SciELO - Scientific Electronic Library Online

 
vol.66 número4Deviation of inverse square law based on Dunkl derivative: deformed Coulomb’s lawDynamics of neutrino wave packet in the Tachyon-like Dirac equation índice de autoresíndice de materiabúsqueda de artículos
Home Pagelista alfabética de revistas  

Servicios Personalizados

Revista

Articulo

Indicadores

Links relacionados

  • No hay artículos similaresSimilares en SciELO

Compartir


Revista mexicana de física

versión impresa ISSN 0035-001X

Rev. mex. fis. vol.66 no.4 México jul./ago. 2020  Epub 31-Ene-2022

https://doi.org/10.31349/revmexfis.66.418 

Research

Gravitation, Mathematical Physics and Field Theory

Double and dual numbers. SU(2) groups, two-component spinors and generating functions

G.F. Torres del Castilloa 

K.C. Gutiérrez-Herrerab 

aInstituto de Ciencias, Universidad Autónoma de Puebla, 72570 Puebla, Pue., México

bFacultad de Ciencias Físico Matemáticas, Universidad Autónoma de Puebla, 72570 Puebla, Pue., México.


Abstract

We explicitly show that the groups of 2×2 unitary matrices with determinant equal to 1 whose entries are double or dual numbers are homomorphic to ${\rm SO}(2,1)$ or to the group of rigid motions of the Euclidean plane, respectively, and we introduce the corresponding two-component spinors. We show that with the aid of the double numbers we can find generating functions for separable solutions of the Laplace equation in the (2+1) Minkowski space, which contain special functions that also appear in the solution of the Laplace equation in the three-dimensional Euclidean space, in spheroidal and toroidal coordinates.

Keywords: Double numbers; dual numbers; unitary groups; spinors; Minkowski (2 + 1) space; Laplace’s equation; spheroidal coordinates; toroidal coordinates

PACS: 02.30.Gp; 02.30.Jr

Resumen

Mostramos explícitamente que los grupos de matrices unitarias 2×2 con determinante igual a 1 cuyas entradas son números dobles o números duales son homomorfos a ${\rm SO}(2,1)$ o al grupo de movimientos rígidos del plano euclideano, respectivamente, e introducimos los espinores de dos componentes correspondientes. Mostramos que con la ayuda de los números dobles podemos hallar funciones generatrices para soluciones separables de la ecuación de Laplace en el espacio de Minkowski (2+1), las cuales contienen funciones especiales que también aparecen en la solución de la ecuación de Laplace en el espacio euclideano tridimensional, en coordenadas esferoidales y toroidales.

Descriptores: Números dobles; números duales; grupos unitarios; espinores; espacio de Minkowski (2 + 1); ecuación de Laplace; coordenadas esferoidales; coordenadas toroidales

1.Introduction

Complex numbers are frequently employed in physics even in areas where the quantities of physical interest are real. For instance, the standard spherical harmonics, Ylm, are complex-valued functions that are commonly used in electromagnetism, in spite of the fact that the potentials and electromagnetic fields are real. In some cases, the complex numbers appear in expressions for real-valued functions. Take, for example, the integral representations

Pl(cosθ)=12π02π(cosθ+isinθcosw)ldw,Jn(x)=12π-ππei(xsint-nt)dt,

for the Legendre polynomials and the Bessel functions of integral order, respectively.

There exist other lesser known types of numbers, somewhat similar to the complex ones, called double and dual numbers (although other names are also employed), which are also useful in various applications (see, e.g., Refs. [1-10] and the references cited therein). The double and the dual numbers (introduced by Clifford in 1873 [5]) are characterized by the presence of “imaginary units” j and ε, respectively, such that j2 = 1 and ε2=0. Most of the existing applications in physics for these numbers are restricted to the double numbers (see, e.g., [1-4, 6, 7], in many cases looking for new theories that might substitute the established ones. As shown in Refs. [8-10], the double and the dual numbers are useful in the solution of well-established systems of real differential equations.

At first sight, it would seem that the use of a unit, j, whose square is +1 is hardly necessary or useful (for instance, it is not required in the solution of algebraic equations with real coefficients), but the examples given in Refs.[1-4, 6-10], together with those presented below, show that having a second unit, besides the real number 1, is indeed very useful.

It is well known that SU(2), the group formed by the 2×2 complex unitary matrices with determinant equal to 1, is homomorphic to SO(3), the group of rotations in the three-dimensional Euclidean space; this homomorphism is useful and physically relevant. For instance, under a rotation, the wavefunctions (spinors) of spin-1/2 particles in non-relativistic quantum mechanics, are transformed by means of a SU(2) matrix. In this paper we explicitly show that the analogs of SU(2), with double or dual numbers as entries, in place of complex numbers, are homomorphic to SO(2,1) (the group of Lorentz transformations in two spatial dimensions) or the group of rigid motions of the Euclidean plane, respectively. In fact, as shown below, the three cases (complex, double and dual) can be treated simultaneously and the corresponding two-component spinors can be defined in a unified manner.

We also show that, making use of the double numbers, we can find a generating function for certain functions analogous to the associated Legendre functions, which arise in the solution of the Laplace equation in the three-dimensional Euclidean space, in spheroidal and toroidal coordinates.

In Sec. 2 we study the special unitary groups formed by 2×2 matrices whose entries are complex, double or dual numbers; we show that they are homomorphic to SO(3), SO(2,1) or SE(2) (the group of rigid motions of the Euclidean plane that preserve the orientation), respectively, and we introduce the corresponding two-component spinors.

In Sec. 3 we study separable solutions of the Laplace equation in the Minkowski (2+1) space, and we find that some of the separated equations coincide with some of the separated equations found in the solution by separation of variables of the Laplace equation in the three-dimensional Euclidean space in spheroidal and toroidal coordinates. In Sec. 4 we show that, making use of the double numbers, one obtains a generating function for the special functions analogous to the associated Legendre functions encountered at Sec. 3. The basic rules for the use of the double and the dual numbers can be found in Refs. [8, 10]; a rigorous and fairly complete discussion of their algebraic properties can be found, e.g., in Refs. [4-7].

2.Special unitary groups

In Sec. 3.1 of Ref. the Hamiltonian

H=12m(px2+py2)-h22mω2(x2+y2), (1)

was considered. Here m and ω are constants, and h may be the usual imaginary unit, i; the hypercomplex unit j (which satisfies the condition j2 = 1); or the hypercomplex unit ε (which satisfies ε2=0). (That is, h2 is equal to -1, +1, or 0, respectively, so that, in all cases, H is real.) This Hamiltonian is invariant under the group of 2×2 matrices of the form

U=(a+hbc+hd-c+hda-hb), (2)

where a,b,c,d are real numbers satisfying the condition

a2+c2-h2(b2+d2)=1. (3)

The matrices of the form (2), fulfilling Eq. (3), form a group with the usual matrix multiplication. In what follows, this group will be denoted by SU(2)h. When h = i this group is the usual SU(2) group which, as is well known, is homomorphic to the rotation group in three dimensions, SO(3).

The invariance of the Hamiltonian (1) under the group SU(2)h is evident if one expresses H in the form

H=12mΨΨ, (4)

where Ψ is the two-component spinor

Ψ(ψ1ψ2)(px+hmωxpy+hmωy) (5)

and the Hermitian adjoint of a matrix is defined as the transpose of the conjugate matrix (in all cases, the conjugate of a + hb is defined by a+hb¯=a-hb). Indeed, as a consequence of (3), the matrix (2) satisfies UU=I and therefore (4) is invariant under the transformation ΨUΨ.

The group SU(2)h is a (real) Lie group of dimension three (the matrix (2) depends on four parameters, which are restricted by one equation) and a basis for its Lie algebra is given by the matrices

σ1(h00-h),σ2(0hh0),σ3(01-10). (6)

The matrices σi are antihermitian because σi=-σi. The trace of each matrix (6) is zero and these matrices form a basis for the real vector space formed by the 2×2 antihermitian matrices with trace equal to zero. The commutators of these matrices are given by [σi,σj]=cijkσk, with sum over repeated indices, where the structure constants cijk are determined by

c123=2h2,c231=-2,c312=-2 (7)

(note that all of them are real, despite the fact that the matrices (6) are not all real).

The products of the matrices (6) can be expressed in the compact form

σiσj=gijI+12cijkσk, (8)

where I is the unit 2×2 matrix and the gij are the entries of the diagonal 3×3 matrix

(gij)=diag(h2,h2,-1). (9)

Thus, the matrix (gij) is singular only if h = ε. With the aid of the matrices (6) we can construct a homomorphism between SU(2)h and a subgroup of SL(3,R) (the group of 3×3 real matrices with determinant equal to 1). In fact, if gSU(2)h, then the product gσig-1 is also a traceless antihermitian matrix. (In fact, tr(gσig-1)=trσi=0 and (gσig-1)=(g-1)σig=g(-σi)g-1=-gσig-1, since the elements of SU(2)h are unitary matrices.) Hence, there exist real numbers aij such that

gσig-1=aijσj. (10)

The mapping g(aji) given by Eq. (10) is a group homomorphism. In fact, if g' is another element of SU(2)h then there exists a matrix (bji) such that g'σig'-1=bijσj and therefore,

gg´σigg´-1=gg´σig´-1g-1=gbijσjg-1=bijajkσk=ajkbijσk

thus showing that the product of the matrix (aij) by (bij) corresponds to the product gg'. ( aij is the entry at the j-th row and i-th column of the matrix (aij).)

In the cases where h is equal to i or to j, the matrices (aji) preserve the metric tensor (gij) (see Eq. (11), below). Indeed, from Eqs. (8) and (10) we have

g(σiσj)g-1=gijI+12cijkakmσm,

which must coincide with (using Eqs. (10) and (8) again)

g(σiσj)g-1=gσig-1gσjg-1=aikσkajlσl=aikajl(gklI+12cklmσm).

By virtue of the linear independence of the set {I,σ1,σ2,σ3}, this amounts to

aikajlgkl=gij (11)

and

aikajlcklm=cijkakm. (12)

Multiplying both sides of Eq. (12) by asngmn we obtain

asngmnaikajlcklm=cijkakmasngmn

or, using Eq. (11),

gmncklmasnaikajl=gkscijk. (13)

We introduce the real constants

csijgkscijk

and from Eqs. (7) and (9) we find that csij=-2h2εsij. Therefore, if hε, Eq. (13) is equivalent to εnklasnaikajl=εsij, which means that det(aji)=1. An explicit computation shows that also in the case where ${\rm h} = \varepsilon$, det(aji)=1 (see Eq. (14) below).

Thus, if h = i or h = j, the matrix (aij) is orthogonal or pseudo-orthogonal; respectively, that is, (aij) belongs to SO(3) or to SO(2,1). In the remaining case, where h=ε, condition (3) reads a2+c2=1, and therefore we can parameterize a and c in the form a=cosθ/2, c=sinθ/2, then, a straightforward computation shows that if gSU(2)h has the form (2), making use of (6) and (10),

(aij)=(cosθsinθ2(bsinθ/2-dcosθ/2)-sinθcosθ2(bcosθ/2+dsinθ/2)001), (14)

which represents an orientation-preserving rigid motion of the Euclidean plane. (If x,y are the coordinates of a point of the Euclidean plane with respect to a set of Cartesian axes, then x',y' given by

x'y'1=cosθsinθx0-sinθcosθy0001xy1

are the coordinates with respect to the same axes of the point obtained by rotating the plane through an angle θ in the clockwise direction about the origin and then translating the points of the plane by the vector (x0,y0).)

2.1.Two-component spinors

Traditionally, spinors are associated with the orthogonal or pseudo-orthogonal groups, SO(n) or SO(p,q), and the most important examples are related to the SO(3) roup (e.g., in the description of the spin for a spin-1/2 particle) and the SO(3,1) group (e.g., in the spinor formalism employed in special or general relativity and in the Dirac equation for the electron). The basic (one-index) spinors form representation spaces for the spin groups, which are covering groups of the orthogonal or pseudo-orthogonal groups. In the standard approach, the spin groups are represented by complex matrices (belonging, e.g., to SU(2) or to SL(2,C)) and the one-index spinors have complex components.

As we shall show below, this can be modified in two ways: instead of spinors with complex components, we can consider spinors whose components are double or dual numbers and, instead of orthogonal or pseudo-orthogonal groups, we can consider “inhomogeneous” groups, specifically, the group of rigid motions of the Euclidean plane (formed by rotations and translations on the Euclidean plane).

A two-component spinor will be represented by a column of the form

Ψ=(ψ1ψ2)(a+hbc+hd), (15)

just as in Eq. (5), where a,b,c,d are real numbers. Under the change of spinor frame given by gSU(2)h, the components (15) transform according to

ΨgΨ. (16)

Hence, ΨΨ is invariant under these transformations. (Note that ΨΨ=ψ1¯ψ1+ψ2¯ψ2 is always real, but only in the case where h = i it is positive definite.)

A non-zero spinor Ψ defines a vector belonging to a real vector space of dimension three with components Ri given by

hRi=Ψ\dagσiΨi=1,2,3. (17)

In fact, since ΨσiΨ is a 1×1 matrix, the conjugate of the right-hand side of Eq. (17) is equal to

(ΨσiΨ)=ΨσiΨ=-ΨσiΨ,

which shows that the numbers Ri are indeed real. Making use of the explicit expression of the matrices σi [see Eqs. (6)], one finds that

R1=ψ1¯ψ1-ψ2¯ψ2,R2+hR3=2ψ1¯ψ2. (18)

According to Eqs. (17) and (10), under (16) the components Ri transform as

hRi(gΨ)\dagσi(gΨ)=Ψ\dagg-1σigΨ=Ψ\daga~ijσjΨ=ha~ijRj,

where (ãij) is the inverse of the matrix (aij). Then, denoting by (gij) the inverse of the matrix (gij) (which exists only when h is i or j), from Eq. (11) it follows that gijRiRj is invariant. In fact, when h = i we find that gijRiRj=-(ΨΨ)2, while in the case where h = j, gijRiRj=(ΨΨ)2.

In the case where h=ε, from Eq. (14) we see that the 2×2 block at the upper left corner of the matrix (aij) represents an ordinary rotation about the origin on the plane R3=0, hence, under these transformations (R1)2+(R2)2 is invariant and one finds that (R1)2+(R2)2=(ΨΨ)2.

Owing to the specific form of the matrices (2) (the entries at the second row are, up to a sign, the conjugates of those at the first row) one readily finds that the components of

Ψ^-ψ2¯ψ1¯ (19)

transform in the same manner as the components of Ψ. The two-component spinor Ψ^ is the mate of Ψ as defined in Ref. [11] for the case where h = i, and differs by a constant factor from the definition of the mate of a spinor in a space with indefinite metric given there. In the applications of the spinors in quantum mechanics (where h = i), the spinor Ψ^ represents a state with the spin in the opposite direction to that corresponding to Ψ. (In the Bloch sphere, Ψ and Ψ^ correspond to diametrically opposite points.)

As an illustration of the differences between the three types of numbers considered here, we look for spinors Ψ which are proportional to their mates (with the proportionality factor being a complex, double, or dual number according to the case at hand): Ψ^=λΨ. According to Eq. (19), we have -ψ2¯=λψ1, ψ1¯=λψ2, which leads to λ¯λ=-1. This condition cannot be satisfied in the case of the complex or the dual numbers; however, in the case of the double numbers it has the general solution λ=±jejα, with αR arbitrary (note that j¯j=(-j)j=-1), furthermore, in that case, ΨΨ=0.

The spinor formalism can be employed in the study of differential geometry (see, e.g., Ref.[11]) and, according to the results of this section, it is possible to develop a unified formalism applicable to three-dimensional Riemannian manifolds of any signature.

The one-index spinors are the basic objects from which vectors and tensors of any rank can be constructed; a field of an arbitrary spin can be expressed in terms of its spinor components; the basic transformation rule (16) determines the transformation of a spinor with any number of indices or, equivalently, of any spin (see, e.g., Ref. [11]).

We close this section with some remarks of a more formal nature. By contrast with the real and the complex numbers, which are fields with the usual operations of sum and product, the double and the dual numbers are only commutative rings with identity. The standard definition of a vector space makes use of a field of scalars; its analog in the case of a ring is called a module (see, e.g., Ref. [12]).

3.The Laplace equation in the Minkowski (2+1) space

In this section we shall consider the Laplace equation for the Minkowski (2+1) space, which is a three-dimensional space with a metric tensor given, e.g., by

ds2=-dx2+dy2+dz2, (20)

in terms of an appropriate coordinate system (similar to the Cartesian coordinate systems of the Euclidean space). Instead of the coordinates (x,y,z) appearing in Eq. (20), we can make use of the local coordinates (r,θ,ϕ) defined by

x=r sinhθ coshϕ,y=r sinhθ sinhϕ,z=r coshθ, (21)

in terms of which the metric tensor (20) takes the form

ds2=-r2dθ2+dr2+r2sinh2θdϕ2. (22)

This last expression shows that the coordinates (r,θ,ϕ) can be regarded as orthogonal, so that the standard formula for the Laplace operator in orthogonal coordinates is applicable, taking care of the minus signs. The Laplace equation in this case (which is just the wave equation in two spatial dimensions) is given by

-2ux2+2uy2+2uz2=0 (23)

or, equivalently,

-1r2sinhθθsinhθuθ+1r2rr2ur+1r2sinh2 θ2uϕ2=0 (24)

Equation (24) admits separable solutions R(r)Θ(θ)Φ(ϕ), where Θ(θ) has to satisfy the equation

1sinhθddθsinhθdΘdθ-ll+1+m2sinh2θΘ=0, (25)

and l,m are separation constants. This equation appears in the solution by separation of variables of the Laplace equation in the three-dimensional Euclidean space in prolate spheroidal equations (with l being an integer) (see, e.g., Ref. [13], Eq. (8.6.7) or Ref. (14), Table 1.06) and in toroidal coordinates (where l is a half-integer) (see, e.g, Ref. [13], Eq. (8.10.11) or Ref. [14], Sec. IV).

Now, in place of (21), we define the local coordinates (r,θ,ϕ) by

x=r coshθ coshϕ,y=r coshθ sinhϕ,z=r sinhθ, (26)

and we find that the metric tensor (20) takes the form

ds2=-dr2+r2dθ2+r2cosh2θdϕ2. (27)

Hence, the Laplace equation (23) is now given by

-1r2rr2ur+1r2coshθθcoshθuθ+1r2coshθ2uϕ2=0 (28)

Equation (28) admits separable solutions R(r)Θ(θ)Φ(ϕ), where Θ(θ) has to satisfy the equation

1coshθddθ(coshθdΘdθ)-#091;l(l+1)-m2cosh2θ#093;Θ=0, (29)

and l,m are separation constants. Equation (29) coincides with one of the separated equations obtained in the solution by separation of variables of the Laplace equation in the three-dimensional Euclidean space, in oblate spheroidal coordinates (see, e.g., Ref. [13], Eq. (8.6.13) or Ref. [14], Table 1.07).

Thus, even though the Minkowski (2+1) space may not seem as interesting as the three-dimensional Euclidean space or the standard Minkowski ( (3+1)) space, as we have shown, the solution of the differential equation (23) is relevant to the solution of the Laplace equation in the three-dimensional Euclidean space.

4.Generating functions

In this section we shall show that one can obtain solutions of Eqs. (25) and (29) by means of generating functions. The starting point is the one employed in Ref. (15): if the metric tensor of the space has the form ds2=gijdxidxj, with sum over repeated indices, i,j,=1,2,,p, and the components gij are constant, then the function (k1x1+k2x2++kpxp)l is a solution of the Laplace equation if and only if the constantsk1,k2,,kp satisfy the condition

gijkikj=0, (30)

where (gij) is the inverse of the matrix (gij). The proof is a straightforward computation, taking into account that, in these coordinates, the Laplacian is given by 2u=gij(/xi)(u/xj).

4.1.Generating solutions of Eq. (25)

In the case of the three-dimensional Euclidean space, with gij=δij, condition (30) reads (k1)2+(k2)2+(k3)2=0, and therefore one is led to make use of complex numbers , but in the case of the Minkowski (2+1) space, in the coordinates (x,y,z) appearing in Eq. (20), condition (30) takes the form

-(k1)2+(k2)2+(k3)2=0, (31)

which can be satisfied by nonzero real numbers k1,k2,k3. However, as we shall see, it is convenient to make use of double numbers:

(k1,k2,k3)=(jcoshv,-jsinhv,1), (32)

where v is an auxiliary parameter. Since, by definition, j2 = 1, the components (32) satisfy condition (31) for all values of v. Then, making use of Eqs. (21) and the fact that ejv=coshv+jsinhv, we have

k1x+k2y+k3z=rj codh vsinhθcoshϕ-jsinhvsinhθsinhϕ+coshθ=rcoshθ+jsinhθ(coshϕcoshv-sinhθsinhv)=rcoshθ+jsinhθ cosh(ϕ-v)=rcoshθ+12 jsinhθ e-j(ϕ-v)+12jsinhθej(ϕ-v) (33)

which shows that k1x+k2y+k3z depends on ϕ and v only through the difference ϕ-v.

With the aid of (33), one can convince oneself that, for l=0,1,2,, the expression (k1x+k2y+k3z)l must be of the form

(k1x+k2y+k3z)l=m=-llj|m|rlflm(θ)ejmϕe-jmv. (34)

Since (k1x+k2y+k3z)l must be a solution of the Laplace equation in the Minkowski (2+1) space and the parameter v is arbitrary, it follows that each term of (34) is a separable solution of the Laplace equation in the Minkowski (2+1) space. In particular, this means that flm(θ) is a solution of Eq. (25). (The factor j|m| is included in order for flm(θ) to be a real-valued function.)

Thus, setting v=0 in Eq. (34), we obtain the generating function

coshθ+12jsinhθe-jϕ+12jsinhθejϕl=m=-lljmflm(θ)ejmϕ (35)

(In the case of the spherical harmonics, considered in Ref. [15], it was useful to keep the factor analogous to e-jmv, in order to obtain integral expressions for the spherical harmonics and the Legendre polynomials.) As a simple example, we have

coshθ+12jsinhθe-jϕ+12jsinhθejϕ2=14sinh2 θ e-2jϕ+jsinhθcoshθ e-jϕ+ cosh2θ+12sinh2θ+jsinhθcoshθejϕ+14sinh2θe2jϕ

which shows that the functions f2,±2(θ), f2,±1(θ), and f20(θ), are proportional to sinh2θ, sinhθcoshθ, and 2cosh2θ+sinh2θ, respectively.

4.2.Generating solutions of Eq. (29)

Following essentially the same steps as in the preceding subsection, with (k1,k2,k3) given again by (32), but using now the coordinates (26) we have

k1x+k2y+k3z=rsinhθ+12jcoshθe-j(ϕ-v)+12jcoshθ ej(ϕ-v) (36)

and, therefore, for l=0,1,2,, we have

rlsinhθ+12jcoshθ e-jϕ+12jcoshθejϕl=m=-lljmrlhlm(θ)ejmϕ (37)

where each term of the sum on the right-hand side is a separable solution of the Laplace equation in the Minkowski (2+1) space and the functions hlm(θ) are solutions of Eq. (29).

5.Final remarks

The real groups found in Sec. 2 are related by means of a contraction in the sense defined in Ref.[16]; however, as we have shown, the use of the complex, double and dual numbers allows us to study these groups simultaneously, without having to take limits.

An advantage of the use double and the dual numbers is that their basic algebraic rules are the same as those of the real or the complex numbers. This means that we can do many computations that are equally applicable to complex, double or dual quantities.

Acknowledgements

The authors acknowledge one of the referees for useful comments and for pointing out Refs.[1, 2, 6, 7] to them.

References

1. A. Crumeyrolle, Variétés différentiables à coordonnées hypercomplexes. Application à une géométrisation et à une généralisation de la théorie d’Einstein-Schrödinger, Ann. Fac. Sci. Toulouse 26 (1962) 105, https://doi.org/10.5802/afst.505. [ Links ]

2. A. Crumeyrolle, Construction d’une nouvelle théorie unitaire en mécanique relativiste. Théorie unitaire hypercomplexe ou complexe hyperbolique, Ann. Fac. Sci. Toulouse 29 (1965) 53, https://doi.org/10.5802/afst.518. [ Links ]

3. P. F. Kelly and R. B. Mann, Ghost properties of algebraically extended theories of gravitation, Class. Quantum Grav. 3 (1986) 705, https://doi.org/10.1088/0264- 9381/3/4/023. [ Links ]

4. F. Antonuccio, Semi-complex Analysis and Mathematical Physics, e-Print: arXiv:grqc/9311032. [ Links ]

5. B. Rosenfeld and B. Wiebe, Geometry of Lie Groups (Springer, Boston, 1997), Chap. 1, https://doi.org/10.1007/978-1-4757-5325-7. [ Links ]

6. F. P. Schuller, Born-Infeld Kinematics, Ann. Phys. 299 (2002) 174, https://doi.org/10.1006/aphy.2002.6273. [ Links ]

7. P. O. Hess, M. Schäfer, and W. Greiner, Pseudo-Complex General Relativity (Springer, Cham, 2016), Chap. 1, 7, https://doi.org/10.1007/978-3-319-25061-8. [ Links ]

8. G. F. Torres del Castillo, Some applications in classical mechanics of the double and the dual numbers, Rev. Mex. Fis. E 65 (2019) 152, https://doi.org/10.31349/RevMexFisE.65.152. [ Links ]

9. G. F. Torres del Castillo, Differentiable Manifolds: A Theoretical Physics Approach, 2nd ed. (Birkhäuser, Cham, 2020), https://doi.org/10.1007/978-3-030-45193-6. [ Links ]

10. G. F. Torres del Castillo , Applications of the double and the dual numbers. The Bianchi models, Rev. Mex. Fis. E 17 (2020) (to be published). [ Links ]

11. G. F. Torres del Castillo , 3-D Spinors, Spin-Weighted Functions and their Applications (Birkhäuser, Basel, 2003), https://doi.org/10.1007/978-0-8176-8146-3. [ Links ]

12. S. Roman, Advanced Linear Algebra, 3rd ed. (Springer-Verlag, New York, 2008), Chap. 4, https://doi.org/10.1007/978-0-387-72831-5. [ Links ]

13. N. N. Lebedev, Special Functions and their Applications (Dover, New York, 1972). [ Links ]

14. P. Moon and D. E. Spencer, Field Theory Handbook, 2nd ed. (Springer-Verlag, Berlin, 1988), https://doi.org/10.1007/978-3-642-83243-7. [ Links ]

15. G. F. Torres del Castillo , A generating function for the spherical harmonics in p dimensions, Rev. Mex. Fis. 59 (2013) 248, https://rmf.smf.mx/ojs. [ Links ]

16. E. Inönü and E. P. Wigner, On the Contraction of Groups and Their Representations, Proc. Natl. Acad. Sci. 39 (1953) 510, https://doi.org/10.1073/pnas.39.6.510. [ Links ]

Received: March 15, 2020; Accepted: March 27, 2020

Creative Commons License This is an open-access article distributed under the terms of the Creative Commons Attribution License