1. Introduction
This paper deals with group contractions, mainly the contraction of one Lie group into another. In physics, we are quite familiar with the proper rotation group
, which is a Lie group. The Lorentz and Poincaré (inhomogeneous Lorentz) groups are Lie groups. The Lorentz group, which serves as the foundational framework for Einstein’s special theory of relativity, can also be considered as an effective mathematical instrument in quantum mechanics. For example, Mukunda et al. [
1] studied the symmetrical behavior of quantized classical equations of motion. Sprung et al. [
2] used the Lorentz transform and the Dirac spinor in 2 + 1 dimensions to give a geometrical interpretation of the action of the transfer matrix on a scattering wave function, and then an analogy between this action and a Lorentz transform on the unit hyperboloid in 2 + 1 dimensions was developed. Blasiak et al. [
3] investigated classical mechanics as related to their quantum analogues, where they illustrated their approach to defining a one-dimensional harmonic oscillator by means of generalized coherent states.
The Lorentz group serves as well in optical sciences, both in the classical and quantum domains. In the classical domain, Simon and Mukunda [
4] considered both coherent and partially coherent beams in first-order optics. Bastiaans and Alieva [
5] found the eigenfunctions of combinations of a magnifier, a lens, and a shearing operator to determine propagation through a first-order optical system. Nazarathy and Shamir [
6] studied the canonical transformation to find relations between the operator representation of wave optics and geometric ray optics. In the quantum domain, Mandel and Wolf [
7], first treated the broad area that deals with the coherence and fluctuation of light; then, they discussed the nonclassical states of light, higher-order squeezing, and quantum effects of down-conversion.
The promising field of quantum computations and information is not exempt from relying on the symmetries generated by
, which is the double cover of the proper Lorentz group. Generally, they turn out to be very effective in dealing with ideas like entanglement. For qubit systems,
is regarded as the invariance group. For n-qubits, it becomes
. Then, as shown in Ðoković and Osterloh [
8], it is possible to construct polynomial invariants for several qubits in order to classify a set of measures of multipartite entanglement. In Heydari [
9], after defining a multilocal Lorentz-group invariant of
, the properties of measures of entanglement were studied. As another example, Teodorescu-Frumosu and Jaeger [
10] showed how the natural quantum Lorentz-group invariant group length can be used to study an arbitrary number of qubits. Additionally, they showed that this invariant length can be utilized to describe entanglement.
Entangled states have also been studied within a similar context. Başkal et al. [
11] showed that in the Lorentz-covariant world, two coupled harmonic oscillators can be transferred to the concept of entanglement. Likewise, Rangamani and Takayanagi [
12] discussed a situation in which holographic entanglement entropy computations respect AdS (anti-de Sitter) isometries and examined how conformal symmetry (including those within
) shapes entanglement structures in quantum field theory.
In this paper, a larger Lorentz group,
, is also studied. Applications of this group are numerous. For instance,
is involved in the extensions of squeezed states [
13,
14]. Arvind and Mukunda and Colas et al. [
15,
16] investigated the two- and four-mode symmetries of squeezed states using the
or the locally isomorphic
group. Since this group contains two coupled harmonic oscillators,
can serve as a physical basis for symmetry decomposition. Furthermore, they also discussed coupled harmonic oscillators in the context of coupled Hamiltonians, as did Fring and Tenney [
17], and in the description of the decoherence process as a symmetry transformation in the
space. The decoherence of two beams of light can be represented by the density matrix, the Poincaré sphere, and the Stokes parameters. The Stokes parameters under various optical transformations can form a Minkowski four-vector, resulting in a two-by-two representation of the Lorentz group. A geometric presentation of the Lorentz group is given by the Poincaré sphere. The Lorentz group preserves the determinant of the density matrix and therefore cannot accommodate the decaying of the off-diagonal elements of the density matrix. This decay results in a decrease in the determinant. The
group contains two Lorentz subgroups. As a consequence, the change in determinant of one Lorentz group is compensated by the change in the other group [
18].
When Einstein formulated his theory of special relativity, he was interested in point particles without internal space-time structures. The focus later shifted to understanding how moving observers perceive internal spacetime structures. To address this question, in 1939, Wigner [
19] considered the subgroups of the inhomogeneous Lorentz group whose transformations leave the momentum of a relativistic particle invariant. These little groups constitute the irreducible representations of the Poincaré group. They transform the internal spacetime structure of the particles. For instance, electrons can have, in addition to energy and momentum, spin degrees of freedom, and photons have helicity and gauge degrees of freedom.
Another interesting manifestation of irreducible representations of the Poincaré group using little groups has recently been introduced through the study of continuous spin particles [
20]. These massless particle states are characterized by a real parameter, often referred to as the continuous spin parameter. They encompass infinitely many helicity states that mix under Lorentz transformations. This mixing reflects the nontrivial action of the Lorentz group on the helicity states. This property distinguishes continuous spin particles from ordinary massless particles of fixed helicity. When the continuous spin parameter is set to zero, the helicity states decouple from one another, reducing to a set of standard massless states in which each helicity appears exactly once [
21,
22]. For a review on this subject, see Bekaert et al. [
23].
In studying the Lorentz group, it is often convenient to transform a given group into a simpler group. For instance, the surface of Earth, because the radius is very large, appears to be flat when we cover short distances, but we know that it is basically spherical when traveling in an airplane. The rotation group applies to spheres, while the two-dimensional Euclidean group operates on flat surfaces. The motions on the plane of
can be reconciled with those on the spherical surface when we transform one group into another, using a technique called group contraction. Inönü and Wigner [
24] first introduced the method of group contraction as a limiting procedure.
Later on, several contraction schemes were introduced from a range of viewpoints for different purposes. For instance, the infinite momentum limit received some considerations from the view point of high-energy particle physics [
25]. We can note [
26,
27,
28,
29], where the infinite momentum has been significant from a group-theoretical perspective. Another proposal by Cattaneo and Wreszubski [
30] includes a complex Hilbert space, where a theory of contractions of Lie algebra representations is applied to
(the proper rotation group in two dimensions) into the
(the three-dimensional Heisenberg group), which has simultaneously yielded to the analysis of the limit
of a quantum system of N identical two-level particles. In Khan [
31], the author contracted the conformal algebra, which is larger than the Poincaré group, to the continuous spin representation of the Poincaré group by compactifying the third transverse direction to a circle of radius
R. Continuous spin representations can also arise from the five-dimensional Poincaré group through a combination of group contraction and Kaluza–Klein radius of the fifth dimension as the contraction parameter [
32]. Contractions of the de Sitter (i.e.,
Lorentz) group were examined by Evans in [
33] and also by Enayati el al. in [
34], where they were studied within the conceptual structure of cosmology.
The formulation of Wigner’s little groups relies on two-by-two matrices, a structure also fundamental to classical ray optics, making the mathematical connection readily apparent. In that context, we discuss the Lie group for the two-by-two representation of the little groups. The principles governing relativistic particles and classical optics may appear to be unrelated, yet the theory of Lie groups and Lie algebras offers a good start for analogous calculations, elucidating the underlying symmetries and transformations that govern both fields.
Before we start the main body of the paper, we make a few definitions that will be useful as we continue. We take
unless it is explicitly stated. A Lie group [
35], is defined as a smooth manifold, where group operations, such as multiplications and inversions, are both smooth maps. The properties of Lie groups are determined by a closed set of generators which is known as the Lie algebra of the group.
Locally, the transformations corresponding to a Lie group can be obtained by exponentiating the generators of its Lie algebra, as explained below. We consider any complex valued square matrix such as
X; then, we have
From this, the correspondence to any parameter
can be derived from
Then, the algebra of a particular Lie group can be obtained by differentiation as
In this paper, we shall consider the four-dimensional spacetime manifold with coordinates
Then, Lorentz transformations are conventionally defined to be the group that preserves the inner product
, with
, where the Minkowski metric is taken to be
. Specifically, we have
where
stands for the components of the transformation matrix. This group is called the proper, orthochronous Lorentz group
when the Lorentz transformations are restricted to the condition that det
, i.e., no space reflections, and to
, i.e., no time inversion.
Our main focus is to illustrate how squeeze transformations appear in the scene as a relatively recent approach to contractions. We are also interested in constructing the covariant four-momentum generators of Einstein’s by contracting the group. To this end, we organize the paper as follows.
Section 2 gives the generators of the Lorentz group and gives the matrix representation of
which is the two-by-two covering group of the Lorentz group.
Section 3 gives an introduction to Wigner’s little groups.
Section 4 gives the two-by-two representation of the little groups. In
Section 5, we give a contraction scheme by using a very large
R. In
Section 6, we give an infinite-momentum/small-mass limit of massive and imaginary mass particles to a massless particle.
Section 7 gives an introduction to squeeze transformations. We contract the two-dimensional representation of the Minkowski vector as well as the rotation matrix on a plane as examples. In
Section 8, a plane and a cylinder tangent to a sphere are considered. We then formulate how the
and
groups can be contracted into the two-dimensional Euclidean group and into the cylindrical group. In
Section 9, we contract the Lorentz group
to the Poincaré group.
In
Section 10, we discuss an application to geometric optics in the form of the contraction of a ray going through a transparent spherical surface to a straight line and develop the lens maker’s formula. In
Section 11, we look at the
matrix of paraxial optics and inquire how we can use this to formulate Wigner’s little groups, namely,
-like,
-like, and
-like, from optical elements. Indeed, we can represent those three little groups using one convex lens.
Section 12 contains conclusions.
Appendix A gives a table of the generators for the Lorentz group
.
2. Generators of the Lorentz Group
The Lorentz group is a six-parameter Lie group where the generators form the Lie algebra. We start with the generators of rotations in the matrix form [
36,
37]:
In this formulation,
is the generator of rotation around the
x-axis,
is the generator of rotation around the
y-axis, and
is the generator of rotation around the
z-axis. These three rotation generators satisfy the commutation relations
which form the Lie algebra of the familiar rotation group.
The matrices which perform Lorentz boosts are
Then,
In this formulation,
is the generator of boosts along the
x-axis.
and
are boosts along the
y-axis and along the
z-axis, respectively. Note that these generators alone do not form the Lie algebra for a group. However, these boost generators, together with the rotation generators from Equation (
6), do form a group. These six generators form the Lie algebra for the Lorentz group
. There are therefore six generators of the Lorentz group, and they satisfy the commutation relations given in Equations (
7) and (
9). The rotation generators are anti-symmetric, while the boost generators are symmetric. Hence, there are two possible four-by-four representations of the Lorentz group: one with
and the other with
.
The Lie algebra of
, which is the double cover of the Lorentz group, can be constructed from the three Pauli matrices [
38,
39,
40], where the Pauli matrices are written as
We define
The explicit forms of the operators in Equation (
11) are
These two-by-two matrices satisfy the following commutation relations:
Apart from an imaginary
i, they are like those given in Equations (
7) and (
9). We note that in the complex representation of the Lorentz group generators given by Weinberg [
36] and others [
37,
41], these commutation relations take the same form. If the signs of the generators in Equation (
13) are changed, the commutation relations remain the same, that is, the commutation relations are invariant. This group, known as
, is the covering group for the Lorentz group, that is, it is homomorphic (onto, but not one-to-one) to the Lorentz group. For one group to be homomorphic to another group, there must be at least two sets of generators of the one group that correspond to one set of generators of the other group. For further details about homomorphism, see Carmeli [
42]. In the present case, since the proper rotation generators are anti-symmetric, but the boost generators are symmetric, there are least two sets of generators in
that correspond to a given set of
generators. For the proper Lorentz group, the sign of the boost generators is unambiguously defined in terms of the time and space variables. This is not the case for
; both signs must be considered.
6. Boosting of Massive and Imaginary Mass Particles to Their Massless Limits
Einstein’s formula,
, shows that in the large-momentum or zero-mass limit the energy–momentum relation for massive particles becomes that for massless particles. In accordance with this line of reasoning, it has been suggested that the internal symmetry of massive particles becomes that of massless particles in the same limit [
45].
The group generated by
,
, and
given in Equation (
6) is the little group for a massive particle at rest. Should this massive particle move along the
z direction comparable to the speed of light, we know that
remains invariant, and the eigenvalue is the helicity. On the other hand, it is to be expected that in the infinite-momentum limit,
and
will go through certain changes.
We saw in
Section 3.2 that the generators,
,
, and
, form the Lie algebra for the
-like little group for massless particles moving along the
z direction and that
values have the form
and thus are related to rotation and boost generators. The four-momentum of massless particles remains invariant under transformations generated by the
,
, and
operators, where
generates rotations and the translation-like generators, while
and
generate gauge transformations [
46,
47,
57].
If a massive particle is at rest, the little group is
, which is generated by
. When this particle is boosted along the
z direction, its four-momentum becomes
, where the boost matrix is [
58]
Under this boost operation,
remains invariant as
However, the boosted
and
become
It is apparent that the Lorentz boosts in Equations (
58) and (
59) are similarity transformations; therefore, the
commutation relations will still be satisfied by the
operators as
In order to obtain contractions, we again resort to Einstein’s equation:
. The right-hand side of this equation goes to zero for massless particles, thereby causing the values of
E and
to approach each other. Thus, while
approaches to one,
approaches to infinity. This suggests that we follow a similar limiting procedure of
Section 5, where the quantity
now serves a function analogous to the radius of the sphere. Therefore, in the same sprit, we rewrite Equation (
51) as
These give the same result as in Equation (
56). This process is similar to the group contraction process described in
Section 5, where in the present case, an
group for a massive particle is contracted to become an
-like group for a massless particle. In passing by, we note that here the contraction parameter goes to one, while the common practice takes it in such a way that it either goes to zero or infinity. However, tending to a finite limit is a strategy that has been discussed long ago by Suskind [
59], where the parameter is
and approaches one as the speed of the particle becomes close to that of light.
Let us now consider the boost matrix of Equation (
15). As noted in
Section 3.3, this will transform the four-momentum
, of an imaginary mass particle, along the
z direction, which becomes
. This vector, normalized with
, approaches its massless counterpart
as
. In fact, the same is true for the massive case, with
approaching
through the same scheme.
As for the generators, when boosted,
remains invariant, while
and
become as in Equation (
36). The generators
, and
satisfy the same Lie algebra as that of Equation (
34), which is an
-like little group.
It is possible to use contraction matrices with different parameters in the components. For instance, we can implement this boost matrix to obtain another
-like group from
. Within the same line of approach as above we simply formulate as follows:
The matrices
and
take the same form as in Equation (
21) in
Section 3.2 and, together with
, satisfy the commutation relations as in Equation (
23) for the
and the
-like little groups.
8. Contraction of SO(3) and SO(2,1) to E(2) and the Cylindrical Group
Let us start with the familiar three-dimensional sphere. Then, we shall consider both a tangent plane at the north pole and a cylindrical surface touching the equatorial belt. The transformation group on these tangential surfaces will be obtained using the squeeze procedure introduced in
Section 7.
We refer to
Figure 2 to compare two contraction schemes with equivalent outcomes in terms of their transformation matrices. The figure on the left represents the spherical surface of the earth. A flat plane tangent at the north pole can be achieved by using the limiting process of
Section 5. Additionally, a cylindrical surface tangent to the equatorial belt can be produced [
44]. As with the figure on the right, it can be seen that a planar surface can also be obtained by shrinking the
z-axis and expanding the area
on the
-plane by implementing squeeze operators. To achieve a cylindrical configuration whose axis is along the z direction, this squeeze process can be used to expand the
z-axis while shrinking the area on the
-plane.
It is possible to construct symmetry properties on these surfaces contracting the
rotation group to
. The Lie algebra of the rotation group was given in Equation (
7), and the generators were defined explicitly in Equation (
6).
On the tangential plane at the north pole, we generate transformations by rotations around the
z direction using the generator
. Translations are then generated along the
x and
y directions using the translation generators in Equation (
25), which represent translations in the
x- and
y-axes. The commutation relations of these three generators, which form the Lie algebra for
, were given in Equation (
24).
We now use a squeeze matrix similar to the one in
Section 7:
and we proceed in the following manner:
The generators
and
have been given in Equations (
6) and (
55).
We can apply the same technique as in Equation (
76) to
. This group is generated by the
,
, and
values that were given explicitly in Equations (
6) and (
8). Of course,
remains the same, and the resulting forms of
and
are similar to those given in Equation (
55).
We can write the transformation matrix for either
or
in exponential form as
For
, the matrix for that result has the form
while for
, the result is
When either the matrix in Equation (
78) or (
79) acts on the vector
, in the space where
is applicable, the form that results is indeed an
transformation.
We now consider a cylindrical plane attached to the equator. As can be seen in
Figure 2, we can accomplish this by expanding the
z-axis while, at the same time, the radius of the circle,
, becomes very small. The inverse of above procedure can be used to obtain explicit forms for
and
as
We find as before that
remains invariant, but
, and
, where they are defined as
As can be seen, this is again the same algebra as for
given in Equation (
55).
For the cylindrical contraction of
, we use the same technique as in Equation (
80). Then,
This set of generators also has the same form as that given in Equation (
81) for contracting the
rotation group to the cylindrical group. The transformations for these two groups are illustrated in
Figure 3.
9. Contraction of SO(3, 2) to the Poincaré Group
The Lorentz group
consists of three space dimensions and two time dimensions. They act on coordinates given as
, where the Minkowski metric on this coordinate space is chosen to be
. We explicitly give the five-by-five matrices representing the ten generators of the
group in
Table A1. From the table, it can be seen that the six
and
matrices, generating rotations and boosts in the
space, contain only zero elements in their first rows and columns. Thus, with respect to them, the
s coordinate remains invariant. This indicates that the the generators
and
constitute the familiar four-by-four matrices of the Lorentz group
operating in Minkowski space. The remaining three boost generators,
, produce boosts with respect to the
s variable. Additionally, the matrix
generates rotations between the two time variables
t and
s. These four matrices have elements only in the first row and column. Here, our goal is converting the
and
generators into the translation generators
and
which can be applied to homogeneous Lorentz group, of the space with
, represented by the six generators
and
to enlarge it into the Poincaré group.
In this section, we intend to obtain the translation-like generators of the Poincaré group by using the procedure introduced in
Section 5. For this purpose, a five-by-five contraction matrix will be employed, where
R is replaced by
. We could as well have employed a five-by-five squeeze matrix for the contractions; however, since here we are only interested in obtaining the translation-like generators, the form of the contraction matrix we use here is sufficiently simple for our purpose. This is illustrated in
Figure 4.
The form of the five-by-five contraction matrix is
We note that this matrix and the inverse leave the last four columns and rows invariant. This results in the invariance of the four-dimensional Minkowski subspace of
. For the generators
and
, this is not true.
The explicit five-by-five matrix for
and
, the translation matrices, as obtained from
and
, have the following forms:
We notice that these translation generators have zero elements in the first rows.
Now, the contraction procedure can be formulated as
These four contracted generators lead to the five-by-five transformation matrix
performing translations
in the four-dimensional Minkowski space. This means that the
group becomes the inhomogeneous Lorentz group governing the Poincaré symmetry for quantum mechanics and quantum field theory. In quantum mechanics, the four-momentum operator is
Thus, we can translate the matrix in Equation (
88) into the language of differential operators as
This contraction procedure produces four translation generators corresponding to the energy-momentum four-vector in the Lorentzian system. The on-shell condition leads to
which is widely known as Einstein’s
.
10. Physical Example I: Contraction of Convex Spherical Surface
We studied group contraction using an approximation of a spherical surface to a flat one ( to /cylindrical). Many branches of physics and engineering employ this procedure. Here, we take an example from geometric optics. Although a lens surface is approximately spherical, if the light rays could be expressed as a linear function, that would be useful. Let us examine that possibility.
The equation for a straight line in the
-coordinate system can be written as
For a given
x variable, the line is completely determined by the
y variable and the slope. Thus, the column vector
completely specifies the line. To move from one coordinate point to another, the transformation can be represented as
The right-hand side matrix above is known as the ray transfer matrix in classical optics [
61,
62]. It has the same triangular form as the ray transfer matrix given in Equation (
103).
Figure 5 illustrates a convex spherical surface of a transparent material that has a radius
R and an index of refraction
n.
Here, the angle of the ray to the axis of the surface will be assumed to be very small. Furthermore, the radius of the spherical surface is assumed to be much larger than the distance from the ray to the
y-axis. Then, from Snell’s law,
where air is assumed to have an index of refraction of 1. Since all the angles are small in the expression above,
and
Although there is no change in
y, the slope changes through
The rays thus have the slopes
a and
, where
a is in air, and
is in the refractive material. We ignore, in the spirit of the already discussed group contraction, the variation in the
x coordinate that is of order
[
63]. Thus, we can represent the relation between
a and
by the matrix
Suppose now that we have another convex surface of refractive material that has a radius of curvature
and that faces in the opposite direction. If the ray comes out from this refractive surface, the applicable matrix is
These two convex surfaces, if the distance between them is small, constitute a thin lens. If we now multiply the matrix of Equation (
99) from the left by the matrix of Equation (
100), the net effect is the matrix
where
The expression in Equation (
102) is known as the lens maker’s formula, and
f is the focal length of the thin lens. The matrix of Equation (
101) and the translation matrix of Equation (
94) are considered to be the basic building blocks comprising a group theoretical approach for modern geometrical optics [
61,
64].
11. Physical Example II: Camera Optics and Little Groups
By way of this example, we aim to show that the two-by-two representation of the little groups in this paper coincides with the matrix representation of the one-lens system. It will then be shown that the focal condition corresponds to the transition from one little group to another. However, the transition from one little group to another is a singular transformation. The transition between little groups will be made possible by considering the correspondence between different little groups. In this way, an analytic transformation of computations on a hyperbolic surface to a spherical surface can be achieved, and the parameters of the lens system with those of the little groups can be formulated.
The simplest lens system consists of the lens matrix and the ray transfer matrix [
61]:
where
f is the focal length, and
d is the separation distance between the two reference planes. Here, we take the focal length to be positive. If the object and image are
and
from the lens, respectively, the optical system is described by
The image becomes focused when the upper-right element of this matrix vanishes with
The diagonal elements are dimensionless; however, the presence of nondimensionless off-diagonal elements renders the
matrix intractable. Nonetheless, it can be decomposed into
with
In the camera configuration, both the image and object distances are larger than the focal length, and both
and
are negative. Thus, we start with the negative of the middle matrix of Equation (
106):
This matrix can further be renormalized so that the two diagonal elements become equal. To this end, it can be written as
with
Then, the core matrix becomes
with
where the latter is written in terms of the
and
x variables.
Now, our main concern is the core matrix
C of Equation (
112), which contains all the parameters of the system, and its upper-right element will vanish when the focal condition is satisfied. This form of the core matrix can be seen to be effective in cavity optics when the
power of the
matrix, which is composed of two mirrors, has to be evaluated [
64]. This form can also be examined in relation to the Lorentz group and Wigner rotations.
Now that we have an equi-diagonal matrix with dimensionless components, we can equate it to (a) squeezed rotation
, (b) an
-like transformation (like in Equation (49)), and (c) a squeezed boost
, with specific constraints defined below:
where the range of the angle variable
is between 0 and
, and
is positive. The expressions given in the right-hand side of Equations (114)–(116) take the same mathematical forms as those of the representations of the
-,
-, and
-like little groups. The two-by-two forms of Wigner’s little groups are given in Equations (48) and (49). Equations (114) and (116) are the squeezed (boosted) forms of the transformation matrices
and
of Equation (
48), respectively. Equation (115) is like Equation (49), which is the little group transformation for a massless particle.
The left-hand sides of Equations (114)–(116) have components such that all the parameters are determined from
, and
f of the lens optics, and if we gradually increase the value of
x, the upper-right element becomes zero, and then it should become positive. The right-hand side of the above expression cannot accommodate this transition, since the transition from one form to another form is a singular transformation. On the other hand, the core matrix of Equation (
112) is analytic in the
x and
variables when both
and
are greater than 1.
Therefore, we seek another form of a little group matrix whose components are analytically well behaved. For this purpose, we consider
, where the rotation and the boost matrices are taken from Equation (
48). This is the two-by-two representation of a transformation that preserves the momentum of a relativistic particle traveling in the minus
z direction [
44]. These little groups can be equated under certain restrictions:
The relations between the parameters of these groups are found to be as follows:
For Equation (117), the relations are
For Equation (118), it is
For Equation (119), they are
Let us note that the quantity , while changing sign from minus to plus, it has to pass through zero. This process is analytic with the parameters and .
Therefore, the matrix in the left-hand side of Equations (117)–(119) will provide the means to define those transformation parameters in terms of the lens parameters of Equation (
112), meanwhile circumventing the singularity problem and emerging from the limiting nature of the contraction procedure. Thus, one can establish a correspondence between lens optics and the transformations of the little groups. Furthermore, transformations from one little group to another can be achieved by adjusting focal conditions.
In terms of these parameters, the core matrix can be written as
Here, both sides have upper-right elements which are analytic as they go through zero. The parameters are now related by
and finally by
Therefore, we have the Lorentz transformation parameters
and
expressed in terms of the parameters of the one-lens system.
Here, we have shown how the physical lens parameters are related to the parameters of the group. The upper-right component of the matrix governs the focal condition. The lens parameters are adjusted to satisfy the focal condition, causing this component to become zero. Thus, it is the realization of this condition that makes the transition from one little group to another possible.
12. Conclusions
In this paper, we discussed a range of approaches to contraction procedures within the context of Wigner’s little groups. We started by introducing the well-known scheme of group contraction using an approximation of a spherical surface to a flat one. We then considered Einstein’s , which shows that in the large-momentum/zero-mass limit, the massive particle becomes a massless particle. Therefore, the limiting procedure is illustrated by converting massive and imaginary mass particles to a massless particle using the infinite-momentum limit.
Our main focus was on the contraction procedure based on squeeze transformations. This approach is relatively recent compared to old-standing limiting procedures. Moreover, as illustrated in
Figure 2, the squeeze transformation offers the advantage of providing a geometric picture comparable to the one with a tangent plane at the north pole of a very large sphere and a tangent cylinder around the equator. In this case, a planar surface can be achieved by shrinking along the
z-axis and expanding the circular area on the
-plane. Conversely, a cylindrical configuration can be obtained, whose axis is along the
z direction, by expanding along the
z-axis while shrinking the area on the
-plane. With this new approach, we contracted Wigner’s
little group to both the
and the cylindrical group. Similarly, Wigner’s little group, the Lorentz group
, was contracted to these same two groups.
As a further example of group contraction, the group was contracted into the Poincaré group, and the result was shown to lead to Einstein’s widely known .
Finally, two physical examples were presented. The first concerned a ray that emanates from a transparent spherical surface into a straight line, wherein we derived the lens maker’s formula. The second physical example illustrates how camera optics can be used to realize Wigner’s little groups from optical elements. It is seen that little groups corresponding to massive, massless, and imaginary mass particles can be transformed from one to another through the focusing condition of the image in a camera.
Despite the fact that laws governing relativistic particles differ significantly from the principles of classical optics, the theory of Lie groups and Lie algebras offers valuable analogous computations that can improve our understanding of symmetries and transformations in both fields.
Although it is a reappraisal of our earlier work, this paper is written from a fresh viewpoint that has not been previously presented.