0% found this document useful (0 votes)
34 views18 pages

Research Article

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
34 views18 pages

Research Article

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 18

International Scholarly Research Network

ISRN Mathematical Physics


Volume 2012, Article ID 760239, 17 pages
doi:10.5402/2012/760239

Research Article
On the Majorana Equation: Relations
between Its Complex Two-Component and
Real Four-Component Eigenfunctions

Eckart Marsch
Institute for Experimental and Applied Physics, Christian Albrechts University at Kiel,
Leibnizstraße 11, 24118 Kiel, Germany

Correspondence should be addressed to Eckart Marsch, [email protected]

Received 2 July 2012; Accepted 24 July 2012

Academic Editors: K. Ammari, A. Herrera-Aguilar, D. Ida, and A. Sanyal

Copyright q 2012 Eckart Marsch. This is an open access article distributed under the Creative
Commons Attribution License, which permits unrestricted use, distribution, and reproduction in
any medium, provided the original work is properly cited.

We first derive without recourse to the Dirac equation the two-component Majorana equation with
a mass term by a direct linearization of the relativistic dispersion relation of a massive particle.
Thereby, we make only use of the complex conjugation operator and the Pauli spin matrices,
corresponding to the irreducible representation of the Lorentz group. Then we derive the complex
two-component eigenfunctions of the Majorana equation and the related quantum fields in a
concise way, by exploiting the so-called chirality conjugation operator that involves the spin-flip
operator. Subsequently, the four-component spinor solutions of the real Majorana equation are
derived, and their intrinsic relations with the spinors of the complex two-component version of
the Majorana equation are revealed and discussed extensively.

1. Introduction
In this paper we first present a new derivation 1 of the two-component Majorana equation
including a mass term by a direct linearization of the relativistic dispersion relation of a
massive particle, in a way similar to that used originally by Dirac 2. We subsequently
derive from the two-component complex the four-component real version of the Majorana
equation 3. We thus obtain some new results and look at the problem from a fresh theoretical
perspective, thus expanding the established theoretical framework as described in modern
monographs 4, 5 on this subject. One intention of this work is to obtain the Majorana
equation “completely on its own rather than as an afterthought when treating the Dirac
equation,” as it was phrased by Case 6, who first reformulated half a century ago the
Majorana theory of the neutrino. Another main objective is to address the key question of
whether Majorana particles are their own antiparticles, and they are according to common
2 ISRN Mathematical Physics

textbook wisdom. We will show that a real four-component or complex two-component


spinor wavefunction does not imply this conclusion necessarily, but following Mannheim
7 we will consistently define antiparticles as being associated with the negative root of the
dispersion relation and particles with the positive one.
The recent paper by Pal 8 titled “Dirac, Majorana, and Weyl fermions” discussed
the definitions of and connections between these three fields and their intrinsic symmetries.
We consider the present paper partly as a complement to his largely tutorial work, yet
here with emphasis on the two-component Majorana equation for a massive fermion, which
is treated without recourse to the Dirac equation 1. Thereby we only make use of the
complex conjugation operator and the Pauli spin matrices, corresponding to the irreducible
representation of the Lorentz group. The Lorentz-invariant complex conjugation 8 involves
the here introduced spin-flip operator. Its connection to chiral symmetry is also discussed.
Finally, we also show that Dirac’s equation in the real Majorana representation is a direct
consequence of the two-component complex Majorana equation as obtained in the following.
We derive in much detail the eigenfunctions of the Majorana equation and the related
quantum fields, yet other than in the paper by Case 6 in a direct way, whereby we will
exploit the spin-flip operator, or as we also may call it perhaps more adequately, the chirality
conjugation operator. Pal 8 gave in the introduction to his paper a good motivation, which
we fully agree with, for dealing again with this theoretical subject. Our main motivation is
to contribute to the ongoing discussion of whether massive neutrinos are Dirac or Majorana
fermions and to better understand the latter theoretically in terms of the two-component
theory and its connection to the four-component real version. In this respect we also
complement the old pedagogical review on Majorana masses by Mannheim 7.

2. New Derivation of the Majorana Equation and Its Symmetries


In this section a new derivation of the two-component Majorana equation is presented, by
linearizing the standard dispersion relation of a massive relativistic particle. Following Dirac
2 historically, one can derive his equation
 in a straightforward way from the relativistic
energy and momentum relation, E  m2  p2 , which in fact is constitutive for all field
theories of massive particles. Here the four-momentum of the particle is denoted as pμ 
E, p, and the units are such that the speed of light c  1, and the particle mass is m.
The contravariant four-momentum operator in space-time is denoted by P μ  P0 , P, and
its covariant form is given by the differential operator Pμ  P0 , −P  i∂μ  i∂/∂t, ∂/∂x
conventionally we use units of   1. The energy-momentum relation is usually written in
manifestly covariant form as mass-shell condition, pμ pμ  m2 . Dirac’s equation results from
its linearization, which is achieved by introducing four new operators usually represented
by matrices, namely, the matrix three-vector α and scalar β, yielding a Hamiltonian in the
linear algebraic form
H  βm  α · P. 2.1

As shown in any text book see, e.g., 9, 10, in terms of matrices one requires at least
four dimensions to represent the operators α and β. Furthermore, to satisfy the mass-shell
requirement they must algebraically obey the indicated by the symbol {, } anticommutator,

 
α, β  0, 2.2
ISRN Mathematical Physics 3

and also have the property that β2  1 and α · p2  p2 . This is readily ensured, for example
in the chiral representation, if we decompose the 4 × 4 matrix as follows, α  ασ, with α2  1.
Thus, α can concisely be expressed in terms of the spin operator σ and α. The 2 × 2 Pauli
matrices are known to have the required property σ·p2  p2 . Similarly, we can write β  βσ0 .
Finally, we quote explicitly the two real matrices α and β and a related one called γ  βα,
which are going to be used later and are defined as

     
1 0 0 1 0 −1
α , β , γ . 2.3
0 −1 1 0 1 0

The three Pauli matrices 11, to which we add here the 2 × 2 unit matrix denoted as σ0 , have
their standard form but are quoted again because of their importance to what follows. They
read

     
0 1 0 −i 1 0
σx  , σy  , σz  2.4
1 0 i 0 0 −1

and build the three-vector σ  σx , σy , σz . We may also introduce the two four-vector forms of
μ
the Pauli matrices. They are defined according to Jehle 12 by the four vectors σ±  σ0 , ±σ
μ
and obey σ±  σ∓μ . Making use of the above matrices, we can obtain Dirac’s equation in
standard chiral form 1, 10 and the classical Hamiltonian representation.
Returning to the basic equation 2.1, the question may then be asked whether there
are other than pure matrix representations of the operators α and β. In a recent paper 1
it was shown that the answer is yes. As this subject is not common knowledge, we shall
repeat here some of the key algebra, making use of the operator of complex conjugation
C, which transforms any complex number z into its complex conjugate z∗ . This appears
naturally in the symmetry operations of time inversion and charge conjugation of Dirac’s
equation 10. Therefore let us define an important operator, which is not of pure algebraic
nature but involves this complex-conjugation operator C. We call in partial nomenclature
analogy to β this operator τ and define it appropriately as τ  σy C, which differs from the
definition in Marsch’s paper 1 by an unimportant phase factor i. The importance of τ was
first recognized by Case 6. It is anti-Hermitian conjugate, τ †  −τ, and other than β obeys the
relation τ 2  −1. Obviously, the operation of τ on the spin vector σ leads to its inversion; that
is, when using the commutation rules of the spin operators, we can show that the operation
τστ −1  −σ yields a spin flip. Also remember from 2.4 that σ ∗  σ T , where the superscript
μ μ
T indicates the transposed matrix. As τ flips the spin, we have τσ±  σ∓ τ, and τi  iτ  0,
because of the action of C. Therefore, τ also anticommutes with the momentum four-vector
operator Pμ , and thus we have τP0  P0 τ  0 and τP  Pτ  0.
Making use of the previously mentioned properties of the spin flip operator τ  σy C,
we can now still use a two-component matrix representation given by the Pauli matrices 2.4.
Therefore, by help of τ we can go a decisive mathematical step beyond pure matrix algebra
and define the linear energy operator

Hτ mσ ·P 2.5


4 ISRN Mathematical Physics

in analogy to 2.1. The operators involved obey the anticommutation rule,

{σ, τ}  0, 2.6

and also have the property that σ · p2  p2 , but τ 2  −1. We stress again that for the space-
time operators {P, τ}  0 and {P0 , τ}  0. When putting P0  H in 2.5, squaring it and
multiplying out, the previous key features of τ must be exploited. Thus we retain, when
inserting the differential operators explicitly, the Klein-Gordon equation.
As a consequence of 2.5, and without recourse to the Dirac equation, we obtain
directly a linear wave equation from 2.5, which is named after Weyl 13 without the mass
term and after Majorana with the mass term. It involves only the Pauli matrix operators
acting on a two-component spinor φ but introduces the complications that are caused by
the operator τ and reads as follows:
 
∂ ∂
i σ · φx, t  mτφx, t. 2.7
∂t ∂x

In 2.7 we have not indicated the 2 × 2-unit matrix explicitly. This equation is nothing but
what is called nowadays the two-component Majorana equation 4, 5, although it was never
written down this way by Majorana himself. As shown by case in 6, this equation can be
also derived from Dirac’s equation in chiral form, or more generally if one imposes on it the
condition of Lorentz-covariant complex conjugation, a procedure which is clearly described
in a recent tutorial paper of Pal 8. However, the view we take here is that equally well we
may consider 2.5 as basic, in the spirit of the citation 6 quoted in the introduction.
Obviously, there is a second version of the Majorana 2.7, which is obtained by the
operation of τ on it. Namely, when we apply τ from the left side, we find that
 
∂ ∂  
i −σ · τφx, t  −mτ τφx, t  mφx, t. 2.8
∂t ∂x

Consequently, if φ solves 2.7 with the plus sign in front of σ, then χ  τφ solves it with the
minus sign, but with a minus sign also at the mass term. Effectively, this amounts to replacing
σ by −σ in 2.7, also implying τ is exchanged by −τ.
Apparently, the two equations 2.7 and 2.8 are closely connected twins. It is
interesting to note that the two together can provide a special but important four-component
Dirac spinor in the form

 
φ
ψ±C  , 2.9
±τφ

which solves the Dirac equation. Moreover, charge or better chirality conjugation 9, 10
is, in the chiral representation, given by the operator C  γτ, which when operating on
ψ±C reproduces it. Since C2  1, this operator has only the eigenvalue ±1. Therefore, the
spinors ψ±C constructed from the solution of the Majorana equation are the two eigenfunctions
with eigenvalue ±1 of the chirality conjugation operator C. Moreover, the chiral projection
operators are given by P±  1/21 ± α, and thus the left- and right-chiral spinor is ψR,L  P± ψ.
ISRN Mathematical Physics 5

As {γ, α}  0, we obtain CP±  P∓ C, and consequently, if one applies C separately to the right-
or left-chiral fields, one obtains

C ψC  ψC , C ψC  ψC . 2.10
R L L R

Thus this operator exchanges the chirality of the chiral projections of the complex-self-
conjugate spinor ψ C , and it transforms left- into right-handed states and vice versa, a virtue
which is essentially associated with the spin-flip operator τ as key ingredient of C.
In conclusion, in this special situation the operator C acts as chirality conjugation or
reversal. While being a composite of the two related and by themselves irreducible Majorana
equations, the reducible chiral Dirac equation has as an intrinsic symmetry of this chirality
conjugation symmetry.
Consequently, in what follows we can interpret the spinor φ as representing a right-
chiral field, and vice versa χ  τφ as representing a left-chiral field, which is obtained from
φ by application of the spin-flip operator. Apparently, it translates from one into the other
irreducible representation of the Lorentz group.
Let us consider now the symmetries of the two-component complex Majorana
equation, in particular the charge exchange C, parity P, and time reversal T operations. Here
we expand the lucid discussion of case in 6. Generally speaking the Majorana equation is
invariant under the symmetry operation O, if the two-component spinor

φO  Oφ 2.11

also fulfils that equation. When applying the operation O from the left and its inverse O−1
from the right, whereby the unit operator is given by the decomposition OO−1  1, we obtain
the result
     
∂ −1 ∂
O i O O σ ·i O−1 − mOτO−1 Oφx, t  0. 2.12
∂t ∂x

We define as usually the time and space coordinate inversion operations T and P on a spinor
φ by

Tφx, t  φx, −t,


2.13
Pφx, t  φ−x, t

and also recall the complex conjugation operation C, which gives CiC−1  −i and yields

Cφx, t  φ∗ x, t. 2.14

Here the asterisk denotes again the complex conjugate number. With these preparations in
mind, it is easy to see which operators provide the various symmetry operations, which are
composed in the Table 1. To complete the operator algebra, it is important to note that the
coordinate reversal operators T and P commute with τ and σ, respectively.
6 ISRN Mathematical Physics

Table 1: Symmetry operations.

Operation Time reversal Parity Chirality charge conjugation


Operator T  τT P  iτ P C  τ  σy C

Let us first consider in 2.12 the time reversal, O  T  τT. Apparently, it does
not affect the mass term with τ and also leaves the first term as well as the kinetic term
iσ unchanged. Therefore, φT  σy φ∗ x, −t solves the Majorana equation 2.7 as well.
Conversely, the parity operation O  P  iτP anticommutes with the mass term and inverts
the sign of the first term in 2.12, and it also does not leave the momentum term invariant,
since as x changes its sign, so does the spatial derivative. Consequently, φP  iσy φ∗ −x, t does
solve the Majorana equation 2.7. Finally, we consider the chirality conjugation, O  C  τ,
which only changes the sign of the first term. Therefore, φC  σy φ∗ x, t does not solve the
Majorana equation 2.7 but solves its conjugate version 2.8. In conclusion, of the symmetry
operations given in Table 1 time reversal and parity are obeyed, but chirality inversion is
maximally broken.
Finally, as the second major topic of our paper we shall reformulate the two-
component Majorana equation in its real form by decomposing the spinor φ into its real and
imaginary part: φ  φR  iφI . Making use of the three real 2 × 2 matrices defined in 2.3 we
can write:
 
∂ ∂ ∂ ∂
γ −α β φR  φI  −mφR , 2.15
∂t ∂x ∂z ∂y

and similarly we obtain

 
∂ ∂ ∂ ∂
γ −α β φI − φR  mφI , 2.16
∂t ∂x ∂z ∂y

both of which can be combined, by introducing φ±  φR ± φI , in a single equation:

   
∂ ∂ ∂ ∂
γ −α β φ±  ± − m φ∓ . 2.17
∂t ∂x ∂z ∂y

2
Note that γ 2  −1, α2  1, and β  1. Furthermore, these matrices all anticommute, and thus
by squaring 2.17 one immediately retains the scalar Klein-Gordon equation reading


∂2 ∂2
−  m φ±  0,
2
2.18
∂t2 ∂x2

which in fact was our starting point, when deriving a relativistic wave equation.
From the two real two-component Majorana equations 2.17, when both are combined
into a single one, the standard Dirac equation in the real Majorana representation follows
immediately. Namely, we may arrange the two spinors φ± into a single four-component real
ISRN Mathematical Physics 7
† †
Dirac spinor and write ψ †  φ , φ− . Then the coupled system of 2.17 transforms into a
4 × 4 real matrix differential equation which reads

        
γ 0 ∂ψ α 0 ∂ψ 0 −1 ∂ψ β 0 ∂ψ 0 1
−    −m ψ. 2.19
0 γ ∂t 0 α ∂x 1 0 ∂y 0 β ∂z 1 0

Consequently, we may now introduce the subsequent real 4 × 4 Majorana matrices in their
natural as deduced from the Pauli matrices representation:

      
μ 0 γ 0 −α 1 0 0 β
γ  , , , . 2.20
γ 0 −α 0 0 −1 β 0

They mutually anticommute and obey γ μ γ ν  γ ν γ μ  −2g μν . Thus the real Dirac equation in
Majorana representation reads

γ μ ∂μ ψ  mψ  0, 2.21

which can, with the help of the purely imaginary Dirac matrices γ μ  iγ μ , easily be brought
into the standard form of the Dirac equation:

iγ μ ∂μ ψ  mψ. 2.22

Therefore, the four-component Dirac equation 2.21 is, in its real Majorana representation, a
direct consequence of the basic two-component complex Majorana equation 2.7, which was
derived here without invoking the Dirac equation in the first place. Subsequently, we shall
also derive the four-component real eigenspinors of 2.21.

3. Eigenfunctions of the Two-Component Majorana Equation


with Mass Term
Returning to the basic equation 2.7, we here derive its eigenfunctions. In order to solve it, we
make the usual plane-wave ansatz, φx, t  up, E exp−iEt  ip · x  vp, E expiEt − ip · x.
But note that we require the plane wave and its complex conjugate as well, because of the
operator τ in 2.7. The two-component spinors u and v cannot be assumed to be complex
conjugated; that is, the wave function cannot be expected to be real. The resulting linked
eigenspinor equations are

E − σ · pup, E  mτvp, E, 3.1

E − σ · pvp, E  −mτup, E. 3.2


8 ISRN Mathematical Physics

By insertion of the first into the second equation, or vice versa, the relativistic dispersion
relation is obtained:

τE − σ · pτE − σ · p  m2 up, E  0, 3.3

which yields the two eigenvalues



E1,2 p  ± m2  p2 , 3.4

which are obtained from the requirement that for nontrivial solutions of the spinors u and
v to exist the determinant associated with the eigenvalue problem 3.3 must vanish. The
negative root in 3.4 cannot be neglected, since as usually it is related to antiparticles. We
can solve 3.2 for v and insert it back into the previous ansatz for φx, t to obtain finally the
solutions of 2.7 and 2.8 in the concise forms:
 
Eσ·p
φx, t  1− τ exp−iEt  ip · xu, 3.5
m
 
E−σ·p
χx, t  τφx, t  τ  exp−iEt  ip · xu. 3.6
m

To validate these solutions by direct differentiation, careful attention must be paid to the
anticommutation rules between τ and i, respectively, σ. We may also solve 3.2 for u instead,
and then insert it back into the previous ansatz for φx, t.
We are free to choose for the eigenspinor u the two standard spin-up and -down
† †
eigenfunctions: u1  1, 0 and u2  0, 1, and similarly for v, but there is a better and
more adequate choice if p is nonzero, as discussed in the following. Similar solutions like
that of 3.5 and 3.6 are obtained for the negative energy root in 3.1, yielding the
antiparticle wavefunctions. Superposition of all the Fourier modes and their summation over
the momentum variable p leads finally to the general Majorana fields, whose quantization
then follows from the canonical rules 2–5 and is given in the following. Above we obtained
the formal solutions of the Majorana equation, which we reiterate here by introducing the
Majorana operator M as follows:
 
∂ ∂
Mi σ· − mτ. 3.7
∂t ∂x

The solution φ obeys Mφ  0. Note, however, that the four-momentum operator Pμ  i∂μ 
i∂/∂t, ∂/∂x does not commute with M, and neither does the spin-flip operator τ. Therefore,
the φ and χ of 3.5 to 3.6 are not eigenfunctions of any of these operators. However, the
helicity operator, σ · P, does commute with M.
Consequently, to chose the eigenfunctions of the helicity operator in φ for the free
functions u or v is most convenient. The eigenvalue equation of the helicity operator in
Fourier space reads

 u± 
σ · p p  ±u± 
p. 3.8
ISRN Mathematical Physics 9

  p/p and are given by


The two eigenvectors depend on the momentum unit vector p

⎛ ⎞ ⎛ ⎞
θ −i/2φ θ −i/2φ
⎜cos e ⎟ ⎜ − sin e ⎟
2 2
p  ⎜
u  ⎝
⎟, p  ⎜
u−  ⎟, 3.9
θ i/2φ ⎠ ⎝ θ i/2φ ⎠
sin e cos e
2 2

in which the half angles of θ and φ appear. The eigenvectors for the same p  are orthogonal
to each other and normalized to a modulus of unity, and they obey the relation u± − p 
u∓ 
p. This is a consequence of the eigenvalue equation 3.8, which implies that u±  p
is an eigenvector of the helicity operator, corresponding to a right-handed, respectively,
left-handed, screw with respect to the momentum direction. According to 3.9 we have
† †
u± 
pu± 
p  1 and u∓ 
pu± 
p  0. The dagger denotes as usually the transposed denoted
by the superscript T  and complex conjugated vector, respectively, matrix. The scalar product
between two vectors spinors v and w is just the sum over the products of their two
components; that is, vw simply stands for v1 w1  v2 w2 .
We emphasize that the spin-flip operator τ  σy C, when operating on the previous
eigenspinors, leads to

p  ±iu∓ 
τu±  p, 3.10

That is, it connects the eigenfunctions of opposite helicity and turns out to be a quite useful
relation in the subsequent considerations. By its application, we can write the two possible
associated eigenfunctions after 3.5 as

i
φ± x, t  exp−iEt  ip · xu± 
p ∓ p,
E  σ · p expiEt − ip · xu∓  3.11
m

where the advantage of the helicity eigenfunctions becomes obvious. Namely, by use of 3.8
we obtain the now normalized to unity eigenfunction
  
1
φ± x, t  √ E ± p exp−iEt  ip · xu± 
p ∓ i E ∓ p expiEt − ip · xu∓ 
p , 3.12
2E

which is a mixed state involving both helicities in a symmetric fashion. Note that for
vanishing mass, m  0 corresponding to the Weyl equation, only the positive helicity remains,
and thus the wavefunction becomes purely right-handed. Its left-handed version is then
obtained by applying the spin-flip operator τ on 3.12 like in 3.6. Operation of M on this
φ± validates that it solves the Majorana equation, that is, Mφ±  0. A simple form of the
eigenspinor is obtained for a particle at rest, that is, p  0, which yields

1  
φ± t  √ exp−imtu± ∓ i expimtu∓ . 3.13
2

The helicity eigenvectors in this case can be chosen to be the standard ones obtained for the
angles φ  0 and θ  0 according to 3.9.
10 ISRN Mathematical Physics

Concerning the adequate


 choice of the eigenspinor u1,2 in 3.5 for the two possible
eigenvalues E1,2 p  ± m  p , we recall that the helicity operator does commute with M,
2 2

and generally we should consider a mixture or superposition of both helicities. But this is
indeed already implied in the solution 3.12. Therefore, as we can only have two linearly
independent eigenvectors to define the basis, we assume that u can be decomposed such that
these two eigenvectors are defined by

u1,2 p  a1,2 pu± 


p, 3.14

with some complex amplitude a1,2 p of module unity, |a1,2 |2  1, to ensure normalization.
The natural but arbitrary association with the sign of the energy 3.4 is to take the positive
sign for positive right-handed helicity and the negative sign for negative left-handed
helicity. Thus we may conventionally refer to particles index 1 and antiparticles index 2.
Upon insertion of this ansatz in 3.5 we obtain a yet more general expression for the two
related eigenfunctions as follows:

 
E1  p E1 − p
φ1 x, t  exp−iE1 t  ip · xa1 u − i expiE1 t − ip · xa∗1 u− ,
2E1 2E1
  3.15
E2 − p E2  p
φ2 x, t  exp−iE2 t  ip · xa2 u−  i expiE2 t − ip · xa∗2 u ,
2E2 2E2

where the obvious momentum arguments in u and a have been suppressed for the sake of
lucidity. Apparently, the previous wavefunction 3.12 is retained for a1,2 p  1, and E1,2  E.
It turns out to be convenient to introduce the two real quantities


  
   E p ±p
ε± p     , 3.16
2E p

the squares of which add up to unity, ε2  ε−2  1. A useful


 property of the epsilons is the
obvious relation E ± pε∓  mε± . From now on Ep  m2  p2 is used for the positive
root obtained from 3.4, and the argument p in the epsilons will subsequently be omitted for
simplicity. Then we can write the two eigenfunctions with a new name concisely as

φ1 x, t  ε exp−iEt  ip · xa1 u − iε− expiEt − ip · xa∗1 u− ,


3.17
φ2 x, t  ε expiEt  ip · xa2 u−  iε− exp−iEt − ip · xa∗2 u .

To make the antiparticle wavefunction having the same standard plane wave phase like
that of the particle, we invert its momentum and note again that the polarization vectors
change according to u± − p  u∓ 
p. This inversion is permitted as we will later sum over
all momenta, and thus p is just a mute index. Consequently, we take φP p  φ1 p for
the particle, but φA p  φ2 −p for the antiparticle, and we redefine the amplitudes as
ISRN Mathematical Physics 11

ap  a1 p for the particle, but bp  ia∗2 −p, respectively, b∗ p  −ia2 −p for the
antiparticle. The new wavefunctions finally read as follows:

φP x, t  ε exp−iEt  ip · xau − iε− expiEt − ip · xa∗ u− ,


3.18
φA x, t  ε− exp−iEt  ip · xbu−  iε expiEt − ip · xb∗ u .


Both eigenfunctions are normalized to unity, yielding the relation φP,A φP,A  1, which
is obtained by using ε2  ε−2  1, exploiting the normalization and orthogonality of the helicity
eigenvectors, and noting that a∗ a  1 and b∗ b  1. They form a complete eigenvector basis
of the Majorana operator M and obey MφP,A  0. However, we emphasize again that they
are neither eigenfunctions of the four-momentum operator P μ  P0 , P nor of the helicity
operator or the spin-reversal operator τ.
By adding up the two contributions we retain the most general solution of the
Majorana equation 2.7 as the two-component spinor in the form

φx, t  φP x, t  φA x, t. 3.19

This superposition of the particle and antiparticle eigenspinors yields the full Majorana
wavefunction in the form

φx, t  exp−iEt  ip · xε u a  ε− u− b − i expiEt − ip · xε− u− a∗ − ε u b∗ . 3.20

4. The Majorana Quantum Field


At this point the transition from 3.20 to a quantum field is quite natural and obvious. We just
have to replace the amplitudes in 3.20 by the canonical anticommuting fermion operators
obeying

  
ap, a† p  δp,p ,
 4.1
 
bp, b† p  δp,p .

Of course, all possible anticommutators between two creation and two annihilation operators
are zero. Mutually, between the as and bs the anticommutators vanish, as they should since
these two degrees of freedom are independent. The operator a† p creates, and vice versa
 the plane-wave state of a particle of positive helicity, with momentum p and
ap annihilates
energy E  m2  p2 . The b-operators do the same, yet for the related antiparticle of the
opposite negative helicity.
We finally get, by keeping explicitly all momentum arguments, the following field
operator for any given momentum p in the concise form

       
Φp x, t  exp −iE p t  ip · x cp − i exp iE p t − ip · x d† p, 4.2
12 ISRN Mathematical Physics

which assumes the standard form similar to what is known from the Dirac equation. The
quantum field operators corresponding to φ are denoted by a capital Φ. We can now sum up
overall momenta to obtain the full Majorana quantum field:

ΦM x, t  Φp x, t. 4.3
p

We introduced as abbreviations two operators, which correspond to the creation,


respectively, annihilation, polarization-vector operators indicated by an underscore for a
given mixed helicity state as follows:
   
cp  ε p u 
pap  ε− p u− 
pbp, 4.4

and similarly

   
d† p  ε− p u− 
pa† p − ε p u 
pb† p. 4.5

They obey the standard anticommutation rule, whereby the inner product between the
original helicity eigenvector is used, as well as the relation ε2  ε−2  1. Thus we obtain, by
implying the standard scalar product between the complex two-component spinors involved,
the result
  
cp, c† p  δp,p , 4.6

and correspondingly
  
dp, d† p  δp,p . 4.7

Again all possible mixed anticommutators between the c and d operators vanish. For the two
particle number operators we obtain from 4.4 and 4.5 the linear combinations

c† c  ε2 a† a  ε−2 b† b,
4.8
d† d  ε−2 a† a  ε2 b† b.

This combined coordinate transformation, involving the original eigenspinors of the


helicity operator and their corresponding state operators classically amplitudes, has a
physically important interpretation. As stressed before, the helicity operator does commute
with M, but one requires a mixed helicity state to represent its eigenfunctions. This mixing
depends on the ratio of the momentum, p, to the total energy, Ep, of a particle. Only if
m  0, we have pure left- or right-handed states, otherwise the massive states are mixed.
That MΦp  0 follows readily from the notion that

     
E − σ · pcp  m ε− p u  pbp  miτd† p,
pap  ε p u−  4.9
ISRN Mathematical Physics 13

respectively, that

   
E − σ · pd† p  m ε p u− 
pa† p − ε− p u 
pb† p  −miτcp, 4.10

where the auxiliary relation E ± pε∓  mε± has been used. We should mention here that
according to its definition the operation of τ on a creation or annihilation operator is defined
such that τa  a† τ, or similarly τa†  aτ.
As an interlude, we may now consider the unmixed massless case m  0, which gives
us the Weyl field. Then ε  1 and ε−  0. The resulting quantum field operator is given by

    
ΦWR x, t  p exp −ipt  ip · x ap  i exp ipt − ip · x b† p .
u  4.11
p

Apparently, this field involves only the polarization vector for positive helicity. Therefore, in
this case the right-chiral Weyl field operator annihilates right-handed particles and creates
left-handed antiparticles. Operating with the spin flip operator τ, according to τaτ −1  a†
and 3.10, on this field yields the left-chiral Weyl field, which reads

    
ΦWL x, t  p exp −ipt  ip · x bp  i exp ipt − ip · x a† p .
u−  4.12
p

Consequently, in the case of the left-chiral Weyl field the operator creates right-handed
particles and annihilates left-handed antiparticles.
Similarly, we obtain for the massive right-chiral Majorana quantum field:

        
ΦMR x, t  exp −iE p t  ip · x cp − i exp iE p t − ip · x d† p , 4.13
p

and by application of τ on it the left-chiral Majorana quantum field:

         
ΦML x, t  exp −iE p t  ip · x iτd† p − i exp iE p t − ip · x iτcp .
p
4.14

The corresponding operators can be read off as in 4.9 and 4.10.


Ultimately, we are interested in the expectation value of a given Hermitian operator
O for the quantum field Φ, for example, the Majorana fields 4.13 and 4.14 or the Weyl
fields of 4.11 and 4.12, which are all not Hermitian fields. For the massless Weyl fields, the
related plane waves are also eigenfunctions of the four-momentum operator P μ . In contrast,
for the massive Majorana field this is not true, and therefore only the expectation value of P μ
can be evaluated. The same comment applies to the kinetic helicity operator, which yet does
commute with M. This interesting feature was discussed at length and clarified by Mannheim
7, who concluded that for the massive Majorana field theory “second quantization is
necessary a priori in order to produce a sensible particle interpretation.” Therefore, we can
14 ISRN Mathematical Physics

only calculate average expectation values, which are defined by volume integrals of the real
part of the binary form


1
O  O  d3 x Φ† OΦ  OΦ† Φ . 4.15
2

This formula shall be now applied to calculate the expectation value of the energy-
momentum four-vector, the helicity and number operator of the right-chiral Majorana
quantum field ΦMR x, t of 4.13. Upon its insertion into 4.15, and after some lengthy but
standard calculations similar to what is usually done with the Dirac equation, e.g., see the
textbooks 10 or 9 we obtain the concise result

μ  
PMR  P0 , PMR  m2  p2 , p a† pap  b† pbp , 4.16
p


where conventionally an infinite constant, − p Ep, has been discarded as irrelevant zero-
point energy. Let us consider the helicity density at a given p of the Majorana operator 4.2.
We find


 Φp  c† pσ · p
Φp σ · p  d† p  a† a − b† b − ε−2  ε2 .
 cp  dpσ · p 4.17

and thus by using these relations and by summing up all contributions of 4.17, we obtain
the total helicity operator HMR , which is within an unimportant constant given by

! 

HMR  σ · P  a† pap − b† pbp .
MR 4.18
p

It involves the difference of the number operators of the particles and antiparticles, which
therefore are expected to contribute oppositely to the net helicity.
Interestingly, we can go through the same procedure with the left-chiral field 4.14
and end up with exactly the same expectation values as calculated previously. One gets a
similar result for the Weyl quantum field, by putting either m  0 in the Majorana field or
by using directly 4.11 and 4.12. In other words it suffices to consider mathematically the
original and constitutive equation 2.5. Chirality conjugation, that is, the replacement of the
spin σ by its negative inverted vector −σ, does not yield new physical information concerning
the above expectation values 4.16 and 4.18.

5. Four-Component Eigenspinors of the Majorana Equation


Following the derivation of the real Majorana equation 2.21, its eigenfunctions can, by
construction according to 2.19, easily be derived. Starting point is the complex two-
component original wave function 3.20, in which the wave amplitudes a and b correspond
after the previous section to particles, respectively, antiparticles, and thus transform into the
ISRN Mathematical Physics 15

related annihilation operators for the quantum fields. The four-component eigenfunction is
therefore given by the spinor
 
φR  φI
ψ , 5.1
φR − φI

where the real and imaginary parts are taken from 3.20. This spinor can be written as a sum
of the particle and antiparticle components and reads

ψx, t  ψP x, t  ψA x, t. 5.2

These two contributions can be expressed, in terms of the complex four-component polar-
ization spinors to be defined below, separately as follows:

αp  expiEt − ip · xa∗ p


ψP x, t  exp−iEt  ip · xap α∗ p, 5.3


ψA x, t  exp−iEt  ip · xbpβp  expiEt − ip · xb∗ pβ∗ p. 5.4


Apparently, ψP,A  ψP,A  and
, and thus the wave functions are real. The polarization spinors α

β can be constructed from the two-component eigenfunctions, given in 3.9, of the helicity
operator 3.8. Thus they can be concisely written as
      
1 p − ε− p u∗− 
ε p u  p 1 − i
 p 
α       , 5.5
2 p  ε− p u∗− 
ε p u  p 1  i
      
1 p  ε p u∗ 
ε− p u−  p 1 − i

βp        . 5.6
2 p − ε p u∗ 
ε− p u−  p 1  i

Use has again been made of the symbols ε± as defined in 3.16. Using their properties and the
  1 and β† β  1, respectively, α
orthogonality of u± , one can readily show that α† α  † β  0  β† α
.
We may now insert the spinors 5.3 and 5.4 into the real Majorana equation 2.21
or complex Dirac equation 2.22, to validate that they are solutions of those equations. With
the covariant four-momentum pμ  E, −p, one finds that the real and imaginary parts of
α
α  R  i  must obey the coupled equations:
αI resp., β

γ μ pμ α
 I  m
αR  0, 5.7

and similarly

γ μ pμ α
 R − m
αI  0, 5.8

which can be combined to yield the complex solution of the Dirac equation as

  m
γ μ pμ α α. 5.9
16 ISRN Mathematical Physics

The same equation must hold true for the polarization spinor β,  and this is indeed the case.
The latter equation can explicitly be written in matrix form as M
α  0, with the non-Hermitian
matrix
⎛ ⎞
−py  im 0 px −E − pz
⎜ 0 −py  im E − pz −px ⎟
M⎜
⎝ px
⎟, 5.10
−E − pz py  im 0 ⎠
E − pz −px 0 py  im

which is the 4 ×4 -matrix analogue of the Majorana operator M in 3.7. The validation of 5.9
requires some lengthy algebraic calculations that shall not be done here. Nontrivial solutions
of α to exist require that the determinant of M vanishes. The corresponding polynomial in the
real variable E is of fourth order, yet only yields the two real roots already quoted in 3.4,
corresponding to particles and antiparticles.
For the sake of completeness we finally give the full four-component polar-
ization spinor α, which in terms of the angles of the momentum unit vector p  
sin θ cos φ, sin θ sin φ, cos θ and p reads as follows:

⎛   θ   θ
 ⎞
−i/2φ i/2φ
⎜ ε p cos 2 e  ε− p sin e 1 − i⎟
⎜ 2 ⎟
⎜  ⎟
⎜   θ   θ ⎟
⎜ ε p sin ei/2φ − ε− p cos e−i/2φ 1 − i⎟
1⎜ 2 2 ⎟
 p  ⎜
α ⎜  ⎟.
⎟ 5.11
2⎜   θ   θ ⎟
⎜ ε p cos e−i/2φ − ε− p sin ei/2φ 1  i⎟
⎜ 2 2 ⎟
⎜  ⎟
⎝   θ   θ ⎠
ε p sin ei/2φ  ε− p cos e−i/2φ 1  i
2 2

Similarly, the polarization spinor β can be derived explicitly from 5.6. This derivation
essentially concludes the section on the solution of the real Majorana equation with mass
term. We have presented its eigenspinor for particles ψP in 5.3 and for antiparticles ψA
in 5.4. The related quantum fields are readily obtained by interpreting a as annihilation
operator, and by replacing a∗ by a† and interpreting it as the particle creation operator both
together obeying the usual fermion anticommutation rule 4.1. The resulting field operator
ΨP is not Hermitian, though, as α  † is not equal to α
 ∗ . The same comments apply to the
antiparticle quantum field described by the creation operator b† and polarization spinor β∗ .

6. Summary and Conclusions


The principal goal of this paper was to rederive and discuss the two-component and
four-component Majorana equations completely on their own rather than as a spinoff of
the chiral Dirac equation. We obtained these equations including a mass term in a new
way by a direct linearization of the relativistic dispersion relation of a massive particle.
Thereby we only made use of the complex conjugation operator and the Pauli spin
matrices, corresponding to the irreducible representation of the Lorentz group. We then
calculated the eigenfunctions of the complex two-component Majorana equation and its
related quantum field in a transparent way, exploiting the spin-flip or related chirality
ISRN Mathematical Physics 17

conjugation operator. Subsequently, we showed the four-component Dirac equation in its


real Majorana representation to be the natural outcome of the genuine, two-component, and
complex Majorana equation 2.7.
As the analysis of this version of the Majorana equation clearly reveals, the two
associated Majorana particles with creation operator a† and similarly b†  represent linearly
independent states and are not their own antiparticles, when being defined like in reference
7 in the sense of having positive or negative energies 3.4. We recall here that the original
eigenfunctions 3.19 and 3.20 result from a superposition of different states and thus do
neither describe a pure helicity state nor do they have a well-defined four-momentum. These
distinct features only emerge by evaluating the expectation values 4.16 and 4.18 of the
quantum fields. They do in fact describe particles and antiparticles with opposite helicities,
consistently with the standard perception of neutrinos and antineutrinos.
The symmetries of the two-component Majorana equation were also analysed. It was
shown to obey time reversal and parity, yet apparently violates chirality conjugation by
construction. As the Majorana quantum field is uncharged, the common term “charge con-
jugation” does not appear adequate here, but better Lorentz-covariant complex conjugation
8. Even more appropriately, while stressing clearly the physical meaning, one may speak of
“chirality conjugation,” which refers to the two conjugate versions of the Majorana equation
having opposite chirality. So, the Majorana equation 2.7, while breaking chiral symmetry
by definition, is as basic as Dirac’s equation and when considered as a quantum field
can describe massive and uncharged right-handed or left-handed particles or antiparticles,
corresponding to massive neutrinos. In our opinion, the approach starting from scratch with
2.5 gives us new insights into the nature of the Majorana equation and shows that it
deserves to be considered in its own right.

References
1 E. Marsch, “The two-component Majorana equation—novel derivations and known symmetries,”
Journal of Modern Physics, vol. 2, pp. 1109–1114, 2011.
2 P. A. M. Dirac, “The quantum theory of the electron,” Proceedings of the Royal Society London, vol. 117,
pp. 610–624, 1928.
3 E. Majorana, “Teoria simmetrica dell’ elettrone e del positrone,” Nuovo Cimento, vol. 14, pp. 171–184,
1937.
4 R. N. Mohapatra and P. B. Pal, Massive Neutrinos in Physics and Astrophysics, World Scientific,
Singapore, 2004.
5 M. Fukugita and T. Yanagida, Physics of Neutrinos and Applications to Astrophysics, Springer, Berlin,
Germany, 2003.
6 K. M. Case, “Reformulation of the Majorana theory of the neutrino,” Physical Review, vol. 107, pp.
307–316, 1957.
7 P. D. Mannheim, “Introduction to Majorana masses,” International Journal of Theoretical Physics, vol.
23, no. 7, pp. 643–674, 1984.
8 P. B. Pal, “Dirac, Majorana, and Weyl fermions,” American Journal of Physics, vol. 79, pp. 485–498, 2011.
9 A. Das, Lectures on Quantum Field Theory, World Scientific Publishing, Singapore, 2008.
10 M. Kaku, Quantum Field Theory. A Modern Introduction, The Clarendon Press Oxford University Press,
New York, NY, USA, 1993.
11 W. Pauli, “Zur quantenmechanik des magnetischen elektrons,” Zeitschrift fur Physik A, vol. 43, pp.
601–623, 1927.
12 H. Jehle, “Two-component wave equations,” Physical Review, vol. 75, p. 1609, 1949.
13 H. Weyl, “Elektron und graviton I,” Zeitschrift fur Physik, vol. 56, pp. 330–352, 1929.
Advances in Advances in Journal of Journal of
Operations Research
Hindawi Publishing Corporation
Decision Sciences
Hindawi Publishing Corporation
Applied Mathematics
Hindawi Publishing Corporation
Algebra
Hindawi Publishing Corporation
Probability and Statistics
Hindawi Publishing Corporation
http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014

The Scientific International Journal of


World Journal
Hindawi Publishing Corporation
Differential Equations
Hindawi Publishing Corporation
http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014

Submit your manuscripts at


http://www.hindawi.com

International Journal of Advances in


Combinatorics
Hindawi Publishing Corporation
Mathematical Physics
Hindawi Publishing Corporation
http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014

Journal of Journal of Mathematical Problems Abstract and Discrete Dynamics in


Complex Analysis
Hindawi Publishing Corporation
Mathematics
Hindawi Publishing Corporation
in Engineering
Hindawi Publishing Corporation
Applied Analysis
Hindawi Publishing Corporation
Nature and Society
Hindawi Publishing Corporation
http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014

International
Journal of Journal of
Mathematics and
Mathematical
Discrete Mathematics
Sciences

Journal of International Journal of Journal of

Hindawi Publishing Corporation Hindawi Publishing Corporation Volume 2014


Function Spaces
Hindawi Publishing Corporation
Stochastic Analysis
Hindawi Publishing Corporation
Optimization
Hindawi Publishing Corporation
http://www.hindawi.com Volume 2014 http://www.hindawi.com http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014

You might also like