a complex matrix, rather than a single number. In short, the discrete symmetries of the Majorana equation are considerably more complicated than those for the Dirac equation, where the electrical charge symmetry constrains and removes these freedoms.

Definition

The Majorana equation can be written in several distinct forms:

These three forms are equivalent, and can be derived from one-another. Each offers slightly different insight into the nature of the equation. The first form emphasises that purely real solutions can be found. The second form clarifies the role of charge conjugation. The third form provides the most direct contact with the representation theory of the Lorentz group.

Purely real four-component form

The conventional starting point is to state that "the Dirac equation can be written in Hermitian form", when the gamma matrices are taken in the Majorana representation. The Dirac equation is then written as [6]

with being purely real 4×4 symmetric matrices, and being purely imaginary skew-symmetric; as required to ensure that the operator (that part inside the parentheses) is Hermitian. In this case, purely real 4‑spinor solutions to the equation can be found; these are the Majorana spinors.

Charge-conjugate four-component form

The Majorana equation is

with the derivative operator written in Feynman slash notation to include the gamma matrices as well as a summation over the spinor components. The spinor is the charge conjugate of By construction, charge conjugates are necessarily given by

where denotes the transpose, is an arbitrary phase factor conventionally taken as and is a 4×4 matrix, the charge conjugation matrix. The matrix representation of depends on the choice of the representation of the gamma matrices. By convention, the conjugate spinor is written as

A number of algebraic identities follow from the charge conjugation matrix [lower-alpha 1] One states that in any representation of the gamma matrices, including the Dirac, Weyl, and Majorana representations, that and so one may write

where is the complex conjugate of The charge conjugation matrix also has the property that

in all representations (Dirac, chiral, Majorana). From this, and a fair bit of algebra, one may obtain the equivalent equation:

Proof

This form is not entirely obvious, and so deserves a proof. Starting with

Expand :

Multiply by use :

Charge conjugation transposes the gamma matrices:

Take the complex conjugate:

The matrix is Hermitian, in all three representations (Dirac, chiral, Majorana):

It is also an involution, taking the Hermitian conjugate:

Multiply by , note that and make use of :

The above is just the definition of the conjugate, so conclude that

A detailed discussion of the physical interpretation of matrix as charge conjugation can be found in the article on charge conjugation. In short, it is involved in mapping particles to their antiparticles, which includes, among other things, the reversal of the electric charge. Although is defined as "the charge conjugate" of the charge conjugation operator has not one but two eigenvalues. This allows a second spinor, the ELKO spinor to be defined. This is discussed in greater detail below.

Complex two-component form

The Majorana operator, is defined as

where

is a vector whose components are the 2×2 identity matrix for and (minus) the Pauli matrices for The is an arbitrary phase factor, typically taken to be one: The is a 2×2 matrix that can be interpreted as the symplectic form for the symplectic group which is a double covering of the Lorentz group. It is

which happens to be isomorphic to the imaginary unit "i" (i.e. and for ) with the matrix transpose being the analog of complex conjugation.

Finally, the is a short-hand reminder to take the complex conjugate. The Majorana equation for a left-handed complex-valued two-component spinor is then

or, equivalently,

with the complex conjugate of The subscript L is used throughout this section to denote a left-handed chiral spinor; under a parity transformation, this can be taken to a right-handed spinor, and so one also has a right-handed form of the equation. This applies to the four-component equation as well; further details are presented below.

Key ideas

Some of the properties of the Majorana equation, its solution and its Lagrangian formulation are summarized here.

Two-component Majorana equation

The Majorana equation can be written both in terms of a real four-component spinor, and as a complex two-component spinor. Both can be constructed from the Weyl equation, with the addition of a properly Lorentz-covariant mass term. [7] This section provides an explicit construction and articulation.

Weyl equation

The Weyl equation describes the time evolution of a massless complex-valued two-component spinor. It is conventionally written as [8] [9] [10]

Written out explicitly, it is

The Pauli four-vector is

that is, a vector whose components are the 2 × 2 identity matrix for μ = 0 and the Pauli matrices for μ = 1, 2, 3. Under the parity transformation one obtains a dual equation

where . These are two distinct forms of the Weyl equation; their solutions are distinct as well. It can be shown that the solutions have left-handed and right-handed helicity, and thus chirality. It is conventional to label these two distinct forms explicitly, thus:

Lorentz invariance

The Weyl equation describes a massless particle; the Majorana equation adds a mass term. The mass must be introduced in a Lorentz invariant fashion. This is achieved by observing that the special linear group is isomorphic to the symplectic group Both of these groups are double covers of the Lorentz group The Lorentz invariance of the derivative term (from the Weyl equation) is conventionally worded in terms of the action of the group on spinors, whereas the Lorentz invariance of the mass term requires invocation of the defining relation for the symplectic group.

The double-covering of the Lorentz group is given by

where and and is the Hermitian transpose. This is used to relate the transformation properties of the differentials under a Lorentz transformation to the transformation properties of the spinors.

The symplectic group is defined as the set of all complex 2×2 matrices that satisfy

where

is a skew-symmetric matrix. It is used to define a symplectic bilinear form on Writing a pair of arbitrary two-vectors as

the symplectic product is

where is the transpose of This form is invariant under Lorentz transformations, in that

The skew matrix takes the Pauli matrices to minus their transpose:

for The skew matrix can be interpreted as the product of a parity transformation and a transposition acting on two-spinors. However, as will be emphasized in a later section, it can also be interpreted as one of the components of the charge conjugation operator, the other component being complex conjugation. Applying it to the Lorentz transformation yields

These two variants describe the covariance properties of the differentials acting on the left and right spinors, respectively.

Differentials

Under the Lorentz transformation the differential term transforms as

provided that the right-handed field transforms as

Similarly, the left-handed differential transforms as

provided that the left-handed spinor transforms as

Proof

These transformation properties are not particularly "obvious", and so deserve a careful derivation. Begin with the form

for some unknown to be determined. The Lorentz transform, in coordinates, is

or, equivalently,

This leads to

In order to make use of the Weyl map

a few indexes must be raised and lowered. This is easier said than done, as it invokes the identity

where is the flat-space Minkowski metric. The above identity is often used to define the elements One takes the transpose:

to write

One thus regains the original form if that is, Performing the same manipulations for the left-handed equation, one concludes that

with [lower-alpha 2]

Mass term

The complex conjugate of the right handed spinor field transforms as

The defining relationship for can be rewritten as From this, one concludes that the skew-complex field transforms as

This is fully compatible with the covariance property of the differential. Taking to be an arbitrary complex phase factor, the linear combination

transforms in a covariant fashion. Setting this to zero gives the complex two-component Majorana equation for the right-handed field. Similarly, the left-chiral Majorana equation (including an arbitrary phase factor ) is

The left and right chiral versions are related by a parity transformation. As shown below, these square to the Klein–Gordon operator only if The skew complex conjugate can be recognized as the charge conjugate form of this is articulated in greater detail below. Thus, the Majorana equation can be read as an equation that connects a spinor to its charge-conjugate form.

Left and right Majorana operators

Define a pair of operators, the Majorana operators,

where is a short-hand reminder to take the complex conjugate. Under Lorentz transformations, these transform as

whereas the Weyl spinors transform as

just as above. Thus, the matched combinations of these are Lorentz covariant, and one may take

as a pair of complex 2-spinor Majorana equations.

The products and are both Lorentz covariant. The product is explicitly

Verifying this requires keeping in mind that and that The RHS reduces to the Klein–Gordon operator provided that , that is, These two Majorana operators are thus "square roots" of the Klein–Gordon operator.

Four-component Majorana equation

The real four-component version of the Majorana equation can be constructed from the complex two-component equation as follows. Given the complex field satisfying as above, define

Using the algebraic machinery given above, it is not hard to show that

Defining a conjugate operator

The four-component Majorana equation is then

Writing this out in detail, one has

Multiplying on the left by

brings the above into a matrix form wherein the gamma matrices in the chiral representation can be recognized. This is

That is,

Applying this to the 4-spinor

and recalling that one finds that the spinor is an eigenstate of the mass term,

and so, for this particular spinor, the four-component Majorana equation reduces to the Dirac equation

The skew matrix can be identified with the charge conjugation operator (in the Weyl basis). Explicitly, this is

Given an arbitrary four-component spinor its charge conjugate is

with an ordinary 4×4 matrix, having a form explicitly given in the article on gamma matrices. In conclusion, the 4-component Majorana equation can be written as

Charge conjugation and parity

The charge conjugation operator appears directly in the 4-component version of the Majorana equation. When the spinor field is a charge conjugate of itself, that is, when then the Majorana equation reduces to the Dirac equation, and any solution can be interpreted as describing an electrically neutral field. However, the charge conjugation operator has not one, but two distinct eigenstates, one of which is the ELKO spinor; it does not solve the Majorana equation, but rather, a sign-flipped version of it.

The charge conjugation operator for a four-component spinor is defined as

A general discussion of the physical interpretation of this operator in terms of electrical charge is given in the article on charge conjugation. Additional discussions are provided by Bjorken & Drell [11] or Itzykson & Zuber. [lower-alpha 3] In more abstract terms, it is the spinorial equivalent of complex conjugation of the coupling of the electromagnetic field. This can be seen as follows. If one has a single, real scalar field, it cannot couple to electromagnetism; however, a pair of real scalar fields, arranged as a complex number, can. For scalar fields, charge conjugation is the same as complex conjugation. The discrete symmetries of the gauge theory follows from the "trivial" observation that

is an automorphism of For spinorial fields, the situation is more confusing. Roughly speaking, however, one can say that the Majorana field is electrically neutral, and that taking an appropriate combination of two Majorana fields can be interpreted as a single electrically charged Dirac field. The charge conjugation operator given above corresponds to the automorphism of

In the above, is a 4×4 matrix, given in the article on the gamma matrices. Its explicit form is representation-dependent. The operator cannot be written as a 4×4 matrix, as it is taking the complex conjugate of , and complex conjugation cannot be achieved with a complex 4×4 matrix. It can be written as a real 8×8 matrix, presuming one also writes as a purely real 8-component spinor. Letting stand for complex conjugation, so that one can then write, for four-component spinors,

It is not hard to show that and that It follows from the first identity that has two eigenvalues, which may be written as

The eigenvectors are readily found in the Weyl basis. From the above, in this basis, is explicitly

and thus

Both eigenvectors are clearly solutions to the Majorana equation. However, only the positive eigenvector is a solution to the Dirac equation:

The negative eigenvector "doesn't work", it has the incorrect sign on the Dirac mass term. It still solves the Klein–Gordon equation, however. The negative eigenvector is termed the ELKO spinor.

Proof

That both eigenstates solve the Klein–Gordon equation follows from the earlier identities for the two-component versions. Defining, as before,

As was previously shown

The four-component spinor requires the introduction of

which also obey

Therefore

The chiral representation requires an extra factor of :

and so one concludes that

That is, both eigenvectors of the charge conjugation operator solve the Klein–Gordon equation. The last identity can also be verified directly, by noting that and that

Parity

Under parity, the left-handed spinors transform to right-handed spinors. The two eigenvectors of the charge conjugation operator, again in the Weyl basis, are

As before, both solve the four-component Majorana equation, but only one also solves the Dirac equation. This can be shown by constructing the parity-dual four-component equation. This takes the form

where

Given the two-component spinor define its conjugate as It is not hard to show that and that therefore, if then also and therefore that

or equivalently

This works, because and so this reduces to the Dirac equation for

To conclude, and reiterate, the Majorana equation is

It has four inequivalent, linearly independent solutions, Of these, only two are also solutions to the Dirac equation: namely and

Solutions

Spin eigenstates

One convenient starting point for writing the solutions is to work in the rest frame way of the spinors. Writing the quantum Hamiltonian with the conventional sign convention leads to the Majorana equation taking the form

In the chiral (Weyl) basis, one has that

with the Pauli vector. The sign convention here is consistent with the article gamma matrices. Plugging in the positive charge conjugation eigenstate given above, one obtains an equation for the two-component spinor

and likewise

These two are in fact the same equation, which can be verified by noting that yields the complex conjugate of the Pauli matrices:

The plane wave solutions can be developed for the energy-momentum and are most easily stated in the rest frame. The spin-up rest-frame solution is

while the spin-down solution is

That these are being correctly interpreted can be seen by re-expressing them in the Dirac basis, as Dirac spinors. In this case, they take the form

and

These are the rest-frame spinors. They can be seen as a linear combination of both the positive and the negative-energy solutions to the Dirac equation. These are the only two solutions; the Majorana equation has only two linearly independent solutions, unlike the Dirac equation, which has four. The doubling of the degrees of freedom of the Dirac equation can be ascribed to the Dirac spinors carrying charge.

Momentum eigenstates

In a general momentum frame, the Majorana spinor can be written as

Electric charge

The appearance of both and in the Majorana equation means that the field  cannot be coupled to a charged electromagnetic field without violating charge conservation, since particles have the opposite charge to their own antiparticles. To satisfy this restriction, must be taken to be electrically neutral. This can be articulated in greater detail.

The Dirac equation can be written in a purely real form, when the gamma matrices are taken in the Majorana representation. The Dirac equation can then be written as [lower-alpha 4]

with being purely real symmetric matrices, and being purely imaginary skew-symmetric. In this case, purely real solutions to the equation can be found; these are the Majorana spinors. Under the action of Lorentz transformations, these transform under the (purely real) spin group This stands in contrast to the Dirac spinors, which are only covariant under the action of the complexified spin group The interpretation is that complexified spin group encodes the electromagnetic potential, the real spin group does not.

This can also be stated in a different way: the Dirac equation, and the Dirac spinors contain a sufficient amount of gauge freedom to naturally encode electromagnetic interactions. This can be seen by noting that the electromagnetic potential can very simply be added to the Dirac equation without requiring any additional modifications or extensions to either the equation or the spinor. The location of this extra degree of freedom is pin-pointed by the charge conjugation operator, and the imposition of the Majorana constraint removes this extra degree of freedom. Once removed, there cannot be any coupling to the electromagnetic potential, ergo, the Majorana spinor is necessarily electrically neutral. An electromagnetic coupling can only be obtained by adding back in a complex-number-valued phase factor, and coupling this phase factor to the electromagnetic potential.

The above can be further sharpened by examining the situation in spatial dimensions. In this case, the complexified spin group has a double covering by with the circle. The implication is that encodes the generalized Lorentz transformations (of course), while the circle can be identified with the action of the gauge group on electric charges. That is, the gauge-group action of the complexified spin group on a Dirac spinor can be split into a purely-real Lorentzian part, and an electromagnetic part. This can be further elaborated on non-flat (non-Minkowski-flat) spin manifolds. In this case, the Dirac operator acts on the spinor bundle. Decomposed into distinct terms, it includes the usual covariant derivative The field can be seen to arise directly from the curvature of the complexified part of the spin bundle, in that the gauge transformations couple to the complexified part, and not the real-spinor part. That the field corresponds to the electromagnetic potential can be seen by noting that (for example) the square of the Dirac operator is the Laplacian plus the scalar curvature (of the underlying manifold that the spinor field sits on) plus the (electromagnetic) field strength For the Majorana case, one has only the Lorentz transformations acting on the Majorana spinor; the complexification plays no role. A detailed treatment of these topics can be found in Jost [12] while the case is articulated in Bleeker. [13] Unfortunately, neither text explicitly articulates the Majorana spinor in direct form.

Field quanta

The quanta of the Majorana equation allow for two classes of particles, a neutral particle and its neutral antiparticle. The frequently applied supplemental condition corresponds to the Majorana spinor.

Majorana particle

Particles corresponding to Majorana spinors are known as Majorana particles, due to the above self-conjugacy constraint. All the fermions included in the Standard Model have been excluded as Majorana fermions (since they have non-zero electric charge they cannot be antiparticles of themselves) with the exception of the neutrino (which is neutral).

Theoretically, the neutrino is a possible exception to this pattern. If so, neutrinoless double-beta decay, as well as a range of lepton-number violating meson and charged lepton decays, are possible. A number of experiments probing whether the neutrino is a Majorana particle are currently underway. [14]

Notes

  1. Caution: Not all authors use the same conventions for charge conjugation, and so there is plenty of room for subtle sign errors. This article, and the article on charge conjugation, use the conventions of Itzykson & Zuber, (Quantum Field Theory, see Chapter 2 and Appendix A). These differ very slightly from Bjorken & Drell Relativistic Quantum Mechanics and so allowances must be made if comparing the two.
  2. The results presented here are identical to those of Aste, op. cit., equations 52 and 57, although the derivation performed here is completely different. The double-covering used here is also identical to Aste equations 48, and to the current version (December 2020) of the article on Lorentz group.
  3. Itzykson and Zuber, op. cit.(Chapter 2-4)
  4. Itzykson & Zuber, (See Chapter 2-1-2, page 49)

Related Research Articles

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 physics, charge conjugation is a transformation that switches all particles with their corresponding antiparticles, thus changing the sign of all charges: not only electric charge but also the charges relevant to other forces. The term C-symmetry is an abbreviation of the phrase "charge conjugation symmetry", and is used in discussions of the symmetry of physical laws under charge-conjugation. Other important discrete symmetries are P-symmetry (parity) and T-symmetry.

In quantum field theory, the Dirac spinor is the spinor that describes all known fundamental particles that are fermions, with the possible exception of neutrinos. It appears in the plane-wave solution to the Dirac equation, and is a certain combination of two Weyl spinors, specifically, a bispinor that transforms "spinorially" under the action of the Lorentz group.

In theoretical physics, the Rarita–Schwinger equation is the relativistic field equation of spin-3/2 fermions in a four-dimensional flat spacetime. It is similar to the Dirac equation for spin-1/2 fermions. This equation was first introduced by William Rarita and Julian Schwinger in 1941.

In differential geometry, the four-gradient is the four-vector analogue of the gradient from vector calculus.

<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.

In mathematical physics, the gamma matrices, also called the Dirac matrices, are a set of conventional matrices with specific anticommutation relations that ensure they generate a matrix representation of the Clifford algebra It is also possible to define higher-dimensional gamma matrices. When interpreted as the matrices of the action of a set of orthogonal basis vectors for contravariant vectors in Minkowski space, the column vectors on which the matrices act become a space of spinors, on which the Clifford algebra of spacetime acts. This in turn makes it possible to represent infinitesimal spatial rotations and Lorentz boosts. Spinors facilitate spacetime computations in general, and in particular are fundamental to the Dirac equation for relativistic spin particles. Gamma matrices were introduced by Dirac in 1928.

In differential geometry and mathematical physics, a spin connection is a connection on a spinor bundle. It is induced, in a canonical manner, from the affine connection. It can also be regarded as the gauge field generated by local Lorentz transformations. In some canonical formulations of general relativity, a spin connection is defined on spatial slices and can also be regarded as the gauge field generated by local rotations.

In mathematical physics, spacetime algebra (STA) is a name for the Clifford algebra Cl1,3(R), or equivalently the geometric algebra G(M4). According to David Hestenes, spacetime algebra can be particularly closely associated with the geometry of special relativity and relativistic spacetime.

In mathematical physics, the Dirac algebra is the Clifford algebra . This was introduced by the mathematical physicist P. A. M. Dirac in 1928 in developing the Dirac equation for spin-1/2 particles with a matrix representation of the gamma matrices, which represent the generators of the algebra.

In mathematics, the spectral theory of ordinary differential equations is the part of spectral theory concerned with the determination of the spectrum and eigenfunction expansion associated with a linear ordinary differential equation. In his dissertation, Hermann Weyl generalized the classical Sturm–Liouville theory on a finite closed interval to second order differential operators with singularities at the endpoints of the interval, possibly semi-infinite or infinite. Unlike the classical case, the spectrum may no longer consist of just a countable set of eigenvalues, but may also contain a continuous part. In this case the eigenfunction expansion involves an integral over the continuous part with respect to a spectral measure, given by the Titchmarsh–Kodaira formula. The theory was put in its final simplified form for singular differential equations of even degree by Kodaira and others, using von Neumann's spectral theorem. It has had important applications in quantum mechanics, operator theory and harmonic analysis on semisimple Lie groups.

In physics, and specifically in quantum field theory, a bispinor is a mathematical construction that is used to describe some of the fundamental particles of nature, including quarks and electrons. It is a specific embodiment of a spinor, specifically constructed so that it is consistent with the requirements of special relativity. Bispinors transform in a certain "spinorial" fashion under the action of the Lorentz group, which describes the symmetries of Minkowski spacetime. They occur in the relativistic spin-1/2 wave function solutions to the Dirac equation.

In quantum field theory, and especially in quantum electrodynamics, the interacting theory leads to infinite quantities that have to be absorbed in a renormalization procedure, in order to be able to predict measurable quantities. The renormalization scheme can depend on the type of particles that are being considered. For particles that can travel asymptotically large distances, or for low energy processes, the on-shell scheme, also known as the physical scheme, is appropriate. If these conditions are not fulfilled, one can turn to other schemes, like the minimal subtraction scheme.

In mathematical physics, the Belinfante–Rosenfeld tensor is a modification of the energy–momentum tensor that is constructed from the canonical energy–momentum tensor and the spin current so as to be symmetric yet still conserved.

<span class="mw-page-title-main">Weyl equation</span> Relativistic wave equation describing massless fermions

In physics, particularly in quantum field theory, the Weyl equation is a relativistic wave equation for describing massless spin-1/2 particles called Weyl fermions. The equation is named after Hermann Weyl. The Weyl fermions are one of the three possible types of elementary fermions, the other two being the Dirac and the Majorana fermions.

<span class="mw-page-title-main">Bargmann–Wigner equations</span> Wave equation for arbitrary spin particles

In relativistic quantum mechanics and quantum field theory, the Bargmann–Wigner equations describe free particles with non-zero mass and arbitrary spin j, an integer for bosons or half-integer for fermions. The solutions to the equations are wavefunctions, mathematically in the form of multi-component spinor fields.

<span class="mw-page-title-main">Matrix representation of Maxwell's equations</span>

In electromagnetism, a branch of fundamental physics, the matrix representations of the Maxwell's equations are a formulation of Maxwell's equations using matrices, complex numbers, and vector calculus. These representations are for a homogeneous medium, an approximation in an inhomogeneous medium. A matrix representation for an inhomogeneous medium was presented using a pair of matrix equations. A single equation using 4 × 4 matrices is necessary and sufficient for any homogeneous medium. For an inhomogeneous medium it necessarily requires 8 × 8 matrices.

<span class="mw-page-title-main">Dirac equation in curved spacetime</span> Generalization of the Dirac equation

In mathematical physics, the Dirac equation in curved spacetime is a generalization of the Dirac equation from flat spacetime to curved spacetime, a general Lorentzian manifold.

<span class="mw-page-title-main">Loop representation in gauge theories and quantum gravity</span> Description of gauge theories using loop operators

Attempts have been made to describe gauge theories in terms of extended objects such as Wilson loops and holonomies. The loop representation is a quantum hamiltonian representation of gauge theories in terms of loops. The aim of the loop representation in the context of Yang–Mills theories is to avoid the redundancy introduced by Gauss gauge symmetries allowing to work directly in the space of physical states. The idea is well known in the context of lattice Yang–Mills theory. Attempts to explore the continuous loop representation was made by Gambini and Trias for canonical Yang–Mills theory, however there were difficulties as they represented singular objects. As we shall see the loop formalism goes far beyond a simple gauge invariant description, in fact it is the natural geometrical framework to treat gauge theories and quantum gravity in terms of their fundamental physical excitations.

In mathematical physics, the Gordon decomposition of the Dirac current is a splitting of the charge or particle-number current into a part that arises from the motion of the center of mass of the particles and a part that arises from gradients of the spin density. It makes explicit use of the Dirac equation and so it applies only to "on-shell" solutions of the Dirac equation.

References

  1. Ettore Majorana, "Teoria Simmetrica Dell' Elettrone E Del Positrone," Nuovo Cimento14 (1937) pp.171–184. PDF Original Italian version
  2. Aste, Andreas (2010). "A direct road to Majorana fields". Symmetry. 2010 (2): 1776–1809. arXiv: 0806.1690 . Bibcode:2010Symm....2.1776A. doi: 10.3390/sym2041776 .
  3. Pal, Palash B. (2011). "Dirac, Majorana, and Weyl fermions". American Journal of Physics. 79 (5): 485–498. arXiv: 1006.1718 . Bibcode:2011AmJPh..79..485P. doi:10.1119/1.3549729. S2CID   118685467.
  4. Marsch, Eckart (2012). "On the Majorana equation: Relations between its complex two-component and real four-component eigenfunctions". ISRN Mathematical Physics. 2012: 1–17. arXiv: 1207.4685 . doi: 10.5402/2012/760239 . Article 760239.
  5. Marsch, Eckart (2013). "A new route to the Majorana equation". Symmetry. 5 (4): 271–286. Bibcode:2013Symm....5..271M. doi: 10.3390/sym5040271 .
  6. Itzykson, Claude; Zuber, Jean-Bernard (1980). Quantum Field Theory. MacGraw-Hill. §2‑1‑2, page 49.
  7. Andreas Aste, (2010) "A Direct Road to Majorana Fields", Symmetry2010(2) 1776-1809; doi:10.3390/sym2041776 ISSN 2073-8994.
  8. Quantum Mechanics, E. Abers, Pearson Ed., Addison Wesley, Prentice Hall Inc, 2004, ISBN   978-0-13-146100-0
  9. The Cambridge Handbook of Physics Formulas, G. Woan, Cambridge University Press, 2010, ISBN   978-0-521-57507-2.
  10. An Introduction to Quantum Field Theory, M.E. Peskin, D.V. Schroeder, Addison-Wesley, 1995, ISBN   0-201-50397-2
  11. James D. Bjorken, Sidney D. Drell, (1964) "Relativistic Quantum Mechanics", McGraw-Hill (See Chapter 5.2, pages 66-70)
  12. Jurgen Jost (2002) "Riemannian geometry and Geometric Analysis (3rd edition) Springer Universitext. (See chapter 1.8 for spin structures, and chapter 3.4 for the Dirac operator.)
  13. David Bleeker, (1981) "Gauge Theory and Variational Principles" Addison-Wesley (See Chapter 6 for the free Dirac field, and Chapter 7 for the interacting field).
  14. A. Franklin, Are There Really Neutrinos?: An Evidential History (Westview Press, 2004), p. 186

Additional reading