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
is the maximal connected compact simple subgroup of the Lorentz group
, and the corresponding time-like energy–momentum hyperboloid is
. In this paper, we show that, in contrast to this, for massless particles, one obtains the line fixgroup
as being the maximal connected noncompact solvable subgroup of the Lorentz group,
Here,
is the Borel subgroup of the Lorentz group, and at least one of the manifolds
is projective (
).
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
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
. 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
, 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
. 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
preserves the indefinite symmetric metric
in the real spacetime
,
As the defining representation of the causality-compatible Lorentz group,
is parametrizable by six real parameters
,
where the first part is compact and the second part noncompact, and the domain of these six parameters is given by
which is homeomorphic to
, where
is the three-dimensional projective space. The generators have the form
where, in general,
.
Here,
,
is the Euclidean basis in
. Therefore,
is a locally compact and doubly connected, path-connected, simple and reductive group with universal coverage
, i.e.,
The last isomorphism means that the representations of
may be seen as representations of the complex rotation group
. For completeness, note that the Lie algebra
is the noncompact real form of
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
acts on the flat spacetime or its dual energy–momentum space by a linear transformation,
Theorem 1 (Chevalley)
. Let G be a linear algebraic group and a closed algebraic subgroup. Then, there is a rational representation and a one-dimensional subspace such thatOtherwise, if spans the line L, i.e., , then the equationdefines a character of H. This character is called the weight to the semi-invariant ℓ, and . Note that, if , then , i.e., 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 be a regular representation of G. Then, there exist characters , and a flagsuch thatfor all . Taking , one obtains the characteristic equationi.e., every solvable group has a common one-dimensional subspace [19,20]. Wigner’s Little Group
What follows was proposed by Eugene Paul Wigner in 1939: for massive particles (
), the point fixgroup (or little group) of the momentum
= (1, 0, 0, 0), and
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
. 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 is a projective variety;
- 4.
, where B is a fixed Borel subgroup of G.
(see Theorem 11.4.7 in Ref. [21], p. 524). Thus, every element
is contained in a Borel subgroup. As mentioned before, a solvable group has a semidirect structure. For the connected solvable group
G, the set
of unipotent elements is a closed connected nilpotent subgroup of
G. There exists a maximal torus
and, for this, an exact sequence
Since the maximal torus
and the maximal connected unipotent subgroup
of
G are those for the Borel subgroup, one has the exact sequence
and
. Here,
, where
is the unipotent component in the Jordan decomposition
.
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 the squared mass, for massive particles, one can move to the rest frame, where . The stabilizer subgroup or fixgroup is given by , 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 . This is the little group that Wigner indicates for massless particles like photons and (massless) neutrinos. For instance, for a momentum vector pointing in z direction, 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 .
Returning to mathematics, let
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
, one has the commutant
. Therefore, the character group
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
with multiplicity one.
The little group of Wigner is
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
,
In this semidirect product, the compact group
acts on the abelian locally compact group
by the multiplication rule
Every irreducible representation of
is equivalent either to a character of
lifted to
or to an induced representation
, where
is the nontrivial character of
. To avoid a continuum of helicity states, one has to require that, for physical states, the noncompact part of
is trivial in all representations, so the little group reduces to
. There are topological considerations that restrict the allowed values of the helicity to integers or half-integers. Thus, the helicity
is Lorentz-invariant for massless particles with the total angular momentum
.
As for any abelian group, the reducible representations of 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 of their helicity.
In nature, there are two classes of particles. The first class consists of particles that can exist in two helicity states
. 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
-doublet
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
and antineutrinos with helicity
.
4. The Borel Subgroup
The key observation in the preceding sections was the character Equation (
3). Solving the character equation
for the light-like standard vector
with
, one obtains [
6,
7]
with
Here,
,
,
As a consequence, the character is given by
The composition rule is
and therefore
All such transformations
with a noncompact parameter space for helicity and gauge, given by
, form the Borel subgroup
Here,
is the maximal torus in
, and
is the unipotent radical of
.
The linearization of
in the neighborhood of the identity
results in the Lie algebra
, with
As a vector space
endowed with commutation relations
the Borel algebra reads
where
is the Lie algebra corresponding to the unipotent radical.
Since the semisimple rank is
, there exists a unique Borel subgroup
, called opposite
, such that [
18,
24,
25,
26]
and
.
To visualize the algebraic structure of
, it is convenient to transform the basis to
with
. This leads to the Kronecker sum decomposition
The Kronecker sum decomposition is easily seen after applying the splitting map
Here,
is the natural Chevalley basis of
.
Again, the linearization of
generates the basis of the underlying vector space as
with nonzero commutation relations
Thus,
where
.
The Kronecker sum decomposition of
into two fundamental solvable groups
is easily extended to an
decomposition for the whole
,
Using the splitting map, for
, one obtains
with commutation relations
For
, one obtains
with commutation relations
and
.
Returning from the defining matrix representation of the Borel algebra to the group matrix representation, by exponentiation, one obtains
with
The Cartan–Killing form in the defining representation is indefinite,
and
For the unipotent radical, one has
Since the algebra
is solvable, and its derived algebra is given by
the Cartan–Killing form is identically zero,
. For
, one obtains
, so, obviously,
nilponent and
unipotent.
5. The Quotients
Given a closed subgroup
H of an algebraic group
G, there is a smooth projection
, where the fibers are precisely the cosets
,
. The projection
has a smooth local injection, given by the compatible section
such that
.
As an example, we consider the Cartan decomposition
, where
is the maximal compact subalgebra of
, and the subspace
consists of the noncompact generators of
. Exponentiating the Lie algebra decomposition into the Lie group,
one obtains the parametrization of the algebraic manifold
.
The map of
onto
G
is a diffeomorphism into
G, i.e.,
.
Different choices of the section give different formulae for the coset representatives. For the Borel subgroup , the factor set is the largest homogeneous space for G, having the structure of a projective variety. Since is complete, the Borel subgroup B has a fixed point in .
5.1. The (3) Parametrization
The Borel decomposition of the Lorentz group
is generated by the decomposition of the algebra
in a natural way by reordering the usual parametrization
More precisely,
Here,
The Cartan–Killing inner product for representatives
is negative, i.e., the representatives are compact operators. For the representatives of the group coset, one has
Therefore, the geometric manifold for the respresentatives of the group coset is compact. Moreover, the parametrization of the
-classes can be given by the three-point
in
as
and
. Then,
with
. Therefore, the
-type coset representatives generate a compact factor set, the two-sphere in
[
27],
Using the Iwasawa decomposition
and
, one obtains the same result.
The projective coordinates for this parametrization are
As mentioned, different choices for the section
provide different parametrizations.
5.2. The (1, 2) Parametrization
Let
Here,
is noncompact, as the Cartan–Killing form is positive,
The group representatives
are noncompact,
as the Cartan–Killing form is positive,
The parametrization of the coset representations for the group can be chosen as the point
in
or as the projective coordinates
,
in the interior of the unit circle of the
plane. In the first case, we have
so that
, and the representatives are
with
and
. As a consequence, the representatives are found on the noncompact on-shell hyperboloid
The projective coordinates
are
Therefore, the two examples for the coset representatives considered up to now are the time-like on-shell hyperboloid and (as a compact partner) the sphere with the common compact subgroup given by .
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
As underlying vector spaces, one has
It is convenient to choose the basis of
as
Then,
where
and the Cartan–Killing product
. This is the second criterion for the solvability of
: an algebra
is solvable if and only if its Cartan–Killing metric tensor is identically zero on its derived algebra
.
The expression for the representatives of the group coset is obtained by exponentiation:
From this, the representations of the group coset can be read as
where
,
,
and
. One observes that the real parameters
describe the subspace of the noncompact elliptic paraboloid. The projective coordinates of this parametrization are
The Cartan–Killing form for the representatives of the group coset is positive:
As a consequence, one has the following:
: a rest frame exists for massive particles ⇒
stabilizer subgroup is the point fixgroup ⇒ leading to spin;
: no rest frame exists for massless particles ⇒
stabilizer subgroup is the line fixgroup ⇒ 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
is a quantum-mechanical problem in the form of Planck’s formula
, rather than a problem of symmetry. In fact, the difference is a mathematical one:
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
. 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
can be extended to a unique linear complex representation
by the holomorphic extension
Although
, the process of holomorphic extension for
arises from the fact that
is the real form of
,
This complexification of
provides the link between the real-valued Lorentz algebra
and the real-valued algebra
, and, using this link, one can construct all the representations of
.
Since a representation
of
comes from a representation
of
, the diagram
has to be commutative. However, for odd
m, there is no such representation
. In order to overcome this problem, one needs
to generate all the finite-dimensional representations of
. The complex representations
(
) may be obtained by holomorphic extension and Weyl’s unitary trick,
Here,
generate the algebra
and
,
are the common representations of
.
Any irreducible finite-dimensional representation of
is isomorphic to
for some
. 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
is defined by the commutative diagram
The representation
is defined as
Here,
,
, with
and
(
) being the Pauli matrices. Since the finite-dimensional representations of
are in one-to-one correspondence with those of
, and the Lie algebra
is noncompact, it is reasonable to define the representations of
in terms of the algebra
. 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
. For example,
contains three such
-isomorphic subalgebras but only one
-isomorphic subalgebra. The following lemma gives the
-structure for
.
Lemma 1. If , are defined bywhere are the generators of the algebra ,then generate the Lorentz algebra , Applying this lemma, one can define the representations of
in terms of the holomorphic extensions of the irreducible representations
of
[
24,
25,
26,
32],
(
). Here,
is the standard representation of
,
with
,
.
In some contexts, it is more convenient to work with the
-basis
,
, and
. Via the lemma, the Borel algebras have the form
for
and
for the opposite
.
The
-dimensional representation
of
is
with
and
k as in Equation (
26).
Theorem 4. Let and let be a simple representation of of dimension . Then,
- 1.
π is equivalent to for some k;
- 2.
the eigenvalues of are ;
- 3.
if satisfies , then , i.e., and have the common eigenvector ;
- 4.
if satisfies , then , i.e., and have the common eigenvector .
(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
, called the helicity states for
. Using the
-decomposition (
27) of
, one obtains
and, with respect to the direct product basis
, where
and
, one obtains the
common eigenvectors for
, given by
for
. The
common eigenvectors for
are given by
for
. Thus, the
-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.
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 and, correpondingly, the proper Lorentz group as a Kronecker sum of two copies of the simplest noncompact algebra . 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.