Next Article in Journal
The Role of Symmetry Aspects in Considering the Spin-1 Particle with Two Additional Electromagnetic Characteristics in the Presence of Both Magnetic and Electric Fields
Previous Article in Journal
A Modified Collocation Technique for Addressing the Time-Fractional FitzHugh–Nagumo Differential Equation with Shifted Legendre Polynomials
Previous Article in Special Issue
Fundamentals of Symmetry and Topology: Applications to Materials Science and Condensed Matter Physics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Zero Mass as a Borel Structure

Institute of Physics, University of Tartu, W. Ostwaldi 1, EE-50411 Tartu, Estonia
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Symmetry 2025, 17(9), 1464; https://doi.org/10.3390/sym17091464
Submission received: 7 June 2025 / Revised: 2 August 2025 / Accepted: 30 August 2025 / Published: 5 September 2025
(This article belongs to the Special Issue The Benefits That Physics Derives from the Concept of Symmetry)

Abstract

The Lorentz group Lor 1 , 3 = SO 0 ( 1 , 3 ) has two point fixgroups, namely SO ( 3 ) for time-like translations and SO 0 ( 1 , 1 ) × I R 2 for light-like translations. However, for light-like translations, it is reasonable to consider a line fixgroup that leads to the Borel structure of the Lorentz group and provides appropriate helicities for massless particles. Therefore, whether a particle is massless or massive is not so much a physical question but rather a question of the underlying Lie group symmetry.

1. Introduction

The basic structure of physics is defined via the representations of Lie groups and Lie algebras. Symmetry arises via both the external or space-like and internal or charge-like degrees of freedom and is represented in complex vector spaces regarding the noncompact–compact dichotomy of the relevant groups. The groups appear in two different varieties, semisimple and solvable, serving as the building blocks for all other Lie groups. Simple Lie groups regenerate themselves under commutation and generate semisimple groups via direct products. Solvable Lie groups do not regenerate themselves under commutation but are constructed in a stepwise way out of Abelian subgroups. Non-semisimple Lie groups are semidirect products of semisimple and solvable subgroups. Note that there is no complete classification for solvable Lie groups and, therefore, for non-semisimple Lie groups.
The subgroup structure of the symmetry group plays a basic role in Wigner’s definition of particles in the electroweak standard model, e.g., the isotropy subgroups of the Lorentz group and their corresponding coset manifolds. In particular, for massive particles, the point fixgroup SO ( 3 ) is the maximal connected compact simple subgroup of the Lorentz group Lor 1 , 3 SO 0 ( 1 , 3 ) , and the corresponding time-like energy–momentum hyperboloid is Y 3 = SO 3 ( 1 , 3 ) / SO ( 3 ) . In this paper, we show that, in contrast to this, for massless particles, one obtains the line fixgroup Bor 1 , 3 as being the maximal connected noncompact solvable subgroup of the Lorentz group,
m 0 maximal connected compact simple point fixgroup SO 3 × × m = 0 maximal connected noncompact solvable line fixgroup Bor 1 , 3
Here, Bor 1 , 3 Lor 1 , 3 is the Borel subgroup of the Lorentz group, and at least one of the manifolds Lor 1 , 3 / Bor 1 , 3 is projective ( = S 2 ).
Scanning the recent literature concerning Lie groups for massless particles, few articles are found. Massless particles are considered within the Poincaré group [1,2,3,4,5] but without any reference to the possible solvability (solubility) of the corresponding stabilizer subgroup. Therefore, our work in this field, in previous publications [6,7] and in this publication, is pioneering and is based mainly on handbooks. As the basis for a deeper understanding, we suggest Refs. [8,9,10,11], from which many of the ideas that led to the research described in this paper were derived. At this point, it is worth emphasizing that, as physicists, we do not apply any rigorous mathematical formalism, consisting of definitions, lemmata, prepositions, theorems, and corollaries, together with the corresponding proofs. Instead, we mention the components of our considerations in these mathematical terms, without providing proofs in most instances; these, instead, can be found in the references that we cite at various points, which also mark the ends of the corresponding theorems. An exception is Lemma 1, which is new and the main outcome of our analysis. This lemma on the SL 2 reconstruction of the proper Lorentz group can be considered as a corollary of what has previously been discussed.
This publication is organized as follows. In Section 2, we present basic facts about the proper Lorentz group, and we present three theorems that are the foundations of our work: the Chevalley theorem, the Lie–Kolchin theorem, and a theorem related to Borel. Having explained Wigner’s concept of a little group, in Section 3, we first deal with Wigner’s result for this, given by the Euclidean group E ( 2 ) . However, the character equation has an additional solution that leads to the Borel subgroup, dealt with in Section 4. In fact, there are a couple of these subgroups, with the union giving the proper Lorentz group, while the cut is the maximal torus. We show explicitly that each of these groups is generated by two elements of the minimal solvable algebra sol 2 , spanning the algebra of the Borel subgroup as a Kronecker sum. Turning to topology, in Section 5, we show that the quotient of the Lorentz group and the Borel subgroup is a projective variety. Section 6 deals with representations, presenting also our Lemma 1, focused on reconstructing the proper Lorentz group by two copies of the simplest noncompact group SL 2 ( I R ) . Via the Weinberg ansatz, in Section 7, we describe a connection back to physics. In Section 8, we present our conclusions.

2. Basics

Here, we give a brief description of the basic theorems and constructions of the theory of semisimple groups with applications to the proper Lorentz group [12,13,14,15,16,17]. The Minkowski representation SO 0 ( 1 , 3 ) preserves the indefinite symmetric metric η = diag ( 1 , 1 , 1 , 1 ) in the real spacetime I R 4 x , y ,
x · y = x μ η μ ν y ν = x T η y .
As the defining representation of the causality-compatible Lorentz group,
Lor 1 , 3 = { Λ T η Λ = η , det Λ = 1 , Λ 0 0 1 , Λ ( Λ μ ν ) GL 4 ( I R ) }
is parametrizable by six real parameters ω μ ν = ω ν μ ,
Λ ( ω ) = exp ( 1 2 ω μ ν e μ ν ) = exp 1 2 ω p ϵ 0 p j k e j k + ω 0 p e 0 p ,
where the first part is compact and the second part noncompact, and the domain of these six parameters is given by
D = { ω 0 p , ω p I R , π < ω p π , p = 1 , 2 , 3 }
which is homeomorphic to I R 3 × P 3 , where P 3 is the three-dimensional projective space. The generators have the form
e j k = 0 0 T 0 e j e k T e k e j T , e 0 j = 0 e j T e j 0 3 ,
where, in general, [ e μ ν , e ρ σ ] = η μ ρ e ν σ + η ν σ e μ ρ η μ σ e ν ρ η ν ρ e μ σ .
Here, e j , j = 1 , 2 , 3 is the Euclidean basis in I R 3 . Therefore, Lor 1 , 3 is a locally compact and doubly connected, path-connected, simple and reductive group with universal coverage SL 2 ( C ) , i.e.,
Lor 1 , 3 SL 2 ( C ) / Z 2 SO ( 3 , C ) .
The last isomorphism means that the representations of Lor 1 , 3 may be seen as representations of the complex rotation group SO ( 3 , C ) . For completeness, note that the Lie algebra lor 1 , 3 log Lor 1 , 3 is the noncompact real form of
so ( 4 , C ) = sl 2 ( C ) sl 2 ( C ) .
To motivate what follows, it is instructive to look at Chevalley’s theorem in the context of the defining representation of the Lorentz group, according to which Lor 1 , 3 acts on the flat spacetime or its dual energy–momentum space by a linear transformation,
Lor 1 , 3 Λ : E 1 , 3 x μ Λ μ ν x ν E 1 , 3 .
Theorem 1 
(Chevalley). Let G be a linear algebraic group and H G a closed algebraic subgroup. Then, there is a rational representation ϕ : G GL ( V ) and a one-dimensional subspace L V such that
H = { g G : ϕ ( g ) L = L } .
Otherwise, if L spans the line L, i.e., L = C , then the equation
ϕ ( h ) = χ ( h ) L , h H
defines a character χ X ( H ) of H. This character is called the weight to the semi-invariant ℓ, and X ( H ) = Mor ( H , GL 1 ) . Note that, if [ H , H ] = H , then X ( H ) = { 0 } , i.e., SL 2 ( C ) has no nontrivial characters and therefore the Lorentz group has no nontrivial characters [18].
Theorem 2 
(Lie–Kolchin). Let G be a connected solvable linear algebraic group, and let ( ϕ , V ) be a regular representation of G. Then, there exist characters χ i X ( G ) , i = 1 , 2 , , n and a flag
V = V 1 V 2 V n V n + 1 = { 0 }
such that
( ϕ ( g ) χ i ( g ) ) V i V i + 1
for all g G . Taking i = n , one obtains the characteristic equation
ϕ ( g ) V n = χ n ( g ) V n ,
i.e., every solvable group has a common one-dimensional subspace L V [19,20].

Wigner’s Little Group

What follows was proposed by Eugene Paul Wigner in 1939: for massive particles ( m 0 ), the point fixgroup (or little group) of the momentum p = (1, 0, 0, 0), and lg ( p )   =   SO ( 3 )     Lor 1 , 3 is maximal connected, compact and simple [4,5,13,14,16]. Proceeding purely mathematically, one finds that, for massless particles (m = 0), the little group is maximal connected, noncompact and solvable, resulting in a Borel subgroup Bor 1 , 3 . By definition, a Borel subgroup of an algebraic group G is a maximal connected solvable subgroup.
Theorem 3. 
Let G be a connected linear algebraic group. Then,
1. 
G contains a Borel subgroup B;
2. 
All other Borel subgroups of G are conjugate to B;
3. 
The homogeneous manifold G / B is a projective variety;
4. 
G = g G g B g 1 , where B is a fixed Borel subgroup of G.
(see Theorem 11.4.7 in Ref. [21], p. 524).
Thus, every element g G is contained in a Borel subgroup. As mentioned before, a solvable group has a semidirect structure. For the connected solvable group G, the set G u of unipotent elements is a closed connected nilpotent subgroup of G. There exists a maximal torus T G G and, for this, an exact sequence
e G G u G u T G = G T G e G .
Since the maximal torus T G and the maximal connected unipotent subgroup G u of G are those for the Borel subgroup, one has the exact sequence
e B B u B u T G = B T G e B .
and B = N G ( B u ) . Here, G u = { g G : g = g u } , where g u is the unipotent component in the Jordan decomposition g = g s g u .

3. The Role of Mass

Before discussing the Borel subgroup in further mathematical detail, let us note some physical considerations regarding the kinematics. Taking the momentum four-vector to be p, with p 2 = m 2 > 0 the squared mass, for massive particles, one can move to the rest frame, where p = ( m , 0 , 0 , 0 ) T . The stabilizer subgroup or fixgroup is given by SO ( 3 ) , the three-dimensional rotations. However, if the particle is massless, such a move to the rest frame is no longer possible. According to Wigner’s classification, the fixgroup is E ( 2 ) . This is the little group that Wigner indicates for massless particles like photons and (massless) neutrinos. For instance, for a momentum vector p = ( p 0 , 0 , 0 , p 0 ) T pointing in z direction, E ( 2 ) consists of rotations about the z axis, translations orthogonal to it and reflections. However, what is not taken into account by this is the interchange of time and space components, which is obviously an additional symmetry transformation. Together with this additional transformation, the fixed point group is given by the Borel subgroup Bor 1 , 3 .
Returning to mathematics, let B G be a Borel subgroup of G and V a finite-dimensional rational G-module. Then, the fixed points of B in V coincide with the fixed points of G. As the Lorentz group has no fixed points, for B, the character Equation (3) is the only one that can be solved. In contrast, for SO ( 3 ) , one has the commutant [ SO ( 3 ) , SO ( 3 ) ] = SO ( 3 ) . Therefore, the character group X ( SO ( 3 ) ) is trivial. The rotation group has no nontrivial characters, and the character Equation (3) is impossible to solve. Moreover, if V is an irreducible rotational G-module (let G be semisimple), then there is a unique Borel-stable one-dimensional subspace spanned by a maximal vector of some weight/character χ with multiplicity one.
The little group of Wigner is E ( 2 ) is nonmaximal, connected, solvable and noncompact [13,14,22,23]. The nonmaximality is explained (or determined) by the requirement to be a point fixgroup of ( 1 ,   0   , 0   , 1 ) T ,
E ( 2 ) I R 2 SO ( 2 ) .
In this semidirect product, the compact group S O ( 2 ) acts on the abelian locally compact group exp I R 2 by the multiplication rule
( x 1 , R 1 ) ( x 2 , R 2 ) = ( x 1 + R 1 x 2 , R 1 R 2 ) .
Every irreducible representation of E ( 2 ) is equivalent either to a character of SO ( 2 ) lifted to E ( 2 ) or to an induced representation ind I R 2 E ( 2 ) χ , where χ is the nontrivial character of I R 2 . To avoid a continuum of helicity states, one has to require that, for physical states, the noncompact part of E ( 2 ) is trivial in all representations, so the little group reduces to SO ( 2 ) . There are topological considerations that restrict the allowed values of the helicity to integers or half-integers. Thus, the helicity
λ = J k | k | = 0 , ± 1 2 , ± 1 ,
is Lorentz-invariant for massless particles with the total angular momentum J .
As for any abelian group, the reducible representations of SO ( 2 ) are one-dimensional. Therefore, according to Wigner’s classification, the free massless particles have only a single degree of freedom and are characterized by the value λ of their helicity.
In nature, there are two classes of particles. The first class consists of particles that can exist in two helicity states ± λ . Such a particle is defined as a representation of the parity-extended Poincaré group [1]. Since the electromagnetic interaction conserves parity, the photon is defined as the SO ( 2 ) -doublet
1 2 , 1 2 SO ( 2 ) ( + 1 ) ( 1 ) 2 × ( 0 ) parity ( ± 1 ) 2 × ( 0 ) .
The second class contains particles for which the parity is not defined, as the interactions that they are involved in violate parity. Such particles are the neutrinos that exist only with helicity 1 2 and antineutrinos with helicity + 1 2 .

4. The Borel Subgroup

The key observation in the preceding sections was the character Equation (3). Solving the character equation
B p = χ ( B ) p
for the light-like standard vector p = ( 1 ,   0 ,   0 ,   1 ) with B Lor 1 , 3 , one obtains [6,7]
B ( + ) ( β ; θ , ω ) = A B T C D
with
A = cosh θ + 1 2 | β | 2 e θ , B = sinh θ 1 2 | β | 2 e θ e 3 + rot 3 ω β , C = sinh θ + 1 2 | β | 2 e θ e 3 + e θ β , D = cosh θ 1 1 2 | β | 2 e θ e 3 e 3 T e θ β e 3 T + ( 1 3 + e 3 β T ) rot 3 ω .
Here, β T = ( β 1 , β 2 , 0 ) = a = 1 2 β a e a T , e 3 T = ( 0 , 0 , 1 ) ,
rot 3 ω = cos ω sin ω 0 sin ω cos ω 0 0 0 1 .
As a consequence, the character is given by
χ ( B ( + ) ) = 1 2 μ , ν = 0 , 3 B μ ν = e θ .
The composition rule is
B ( + ) ( β 1 ; θ 1 , ω 1 ) B ( + ) ( β 2 ; θ 2 , ω 2 ) = B ( + ) ( β 1 + e θ 1 rot 3 ω 1 β 2 ; θ 1 + θ 2 , ω 1 + ω 2 )
and therefore
B ( + ) ( β ; θ , ω ) = B u ( + ) ( β ) T ( θ , ω ) .
All such transformations B ( + ) ( β ; θ , ω ) with a noncompact parameter space for helicity and gauge, given by { β I R 2 , θ > 0 , 0 < ω π } , form the Borel subgroup
Bor 1 , 3 ( + ) = ( Bor 1 , 3 ( + ) ) u Tor 1 , 3 Lor 1 , 3 .
Here, Tor 1 , 3 = SO 0 ( 1 , 1 ) × SO ( 2 ) is the maximal torus in Lor 1 , 3 , and ( Bor 1 , 3 ( + ) ) u is the unipotent radical of Bor 1 , 3 ( + ) .
The linearization of Bor 1 , 3 ( + ) in the neighborhood of the identity
B ( + ) ( β ; θ , ω ) = 1 4 + β 1 b 1 + β 2 b 2 + θ b 0 + ω b 3 = 1 4 1 2 ω μ ν e μ ν
results in the Lie algebra bor 1 , 3 ( + ) = log Bor 1 , 3 ( + ) , with
b 0 = e 03 , b 1 = e 01 + e 31 , b 2 = e 02 + e 32 , b 3 = e 21 .
As a vector space span I R { b μ } 0 3 endowed with commutation relations
[ b 0 , b a ] = b a , [ b 3 , b a ] = ϵ 3 a b b b , a = 1 , 2
the Borel algebra reads
bor 1 , 3 ( + ) = ( bor 1 , 3 ( + ) ) u tor 1 , 3 = so ( 1 , 1 ) so ( 2 ) I R 2 ,
where ( bor 1 , 3 ( + ) ) u = rad u ( + ) is the Lie algebra corresponding to the unipotent radical.
Since the semisimple rank is rank ss SO ( 1 , 3 ) = 2 , there exists a unique Borel subgroup Bor 1 , 3 ( ) Lor 1 , 3 , called opposite Bor 1 , 3 , such that [18,24,25,26]
Bor 1 , 3 ( ) Bor 1 , 3 ( + ) = Tor 1 , 3 = SO 0 ( 1 , 1 ) × SO ( 2 )
and Lor 1 , 3 = Bor 1 , 3 ( ) Bor 1 , 3 ( + ) .
To visualize the algebraic structure of bor 1 , 3 ( + ) , it is convenient to transform the basis to
t 0 = 1 2 ( b 0 + i b 3 ) , t + = 1 2 ( i b 1 b 2 ) , [ t 0 , t + ] = t +
u 0 = 1 2 ( b 0 i b 3 ) , u + = 1 2 ( i b 1 b 2 ) , [ u 0 , u + ] = u +
with [ t 0 , + , u 0 , + ] = 0 . This leads to the Kronecker sum decomposition
bor 1 , 3 ( + ) = sol 2 ( e ) sol 2 ( f ) : = sol 2 ( e ) 1 2 + 1 2 sol 2 ( f ) .
The Kronecker sum decomposition is easily seen after applying the splitting map
t 0 = 1 2 h 1 2 , t + = ( i e ) 1 2 for sol 2 ( e ) , u 0 = 1 2 ( 1 2 h ) , u + = 1 2 ( i f ) for sol 2 ( f ) .
Here,
h = 1 0 0 1 , e = 0 1 0 0 , f = 0 0 1 0
is the natural Chevalley basis of sl 2 ( C ) .
Again, the linearization of Bor 1 , 3 ( ) generates the basis of the underlying vector space as
k 0 = e 03 , k 3 = e 21 , k a = e 0 a + e 3 a , a = 1 , 2
with nonzero commutation relations
[ k 0 , k a ] = k a , [ k 3 , k a ] = ϵ 3 a b k b .
Thus,
lor 1 , 3 = bor 1 , 3 ( ) + bor 1 , 3 ( + ) = ( bor 1 , 3 ( ) ) u ( bor 1 , 3 ( + ) ) u + tor 1 , 3 ,
where tor 1 , 3 = log Tor 1 , 3 = log SO 0 ( 1 , 1 ) log SO ( 2 ) .
The Kronecker sum decomposition of bor 1 , 3 ( + ) into two fundamental solvable groups sol 2 is easily extended to an sl 2 ( I R ) decomposition for the whole lor 1 , 3 ,
lor 1 , 3 sl 2 ( I R ) e sl 2 ( R ) f .
Using the splitting map, for sl 2 ( I R ) e , one obtains
t 0 = 1 2 ( b 0 + i b 3 ) 1 2 h 1 2 t + = 1 2 ( i b 1 b 2 ) ( i e ) 1 2 t = 1 2 ( i k 1 k 2 ) ( i f ) 1 2
with commutation relations
[ t + , t ] = 2 t 0 , [ t 0 , t ε ] = ε t ε , ε = ±
For sl 2 ( I R ) f , one obtains
u 0 = 1 2 ( b 0 i b 3 ) 1 2 1 2 h u + = 1 2 ( i b 1 b 2 ) 1 2 ( i f ) u = 1 2 ( i k 1 k 2 ) 1 2 ( i e )
with commutation relations
[ u + , u ] = 2 t 0 , [ u 0 , u ε ] = ε t ε , ε = ±
and [ sl 2 ( I R ) e , sl 2 ( I R ) f ] = 0 .
Returning from the defining matrix representation of the Borel algebra to the group matrix representation, by exponentiation, one obtains
tor 1 , 3 ϑ b 0 + ω b 3 exp exp ( ϑ b 0 + ω b 3 ) Tor 1 , 3
with
exp ( ϑ b 0 + ω b 3 ) = 1 4 + sinh ϑ b 0 + ( cosh ϑ 1 ) b 0 2 1 4 + sin ω b 3 + ( 1 cos ω ) b 3 2 = = 1 + x T x x T x rot 3 ω 1 3 + x x T , x = sinh ϑ e 3 .
The Cartan–Killing form in the defining representation is indefinite,
( ϑ b 0 + ω b 3 , ϑ b 0 + ω b 3 ) = 2 ( ϑ 2 ω 2 )
and
exp ( ϑ b 0 + ω b 3 ) , exp ( ϑ b 0 + ω b 3 ) = 2 cosh ϑ + cos ω 0 .
For the unipotent radical, one has
rad u ( + ) b = β 1 b 1 + β 2 b 2 exp exp b = 1 4 + exp ( β 1 b 1 ) + exp ( β 2 b 2 ) .
Since the algebra bor 1 , 3 ( + ) is solvable, and its derived algebra is given by
Der bor 1 , 3 ( + ) = rad u ( + ) ,
the Cartan–Killing form is identically zero, ( rad u ( + ) , rad u ( + ) ) 0 . For Rad u ( + ) , one obtains ( exp b , exp b ) = 4 , so, obviously, b rad u ( + ) nilponent and exp b Rad u ( + ) unipotent.

5. The Quotients

Given a closed subgroup H of an algebraic group G, there is a smooth projection π : G G / H , where the fibers are precisely the cosets g H , g G . The projection π has a smooth local injection, given by the compatible section
γ : G / H G   id π     G / H
such that π γ = id G / H .
As an example, we consider the Cartan decomposition g = log G = log K p , where log K is the maximal compact subalgebra of g , and the subspace p = g mod log K consists of the noncompact generators of g . Exponentiating the Lie algebra decomposition into the Lie group,
g = log K p compact noncompact exp       G = K exp p compact coset subgroup representatives
one obtains the parametrization of the algebraic manifold G / K = exp p .
The map of K × p onto G
K × p ( k , X ) k exp X G
is a diffeomorphism into G, i.e., G = K × exp p .
Different choices of the section γ give different formulae for the coset representatives. For the Borel subgroup B G , the factor set G / B is the largest homogeneous space for G, having the structure of a projective variety. Since G / B is complete, the Borel subgroup B has a fixed point in G / B .

5.1. The SO (3) Parametrization

The Borel decomposition of the Lorentz group Lor 1 , 3 = SO 0 ( 1 , 3 ) is generated by the decomposition of the algebra lor 1 , 3 = log SO 0 ( 1 , 3 ) in a natural way by reordering the usual parametrization
lor 1 , 3 1 2 ω μ ν e μ ν = ω 1 e 31 + ω 2 e 32 + β 1 b 1 + β 2 b 2 + ϑ e 03 + ω e 21 coset   representatives bor 1 , 3 exp Lor 1 , 3 exp a = 1 2 ω a e 3 a × Bor 1 , 3 ( + ) ( β 1 , β 2 ; ϑ , ω )
More precisely,
P ( 2 ) a = 1 2 ω a e 3 a = 0 0 T 0 e 3 ω ( 3 ) T ω ( 3 ) e 3 T exp P ( 2 ) = 1 0 T 0 0 1 2 x ( 2 ) x ( 2 ) T x ( 2 ) 0 x ( 2 ) T 1 x ( 2 ) T x ( 2 ) .
Here,
x ( 2 ) = sin ω ( 2 ) ω ( 2 ) ω ( 2 ) , ω ( 2 ) = ω 1 ω 2 , ω ( 2 ) 2 = ω 1 2 + ω 2 2 , ω ( 3 ) = ω ( 2 ) 0 .
The Cartan–Killing inner product for representatives
( P ( 2 ) , P ( 2 ) ) = tr P ( 2 ) 2 = 2 ω ( 2 ) 2 < 0
is negative, i.e., the representatives are compact operators. For the representatives of the group coset, one has
( exp P ( 2 ) , exp P ( 2 ) ) = 4 cos 2 ω ( 2 ) > 0 .
Therefore, the geometric manifold for the respresentatives of the group coset is compact. Moreover, the parametrization of the Bor 1 , 3 -classes can be given by the three-point x = ( x 1 , x 2 , x 3 ) T in I R 3 as
x ( 2 ) = sin ω ( 2 ) ω ( 2 ) ω ( 2 ) = x 1 x 2
and x 3 = cos ω ( 2 ) . Then,
exp P ( 2 ) = 1 0 T 0 0 T 1 2 1 1 + x 3 x ( 2 ) x ( 2 ) T x ( 2 ) 0 x ( 2 ) T x 3
with det exp P ( 2 ) = x 1 2 + x 2 2 + x 3 2 = 1 . Therefore, the SO -type coset representatives generate a compact factor set, the two-sphere in I R 3 [27],
Lor 1 , 3 / Bor 1 , 3 ( + ) = SO ( 3 ) / SO ( 2 ) = S 2 .
Using the Iwasawa decomposition
Lor 1 , 3 = SO ( 3 ) × SO 0 ( 1 , 1 ) × exp I R 2 = ( SO ( 3 ) / SO ( 2 ) ) × SO ( 2 ) × S O 0 ( 1 , 1 ) × exp I R 2 Bor 1 , 3
and Bor 1 , 3 ( + ) = ( SO ( 2 ) × SO 0 ( 1 , 1 ) ) exp I R 2 , one obtains the same result.
The projective coordinates for this parametrization are
< z a = x a 1 x 1 2 x 2 2 = x a x 3 < , a = 1 , 2 .
As mentioned, different choices for the section γ provide different parametrizations.

5.2. The SO (1, 2) Parametrization

Let
lor 1 , 3 1 2 ω μ ν e μ ν = ϑ 1 e 01 + ϑ 2 e 02 + β 1 b 1 + β 2 b 2 + ϑ e 03 + ω e 21 coset   representatives   bor 1 , 3 ( + ) exp   Lor 1 , 3 exp ϑ ( 2 ) × Bor 1 , 3 ( + )
Here, ϑ ( 2 ) = a = 1 2 ϑ a e 0 a is noncompact, as the Cartan–Killing form is positive,
( ϑ ( 2 ) , ϑ ( 2 ) ) = 2 ϑ ( 2 ) 2 = 2 ( ϑ 1 2 + ϑ 2 2 ) > 0 .
The group representatives exp ϑ ( 2 ) are noncompact,
SO 0 ( 1 , 2 ) exp ϑ ( 2 ) = 1 + y ( 3 ) T y ( 3 ) y ( 3 ) T y ( 3 ) 1 3 + y ( 3 ) y ( 3 ) T , y ( 3 ) = sinh ϑ ( 2 ) ϑ ( 2 ) ϑ 1 ϑ 2 0 ,
as the Cartan–Killing form is positive,
( exp ϑ ( 2 ) , exp ϑ ( 2 ) ) = 4 cosh 2 ϑ ( 2 ) > 0 .
The parametrization of the coset representations for the group can be chosen as the point y = ( y 0 , y 1 , y 2 ) in I R 3 or as the projective coordinates z a = y a / y 0 , a = 1 , 2 in the interior of the unit circle of the y 0 = 1 plane. In the first case, we have
y 0 = cosh ϑ ( 2 ) , y a = sinh ϑ ( 2 ) ϑ ( 2 ) ϑ a , a = 1 , 2 ,
so that y 0 2 y 1 2 y 2 2 = 1 , and the representatives are
Q ( y 0 , y 1 , y 2 ) = y 0 y ( 3 ) T y ( 3 ) 1 3 + 1 1 + y 0 y ( 3 ) y ( 3 ) T
with det Q = ( y 0 2 y 1 2 y 2 2 ) / ( 1 + y 0 ) = 1 and ( Q , Q ) = 4 y 0 2 > 0 . As a consequence, the representatives are found on the noncompact on-shell hyperboloid
Y 2 = Lor 1 , 3 / Bor 1 , 3 ( + ) = SO 0 ( 1 , 2 ) / SO ( 2 ) .
The projective coordinates ( q 1 , q 2 ) are
1 < q a = y a 1 + y 1 2 + y 2 2 = y a y 0 < 1 , a = 1 , 2 .
Therefore, the two examples for the coset representatives considered up to now are the time-like on-shell hyperboloid Y 2 and (as a compact partner) the sphere S 2 with the common compact subgroup given by SO ( 2 ) Tor 1 , 3 .

5.3. The Borel Parametrization

Finally, to be systematic, the Borel structure of the Lorentz group provides a constructive procedure to determine representatives of different cosets. To begin with, we recall the Borel decomposition
lor 1 , 3 = bor 1 , 3 ( ) bor 1 , 3 ( + ) = rad u ( ) ( rad u ( + ) tor 1 , 3 ) = rad u ( + ) ( rad u ( ) tor 1 , 3 ) .
As underlying vector spaces, one has
lor 1 , 3 = span I R { e μ ν = e ν μ } 0 3 = span I R { b 0 , b 3 , b a , k a } a = 1 2 = span I R { t 0 , t ε ; u 0 , u ε } ε = ± 1 .
It is convenient to choose the basis of rad u ( ) as
k a = 0 e a T e a e 3 e a T e a e 3 T .
Then,
rad u ( ) k = κ 1 k 1 + κ 2 k 2 = 0 κ ( 3 ) T κ ( 3 ) e 3 κ ( 3 ) T κ ( 3 ) e 3 T ,
where
κ ( 3 ) = κ 1 κ 2 0 , κ ( 3 ) 2 = κ 1 2 + κ 2 2 ,
and the Cartan–Killing product ( k , k ) = tr k 2 = 0 . This is the second criterion for the solvability of bor 1 , 3 ( ) : an algebra g is solvable if and only if its Cartan–Killing metric tensor is identically zero on its derived algebra D 1 g .
The expression for the representatives of the group coset is obtained by exponentiation:
lor 1 , 3 rad u ( ) bor 1 , 3 ( + ) exp       Lor 1 , 3 Rad u ( ) × Bor 1 , 3 ( + )
From this, the representations of the group coset can be read as
Rad u ( ) exp k = 1 4 + k + 1 2 k 2 = 1 + t x y t x 1 0 x y 0 1 y t x y 1 t ,
where t = 1 2 κ ( 3 ) 2 , x = κ 1 , y = x 2 and 2 t ( x 2 + y 2 ) = 0 . One observes that the real parameters ( t , x , y ) describe the subspace of the noncompact elliptic paraboloid. The projective coordinates of this parametrization are
( p 1 , p 2 ) = x t , y t = x x 2 + y 2 , y x 2 + y 2 , p 1 2 + p 2 2 > 0 .
The Cartan–Killing form for the representatives of the group coset is positive:
( exp k , exp k ) = 4 > 0 .
As a consequence, one has the following:
  • m 0 : a rest frame exists for massive particles ⇒
    stabilizer subgroup is the point fixgroup SO ( 3 ) ⇒ leading to spin;
  • m = 0 : no rest frame exists for massless particles ⇒
    stabilizer subgroup is the line fixgroup Bor 1 , 3 ⇒ leading to helicity.
Indeed, accepting the undulatory theory of light, the plane wave, as the most elementary type of wave, cannot be localized in space. Moreover, the characteristic Equation (3) suggests that the massless particle can be enclosed on a line. Therefore, the question of how a massless particle with energy E differs from the same particle with energy λ E is a quantum-mechanical problem in the form of Planck’s formula E = h ν , rather than a problem of symmetry. In fact, the difference is a mathematical one:
  • m 0 : semisimple compact;
  • m = 0 : solvable noncompact.

6. The Representations

In general, Lie algebras play their role in physics not as abstract algebras but through their representations that act on suitable representation spaces. For example, spin and helicity are determined by the stabilizer subgroups of Lor 1 , 3 . For mathematical convenience, it is reasonable to consider representations in vector spaces over complex number fields. This is because, in physics, the concept of reducibility is of fundamental importance, and the mathematical structure of quantum mechanics works with complex Hilbert spaces [13,14,15,16,22,28,29,30,31].
A representation D of a real algebra g can be extended to a unique linear complex representation D ^ by the holomorphic extension
D ^ ( A + i B ) = D ( A ) + i D ( B ) , A , B g .
Although Lor 1 , 3 SO ( 3 , C ) , the process of holomorphic extension for lor 1 , 3 arises from the fact that lor 1 , 3 is the real form of so 4 ( C ) ,
lor 1 , 3 C I R lor 1 , 3 so 4 ( C ) = sl 2 ( C ) sl 2 ( C ) = C I R su ( 2 ) C I R su ( 2 ) .
This complexification of lor 1 , 3 provides the link between the real-valued Lorentz algebra lor 1 , 3 and the real-valued algebra su ( 2 ) , and, using this link, one can construct all the representations of lor 1 , 3 .
Since a representation τ m of so ( 3 ) comes from a representation D ( m ) of SO ( 3 ) Lor 1 , 3 , the diagram
τ m : so 3 x τ m ( x ) exp T ( m ) : SO ( 3 ) exp x D ( m ) ( exp x ) = exp τ m ( x )
has to be commutative. However, for odd m, there is no such representation D ( m ) . In order to overcome this problem, one needs su ( 2 ) to generate all the finite-dimensional representations of lor 1 , 3 . The complex representations T ( k , l ) ( k , l = 0 , 1 2 , 1 , ) may be obtained by holomorphic extension and Weyl’s unitary trick,
{ e μ ν } 0 3 unitary trick Weyl s 1 2 1 2 ϵ p n q e n q ± i e 0 p 1 3         map splitting D p = m p m p , B p = i ( m p ( m p ) ) 1 3               T ( k , l ) T ( k , l ) ( D p ) = D ( k ) ( m p ) D ( l ) ( m p ) , T ( k , l ) ( B p ) = i D ( k ) ( m p ) ( D ( l ) ( m p ) ) 1 3 .
Here, { m p } 1 3 generate the algebra su ( 2 ) and D ( k ) , k = 0 , 1 2 , 1 , are the common representations of su ( 2 ) .
Any irreducible finite-dimensional representation of lor 1 , 3 is isomorphic to T ( k , l ) for some ( k , l ) . As a special case, there are two inequivalent fundamental representations from which all others can be obtained by reducing the tensor products. The two-dimensional spinor representation ( 1 2 , 0 ) is defined by the commutative diagram
Lor 1 , 3 Λ : E 1 , 3 p     Λ p T ( 1 / 2 , 0 ) σ σ SL 2 ( C ) T ( 1 / 2 , 0 ) ( Λ ) A Λ : H 2 σ ( p )     A Λ σ ( p ) A Λ = σ ( Λ p )
The representation ( 0 , 1 2 ) is defined as
Lor 1 , 3 Λ : E 1 , 3 p     Λ p T ( 0 , 1 / 2 ) σ ˜ σ ˜ SL 2 ( C ) T ( 0 , 1 / 2 ) ( Λ ) A ˜ Λ : H 2 σ ˜ ( p )     A ˜ Λ σ ˜ ( p ) A ˜ Λ = σ ˜ ( Λ p )
Here, σ ˜ ( p ) = p μ σ ˜ μ , σ ( p ) = p μ σ μ , with σ ˜ 0 = σ 0 = 1 2 and σ ˜ p = σ p ( p = 1 , 2 , 3 ) being the Pauli matrices. Since the finite-dimensional representations of su ( 2 ) are in one-to-one correspondence with those of sl 2 ( I R ) , and the Lie algebra lor 1 , 3 is noncompact, it is reasonable to define the representations of lor 1 , 3 in terms of the algebra sl 2 ( I R ) . Moreover, the most important technique by which to study the representations of linear noncompact groups is to reduce the problem to the subgroups isomorphic to the simplest noncompact group SL 2 ( I R ) . For example, lor 1 , 3 contains three such sl -isomorphic subalgebras but only one su -isomorphic subalgebra. The following lemma gives the sl -structure for lor 1 , 3 .
Lemma 1. 
If e μ ν = e ν μ , μ , ν = 0 , 1 , 2 , 3 are defined by
e 01 = e 2 ( e 2 ) e 02 = i ( e 1 ( e 1 ) ) e 03 = e 3 ( e 3 ) e 31 = e 1 e 1 e 32 = i ( e 2 e 2 ) e 21 = i ( e 3 e 3 ) ,
where { e k } 1 3 are the generators of the algebra sl 2 ( I R ) ,
[ e 3 , e 1 ] = e 2 , [ e 3 , e 2 ] = e 1 , [ e 1 , e 2 ] = e 3 ,
then e μ ν generate the Lorentz algebra lor 1 , 3 ,
[ e μ ν , e ρ σ ] = η μ ρ e ν σ + η ν σ e μ ρ η μ σ e ν ρ η ν ρ e μ σ .
Applying this lemma, one can define the representations of lor 1 , 3 in terms of the holomorphic extensions of the irreducible representations π ( n ) of sl 2 ( I R ) [24,25,26,32],
π ( k , l ) ( e 01 ) = π ( k ) ( e 2 ) π ( l ) ( e 2 ) π ( k , l ) ( e 31 ) = π ( k ) ( e 1 ) π ( l ) ( e 1 ) π ( k , l ) ( e 21 ) = i π ( k ) ( e 3 ) π ( l ) ( e 3 )
( k , l = 0 , 1 2 , 1 , ). Here, π ( k ) is the standard representation of sl 2 ( I R ) ,
π ( k ) ( e 1 ) | k , m = 1 2 ρ m ( k ) | k , m + 1 1 2 ρ m 1 ( k ) | k , m 1 , π ( k ) ( e 2 ) | k , m = 1 2 ρ m ( k ) | k , m + 1 + 1 2 ρ m 1 ( k ) | k , m 1 , π ( k ) ( e 3 ) | k , m = m | k , m ,
with ρ m ( k ) = ( k + m + 1 ) ( k m ) , m = k , k + 1 , , k .
In some contexts, it is more convenient to work with the sl -basis e = e 1 + e 2 , f = e 1 + e 2 , and h = 2 e 3 . Via the lemma, the Borel algebras have the form
b 0 = 1 2 ( h ( h ) ) b 1 = e ( f ) b 2 = i ( e f ) b 3 = i 2 ( h h )
for bor 1 , 3 ( + ) and
k 0 = 1 2 ( h ( h ) ) k 1 = ( f ) e k 2 = i ( f e ) k 3 = i 2 ( h h )
for the opposite bor 1 , 3 ( ) .
The ( 2 k + 1 ) -dimensional representation π ( k ) of sl 2 ( C ) is
π ( k ) ( h ) | k , m = 2 m | k , m π ( k ) ( e ) | k , m = ρ m ( k ) | k , m + 1 π ( k ) ( f ) | k , m = ρ m 1 ( k ) | k , m 1
with ρ and k as in Equation (26).
Theorem 4. 
Let 2 k N and let ( π , V ) be a simple representation of sl 2 ( C ) of dimension 2 k + 1 . Then,
1. 
π is equivalent to π ( k ) for some k;
2. 
the eigenvalues of π ( k ) ( h ) are { 2 k , 2 k 2 , , 2 k } = Spec π ( k ) ( h ) ;
3. 
if 0 v V satisfies π ( k ) ( e ) v = 0 , then π ( k ) ( h ) v = 2 k v , i.e., π ( k ) ( h ) and π ( k ) ( e ) have the common eigenvector | k , k ;
4. 
if 0 v V satisfies π ( k ) ( f ) v = 0 , then π ( k ) ( h ) v = 2 k v , i.e., π ( k ) ( h ) and π ( k ) ( f ) have the common eigenvector | k , k .
(Theorem 19.2.5 in Ref. [33], p. 281).
As a matter of fact, points 3 and 4 generate/define the eigenvectors of the representation π ( k , l ) , called the helicity states for bor 1 , 3 . Using the sl -decomposition (27) of bor 1 , 3 , one obtains
π ( k , l ) ( t 0 ) = 1 2 π ( k ) ( h ) 1 2 l + 1 π ( k , l ) ( t + ) = i π ( k ) ( e ) 1 2 l + 1 π ( k , l ) ( u 0 ) = 1 2 k + 1 1 2 π ( l ) ( h ) π ( k , l ) ( u + ) = 1 2 k + 1 i π ( l ) ( f )
and, with respect to the direct product basis | k , l ; m k , m l = | k , m k | l , m l , where k m k k and l m l l , one obtains the 2 l + 1 common eigenvectors for sol 2 ( e ) , given by
π ( k , l ) ( t 0 ) | k , l ; k , m l = k | k , l ; k , m l , π ( k , l ) ( t + ) | k , l ; k , m l = 0
for m l = l , l + 1 , , l . The 2 k + 1 common eigenvectors for sol 2 ( f ) are given by
π ( k , l ) ( u 0 ) | k , l ; m k , l = l | k , l ; m k , l , π ( k , l ) ( u + ) | k , l ; m k , l = 0
for m k = k , k + 1 , , k . Thus, the sol 2 -invariant subspaces of the representations of the proper Lorentz group, represented by the two components of the Kronecker sum as “left-handed” and “right-handed” states, lead to the concept of helicity. Accordingly, the (group-theoretical version of the) “Weinberg ansatz” is based on the concept of helicity.

7. The Weinberg Ansatz

Considering the states ( m k , m l ) as points on a lattice of dimension ( 2 k + 1 ) × ( 2 l + 1 ) , according to Equations (30) and (31), the eigenstates of sol 2 are found with the values m k = k and m l = l , i.e., at the boundary of this lattice. This can be interpreted physically as a constraint on the spin degrees of freedom of the massless particle to only one of the helicity states. It is even possible to show that, for particles, this is the left-handed helicity state [7]. If, however, the particle is equal to its antiparticle, the full spectrum of helicity states is available. In general terms, this is formulated in Refs. [15,16] in the following way:
  • If a massless particle is equal to its antiparticle, it is described by the irreducible representation ( k , k ) of the proper orthochroneous Lorentz group (Majorana case);
  • If a massless particle is not equal to its antiparticle, the particle is described by the irreducible representation ( k , 0 ) of the proper orthochroneous Lorentz group, while the antiparticle is described by the irreducible representation ( 0 , k ) of the proper orthochroneous Lorentz group (Dirac case).
Note that the massless particle is defined via the Borel subgroup by the irreducible representation of the proper orthochroneous Lorentz group, without the need to introduce space inversion.

7.1. The Majorana Case

For the representation ( k , k ) , the ansatz yields 2 k + 1 helicity states associated with sol 2 ( e ) ,
| k , k ; k , k + p , p = 0 , 1 , , 2 k ,
and 2 k + 1 helicity states associated with sol 2 ( f ) ,
| k , k ; k + p , k , p = 0 , 1 , , 2 k .
Since the state | k , k ; k , k is twice and at the same time excluded by the condition
D 3 ( k , k ) | k , k ; k , k = 0 , B 3 ( k , k ) | k , k ; k , k = 2 k | k , k ; k , k ,
the particle with zero mass and helicity λ = 2 k has 4 k helicity states. In particular, the defining representation ( 1 2 , 1 2 ) describes a massless particle with helicity 1 and two helicity states,
| 1 2 , 1 2 ; 1 2 , 1 2 and | 1 2 , 1 2 ; 1 2 , 1 2 .

7.2. The Dirac Case

According to Weinberg, the Dirac case ( k , 0 ) ( 0 , k ) is a particular case of the general situation ( k , l ) with l = 0 or k = 0 , respectively. For the representation ( k , 0 ) , there exists only a single eigenvector | k , 0 ; k , 0 of the Borel algebra bor 1 , 3 ( k , 0 ) , i.e., only a single helicity state | k , 0 ; k , 0 with
t 0 | k , 0 ; k , 0 = k | k , 0 ; k , 0 , t + | k , 0 ; k , 0 = u 0 | k , 0 ; k , 0 = u + | k , 0 ; k , 0 = 0 .
Similarly, the only helicity state for the representation ( 0 , k ) is | 0 , k ; 0 , k with
u 0 | 0 , k ; 0 , k = k | 0 , k ; 0 , k , u + | 0 , k ; 0 , k = t 0 | 0 , k ; 0 , k = t + | 0 , k ; 0 , k = 0 .
For example, the fundamental representation bor 1 , 3 ( 1 2 , 0 ) can be expressed as
b 0 ( 1 2 , 0 ) = 1 2 h , b 1 ( 1 2 , 0 ) = e , b 2 ( 1 2 , 0 ) = i e b 3 ( 1 2 , 0 ) 1 2 i h .
The corresponding representation of the algebra sol 2 ( e ) has the form
t 0 ( 1 2 , 0 ) = 1 2 h , t + ( 1 2 , 0 ) = i e
with sol 2 ( f ) being trivial. Therefore, in the case of the irreducible representation ( 1 2 , 0 ) , there exists only a single solution e 1 = ( 1 , 0 ) T , i.e., a helicity state λ = 1 / 2 , and this helicity state is equal to the solution of the Weyl equation
σ ˜ μ p μ ψ ( p ) = 0 .
In case of the representation ( 0 , 1 2 ) , sol 2 ( e ) is trivial, and the nontrivial algebra sol 2 ( f ) is of the form
u 0 ( 0 , 1 2 ) = 1 2 h , u + ( 0 , 1 2 ) = i f
with only a single common eigenvector e 1 = ( 0 , 1 ) T .

8. Conclusions

With this work, we have delved into the rich solvable structure of the proper Lorentz group. As for a massless particle, the stabilizer subgroup of the momentum four-vector is given by the Borel subgroup as the maximal noncompact subgroup of the Lorentz group, of which the Lorentz group contains two copies. Thus, we can generate the Borel subgroup as a Kronecker sum of two copies of the simplest solvable algebra sol 2 sl 2 and, correpondingly, the proper Lorentz group as a Kronecker sum of two copies of the simplest noncompact algebra sl 2 . This is formulated in Lemma 1. From our investigation in this paper, we conclude that, if there is a particle state with pure helicity or spin, the mass of this particle is zero and the stabilizer subgroup is the Borel subgroup, fixing the line of the light-like propagation. Therefore, at least for the electromagnetic field, the symmetry determines the dynamics.

Author Contributions

Conceptualization, R.S.; methodology, R.S. and S.G.; validation, S.G.; formal analysis, R.S.; investigation, R.S. and S.G.; writing—original draft preparation, R.S.; writing—review and editing, S.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

All data produced can be found in the manuscript.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
Lor 1 , 3 proper Lorentz group SO 0 ( 1 , 3 )
lor 1 , 3 proper Lorentz algebra so 0 ( 1 , 3 )
Bor 1 , 3 Borel subgroup of the Lorentz group SO 0 ( 1 , 3 )
bor 1 , 3 Borel subalgebra of the Lorentz algebra so 0 ( 1 , 3 )
( Bor 1 , 3 ) u unipotent radical of Bor 1 , 3
Der g derived Lie algebra of the Lie algebra g
rad u ( ± ) Lie algebra of the unipotent radical ( bor 1 , 3 ) u = Der bor 1 , 3 ( ± )
Rad u ( ± ) exponential of rad u ( ± )
Tor 1 , 3 maximal torus group of the Lorentz group SO 0 ( 1 , 3 )
tor 1 , 3 maximal torus algebra of the Lorentz algebra so 0 ( 1 , 3 )
lg Wigner’s little group (point fixgroup)
T translational group
GL n ( K ) general linear group of dimension n over the field K
SL n ( K ) special linear group of dimension n over the field K
sl n ( K ) Lie algebra corresponding to SL n ( K )
sol 2 two-dimensional solvable Lie algebra

References

  1. Choi, T.; Cho, S.Y. Spin operators and representations of the Poincaré group. arXiv 2018, arXiv:1807.06425. [Google Scholar]
  2. Nisticò, G. Group theoretical derivation of consistent massless particle theories. arXiv 2021, arXiv:2106.12206. [Google Scholar] [CrossRef]
  3. Dragon, N. Bundle Structure of Massless Unitary Representations of the Poincaré Group. Int. J. Theor. Phys. 2024, 63, 149. [Google Scholar] [CrossRef]
  4. Sazdović, B. Poincare field theory for massive particles. arXiv 2024, arXiv:2410.12549. [Google Scholar] [CrossRef]
  5. Sazdović, B. Poincare field theory for massless particles. arXiv 2024, arXiv:2410.12969. [Google Scholar] [CrossRef]
  6. Saar, R.; Groote, S. Mass, zero mass and … nophysics. Adv. Appl. Clifford Algebras 2017, 27, 2739–2768. [Google Scholar] [CrossRef]
  7. Groote, S.; Saar, R. A solvable algebra for massless fermions. Symmetry 2024, 16, 97. [Google Scholar] [CrossRef]
  8. Joos, H. On the Representation theory of inhomogeneous Lorentz groups as the foundation of quantum mechanical kinematics. Fortsch. Phys. 1962, 10, 65–146. [Google Scholar] [CrossRef]
  9. Niederer, U.H.; Raifeartaigh, L.O. Realizations of the unitary representations of the inhomogeneous space-time groups. 1. general structure. Fortsch. Phys. 1974, 22, 111–129. [Google Scholar] [CrossRef]
  10. Niederer, U.H.; Raifeartaigh, L.O. Realizations of the unitary representations of the inhomogeneous space-time groups. 2. covariant realizations of the poincare group. Fortsch. Phys. 1974, 22, 131–157. [Google Scholar] [CrossRef]
  11. Borel, A. Linear Algebraic Groups; Springer: New York, NY, USA, 1991. [Google Scholar]
  12. Poincaré, H. Sur la dynamique de l’électron. Rend. Del Circ. Mat. Di Palermo 1906, 21, 129–176. [Google Scholar] [CrossRef]
  13. Wigner, E.P. On Unitary Representations of the Inhomogeneous Lorentz Group. Annals Math. 1939, 40, 149–204, Reprint in Nucl. Phys. Proc. Suppl. 1989, 6, 9. [Google Scholar] [CrossRef]
  14. Wigner, E.P. Unitary Representations of the inhomogeneous Lorentz Group including Reflections. In Lecture at the Istanbul Summer School of Theoretical Physics; Gürsey, F., Ed.; Gordan and Breach: New York, NY, USA; London, UK, 1962; pp. 37–80. [Google Scholar]
  15. Ohnuki, Y. Unitary Representations of the Poincaré Group and Relativistic Wave Equations; World Scientific: Singapore, 1976. [Google Scholar]
  16. Weinberg, S. The Quantum Theory of Fields; Cambridge University Press: Cambridge, UK, 1995. [Google Scholar]
  17. McRae, C. A convenient Prepresentation Theory of Lorentzian Pseudo-Tensors: P and T in O(1, 3). arXiv 2025, arXiv:2501.05400. [Google Scholar]
  18. Humphreys, J.E. Linear Algebraic Groups; Springer: New York, NY, USA, 1995. [Google Scholar]
  19. Lie, S. Theorie der Transformationsgruppen. Abhandlung II. Arch. Math. Naturvidenskab 1876, 1, 152–193. [Google Scholar]
  20. Kolchin, E.R. Algebraic matric groups and the Picard–Vessiot theory of homogeneous linear ordinary differential equations. Ann. Math. 1948, 49, 1–42. [Google Scholar] [CrossRef]
  21. Goodman, R.; Wallach, N.R. Symmetry, Representations, and Invariants; Springer: New York, NY, USA, 2009. [Google Scholar]
  22. Tung, W.-K. Group Theory in Physics; World Scientific: Singapore, 1999. [Google Scholar]
  23. Croft, M.; Todd, H.; Corbett, E. The Wigner Little Group for Photons Is a Projective Subalgebra. Adv. Appl. Clifford Algebras 2025, 35, 11. [Google Scholar] [CrossRef]
  24. Shaw, R. Unitary representations of the inhomogeneous Lorentz group. Nuovo Cim. 1964, 33, 1074–1090. [Google Scholar] [CrossRef]
  25. Bump, D. Lie Groups. In Graduate Text in Mathematics; Springer: New York, NY, USA, 2004. [Google Scholar]
  26. Saller, H. Operational Spacetime. In Interactions and Particles; Springer: New York, NY, USA, 2010. [Google Scholar]
  27. Mendes, R.V. Complex hidden symmetries in real spacetime and their algebraic structures. arXiv 2025, arXiv:2501.12960. [Google Scholar]
  28. Kwoh, D. Princeton University, Princeton, NJ, USA. Unpublished Work. 1970. [Google Scholar]
  29. Wightman, A.S. Relativistic wave equations as singular hyperbolic systems. Proc. Symp. Pure Math. 1973, 23, 441. [Google Scholar] [CrossRef]
  30. Bäuerle, G.G.A.; de Kerf, E.A. Lie algebras—Finite and Infinite Dimensional Lie Algebras and Applications in Physics; Elsevie: Amsterdam, The Netherlands, 1990. [Google Scholar]
  31. Barut, A.O.; Raczka, R. Theory of Group Representations and Applications; World Scientific: Singapore, 1986. [Google Scholar]
  32. Campoamor-Stursberg, R.; Marrani, A.; de Traubenberg, M.R. An infinite-rank Lie algebra associated to SL(2,R) and SL(2,R)/U(1). J. Math. Phys. 2024, 65, 081702. [Google Scholar] [CrossRef]
  33. Tauvel, P.; Yu, R.W.T. Lie Algebras and Algebraic Groups; Springer: New York, NY, USA, 2005. [Google Scholar]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Saar, R.; Groote, S. Zero Mass as a Borel Structure. Symmetry 2025, 17, 1464. https://doi.org/10.3390/sym17091464

AMA Style

Saar R, Groote S. Zero Mass as a Borel Structure. Symmetry. 2025; 17(9):1464. https://doi.org/10.3390/sym17091464

Chicago/Turabian Style

Saar, Rein, and Stefan Groote. 2025. "Zero Mass as a Borel Structure" Symmetry 17, no. 9: 1464. https://doi.org/10.3390/sym17091464

APA Style

Saar, R., & Groote, S. (2025). Zero Mass as a Borel Structure. Symmetry, 17(9), 1464. https://doi.org/10.3390/sym17091464

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop