This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

The intermediate fermionic species created by SO(3)SO(3) rotation in the representation of the Dirac equation

H. Moaiery Jundi-Shapur University of Technology, Desful, Iran    M. N. Najafi Department of Physics, University of Mohaghegh Ardabili, P.O. Box 179, Ardabil, Iran
Abstract

The question of how does the Dirac equation depend on the choice of the γ\gamma matrices has partially been addressed and explored in the literature. In this paper we focus on this question by considering a general form of γ\gamma matrices, and call the resulting spin 12\frac{1}{2} fermions as intermediate fermion species (IFS). By inspecting the properties of IFS, we find that all species transform to each other by a SO(3)SO(3) similarity transformation in the space of parameters, that are the entities of the γ\gamma matrices. Many properties, like eigenvalue problem and boost are tested for IFS. We find also sub-representations that generate Majorana fermions, which is isomorphism to U(1)U(1) group.

Dirac equation, S(3) symmetry, Boost and rotation
pacs:
05., 05.20.-y, 05.10.Ln, 05.45.Df

I Introduction

The representations of fermions governed by the Dirac equation have vast applications in various fields in the fundamental and theoretical physics, ranging from elementary particles Gasiorowicz (1966) and quantum chromodynamics Ioffe et al. (2010) to condensed matter Morandi et al. (2013), photonics  Sansoni et al. (2012), and superconductivity Chamon et al. (2010). Three important representations of the Dirac equation are the Dirac fermions, the Weyl fermions and the Majorana fermions Leijnse and Flensberg (2012), depending on the choice of the matrices in the Dirac equation (namely the γ\gamma matrices), which show different properties in some aspects Ryder (1996). The examples in condensed matter physics are intrinsic graphene for which the electrons show linear (gapless) dispersion realizing the massless Dirac fermions Neto et al. (2009); Sarma et al. (2011); Peres (2010), and the edge states in quantum Hall effect which are Majorana fermions Nayak et al. (2008). Weyl fermions may be realized as emergent quasiparticles in a low-energy condensed matter system. In Weyl semimetals Xu et al. (2015a, b, c); Armitage et al. (2018) (as a topologically nontrivial phase of matter) the low energy excitations are Weyl fermions that carry electrical charge, with distinct chiralities. Symmetry-protected Dirac fermions in topological insulators Moore (2010); Hasan and Moore (2010); Kane and Mele (2005); Gu and Wen (2009); Pollmann et al. (2012); Chen et al. (2011), Majorana fermions in low energy excitations of condensed matter systems Elliott and Franz (2015); Beenakker (2015), like quantum Hall effect Nayak et al. (2008) and superconductivity Elliott and Franz (2015), are the other examples of emergent fermions with vast applications. Not surprisingly, the representation of fermions in the standard model of particles also plays a dominant role Avignone III et al. (2008); Kuno and Okada (2001)

Despite of this huge applications which stimulated a huge rate of studies on emergent fermionic systems (putting the subject in a large class “Dirac material” Chiu et al. (2016)), there is no comprehensive systematic investigation on the possible intermediate fermionic species (IFS) which arises by inspecting the possible representations of the Dirac equation. To be more precise, only a few specific and useful forms of fermions have been was considered, such as the standard or the Dirac-Pauli Representation (shown here by an S-index), supersymmetric representation, Weyl or spinor representation, Majorana representation etc., corresponding to appropriate choice of γ\gamma matrices Ryder (1996). Here we show that the Dirac equation admits the existence of IFSs by a systematic continuous SO(3)SO(3) rotation in the representation space of the Dirac equation, i.e. introducing and analyzing the general form of the γ\gamma matrices. We argue that, although the “type” of the fermions change by this rotation, some properties of the particles like the dispersion relation and the Klein tunneling do not change, leading us to consider it as a symmetry of the Dirac equation. Many properties of the IFS particles are investigated in terms of rotation parameters, like the helicity and boost properties, parity. We also investigate the intermediate Majorana species as a sub-representation of SO(3)SO(3) which is homomorphic to U(1)U(1) group.

The paper has been organized as follows: In the next section we introduce the IFS particles as the solution of the Dirac equation with general γ\gamma matrices. Besides finding the wave functions, we investigate the behavior of the IFS under boost transformation. In SEC. III we introduce SO(3)SO(3) rotations which transform the IFS particles. Section IV is devoted to sub-representations of the SO(3)SO(3) group, i.e. the U(1)U(1) group. In SEC. V we explain how to find IFS particles from the Dirac fermions. We close the paper by a conclusion.

II General representation of linearized relativistic particles

In the natural units =c=1\hbar=c=1 the Dirac equation in d+1d+1 dimensions is

(γμpμm)Ψ(x,t)=0\left(\gamma^{\mu}p_{\mu}\mp m\right)\Psi\left(\vec{x},t\right)=0 (1)

where pμp^{\mu} is four- [or generally (d+1)(d+1)-] momentum (μ=0,1,,d\mu=0,1,...,d, in which 0 stands for the time component, and the others for the spatial components), and γμ\gamma^{\mu}s are the (even dimensional) Dirac matrices that are not unique (the possible minimum dimension of which depends on ddRyder (1996). As a well-known fact, the above equation gives rise to the Klein-Gordon equation, imposing some strong limitations on the choice (representation) of the γ\gamma matrices. Having chosen the representation of the γ\gamma matrices, one may reach to the other representations by a simple similarity transformation. To be more precise, let us suppose that we have

γμ=TγDμT1\gamma^{\mu}=T\gamma^{\mu}_{D}T^{-1} (2)

where γDμ\gamma^{\mu}_{D} are the Dirac gamma matrices, and TT is a general transformation. Then obviously the same Dirac equation is valid, i.e. (γDμpμm)ΨD(x,t)=0\left(\gamma^{\mu}_{D}p_{\mu}\mp m\right)\Psi_{D}\left(\vec{x},t\right)=0, where ΨTΨD\Psi\equiv T\Psi_{D} is the solution of the Dirac equation in the Dirac representation. It is the aim of the present paper to study systematically this problem by considering an arbitrary form of γ\gamma matrices and find the possible forms that TT’s can have. We argue about some possible non-trivial outcomes and consequences of this “generalization”, postponing any further investigations and uncovering any possible consequences to the community and also our future works.

II.1 General γ\gammas and wave functions

Let us start with the general expectation that the Klein-Gordon equation casts to a product of two copies of Eq. 1 with the requirement Ryder (1996)

γνγμpνpμ=gMμνpνpμ=m2,\gamma^{\nu}\gamma^{\mu}p_{\nu}p_{\mu}=g_{\text{M}}^{\mu\nu}p_{\nu}p_{\mu}=m^{2}, (3)

where gμνM=diag(+1,1,,1)g_{\mu\nu}^{\text{M}}=\text{diag}\left(+1,-1,\cdots,-1\right) is the symmetric Minkowski metric, giving rise to

{γμ,γν}=2gμνI,\left\{\gamma^{\mu},\gamma^{\nu}\right\}=2g^{\mu\nu}I, (4)

where II is the d+1d+1 dimensional identity matrix. Throughout this paper we use the following Hermitian matrices:

γ0γ¯0,γjiγ¯j;j=1,2,,d\gamma^{0}\equiv\bar{\gamma}^{0}\ \ ,\ \ \gamma^{j}\equiv i\bar{\gamma}^{j}\ \ ;\ \ j=1,2,\cdots,d (5)

for which the following relations hold

{γ¯μ,γ¯ν}=2δμν,(γ¯μ)=γ¯μ,Tr(γ¯μ)=0\left\{\bar{\gamma}^{\mu},\bar{\gamma}^{\nu}\right\}=2\delta^{\mu\nu}\ ,\ \left(\bar{\gamma}^{\mu}\right)^{{\dagger}}=\bar{\gamma}^{\mu}\ ,\ \text{Tr}\left(\bar{\gamma}^{\mu}\right)=0 (6)

Here after, we consider the case d=1d=1, for which the gamma matrices are 2×22\times 2. The generalization of the formalism to higher dimension is straightforward. For this case, one can easily show that the γ¯\bar{\gamma} matrices have to be in the following form (see Eq. 41 Appendix):

γ¯μ=(cμaμibμaμ+ibμcμ)=aμσx+bμσy+cμσz\begin{split}\bar{\gamma}^{\mu}&=\left(\begin{matrix}c_{\mu}&a_{\mu}-ib_{\mu}\\ a_{\mu}+ib_{\mu}&-c_{\mu}\end{matrix}\right)\\ &=a_{\mu}\sigma_{x}+b_{\mu}\sigma_{y}+c_{\mu}\sigma_{z}\end{split} (7)

where μ=0,1\mu=0,1, (aμ,bμ,cμ)(a_{\mu},b_{\mu},c_{\mu})\in\Re, and σx=(0110)\sigma_{x}=\left(\begin{matrix}0&1\\ 1&0\end{matrix}\right), σy=i(0110)\sigma_{y}=i\left(\begin{matrix}0&-1\\ 1&0\end{matrix}\right), σz=(1001)\sigma_{z}=\left(\begin{matrix}1&0\\ 0&-1\end{matrix}\right) are Pauli matrices. For later convenience, let us define γ¯2\bar{\gamma}^{2} with the same definition as above, so that μ=0,1,2\mu=0,1,2 in the above equation. Using the anti-commutation relation of γ¯\bar{\gamma} matrices, one can show that the general relation aμaν+bμbν+cμcν=δμνa_{\mu}a_{\nu}+b_{\mu}b_{\nu}+c_{\mu}c_{\nu}=\delta_{\mu\nu} holds. One easily retrieves the standard representation (SR) limit by setting c0=b1=a2=1c_{0}=b_{1}=a_{2}=1, and zero for the others.

By constructing the general Dirac equation using this general γ\gamma matrices one can readily find the plane wave solutions to be (see Eq. 43)

DIFSψ(E,k)(γ¯0Eiγ¯1kmI)ψ(E,k)=0,D_{\text{IFS}}\psi(E,k)\equiv\left(\bar{\gamma}^{0}E-i\bar{\gamma}^{1}k-mI\right)\psi(E,k)=0, (8)

where kk is the momentum of particle, and the explicit form of DIFSD_{\text{IFS}} is

DIFS(c0Eic1km(a0ib0)Ei(a1ib1)k(a0+ib0)Ei(a1+ib1)kc0E+ic1km)D_{\text{IFS}}\equiv\begin{pmatrix}c_{0}E-ic_{1}k-m&(a_{0}-ib_{0})E-i(a_{1}-ib_{1})k\\ (a_{0}+ib_{0})E-i(a_{1}+ib_{1})k&-c_{0}E+ic_{1}k-m\end{pmatrix} (9)

By setting the determinant of DIFSD_{\text{IFS}} to zero, we recover the dispersion relation E2=k2+m2E^{2}=k^{2}+m^{2} (E=±E0E=\pm E_{0}, where E0=k2+m2E_{0}=\sqrt{k^{2}+m^{2}}) as expected. The eigenfunctions are then given by

ψIFS+E0=ζ(1f+),ψIFSE0=ζ(f1)\psi_{\text{IFS}}^{+E_{0}}=\zeta\begin{pmatrix}1\\ f^{+}\end{pmatrix},\ \ \psi_{\text{IFS}}^{-E_{0}}=\zeta\begin{pmatrix}f^{-}\\ 1\end{pmatrix} (10)

where f±=i(b0E0a1k)±(a0E0+b1k)c0E0+mic1kf^{\pm}=\dfrac{i(b_{0}E_{0}-a_{1}k)\pm(a_{0}E_{0}+b_{1}k)}{c_{0}E_{0}+m\mp ic_{1}k}, and ζ2=(c0E0+m)2+c12k22E0(E0c2k+c0m)\zeta^{2}=\dfrac{(c_{0}E_{0}+m)^{2}+c_{1}^{2}k^{2}}{2E_{0}(E_{0}-c_{2}k+c_{0}m)}. This is a compact form of Eq. 44, and gives the correct solution for the standard representation (SR), i.e. Eq. 45 as the SR limit is taken. The other approach to get the above result is going to the moving reference (in which the particle is at rest, i.e. k=0k=0), see Eqs. 46 and 47, which leads consistently to a same result as Eq. 10 after the appropriate boost.
In the moving reference the eigenstates have to be simultaneously the eigenstates of the spin operator SzS_{z}, through which its shape can be found in this general representation. By requiring that Szψ±=±12ψ±S_{z}\psi_{\pm}=\pm\frac{1}{2}\psi_{\pm}, it is not hard to find out that Sz=12γ0=12γ¯0S_{z}=\dfrac{1}{2}\gamma^{0}=\dfrac{1}{2}\bar{\gamma}^{0}. By going to the SR limit, one easily finds that SyS_{y} has no chance but following the relation Sy=12γ¯1S_{y}=\frac{1}{2}\bar{\gamma}^{1}. Then using the fundamental commutation relation [Si,Sj]=iϵijkSk\left[S_{i},S_{j}\right]=i\epsilon_{ijk}S_{k}, we can find SxS_{x} as Eq. 48, which casts to

Sx12γ¯2=12(c2a2ib2a2+ib2c2);S_{x}\equiv\frac{1}{2}\bar{\gamma}^{2}=\frac{1}{2}\begin{pmatrix}c_{2}&a_{2}-ib_{2}\\ a_{2}+ib_{2}&-c_{2}\end{pmatrix}; (11)

with the following definitions

{a2=c0b1b0c1b2=a0c1c0a1c2=b0a1a0b1,\begin{split}\begin{cases}a_{2}=c_{0}b_{1}-b_{0}c_{1}\\ b_{2}=a_{0}c_{1}-c_{0}a_{1}\\ c_{2}=b_{0}a_{1}-a_{0}b_{1}\end{cases},\end{split} (12)

using of which one can easily show that

{iγ¯0γ¯1γ¯2=iγ0γ1γ2=Iγ¯μγ¯ν=iϵμνθγ¯θ.\begin{cases}i\bar{\gamma}^{0}\bar{\gamma}^{1}\bar{\gamma}^{2}=-i\gamma^{0}\gamma^{1}\gamma^{2}=I\\ \bar{\gamma}^{\mu}\bar{\gamma}^{\nu}=-i\epsilon^{\mu\nu\theta}\bar{\gamma}^{\theta}.\end{cases} (13)

where ϵμνθ\epsilon^{\mu\nu\theta} is totally antisymmetric symbol, and μ,ν,θ=0,1,2\mu,\nu,\theta=0,1,2. It should be noted that it is Hermitian and traceless, and satisfies the following relations

{Sx,γ¯0}={Sx,γ¯1}=0,Sx2=14I\left\{S_{x},\bar{\gamma}^{0}\right\}=\left\{S_{x},\bar{\gamma}^{1}\right\}=0\ ,\ S_{x}^{2}=\dfrac{1}{4}I (14)

These all can be easily generalized to 2+12+1-dimensional space-time, using the same γ¯\bar{\gamma}s.

The other important question is concerning the spin representation of the particles. Based on the above-mentioned generalizations, we find that the general form of the spin operator is Sμ=12γ¯μS_{\mu}=\dfrac{1}{2}\bar{\gamma}^{\mu} with the following eigenstates

|Sμ±=12(1±cμ)(1±cμaμ+ibμ,)\left|S_{\mu}\pm\right\rangle=\dfrac{1}{\sqrt{2(1\pm c_{\mu})}}\begin{pmatrix}1\pm c_{\mu}\\ a_{\mu}+ib_{\mu},\end{pmatrix} (15)

where μ=0\mu=0, μ=1\mu=1, and μ=2\mu=2 represent SzS_{z}, SyS_{y} and SxS_{x} respectively. This also helps to find the helicity operator for the rest frame (k0k\neq 0), i.e. (h=Sp/p)(h=\textbf{S}\cdot\textbf{p}/\mid\textbf{p}\mid), which is h=Szh=S_{z} when the particle moves in the zz direction. Consequently, the right-hand and the left-hand side wave functions are ++ and - eigenstates of SzS_{z} respectively.

One can easily prove that three independent matrices at most can be constructed for the case d=1d=1, i.e. two dimensional γ\gamma matrices.

Before finishing this section, let us summarize the relationships between the elements

aμaν+bμbν+cμcν=δμνaμaμ=bμbμ=cμcμ=1aμbμ=aμcμ=bμcμ=0\begin{split}a_{\mu}a_{\nu}+b_{\mu}b_{\nu}+c_{\mu}c_{\nu}=\delta_{\mu\nu}\\ a_{\mu}a_{\mu}=b_{\mu}b_{\mu}=c_{\mu}c_{\mu}=1\\ a_{\mu}b_{\mu}=a_{\mu}c_{\mu}=b_{\mu}c_{\mu}=0\end{split} (16)

which is eqivalent to

aμ=bνcθϵμνθ,bμ=cνaθϵμνθ,cμ=aνbθϵμνθ\begin{split}&a_{\mu}=-b_{\nu}c_{\theta}\epsilon_{\mu\nu\theta},\\ &b_{\mu}=-c_{\nu}a_{\theta}\epsilon_{\mu\nu\theta},\\ &c_{\mu}=-a_{\nu}b_{\theta}\epsilon_{\mu\nu\theta}\end{split} (17)

where (μ,ν,θ=0,1,2)(\mu,\nu,\theta=0,1,2), and Einstein summation rule was used.
In the next section we re-shape the above equations in a single clean form, which is the main achievement of the present paper.

II.2 Boost of IFSs

A crucial question for any fermion that is governed by the Dirac equation is its behavior under the boost. Let us denote the space-time Lorentz transformation as xμ=Λμνxνx^{\prime}_{\mu}=\Lambda_{\mu\nu}x_{\nu}, then the wave functions transform as ψ(x)=S(Λ)ψ(x)\psi^{\prime}(x^{\prime})=S(\Lambda)\psi(x), where S(Λ)S(\Lambda) is a representation of the Lorentz transformation. Here the prime means inertial system OO^{\prime} that moves with velocity v=tanhθβv=\tanh\theta\equiv\beta relative to the system OO. Therefore one can easily verify that Λ00=Λ11=coshθ\Lambda_{0}^{0}=\Lambda_{1}^{1}=\cosh\theta and Λ01=Λ10=sinhθ\Lambda_{0}^{1}=\Lambda_{1}^{0}=\sinh\theta. In the 1+1-dimensional system we have just one boost direction, so that

Em=kmβ=sinhθβ=coshθΓ\frac{E}{m}=\frac{k}{m\beta}=\frac{\sinh\theta}{\beta}=\cosh\theta\equiv\varGamma (18)

so that coshθ2=Γ+12=E0+m2m\cosh\frac{\theta}{2}=\sqrt{\frac{\varGamma+1}{2}}=\sqrt{\frac{E_{0}+m}{2m}} and sinhθ2=Γ12=E0m2m\sinh\frac{\theta}{2}=\sqrt{\frac{\varGamma-1}{2}}=\sqrt{\frac{E_{0}-m}{2m}}. In analogy with the boost of standard fermions, we examine the representation

S=exp(γ0γ1θ/2)=exp(iγ¯0γ¯1θ/2)=exp(γ¯2θ/2)=Icoshθ/2+γ¯2sinhθ/2=E0+m2m+γ¯2E0m2m\begin{split}S&=\exp(\gamma^{0}\gamma^{1}\theta/2)=\exp(-i\bar{\gamma}^{0}\bar{\gamma}^{1}\theta/2)=\exp(-\bar{\gamma}^{2}\theta/2)\\ &=I\cosh\theta/2+\bar{\gamma}^{2}\sinh\theta/2=\sqrt{\frac{E_{0}+m}{2m}}+\bar{\gamma}^{2}\sqrt{\frac{E_{0}-m}{2m}}\end{split} (19)

which gives us the final result for the boost of the IFSs

SIFS=(E0+m2m+c2E0m2m(a2ib2)E0m2m(a2+ib2)E0m2mE0+m2mc2E0m2m)S_{\text{IFS}}=\begin{pmatrix}\sqrt{\frac{E_{0}+m}{2m}}+c_{2}\sqrt{\frac{E_{0}-m}{2m}}&(a_{2}-ib_{2})\sqrt{\frac{E_{0}-m}{2m}}\\ (a_{2}+ib_{2})\sqrt{\frac{E_{0}-m}{2m}}&\sqrt{\frac{E_{0}+m}{2m}}-c_{2}\sqrt{\frac{E_{0}-m}{2m}}\end{pmatrix} (20)

The general Dirac equation can be obtained using the above formula for the boost of IFS, see Appendix B for the details. To see if this formulation works, let us boost the solution in the rest reference (ψ0IFS\psi_{0\text{IFS}}), for which we use Eq. 10. The result is abbreviated as follows (see Eq. 47)

ψIFS±m(k=0)=12(1+c0)(ψ1±ψ2±).\psi_{\text{IFS}}^{\pm m}(k=0)=\dfrac{1}{\sqrt{2(1+c_{0})}}\begin{pmatrix}\psi_{1}^{\pm}\\ \psi_{2}^{\pm}\end{pmatrix}. (21)

where ψ1+=ψ2=1+c0\psi_{1}^{+}=\psi_{2}^{-}=1+c_{0}, ψ2+=a0+ib0\psi_{2}^{+}=a_{0}+ib_{0}, and ψ1=a0+ib0\psi_{1}^{-}=-a_{0}+ib_{0}. It should be taken into account that ψIFS±(k)=S(Λ)ψIFS±m(k=0)\psi_{\text{IFS}}^{\pm}(k)=S(\Lambda)\psi_{\text{IFS}}^{\pm m}(k=0) which is exactly the Eq. 10. Now let us find a matrix which satisfies DSIFSψIFS+m(k=0)=0DS_{\text{IFS}}\psi_{\text{IFS}}^{+m}(k=0)=0 which is the Dirac equation. To this end, we notice that

SIFSψIFS+m(k=0)=ϱ(c0E0+mic1k(a0+ib0)E0i(a1+ib1k))S_{\text{IFS}}\psi_{\text{IFS}}^{+m}(k=0)=\varrho\begin{pmatrix}c_{0}E_{0}+m-ic_{1}k\\ (a_{0}+ib_{0})E_{0}-i(a_{1}+ib_{1}k)\end{pmatrix} (22)

where

ϱ=[(1+c0)(E0+m)(c2ic1)k]E02m(1+c0)(E0+m)(E0+c0mc2k).\varrho=\dfrac{\left[(1+c_{0})(E_{0}+m)-(c_{2}-ic_{1})k\right]\sqrt{E_{0}}}{\sqrt{2m(1+c_{0})(E_{0}+m)(E_{0}+c_{0}m-c_{2}k)}}. (23)

Then, by requiring that detD=E2k2m2\det D=E^{2}-k^{2}-m^{2} one readily finds D=DIFSD=D_{\text{IFS}}. This shows that one can reach to the wave function in general frame by a boost from the rest frame.

III SO(3)SO(3) symmetry in the representation of γ\gamma matrices

In this section we aim to find the structure of the parameters that were obtained in the previous section, i.e. the relation between aμ,bμa_{\mu},b_{\mu} and cμc_{\mu}, μ=0,1,2\mu=0,1,2. To this end, let us put the parameters into a 3×33\times 3 matrix OO as follows.

O(a2a1a0b2b1b0c2c1c0)O\equiv\begin{pmatrix}a_{2}&a_{1}&a_{0}\\ b_{2}&b_{1}&b_{0}\\ c_{2}&c_{1}&c_{0}\end{pmatrix} (24)

Note that OS=IO_{S}=I. At the first glance, it may seem ad hoc, but as will become clear soon, it helps much to view the transformation between IFS (representations of the Dirac equation) as a matrix operation. The interesting fact is that the conditions depicted in Eq. LABEL:Eq:GeneralRelations can actually be written in the form

OOT=OTO=I,OO^{T}=O^{T}O=I, (25)

i.e. the matrix OO is orthogonal and reversible. As a result, the matrix OO is a member of SO(3)SO(3) group, so that various IFSs can be reached via rotation in this space. Let us show a rotation matrix with the angle φ\varphi around the unit vector n^=nxi^+nyj^+nzk^\hat{n}=n_{x}\hat{i}+n_{y}\hat{j}+n_{z}\hat{k} as follows:

Rn(φ)=eiJn^φ=(cosφ+nx2(1cosφ)nxny(1cosφ)nzsinφnxnz(1cosφ)+nysinφnynx(1cosφ)+nzsinφcosφ+ny2(1cosφ)nynz(1cosφ)nxsinφnznx(1cosφ)nysinφnzny(1cosφ)+nxsinφcosφ+nz2(1cosφ))R_{n}(\varphi)=e^{-i\textbf{J}\cdotp\hat{n}\varphi}=\begin{pmatrix}\cos\varphi+n_{x}^{2}(1-\cos\varphi)\ \ &n_{x}n_{y}(1-\cos\varphi)-n_{z}\sin\varphi\ \ &n_{x}n_{z}(1-\cos\varphi)+n_{y}\sin\varphi\\ n_{y}n_{x}(1-\cos\varphi)+n_{z}\sin\varphi\ \ &\cos\varphi+n_{y}^{2}(1-\cos\varphi)\ \ &n_{y}n_{z}(1-\cos\varphi)-n_{x}\sin\varphi\\ n_{z}n_{x}(1-\cos\varphi)-n_{y}\sin\varphi\ \ &n_{z}n_{y}(1-\cos\varphi)+n_{x}\sin\varphi\ \ &\cos\varphi+n_{z}^{2}(1-\cos\varphi)\end{pmatrix} (26)

so that Jn^=idRn(φ)dφ|φ=0\textbf{J}\cdotp\hat{n}=i\dfrac{dR_{n}(\varphi)}{d\varphi}|_{\varphi=0}. By matching elements of OO matrix with Rn(φ)R_{n}(\varphi) we obtain

2cosφ=a2+b1+c01\displaystyle 2\cos\varphi=a_{2}+b_{1}+c_{0}-1\ \ \ \ \ \

in such a way that if a2=b1=c0=1a_{2}=b_{1}=c_{0}=1 then φ=0\varphi=0, and if the other parameters are set to zero, then Rn(φ)=IR_{n}(\varphi)=I as expected. In general

{2nzsin(φ)=b2a12nysin(φ)=a0c22nxsin(φ)=c1b0\begin{cases}2n_{z}\sin(\varphi)=b_{2}-a_{1}\\ 2n_{y}\sin(\varphi)=a_{0}-c_{2}\\ 2n_{x}\sin(\varphi)=c_{1}-b_{0}\end{cases} (27)

and also

{nxny=b2+a13a2b1c0nxnz=a0+c23a2b1c0nynz=c1+b03a2b1c0.\begin{cases}n_{x}n_{y}=\dfrac{b_{2}+a_{1}}{3-a_{2}-b_{1}-c_{0}}\\ n_{x}n_{z}=\dfrac{a_{0}+c_{2}}{3-a_{2}-b_{1}-c_{0}}\\ n_{y}n_{z}=\dfrac{c_{1}+b_{0}}{3-a_{2}-b_{1}-c_{0}}.\end{cases} (28)

The above equations give us the full correspondence between the space of representation of the Dirac equation (shown by OO matrices) and the general representation of SO(3)SO(3) group. Using the correspondence between SO(3)SO(3) and SU(2) groups, one can associate the representation of the IFSs with SU(2) group. We make this correspondence using the (Sμ)(S_{\mu}) that we found in the previous section as the generators of SU(2)SU(2). More precisely, let us define

U=eiSn^φ=(cosφ/2inμcμsinφ/2nμ(bμ+iaμ)sinφ/2nμ(bμiaμ)sinφ/2cosφ/2+inμcμsinφ/2)\begin{split}U&=e^{-i\textbf{S}\cdot\hat{n}\varphi}\\ &=\begin{pmatrix}\cos\varphi/2-in_{\mu}c_{\mu}\sin\varphi/2&-n_{\mu}(b_{\mu}+ia_{\mu})\sin\varphi/2\\ n_{\mu}(b_{\mu}-ia_{\mu})\sin\varphi/2&\cos\varphi/2+in_{\mu}c_{\mu}\sin\varphi/2\end{pmatrix}\end{split} (29)

where det(U)=1\det(U)=1. An example is nμ=aμn_{\mu}=a_{\mu} for which U=(cosφ/2isinφ/2isinφ/2cosφ/2)U=\begin{pmatrix}\cos\varphi/2&-i\sin\varphi/2\\ -i\sin\varphi/2&\cos\varphi/2\end{pmatrix}. Generally if we define

M=(cνxν(aνibν)xν(aν+ibν)xνcνxν)M=\begin{pmatrix}c_{\nu}x_{\nu}&(a_{\nu}-ib_{\nu})x_{\nu}\\ (a_{\nu}+ib_{\nu})x_{\nu}&-c_{\nu}x_{\nu}\end{pmatrix} (30)

then the transformed matrix M=UMU=(cνxν(aνibν)xν(aν+ibν)xνcνxν)M^{\prime}=UMU^{{\dagger}}=\begin{pmatrix}c_{\nu}x^{\prime}_{\nu}&(a_{\nu}-ib_{\nu})x^{\prime}_{\nu}\\ (a_{\nu}+ib_{\nu})x^{\prime}_{\nu}&-c_{\nu}x^{\prime}_{\nu}\end{pmatrix} is such that (xyz)=(xyz)Rn(φ)\begin{pmatrix}x^{\prime}&y^{\prime}&z^{\prime}\end{pmatrix}=\begin{pmatrix}x&y&z\end{pmatrix}R_{n}^{\top}(\varphi).

IV sub-representations of IFS

By “sub-representation”, we mean restricted γ\gamma representations. For instance, let us consider the Majorana representation for which fermions and antifermions are the same, limiting strongly the range of the entities of γ\gamma matrices. Fermions (ψ\psi) and antifermions (ψc\psi_{c} obtained by charge conjugation) satisfy the Dirac equation in the presence of electromagnetic field (AμA_{\mu})

{[γμ(pμeAμ)m]ψ=0[γμ(pμ+eAμ)m]ψc=0\begin{cases}\left[\gamma^{\mu}(p_{\mu}-eA_{\mu})-m\right]\psi=0\\ \left[\gamma^{\mu}(p_{\mu}+eA_{\mu})-m\right]\psi_{c}=0\end{cases} (31)

If there is a transformation UU such that U(γμ)U1=γμU(\gamma^{\mu})^{*}U^{-1}=-\gamma^{\mu}, then one can show by inspection that UψU\psi^{*} is a solution of the second equation, giving us no chance but ψc=eiαUψ\psi_{c}=e^{i\alpha}U\psi^{*} where α\alpha is an arbitrary phase. These fermions are Majorana, in which, for the simple case U=IU=I (identity matrix), the wave function of fermions and antifermions are the same. Without loss of generality, we set U=IU=I in this paper (in other cases we always can transform γ\gamma so that it applies). Let us call the γ\gamma matrices that satisfy this condition constitute the general Majorana representation, which are

γM±0=±i(0110),γM±1=i(c1±1c12±1c12c1).\gamma^{0}_{\text{M}\pm}=\pm i\begin{pmatrix}0&-1\\ 1&0\end{pmatrix}\ ,\ \gamma^{1}_{\text{M}\pm}=i\begin{pmatrix}c_{1}&\pm\sqrt{1-c_{1}^{2}}\\ \pm\sqrt{1-c_{1}^{2}}&-c_{1}\end{pmatrix}. (32)

were “M” stands for “Majorana”, and the “±\pm” for γM0\gamma^{0}_{\text{M}} is independent of “±\pm” in γM1\gamma^{1}_{\text{M}}, so that we have four possibilities for selecting γ\gammas, i.e. (γM+0,γM+1)(\gamma_{\text{M}+}^{0},\gamma_{\text{M}+}^{1}), (γM+0,γM1)(\gamma_{\text{M}+}^{0},\gamma_{\text{M}-}^{1}), etc. Then, for example for the (+,+)(+,+) case, using these matrices the eigenstates are calculate to be

ψη,ηM(E,k)=A0(mic1ki(ηEη1c12k))\psi_{\eta,\eta^{\prime}}^{\text{M}}(E,k)=A_{0}\begin{pmatrix}m-ic_{1}k\\ i\left(\eta E-\eta^{\prime}\sqrt{1-c_{1}^{2}}k\right)\end{pmatrix} (33)

where A0=12E(Eηη1c12k)A_{0}=\dfrac{1}{\sqrt{2E(E-\eta\eta^{\prime}\sqrt{1-c_{1}^{2}}k)}} and η,η=±\eta,\eta^{\prime}=\pm refer to the signs of γ0\gamma^{0} and γ1\gamma^{1} respectively.

Refer to caption
(a)
Refer to caption
(b)
Figure 1: Three-dimensional representation of the rotation of Eq. 36 from two views in terms of one parameter c1c_{1} for η=η=1\eta=\eta^{\prime}=1. The start point (x0,y0,z0)(x_{0},y_{0},z_{0}) is represented by red circle.

For Weyl-Majorana (WM) fermions, one sets m=0m=0, for which the γ\gamma matrices have to satisfy γμ=±γμ{\gamma^{\mu}}^{*}=\pm\gamma^{\mu} (note that “-” sign is also permissible), and ΨWMΨWM\Psi_{\text{WM}}^{*}\propto\Psi_{\text{WM}}. If (γμ)=γμ\left(\gamma^{\mu}\right)^{*}=-\gamma^{\mu} we get previous representation just above, but for the opposite case we have

γWM±0=(c0±1c02±1c02c0),γWM±1=(0±110)\begin{split}\gamma^{0}_{\text{WM}\pm}&=\begin{pmatrix}c_{0}&\pm\sqrt{1-c_{0}^{2}}\\ \pm\sqrt{1-c_{0}^{2}}&-c_{0}\end{pmatrix}\ ,\ \gamma^{1}_{\text{WM}\pm}=\begin{pmatrix}0&\pm 1\\ \mp 1&0\end{pmatrix}\end{split} (34)

for which the eigenstates are

ψη,ηWM=12(1ηη1c02)(c0η1c02η)\psi^{\text{WM}}_{\eta,\eta^{\prime}}=\dfrac{1}{\sqrt{2(1-\eta\eta^{\prime}\sqrt{1-c_{0}^{2}})}}\begin{pmatrix}c_{0}\\ \eta\sqrt{1-c_{0}^{2}}-\eta^{\prime}\end{pmatrix} (35)

Note that in the above equation EE and kk cancel out, so that its form is simpler than Eq. 33.

It is worth mentioning that we can reach the Majorana representation starting from the Dirac equation by a rotation. To be more precise, if we “rotate” the γ\gamma matrices in the Dirac representation about n^\hat{n} by the angle φ\varphi, which are

nx=±η1ηc13+ηc1,ny=±1+ηc13+ηc1,nz=±η1+ηc13+ηc1cosφ=12(1+ηc1),sinφ=ηη232ηc1c12\begin{split}&n_{x}=\pm\eta\sqrt{\frac{1-\eta^{\prime}c_{1}}{3+\eta^{\prime}c_{1}}},n_{y}=\pm\sqrt{\frac{1+\eta^{\prime}c_{1}}{3+\eta^{\prime}c_{1}}},n_{z}=\pm\eta^{\prime}\sqrt{\frac{1+\eta^{\prime}c_{1}}{3+\eta^{\prime}c_{1}}}\\ &\cos\varphi=\frac{-1}{2}(1+\eta^{\prime}c_{1}),\sin\varphi=\frac{\mp\eta\eta^{\prime}}{2}\sqrt{3-2\eta^{\prime}c_{1}-c_{1}^{2}}\end{split} (36)

then we reach the γ\gamma matrices given in Eq. 32. We notice here that all quantities depend on a single parameter, i.e. c1c_{1}, showing that the transformation is isomorphism to U(1)U(1) group, which is a subgroup of SU(2)SU(2), which itself is homomorphism to SO(3)SO(3). In the Fig. 1 we show a three-dimensional representation of this transformation. In this figure a starting point (x0,y0,z0)(x_{0},y_{0},z_{0}) (which is represented by a red bold circle) under the effect of Eq. 36 and Eq. 26. This can also be done easily for WM fermions.

V Transformation of Dirac to generalized particles

In this section we find the general transformations TT using of which IFSs are obtained from standard (Dirac) representation. For the definition of TT see Eq. 2. Using the calculations presented in Appendix C one can show that

T=ϱ(1+q1q2eiα(q1q2eiα)q1q2eiα1+q1q2eiα)T=\varrho^{\prime}\begin{pmatrix}1+q_{1}q_{2}e^{i\alpha}&-(q_{1}-q_{2}e^{-i\alpha})\\ q_{1}-q_{2}e^{i\alpha}&1+q_{1}q_{2}e^{-i\alpha}\end{pmatrix} (37)

where ϱ12E01(E0+m)(E0+c0m+c2k)\varrho^{\prime}\equiv\frac{1}{2}E_{0}^{-1}\sqrt{\left(E_{0}+m\right)\left(E_{0}+c_{0}m+c_{2}k\right)} and α=tan1b0m+b2ka0m+a2k\alpha=\tan^{-1}\frac{b_{0}m+b_{2}k}{a_{0}m+a_{2}k}, and

q1=E0mE0+m,q2=E0c0mc2kE0+c0m+c2kq_{1}=\sqrt{\dfrac{E_{0}-m}{E_{0}+m}}\ ,\ q_{2}=\sqrt{\dfrac{E_{0}-c_{0}m-c_{2}k}{E_{0}+c_{0}m+c_{2}k}} (38)

One can easily show that TT=TT=ITT^{{\dagger}}=T^{{\dagger}}T=I and detT=1\det T=1, showing that they are unitary transformations. These matrices are also represented by T=eiσn^φ/2T=e^{-i\sigma\cdot\hat{n}\varphi/2}, were φ\varphi is a real parameter, rotating ΨS\Psi_{{}_{S}} to ΨIFS\Psi_{{}_{IFS}}, i.e. TψS=ψIFST\psi_{\text{S}}=\psi_{\text{IFS}}. Using this notation, one can easily find the rotation parameters, represented by n^=(nx,ny,nz)\hat{n}=(n_{x},n_{y},n_{z}), satisfying the following identities

nxsinφ/2=12sinα(1+mE)(1c0mEc2kE)nysinφ/2=12[(1mE)(1+c0mE+c2kE)cosα(1+mE)(1c0mEc2kE)]nzsinφ/2=12sinα(1mE)(1c0mEc2kE)cosφ/2=12[(1+mE)(1+c0mE+c2kE)+cosα(1mE)(1c0mEc2kE)]\begin{split}&n_{x}\sin\varphi/2=\frac{1}{2}\sin\alpha\sqrt{\left(1+\frac{m}{E}\right)\left(1-c_{0}\frac{m}{E}-c_{2}\frac{k}{E}\right)}\\ &n_{y}\sin\varphi/2=\frac{1}{2}\left[\sqrt{\left(1-\frac{m}{E}\right)\left(1+c_{0}\frac{m}{E}+c_{2}\frac{k}{E}\right)}-\cos\alpha\sqrt{\left(1+\frac{m}{E}\right)\left(1-c_{0}\frac{m}{E}-c_{2}\frac{k}{E}\right)}\right]\\ &n_{z}\sin\varphi/2=-\frac{1}{2}\sin\alpha\sqrt{\left(1-\frac{m}{E}\right)\left(1-c_{0}\frac{m}{E}-c_{2}\frac{k}{E}\right)}\\ &\cos\varphi/2=\frac{1}{2}\left[\sqrt{\left(1+\frac{m}{E}\right)\left(1+c_{0}\frac{m}{E}+c_{2}\frac{k}{E}\right)}+\cos\alpha\sqrt{\left(1-\frac{m}{E}\right)\left(1-c_{0}\frac{m}{E}-c_{2}\frac{k}{E}\right)}\right]\end{split} (39)

from which one can show nx2+ny2+nz2=1n_{x}^{2}+n_{y}^{2}+n_{z}^{2}=1. As an example, let us consider the transformation TS-MT_{\text{S-M}} that converts a standard Dirac particle to a Majorana particle

TS-M=ϱ′′(1+q1q2eiα(q1q2eiα)q1q2eiα1+q1q2eiα)T_{\text{S-M}}=\varrho^{\prime\prime}\begin{pmatrix}1+q^{\prime}_{1}q^{\prime}_{2}e^{i\alpha^{\prime}}&-(q^{\prime}_{1}-q^{\prime}_{2}e^{-i\alpha^{\prime}})\\ q^{\prime}_{1}-q^{\prime}_{2}e^{i\alpha^{\prime}}&1+q^{\prime}_{1}q^{\prime}_{2}e^{-i\alpha^{\prime}}\end{pmatrix} (40)

were ϱ′′=12E01(E0+m)(E0+c2k)\varrho^{\prime\prime}=\frac{1}{2}E_{0}^{-1}\sqrt{(E_{0}+m)(E_{0}+c_{2}k)}, α=tan1(m/c1)\alpha^{\prime}=\tan^{-1}(-m/c_{1}), q1=q1q^{\prime}_{1}=q_{1}, and q2=E0c2kE0+c2kq^{\prime}_{2}=\sqrt{\dfrac{E_{0}-c_{2}k}{E_{0}+c_{2}k}}.

VI Conclusion

In this paper we considered a general form for the γ\gamma matrices. Motivated by the fact that the resulting fermions are “intermediate” in the sense of normal representation (i.e. standard representation, Weyl representation, etc.) we call them the “intermediate fermion species” (IFS). Many properties of the IFS were calculated and explored, like the eigenvalue problem, boost and rotation, and transformation between species. We observed that the latter (the transformation between species) corresponds to SO(3)SO(3) rotations in the space of the parameters of the problem (the entities of the γ\gamma matrices). Therefore any arbitrary representation of spin 12\frac{1}{2} fermions is obtained by a SO(3)SO(3) rotation in the parameters of the γ\gamma matrices. Based on this, we calculated the sub-representations which admits the Majorana fermions. Importantly, we clearly established that any IFS can be obtained from the Dirac spinors by a SU(2) similarity transformation.

It is worth mentioning that this transformation does not change the transport properties of particles. For instance, we measured the transport parameters of the Klein tunneling, and noticed that none of these parameters (reflection and transmission coefficients) change under the mentioned SO(3)SO(3) transformation in normal incidence. This motivated us to call it “the symmetry” of the Dirac equation.

According to the Noether theorem this symmetries leads to some conservations between IFS particles. In our future research we intend to concentrate on this topic, and also finding the other aspects of this transformation, such as Andreev reflection, to see if we can design an experiment which distinguish between these particles.

Appendix A The properties of γ¯\bar{\gamma} matrices

For interested readers a detailed description of first subsection in section two is presented in this Appendix. In 1+11+1, γ¯\bar{\gamma} matrices should be of the following form

γ¯μ=(cμaμibμaμ+ibμcμ);aμ,bμ,cμ\bar{\gamma}^{\mu}=\begin{pmatrix}c_{\mu}&a_{\mu}-ib_{\mu}\\ a_{\mu}+ib_{\mu}&-c_{\mu}\end{pmatrix}\ ;\ a_{\mu},b_{\mu},c_{\mu}\in\Re (41)

for which μ,ν=0,1\mu,\nu=0,1. Using the anticommutation relation of γ¯\bar{\gamma} matrices (Eq. 3), one can generally show that

aμaν+bμbν+cμcν=δμνa_{\mu}a_{\nu}+b_{\mu}b_{\nu}+c_{\mu}c_{\nu}=\delta_{\mu\nu} (42)

In the 1+1 dimension, according to the what had been said, we only need γ0\gamma^{0} and γ1\gamma^{1}

(γ0Eγ1kmI)ψ(E,k)ei(kxEt)=(γ¯0Eiγ¯1kmI)ψ(E,k)ei(kxEt)=0\begin{split}&\left(\gamma^{0}E-\gamma^{1}k-mI\right)\psi(E,k)\ e^{i(kx-Et)}=\\ &\left(\bar{\gamma}^{0}E-i\bar{\gamma}^{1}k-mI\right)\psi(E,k)\ e^{i(kx-Et)}=0\end{split} (43)

therefore, the non-differential form of intermediate fermionic species (IFS) Dirac equation will be Eq. 8. We have non-trivial answer, if E2=k2+m2E=±E0=±k2+m2E^{2}=k^{2}+m^{2}\Rightarrow E=\pm E_{0}=\pm\sqrt{k^{2}+m^{2}} , which we also expect it before. General shape of free particle’s wave function in 1+11+1 dimensions is:

{ψIFS+E0=ζ(1i(b0E0a1k)+(a0E0+b1k)c0E0+mic1k)ψIFSE0=ζ(i(b0E0a1k)(a0E0+b1k)c0E0+m+ic1k1)\begin{cases}\psi_{\text{IFS}}^{+E_{0}}=\zeta\begin{pmatrix}1\\ \dfrac{i(b_{0}E_{0}-a_{1}k)+(a_{0}E_{0}+b_{1}k)}{c_{0}E_{0}+m-ic_{1}k}\end{pmatrix}\\ \\ \psi_{\text{IFS}}^{-E_{0}}=\zeta\begin{pmatrix}\dfrac{i(b_{0}E_{0}-a_{1}k)-(a_{0}E_{0}+b_{1}k)}{c_{0}E_{0}+m+ic_{1}k}\\ 1\end{pmatrix}\end{cases} (44)

where ζ=(c0E0+m)2+c12k22E0(E0c2k+c0m)\zeta=\sqrt{\dfrac{(c_{0}E_{0}+m)^{2}+c_{1}^{2}k^{2}}{2E_{0}(E_{0}-c_{2}k+c_{0}m)}}. For Standard Representation limit (c0=b1=a2=1,etc=0)(c_{0}=b_{1}=a_{2}=1\ ,\ etc=0) they Turns into:

{ψS+=E0+m2E0(1+kE+m)ψS=E0+m2E0(kE+m1)\begin{cases}\psi_{{}_{S+}}=\sqrt{\frac{E_{0}+m}{2E_{0}}}\begin{pmatrix}1\\ \frac{+k}{E+m}\end{pmatrix}\\ \psi_{{}_{S-}}=\sqrt{\frac{E_{0}+m}{2E_{0}}}\begin{pmatrix}\frac{-k}{E+m}\\ 1\end{pmatrix}\end{cases} (45)

The non-differential Dirac equation in k=0k=0 reduces as follows:

(c0Em(a0ib0)E(a0+ib0)Ec0Em)ψ0IFS=0E=±m\begin{pmatrix}c_{0}E-m&(a_{0}-ib_{0})E\\ (a_{0}+ib_{0})E&-c_{0}E-m\end{pmatrix}\psi_{{}_{0IFS}}=0\Rightarrow E=\pm m (46)
{E+=+mψ0IFS+=1+c02(1a0+ib01+c0)E=mψ0IFS=1+c02((a0ib01+c0)1).\Rightarrow\begin{cases}E_{+}=+m\Rightarrow\psi_{{}_{0IFS+}}=\sqrt{\dfrac{1+c_{0}}{2}}\begin{pmatrix}1\\ \dfrac{a_{0}+ib_{0}}{1+c_{0}}\end{pmatrix}\\ E_{-}=-m\Rightarrow\psi_{{}_{0IFS-}}=\sqrt{\dfrac{1+c_{0}}{2}}\begin{pmatrix}-(\dfrac{a_{0}-ib_{0}}{1+c_{0}})\\ 1\end{pmatrix}.\end{cases} (47)

We know the answers of Dirac equation in the case k=0, namely ψ±\psi_{\pm}, must also be the eigenstates of the operator SzS_{z}. So, we look for the matrix form of operator SzS_{z} such that Szψ±=±12ψ±S_{z}\psi_{\pm}=\pm\frac{1}{2}\psi_{\pm} are its eigenstates. Then Sz=12γ0=12γ¯0S_{z}=\dfrac{1}{2}\gamma^{0}=\dfrac{1}{2}\bar{\gamma}^{0}. By respecting that γ¯1\bar{\gamma}^{1} in the limit SR, becomes similar to SyS_{y}, we can assume that Sy=12γ¯1S_{y}=\frac{1}{2}\bar{\gamma}^{1}. Now, by using the fundamental commutation relation [Si,Sj]=iijkSk\left[S_{i},S_{j}\right]=i\in_{ijk}S_{k}, we can also obtain SxS_{x}:

2Sx=((b0a1a0b1)(c0b1b0c1)i(a0c1c0a1)(c0b1b0c1)+i(a0c1c0a1)(b0a1a0b1)).2S_{x}=\begin{pmatrix}(b_{0}a_{1}-a_{0}b_{1})&(c_{0}b_{1}-b_{0}c_{1})-i(a_{0}c_{1}-c_{0}a_{1})\\ (c_{0}b_{1}-b_{0}c_{1})+i(a_{0}c_{1}-c_{0}a_{1})&-(b_{0}a_{1}-a_{0}b_{1})\end{pmatrix}. (48)

It is clear that the matrix SxS_{x} is Hermitian and traceless. By performing the respected calculations, we can see that:

{Sx,γ¯0}={Sx,γ¯1}=0,Sxγ¯00Sxγ¯1,Sx2=14I\left\{S_{x},\bar{\gamma}^{0}\right\}=\left\{S_{x},\bar{\gamma}^{1}\right\}=0,S_{x}\bar{\gamma}^{0}\neq 0\neq S_{x}\bar{\gamma}^{1},S_{x}^{2}=\dfrac{1}{4}I (49)

According to (4) condition, it can be say that 2Sx2S_{x} has the all conditions of a γ¯\bar{\gamma} (Eq. 10)

Appendix B Obtaining Standard Dirac equation by Lorentz operator

In this Appendix we want to obtain the non-differential form of Dirac equation in Standard Representation by acting of Lorentz operator on wave function in rest framework i.e. ψS0+=(10),ψS0=(01)\psi_{{}_{S0+}}=\begin{pmatrix}1\\ 0\end{pmatrix},\psi_{{}_{S0-}}=\begin{pmatrix}0\\ 1\end{pmatrix}. Lorentz operator in (1+1) dimension in Standard Representation is

SS=(E0+m2mE0m2mE0m2mE0+m2m),S_{{}_{S}}=\begin{pmatrix}\sqrt{\frac{E_{0}+m}{2m}}&\sqrt{\frac{E_{0}-m}{2m}}\\ \sqrt{\frac{E_{0}-m}{2m}}&\sqrt{\frac{E_{0}+m}{2m}}\end{pmatrix}, (50)

then

SSψS0+=E0mψS+(E0+mk).S_{{}_{S}}\psi_{{}_{S0+}}=\sqrt{\frac{E_{0}}{m}}\psi_{{}_{S+}}\sim\begin{pmatrix}E_{0}+m\\ k\end{pmatrix}. (51)

If we introduce Dirac equation with DSD_{{}_{S}} then DSSSψS0+=DSψS+=0D_{{}_{S}}S_{{}_{S}}\psi_{{}_{S0+}}=D_{{}_{S}}\psi_{{}_{S+}}=0. By requiring that det(DS)=E02k2m2=0det(D_{{}_{S}})=E_{0}^{2}-k^{2}-m^{2}=0 one readily find

DS=(Emk+k(E+m))ψS=0D_{{}_{S}}=\begin{pmatrix}E-m&-k\\ +k&-(E+m)\end{pmatrix}\psi_{{}_{S}}=0 (52)

that is exactly the Dirac equation known in (1+1).

Appendix C Transformation matrix of various spinors

Standard Dirac Hamiltonian and Intermediate Fermionic Species Dirac Hamiltonian are

HS=(mkkm),HIFS=(c0m+c2k(a2ib2)k+(a0ib0)m(a2+ib2)k+(a0+ib0)m(c0m+c2k)).\begin{split}H_{{}_{S}}=\begin{pmatrix}m&k\\ k&-m\end{pmatrix}\qquad,\qquad H_{{}_{IFS}}=\begin{pmatrix}c_{0}m+c_{2}k&(a_{2}-ib_{2})k+(a_{0}-ib_{0})m\\ (a_{2}+ib_{2})k+(a_{0}+ib_{0})m&-(c_{0}m+c_{2}k)\end{pmatrix}\end{split}. (53)

We know that eigenvalues of both Hamiltonians are the same (±E0)(\pm E_{0}). Then in according to eigenstates of those, one can obtain nonsingular matrices ΘS\Theta_{{}_{S}} andΘIFS\Theta_{{}_{IFS}} as follows

ΘS=(E0+m2E0E0m2E0E0m2E0E0+m2E0)ΘIFS=(E0+c0m+c2k2E0eiαE0c0mc2k2E0eiαE0c0mc2k2E0E0+c0m+c2k2E0)\begin{split}\Theta_{{}_{S}}=\begin{pmatrix}\sqrt{\frac{E_{0}+m}{2E_{0}}}&-\sqrt{\frac{E_{0}-m}{2E_{0}}}\\ \sqrt{\frac{E_{0}-m}{2E_{0}}}&\sqrt{\frac{E_{0}+m}{2E_{0}}}\end{pmatrix}\qquad\qquad\\ \Theta_{{}_{IFS}}=\begin{pmatrix}\sqrt{\frac{E_{0}+c_{0}m+c_{2}k}{2E_{0}}}&-e^{-i\alpha}\sqrt{\frac{E_{0}-c_{0}m-c_{2}k}{2E_{0}}}\\ e^{i\alpha}\sqrt{\frac{E_{0}-c_{0}m-c_{2}k}{2E_{0}}}&\sqrt{\frac{E_{0}+c_{0}m+c_{2}k}{2E_{0}}}\end{pmatrix}\end{split} (54)

where α=tan1(b0m+b2ka0m+a2k)\alpha=tan^{-1}\left(\dfrac{b_{0}m+b_{2}k}{a_{0}m+a_{2}k}\right), then

ΘSHSΘS=ΘIFSHIFSΘIFS=(E000E0).\Theta_{{}_{S}}^{{\dagger}}H_{{}_{S}}\Theta_{{}_{S}}=\Theta_{{}_{IFS}}^{{\dagger}}H_{{}_{IFS}}\Theta_{{}_{IFS}}=\begin{pmatrix}E_{0}&0\\ 0&-E_{0}\end{pmatrix}. (55)

This suggest that ΘIFSΘSHSΘSΘIFS=HIFS\Theta_{{}_{IFS}}\Theta_{{}_{S}}^{{\dagger}}H_{{}_{S}}\Theta_{{}_{S}}\Theta_{{}_{IFS}}^{{\dagger}}=H_{{}_{IFS}}. With definition ΘSΘIFSϑ\Theta_{{}_{S}}\Theta_{{}_{IFS}}^{{\dagger}}\equiv\vartheta one can see that ϑψS=ψIFS\vartheta\psi_{{}_{S}}=\psi_{{}_{IFS}}, so that

ϑ=ϱ(1+q1q2eiα(q1q2eiα)q1q2eiα1+q1q2eiα)\vartheta=\varrho^{\prime}\begin{pmatrix}1+q_{1}q_{2}e^{i\alpha}&-(q_{1}-q_{2}e^{-i\alpha})\\ q_{1}-q_{2}e^{i\alpha}&1+q_{1}q_{2}e^{-i\alpha}\end{pmatrix} (56)

where

ϱ=(E0+m)(E0+c0m+c2k)2E0α=tan1(b0m+b2ka0m+a2k)q1=E0mE0+m,q2=E0c0mc2kE0+c0m+c2k\begin{split}\varrho^{\prime}=\dfrac{\sqrt{(E_{0}+m)(E_{0}+c_{0}m+c_{2}k)}}{2E_{0}}\quad\\ \alpha=tan^{-1}\left(\dfrac{b_{0}m+b_{2}k}{a_{0}m+a_{2}k}\right)\qquad\\ q_{1}=\sqrt{\dfrac{E_{0}-m}{E_{0}+m}}\ ,\ q_{2}=\sqrt{\dfrac{E_{0}-c_{0}m-c_{2}k}{E_{0}+c_{0}m+c_{2}k}}\end{split} (57)

References

  • Gasiorowicz (1966) S. Gasiorowicz, Elementary particle physics (Wiley New York, 1966).
  • Ioffe et al. (2010) B. L. Ioffe, V. S. Fadin,  and L. N. Lipatov, Quantum chromodynamics: Perturbative and nonperturbative aspects, 30 (Cambridge University Press, 2010).
  • Morandi et al. (2013) G. Morandi, P. Sodano, A. Tagliacozzo,  and V. Tognetti, Field Theories for Low-Dimensional Condensed Matter Systems: Spin Systems and Strongly Correlated Electrons, Vol. 131 (Springer Science & Business Media, 2013).
  • Sansoni et al. (2012) L. Sansoni, F. Sciarrino, G. Vallone, P. Mataloni, A. Crespi, R. Ramponi,  and R. Osellame, Physical review letters 108, 010502 (2012).
  • Chamon et al. (2010) C. Chamon, R. Jackiw, Y. Nishida, S.-Y. Pi,  and L. Santos, Physical Review B 81, 224515 (2010).
  • Leijnse and Flensberg (2012) M. Leijnse and K. Flensberg, Semiconductor Science and Technology 27, 124003 (2012).
  • Ryder (1996) L. H. Ryder, Quantum field theory (Cambridge university press, 1996).
  • Neto et al. (2009) A. C. Neto, F. Guinea, N. M. Peres, K. S. Novoselov,  and A. K. Geim, Reviews of modern physics 81, 109 (2009).
  • Sarma et al. (2011) S. D. Sarma, S. Adam, E. Hwang,  and E. Rossi, Reviews of modern physics 83, 407 (2011).
  • Peres (2010) N. Peres, Reviews of modern physics 82, 2673 (2010).
  • Nayak et al. (2008) C. Nayak, S. H. Simon, A. Stern, M. Freedman,  and S. D. Sarma, Reviews of Modern Physics 80, 1083 (2008).
  • Xu et al. (2015a) S.-Y. Xu, I. Belopolski, N. Alidoust, M. Neupane, G. Bian, C. Zhang, R. Sankar, G. Chang, Z. Yuan, C.-C. Lee, et al., Science 349, 613 (2015a).
  • Xu et al. (2015b) S.-Y. Xu, N. Alidoust, I. Belopolski, Z. Yuan, G. Bian, T.-R. Chang, H. Zheng, V. N. Strocov, D. S. Sanchez, G. Chang, et al., Nature Physics 11, 748 (2015b).
  • Xu et al. (2015c) S.-Y. Xu, I. Belopolski, D. S. Sanchez, C. Zhang, G. Chang, C. Guo, G. Bian, Z. Yuan, H. Lu, T.-R. Chang, et al., Science advances 1, e1501092 (2015c).
  • Armitage et al. (2018) N. Armitage, E. Mele,  and A. Vishwanath, Reviews of Modern Physics 90, 015001 (2018).
  • Moore (2010) J. E. Moore, Nature 464, 194 (2010).
  • Hasan and Moore (2010) M. Z. Hasan and J. E. Moore, arXiv preprint arXiv:1011.5462  (2010).
  • Kane and Mele (2005) C. L. Kane and E. J. Mele, Physical review letters 95, 146802 (2005).
  • Gu and Wen (2009) Z.-C. Gu and X.-G. Wen, Physical Review B 80, 155131 (2009).
  • Pollmann et al. (2012) F. Pollmann, E. Berg, A. M. Turner,  and M. Oshikawa, Physical review b 85, 075125 (2012).
  • Chen et al. (2011) X. Chen, Z.-C. Gu,  and X.-G. Wen, APS 2011, K1 (2011).
  • Elliott and Franz (2015) S. R. Elliott and M. Franz, Reviews of Modern Physics 87, 137 (2015).
  • Beenakker (2015) C. Beenakker, Reviews of Modern Physics 87, 1037 (2015).
  • Avignone III et al. (2008) F. T. Avignone III, S. R. Elliott,  and J. Engel, Reviews of Modern Physics 80, 481 (2008).
  • Kuno and Okada (2001) Y. Kuno and Y. Okada, Reviews of Modern Physics 73, 151 (2001).
  • Chiu et al. (2016) C.-K. Chiu, J. C. Teo, A. P. Schnyder,  and S. Ryu, Reviews of Modern Physics 88, 035005 (2016).