Paper The following article is Open access

Kinematical versus dynamical contractions of the de Sitter Lie algebras

Published 7 October 2019 © 2019 The Author(s). Published by IOP Publishing Ltd
, , Citation Joachim Nzotungicimpaye 2019 J. Phys. Commun. 3 105003 DOI 10.1088/2399-6528/ab4683

2399-6528/3/10/105003

Abstract

We present two kinematical Lie algebras contraction processes to improve the Bacry and Lévy-Leblond contractions (H Bacry, et al, 1968 J. Math. Phys., 9, 1605–1614) :(speed-time, speed-space and space-time contraction). For the first one, we introduce kinematical parameters, namely the radius r of the Universe, the period τ of the Universe and the speed of light $c=r{\tau }^{-1}$. Next we present them as static, Newtonian and flat limits through the use of the dynamical parameters, namely the mass, m, the energy, E0 and the compliance C, all depending on mass as well as length and time. We consider that the second one as the best. To give a little physical taste for each kinematical Lie algebra, we set up the equations of change with respect each group parameter through the use of the Poisson brackets defined by the Kirillov form.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 3.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Group (algebra) contraction is a method which allows to construct a new group (algebra) from an old one. Contraction of Lie groups and Lie algebras started sixty six years ago with nonu and Wigner [1] in 1953, when they were trying to connect Galilean relativity and special relativity. Eight years later, in 1961, Saletan [2] provided a mathematical foundation for the Inonu-Wigner method. Since then, various papers have been produced and the method of contraction has been applied to various Lie groups and Lie algebras [310].

The method has also been used by Bacry and Lévy-Leblond [11] to connect the de Sitter Lie algebras to all other kinematical Lie algebras through three kinds of contractions: speed-space contractions, speed-time contractions and space-time contractions. The terminology is related to the fact that Bacry and Lévy-Leblond have, first of all, scaled the velocity-space generators, the velocity-time translation generators and the space-time translation generators by a parameter epsilon to obtain, in the limit $\epsilon \to 0$, the respective contractions that we prefer to call velocity-space contractions, velocity-time contractions and space-time contractions. The Lévy-Leblond contraction approach has been also extended to supersymmetry [12] and kinematical superalgebras [1315].

Within the corresponding eleven Lie groups, four of them, namely the Galilei group G governing the Newtonian physics (Galilean relativity), the Poincaré group P governing the Einstein physics (special relativity), the Newton-Hooke groups NH± describing Galilean relativity in the presence of a cosmological constant and the de Sitter Lie groups dS± governing the de Sitter relativity of a space-time in expansion or oscillating universe, are well known in physics literature.

Within the remaining five ones, the Para-Poincaré groups P± and the Static S are still unknown in physics, but the Para-Galilei group G± and the Carroll group C are gaining more interest in recent times.

The Para-Galilei group has been identified as governing a light spring [16].

The Carroll group has been associated to tachyon dynamics [1719], to Carrollian electromagnetism [20] versus Galilean electromagnetism [21] or to the dynamics of Carroll particles [22] and Carroll strings [23]. The anisotropic Carroll group in two space dimensions (i.e. without rotations) has been identified as the isometry group of gravitational plane waves [24, 25]. The Carroll group has also been used recently in the study of ultra-relatistic gravity [26] and for the generalization of Newton-Cartan gravity [27, 28]. The Carroll group has been compared to the Galilei group in the study of gravitational waves [29, 30], of confined dynamical systems [31], of gravity [32] and of covariant hydrodynamics [33].

The purpose of this paper is, first of all, to clarify the origin of the names given to the three Lévy-Leblond types of contraction and then improve the Levy-Leblond method further.

The main purpose of this paper is to improve the contraction process conducting from the de Sitter Lie algebra to other kinematical Lie algebras. It presents a contraction process in terms of a new set of dynamical parameters (a mass m, an energy E0 and a compliance C) related to the kinematical parameters (a speed c, a radius r and a period τ) by ${r}^{2},{\tau }^{2})=C({E}_{0},m)$, where $c=r{\tau }^{-1}$. Note that the kinematic descriptions are associated only with lengths and times, while the dynamic descriptions are associated with the mass as well as with lengths and time.

In section 2 we recall the Inonu-Wigner contraction, while section 3 recalls the Bacry—Lévy-Leblond method and uses it to establish the twelve kinematical Lie algebras as obtained by Ngendakumana et al [10]. We end the section by setting the raison d'être of a need of improvement the Lévy-Leblond contraction process. It is the purpose of the section 4. With the subsection 4.2, we revisit the Lévy-Leblond method by replacing the dimensional basis by a dimensioned one by scaling some of the vector basis, according the kind of Lie algebra we want to obtain, by either $\tfrac{1}{c}$, $\omega =\tfrac{1}{\tau }$ or $\kappa =\tfrac{1}{r}$. We recover the all the kinematical Lie, except the static Lie algebra, as results of a velocity-space contraction, velocity-time contraction or a space-time contraction. The static Lie algebra is obtained as a velocity-space-time contraction (that Levy-Leblond call the general one) of the de Sitter Lie algebras. The process of this section fails to find the Static Lie algebra as a velocity-space contraction of the Carroll Lie algebra, as a velocity-time contraction of the Galilei Lie algebra or as a space-time contraction of the Para-Galilei Lie algebra (see figure 1 in [11]). With the subsection 4.3 we solve the problem by working with the kinematical parameters which are radius r of the Universe, related to the cosmological constant by ${r}^{2}=\tfrac{3}{{\rm{\Lambda }}}$, the period τ of the Universe, and the velocity c of light defined by $c=r{\tau }^{-1}$. In doing so, all the kinematical Lie algebras are found by the contraction process which consists in keeping one parameter finite and letting the remaining two tend to infinity, their ratio being kept finite. The results are summarized in the figure 1 in this paper. However, with the method used in subsection 4.3, we have to keep one parameter finite and let the remaining two tend to infinity, their ratio being kept finite.

Figure 1.

Figure 1. Vertical, horizontal and oblique arrows indicate velocity-space, space-time and velocity-time contractions respectively.

Standard image High-resolution image

To not have to take the precaution above we introduce with section 5 the dynamical contractions by first parameterizing the de Sitter Lie algebras by the dynamical parameters mass m, compliance C (inverse of stiffness or of Hooke constant or of force constant), and energy, E0. The dynamical parameters and the kinematical parameters are related by $({r}^{2},{\tau }^{2})=C({E}_{0},m)$ implying that ${E}_{0}={{mc}}^{2}$. The corresponding contraction consist in letting only of the dynamical parameters go to infinity without constraining the remaining ones, contrary to the kinematical contractions process. The three Bacry-Lévy-Leblond contractions, i.e. the velocity-space contraction, the velocity-time contraction and the space-time contraction correspond then respectively to an infinite energy E0, an infinite mass m and an infinite compliance C. They are the Newtonian limit, the static limit and the flat limit of Dyson [34]. This why we claim that the dynamical contraction process is the best one.

Finally in section 6, the Kirillov method is used to establish, for each kinematical Lie algebra, a Poisson-Lie algebra and the equations of change with respect any parameter of the Lie group. Those equations clarify the relationships and differences between the twelve kinematical Lie algebras according the up-down, right-left and frontward-backward contractions (see figure 2). They also permit to split the dual vector space of a kinematical Lie algebra in direct sum of irreducible vector subspaces with respect the operator $\tfrac{d}{{ds}}$, s being a parameter of the Lie group.

Figure 2.

Figure 2. Vertical arrows, horizontal arrows and oblique arrows indicate energy, mass and compliance tending to infinity respectively.

Standard image High-resolution image

2. Inonu-Wigner contractions of Lie algebras

2.1. Contraction of Lie algebras

We start with a Lie algebra $({ \mathcal G },\varphi )$ where ${ \mathcal G }$ is a vector space generated by Xi and φ is a skew symmetric mapping $\varphi :{ \mathcal G }\times { \mathcal G }\to { \mathcal G }$ defined by $\varphi ({X}_{i},{X}_{j})={X}_{k}{C}_{{ij}}^{k}$ and satisfying the Jacobi identity

Equation (1)

The Ckij are called the structure constants of the Lie algebra $({ \mathcal G },\varphi )$. The Jacobi identity shows that a Lie algebra is non associative algebra.

If the mapping ${\psi }_{\epsilon }:{ \mathcal G }\to { \mathcal G }$ is singular for a certain value epsilon0 of epsilon and if the mapping ${\varphi }^{{\prime} }:{ \mathcal G }\times { \mathcal G }\to { \mathcal G }$ is defined by

Equation (2)

then $({ \mathcal G },{\varphi }^{{\prime} })$ is a new Lie algebra called the contraction of the Lie algebra $({ \mathcal G },\varphi )$ [5].

2.2. Inonu-Wigner contrations

The pioneering contraction method is that of Inonu and Wigner [1] which starts with a Lie algebra ${ \mathcal G }={ \mathcal H }+{ \mathcal P }$ where ${ \mathcal H }$ is generated by Xa, ${ \mathcal P }$ is generated by Xα; the structure of ${ \mathcal G }$ being a priori given by

where $a,b,c=1,\ldots ,{\dim }({ \mathcal H })$ and $\alpha ,\beta ,\gamma =1,\ldots ,{\dim }({ \mathcal P })$.

The Inonu-Wigner method uses the parameterized change of basis ${\psi }_{\epsilon }:({X}_{a},{X}_{\alpha })\to ({Y}_{a},{Y}_{\alpha })$ defined by ${Y}_{a}={X}_{a}\,,\,{Y}_{\alpha }=\epsilon {X}_{\alpha }$. The structure of the Lie algebra ${ \mathcal G }$ is given in the new basis by

In which condition an Inonu-Wigner contraction is it possible? In the limit $\epsilon \to 0$ , the term ${\epsilon }^{-1}{Y}_{\gamma }{C}_{{ab}}^{\gamma }$ diverges. A limit will exist if only if the structure constants ${C}_{{ab}}^{\gamma }$ vanish. Hence to get a Inonu-Wigner contraction, the structure of ${ \mathcal G }$ in the basis $({X}_{a},{X}_{\alpha })$ must be

Equation (3)

and

in the basis (Ya, Yα), i.e.that ${ \mathcal H }$ must be a subalgebra of ${ \mathcal G }$. The structure of the contracted Lie algebra is then

Equation (4)

The Lie algebra $({ \mathcal G },{\varphi }^{{\prime} })$ defined by (4) is a Inonu-Wigner contraction of the mother Lie algebra $({ \mathcal G },\varphi )$ with respect to the Lie subalgebra ${ \mathcal H }$. It is a semi-direct sum of $({ \mathcal H },{\varphi }^{{\prime} })$ and the abelian Lie algebra $({ \mathcal P },{\varphi }^{{\prime} })$.

This process has been used by Bacry and Levy-Leblond [11]. In the next section we briefly recall the results and point why we need to improve some aspect of that paper.

3. Possible kinematical Lie algebras à la Lévy-Leblond

According to Bacry and Lévy-Leblond [11], a kinematical group is a space-time transformation group which keeps laws of physics invariant. Due to the assumptions of space isotropy, space-time homogeneity and existence of inertial transformations, a kinematical group is a ten dimensional Lie group whose Lie algebra is generated by three rotation generators Ji (isotropy of space), three space translation generators Pi (homogeneity of space), a time translation generator H (homogeneity of time) and three inertial transformation generators Ki. Following Bacry and Lévy-Leblond [11], Ngendakumana and coauthors [10] have shown that under some mathematical physics assumptions only twelve kinematical Lie algebras exist. Their Lie algebraic structures have in common the Lie brackets defining the adjoint representation of the rotation generators

The remaining Lie brackets are given by the table 1 [10]. The ParaPoincaré Lie algebra ${ \mathcal P }{,}_{+}$ which is isomorphic to the Euclidean Lie algebra ${ \mathcal E }(4)$ where the 'translations' generated by Ki and H form an abelian Lie subalgebra does not appear in the list of kinematical ones by Bacry and Lévy-Leblond [11]. The argument is that the inertial transformations are compact. However they are noncompact and only space translations are compact.

Table 1.  The kinematical Lie algebras in term of c,r and τ.

Lie symbol Lie algebra name $[{K}_{i},H]$ $[{K}_{i},{K}_{j}]$ $[{K}_{i},{P}_{j}]$ $[{P}_{i},{P}_{j}]$ $[{P}_{i},H]$
${{dS}}_{\pm }$ ${de}\,{Sitter}$ Pi $-\tfrac{1}{{c}^{2}}{J}_{k}{\epsilon }_{{ij}}^{k}$ $\tfrac{1}{{c}^{2}}H{\delta }_{{ij}}$ $\pm \tfrac{1}{{r}^{2}}{J}_{k}{\epsilon }_{{ij}}^{k}$ $\pm \tfrac{1}{{\tau }^{2}}{K}_{i}$
P Poincare Pi $-\tfrac{1}{{c}^{2}}{J}_{k}{\epsilon }_{{ij}}^{k}$ $\tfrac{1}{{c}^{2}}H{\delta }_{{ij}}$ 0 0
${{NH}}_{\pm }$ NewtonHooke Pi 0 0 0 $\pm \tfrac{1}{{\tau }^{2}}{K}_{i}$
P± ${Para}\,{Poincare}$ 0 0 $\tfrac{1}{{c}^{2}}H{\delta }_{{ij}}$ $\pm \tfrac{1}{{r}^{2}}{J}_{k}{\epsilon }_{{ij}}^{k}$ $\pm \tfrac{1}{{\tau }^{2}}{K}_{i}$
G Galilei Pi 0 0 0 0
G± ParaGalilei 0 0 0 0 $\pm \tfrac{1}{{\tau }^{2}}{K}_{i}$
C Carroll 0 0 $\tfrac{1}{{c}^{2}}H{\delta }_{{ij}}$ 0 0
S Static 0 0 0 0 0

Using the Inonu-Wigner contraction method [1], Bacry and Lévy-Leblond [11] have established that these Lie algebras are approximations of the de Sitter Lie algebras. Their links are summarized by the contractions scheme (see figure 1 on page 1610 of [11]). We will refer to these nomenclature in the next two sections.

The parameter used by Bacry and Levy-Leblond to scale the subalgebra ${ \mathcal P }$ is dimensionless. This is certainly due to the fact those authors have set c = 1 and r = 1, c being the speed of light while r is the radius of the Universe. The table at page 1608 shows that the generators of the kinematical Lie algebras in question seem to be dimensionless. This physics interpretation behind is then a bit difficult to follow. We propose to improve this situation in the next two sections.

4. Kinematical improvement of the Levy-Leblond approach

4.1. The Lie algebras O± (5)

We propose to recover the Bacry-Lévy-Leblond contractions scheme by using the kinematical parameters r, τ and c which are respectively the radius of universe, the period of the Universe and speed of light. We first introduce the de Sitter Lie algebras dS± as isomorphic to the pseudo-orthogonal Lie algebras ${O}_{\pm }(5)$, i.e. that ${{dS}}_{+}(3)$ $[{{dS}}_{-}(3)]$ is isomorhic to $O(1,4)$ [O(2,3)] Lie algebra. The aim of this section is to better clarify velocity-space contractions, velocity-time contractions and space-time contractions of Bacry and Lévy-Leblond [11]. Let V be a five dimensional manifold equipped with the metric

Equation (5)

where the dimension of the xa is that of length. The matrix elements ηab form the diagonal matrix ${\eta }_{\pm }={diag}({I}_{3\times 3},-1,\pm 1)$. Let denote by G± the group of transformations ${x}^{{\prime} a}={g}_{b}^{a}{x}^{b}$ keeping ds2 invariant. It is the group of real square matrices g with order five satisfying ${g}^{t}{\eta }_{\pm }g={\eta }_{\pm }$. The Lie group ${{SO}}_{0}(4,1)$ is the connected component of G+ while ${{SO}}_{0}(3,2)$ is the connected component of ${G}_{-}$. The Lie algebra ${{ \mathcal O }}_{\pm }(5)$ is the set of the real square matrices X of order 5 satisfying ${}^{t}X{\eta }_{\pm }+{\eta }_{\pm }X=0$. We easily verify that $X={J}_{k}{\theta }^{k}+{A}_{k}{\alpha }^{k}+{B}_{k}{\beta }^{k}+\gamma {\rm{\Gamma }},\,\,\,k=1,2,3,\,$ is the dimensionless matrix

Equation (6)

and that $({J}_{k},\,{A}_{k}\,{B}_{k},\,{\rm{\Gamma }})$ is a basis of ${{ \mathcal O }}_{\pm }(5)$. The Lie algebra ${{ \mathcal O }}_{\pm }(5)$ structure is defined by the Lie brackets

Equation (7)

Equation (8)

Equation (9)

The dimensionless generator Ai and Bi play the role of velocity generator and space translation generator in the ith direction respectively while the dimensionless Γ plays the role of time translations generator.

4.2. Inonu-Wigner approach revisited

4.2.1. Relative space Lie algebras:de Sitter, Newton-Hooke, Poincare and Galilei

Let set ${K}_{i}=\tfrac{1}{c}{A}_{i},\,{P}_{i}=\tfrac{\omega }{c}{B}_{i}$ and H = ω Γ where c is a speed while $\omega =\tfrac{1}{\tau }$ is a frequency. The Lie brackets (7), (8) and (8) become

Equation (10)

Equation (11)

Equation (12)

They define the de Sitter Lie algebras dS± in the basis $({J}_{i},{K}_{i},{P}_{i},H)$. The general element of the Lie algebra dS± is $X={J}_{i}{\theta }^{i}+{K}_{i}{v}^{i}+{P}_{i}{x}^{i}+{Ht}$, θi being dimmensionless, the physical dimensions of vi, xi and t being a velocity, a length and a time respectively. Also the general element of the dual of dS± is $\alpha ={j}_{i}{J}^{* i}+{k}_{i}{K}^{* i}+{p}_{i}{P}^{* i}+{{EH}}^{* }$ where ji is the ith component of the angular momentum, ki is the ith component of the static momentum, pi is the ith component of the linear momentum and E is an energy. In the limit $\epsilon =\tfrac{1}{c}\to 0$, the Lie brackets (10), (11) and (12) become

Equation (13)

Equation (14)

Equation (15)

They define the Newton-Hooke Lie algebras ${ \mathcal N }{{ \mathcal H }}_{\pm }$ in the basis $({J}_{i},{K}_{i},{P}_{i},H)$.

In the limit $\omega =\tfrac{1}{\tau }\to 0$ , the Lie brackets (10), (11) and (12) become

Equation (16)

Equation (17)

Equation (18)

They define the Poincare Lie algebra ${ \mathcal P }$ in in the basis (Ji, Ki, Pi, H).

As $\tfrac{1}{c}$ multiplies Ai and Bi while ω mulipliplies Bi and Γ, it follows that the Newton-Hooke lie algebras and the Poincare Lie algebra are velocity-space contractions and space-time contraction of the de Sitter Lie algebras respectively.

The Lie algebra structure

Equation (19)

Equation (20)

Equation (21)

which defines the Galilei Lie algebra ${ \mathcal G }$ in the basis (Ji, Ki, Pi, H), is obtained from (13), (14) and (15) when $\omega \to 0$ or from (16), (17) and (18) when $\epsilon =\tfrac{1}{c}\to 0$. The Galilei Lie algebra is then a velocity-space contraction of the Poincare Lie algebra and a space-time contraction of the Newton-Hooke Lie algebras.

4.2.2. Relative time Lie algebras:de Sitter, Poincare, Para-Poincare and Carroll

Let set ${K}_{i}=\tfrac{1}{c}{A}_{i},\,{P}_{i}=\kappa {B}_{i}$ and $M=\tfrac{\kappa }{c}{\rm{\Gamma }}$ where c is a speed while $\kappa =\tfrac{1}{r}$ is a curvature. The Lie brackets (7), (8) and (8) become

Equation (22)

Equation (23)

Equation (24)

They define the de Sitter Lie algebras dS± in the basis (Ji, Ki, Pi, M). The general element of the Lie algebra dS± is $X={J}_{i}{\theta }^{i}+{K}_{i}{v}^{i}+{P}_{i}{x}^{i}+M\xi $, θi being dimmensionless, the physical dimensions of vi, xi and ξ being a velocity, a length and a a specific action respectively. Also the general element of the dual of dS± is $\alpha ={j}_{i}{J}^{* i}+{k}_{i}{K}^{* i}+{p}_{i}{P}^{* i}+{{mM}}^{* }$ where ji is the ith component of the angular momentum, ki is the ith component of the static momentum, pi is the ith component of the linear momentum and m is a mass.

In the limit $\epsilon =\tfrac{1}{c}\to 0$,the Lie brackets (22), (23) and (24) become

Equation (25)

Equation (26)

Equation (27)

They define the Para-Poincare Lie algebras ${ \mathcal N }{{ \mathcal H }}_{\pm }$ in the basis (Ji, Ki, Pi, M).

In the limit $\kappa \to 0$ , the Lie brackets (22), (23) and (24) become

Equation (28)

Equation (29)

Equation (30)

They define the Poincare Lie algebra ${ \mathcal P }$ in the basis (Ji, Ki, Pi, M).

As $\tfrac{1}{c}$ multiplies Ai and Γ while κ mulipliplies Bi and Γ, it follows that the Para-Poincare lie algebras and the Poincare Lie algebra are velocity-time contractions and space-time contraction of the de Sitter Lie algebras respectively.

The Lie algebra structure

Equation (31)

Equation (32)

Equation (33)

which defines the Carroll Lie algebra ${ \mathcal C }$ in the basis (Ji, Ki, Pi, M), is obtained from (25), (26) and (27) when $\kappa \to 0$ or from (28), (29) and (30) when $\epsilon =\tfrac{1}{c}\to 0$. The Carroll Lie algebra is then a velocity-time contraction of the Poincare Lie algebra and a space-time contraction of the Para-Poincare Lie algebras.

4.2.3. Cosmological Lie algebras:de Sitter, Newton-Hooke, Para-Poincare and Para-Galilei

Let set ${F}_{i}=\tfrac{\omega }{r}{A}_{i},\,{P}_{i}=\tfrac{1}{r}{B}_{i}$ and H = ω Γ where r is a radius while $\omega =\tfrac{1}{\tau }$ is a frequency. The Lie brackets (7), (8) and (8) become

Equation (34)

Equation (35)

Equation (36)

They define the de Sitter Lie algebras dS± in the basis (Ji, Fi, Pi, H). The general element of the Lie algebra dS± is $X={J}_{i}{\theta }^{i}+{F}_{i}{\zeta }^{i}+{P}_{i}{x}^{i}+{Ht}$, θi being dimmensionless, the physical dimensions of ${\zeta }^{i},{x}^{i}$ and t being a velocity, a length and a time respectively. Also the general element of the dual of dS± is $\alpha ={j}_{i}{J}^{* i}+{f}_{i}{F}^{* i}+{p}_{i}{P}^{* i}+{{EH}}^{* }$ where ji is the ith component of the angular momentum, fi is the ith component of a force, pi is the ith component of the linear momentum and E is an energy.

In the limit $\omega =\tfrac{1}{\tau }\to 0$,the Lie brackets (34), (35) and (36) become

Equation (37)

Equation (38)

Equation (39)

They define the Para-Poincare Lie algebras ${{ \mathcal P }}_{\pm }$ in the basis (Ji, Fi, Pi, H)

In the limit $\kappa =\tfrac{1}{r}\to 0$ , the Lie brackets (34), (35) and (36) become

Equation (40)

Equation (41)

Equation (42)

They define the Newton-Hooke Lie algebra ${ \mathcal N }{{ \mathcal H }}_{\pm }$ in the basis $({J}_{i},{F}_{i},{P}_{i},H)$.

As ω multiplies Ai and Γ while $\kappa =\tfrac{1}{r}$ mulipliplies Ai and Bi, it follows that the Newton-Hooke lie algebras and the Para-Poincare Lie algebras are velocity-space contractions and space-time contraction of the de Sitter Lie algebras respectively.

The Lie algebra structure

Equation (43)

Equation (44)

Equation (45)

which defines the Para-Galilei Lie algebra ${ \mathcal G }\pm $ in the basis (Ji, Fi, Pi, H), is obtained from (37), (38) and (39) when $\kappa =\tfrac{1}{r}\to 0$ or from (40), (41) and (42) when $\omega =\tfrac{1}{\tau }\to 0$. The Para-Galilei Lie algebra is then a velocity-space contraction of the Newton-Hooke Lie algebra and a velocity-space contraction of the Para-Poincare Lie algebras.

4.2.4. The Static Lie algebra

By setting ${K}_{i}=\tfrac{\kappa }{\omega }{A}_{i}$, ${P}_{i}=\kappa {B}_{i}$ and $M=\tfrac{\kappa }{\omega }{\rm{\Gamma }}$, the structure of the de Sitter Lie dS± becomes

Equation (46)

Equation (47)

Equation (48)

In the limit $\kappa \to 0$, the structure defined by (46), (47) and (48) becomes the structure

Equation (49)

Equation (50)

Equation (51)

defining the Static Lie algebra ${ \mathcal S }$ in the relative time kinematical Lie algebras. As κ multiplies Ai, Bi and Γ, the static Lie algebra is a velocity-space-time (the general according Levy-Leblond [11]) contraction of the de Sitter Lie algebras.

4.3. Kinematical improvment

Even if the previous section clarifies better the Inonu-Wigner contractions of the de Sitter Lie algebras, there remain the following problem: the structures of relative space Lie algebras, the relative time Lie algebras and the cosmological Lie algebras are defined in different bases. We propose to remove that situation in the next section. We use three kinematical parameters: the speed c, the radius r and the period τ. We ignore the relation $c=r{\tau }^{-1}$ during all the process.

4.3.1. The de Sitter Lie algebras

Let ${K}_{i}=\tfrac{1}{c}{A}_{i}$, ${P}_{i}=\tfrac{1}{r}{B}_{i}$ and $H=\tfrac{1}{\tau }{\rm{\Gamma }}$ where $\sigma =\tfrac{1}{c}$ is a slowness, $\kappa =\tfrac{1}{r}$ is a curvature while $\omega =\tfrac{1}{\tau }$ is a frequency. In the new basis $({J}_{k},\,{K}_{k}\,{P}_{k},\,H)$ of ${{ \mathcal O }}_{\pm }(5)$ the matrix X above becomes

Equation (52)

where ${\alpha }^{i}=\tfrac{{v}^{i}}{c}\,,{\beta }^{i}=\tfrac{{x}^{i}}{r}\,,\gamma =\tfrac{t}{\tau }$. Hence the parameters associated with Ki ,Pi and H have velocity, length and time as respective physical dimension. The Lie brackets (7), (8) and (8) become then

Equation (53)

Equation (54)

Equation (55)

Let us now study the limits of the de Sitter Lie algebras as the constants tend to infinity. Normally the three constants are constrained by $c=r{\tau }^{-1}$. However, we ignore it for a moment. We use it at the end of the section to show that our way of doing has recovered the results of table 1. It is first of all evident that (53) does not change. We are then only interested in the behavior of (54) and (55).

4.3.2. The Newton-Hooke, Poincaré and Para-Poincaré Lie algebras

In this section we look for the limits of the de Sitter Lie algebras as two of the constants tend to infinity while their ratio is kept finite.

  • (a)  
    Newton-Hooke Lie algebras. We verify that the limits of (54) and (55), as the speed c and the radius r tend to infinity while their ratio $\tfrac{r}{c}$ and τ are kept finite, are
    Equation (56)
    Equation (57)
    The Lie brackets (53), (56) and (57) define the Newton-Hooke Lie algebra ${ \mathcal N }{{ \mathcal H }}_{\pm }$.
  • (b)  
    Poincaré Lie algebra. If the period τ and the radius r tend to infinity while their ratio $\tfrac{r}{\tau }$ and c are kept finite, the brackets (54) and (55) become
    Equation (58)
    Equation (59)
    The Lie brackets (53), (58) and (59) define the Poincaré Lie algebra ${ \mathcal P }$.
  • (c)  
    Para-Poincaré Lie algebras. Similarly if the speed c and the time τ tend to infinity while their ratio $\tfrac{c}{\tau }$ and r are kept finite then the Lie brackets (54) and (55) become
    Equation (60)
    Equation (61)
    The Lie brackets (53), (60) and (61) define the Para-Poincaré Lie algebra ${{ \mathcal P }}_{\pm }$. We then notice that the Newton-Hooke Lie algebras, the Poincaré Lie algebra and the Para-Poincaré Lie algebras are respectively the velocity-space, space-time and velocity-time contractions of the de Sitter Lie algebras as in [11].

4.3.3. The Galilei, Para-Galilei and Carroll Lie algebras

  • (a)  
    Galilei Lie algebra. The limit of the Lie brackets (56) and (57) as the radius r and the period τ tend to infinity while $\tfrac{r}{\tau }$ and c are kept finite and the limit (58) and (59) as the radius r and the speed c tend to infinity while $\tfrac{c}{r}$ and τ are kept finite are the same, i.e.
    Equation (62)
    Equation (63)
    The Lie brackets (53), (62) and (63) define the Galilei Lie algebra ${ \mathcal G }$.
  • (b)  
    Para-Galilei Lie algebras. The limit of the Lie brackets (56) and (57) as the speed c and the period τ tend to infinity while $\tfrac{c}{\tau }$ and r are kept finite and the limit (60) and (61) as the radius r and the speed c tend to infinity while $\tfrac{c}{r}$ and τ are kept finite are the same, i.e.
    Equation (64)
    Equation (65)
    The Lie brackets (53), (64) and (65) define the Para-Galilei Lie algebras ${{ \mathcal G }}_{\pm }$.
  • (c)  
    Carroll Lie algebra. The limit of the Lie brackets (58) and (59) as the speed c and the period τ tend to infinity while $\tfrac{c}{\tau }$ is kept finite and the limit (60) and (61) as the radius r and the period τ tend to infinity while $\tfrac{r}{\tau }$ is kept finite are the same, i.e.
    Equation (66)
    Equation (67)
    The Lie brackets (53), (66) and (67) define the Carroll Lie algebra ${ \mathcal C }$. Hence the Galilei, the Para-Galilei, the Carroll Lie algebras are respective contractions of the Newton-Hooke or Poincaré Lie algebras, the Newton-Hooke or the Para-Poincaré Lie algebras, the Poincaré or the Para-Poincaré Lie algebras respectively.

4.3.4. The static Lie algebra

The limit of the Lie brackets (62) and (63) as the speed c and the period τ tend to $\infty $ while $\tfrac{c}{\tau }$ and r are kept finite ,the limit (64) and (65) as the radius r and the period τ tend to $\infty $ while $\tfrac{r}{\tau }$ and c are kept finite and the limit of the Lie brackets (66) and (67) as the speed c and the radius r tend infinity while $\tfrac{c}{r}$ and τ are kept finite are the same; i.e

Equation (68)

Equation (69)

The Lie brackets (53), (68) and (69) define the Static Lie algebra ${ \mathcal S }$.

When the constraint c = −1 is taken in account, the Lie brackets in the table 1 are recovered.

These approximations through kinematical parameters are summarized in the following cube (see figure 1). On the cube, the horizontal arrows represent the contractions as $c,\tau \to \infty $, $\tfrac{c}{\tau }$ and r finite (velocity-time contractions ), the vertical arrows represent the contractions as $c,r\to \infty $, $\tfrac{r}{c}$ and τ finite (velocity-space contractions) and the oblique arrows represent the contractions as $r,\tau \to \infty $, $\tfrac{r}{\tau }$ and c finite (space-time contractions).

5. Dynamical improvement of the Levy-Leblond approach

In the process of the previous section, we were sending two parameters at infinity while keeping finite their ratio and the third one. We introduce in this section the dynamical parameters compliance C, mass m and energy E0. These dynamical parameters enter the de Sitter Lie algebras structure by replacing the boost generators Ki by the momentum generators ${Q}_{i}=\tfrac{1}{m}{K}_{i}$, m being a mass and by defining the compliance C and the energy E0 respectively by $C=\tfrac{{\tau }^{2}}{m}$ and ${E}_{0}={{mc}}^{2}$. In the contraction process, only one parameter will be sent to infinity without any precaution on the other two. Similarly to the kinematical contraction process, at the end of the dynamical contraction process, the kinematical Lie algebras will distributed on a cube. Two opposite faces with finite versus infinite mass, two opposite faces with a finite versus infinite energy and two faces with a finite versus infinite compliance.

5.1. Three finite parameters Lie algebras: the de Sitter

The de Sitter Lie algebras dS± are then defined in the basis (Ji, Qi, Pi, H), by the Lie brackets

Equation (70)

Equation (71)

Equation (72)

The de Sitter Lie algebras dS± are then characterized by the three dynamical parameters m, C and E0. They are at the edge of the finite mass, finite energy and finite compliance. The de Sitter Lie algebras are then characterized by three finite kinematical parameters: the frequency $\omega =\tfrac{1}{\sqrt{{mC}}}$ (time $\tau =\sqrt{{mC}}$), the speed $c=\sqrt{\tfrac{{E}_{0}}{m}}$ (slowness $s=\sqrt{\tfrac{m}{{E}_{0}}}$) and the curvature $\kappa =\tfrac{1}{\sqrt{{{CE}}_{0}}}$ (the radius $r=\sqrt{{{CE}}_{0}}$).

5.2. Two finite parameters Lie algebras: Newton-Hooke, Poincare and Para-Poioncare

When one of the three parameters becomes infinite, the structure of the de Sitter Lie algebras gives rise to a Lie algebra characterized by the remaining two. In one dimension of space (case of ${{ \mathcal O }}_{\pm }(3)$), these algebras are the solvable ones.

5.2.1. Mass-Compliance Lie algebras: Newton-Hooke

When the energy E0 tends to infinity, the structure of the de Sitter Lie algebras given by (70), (71) and (72) becomes

Equation (73)

Equation (74)

Equation (75)

and defines the Newton-Hooke Lie algebras characterized by the mass m and the compliance C related to the frequency $\omega =\tfrac{1}{\sqrt{{mC}}}$ (time $\tau =\sqrt{{mC}}$). The Newton-Hooke Lie group is then a semi direct product of the direct product of rotations and time translations on the abelian group of impulses-positions.

5.2.2. Mass-Energy Lie algebra: Poincare

When the compliance C tends to infinity, the structure of the de Sitter Lie algebras given by (70), (71) and (72) becomes

Equation (76)

Equation (77)

Equation (78)

and defines the Poincare Lie algebra charcterized by the mass m and the energy E0 related to the speed $c=\sqrt{\tfrac{{E}_{0}}{m}}$ (slowness $s=\sqrt{\tfrac{m}{{E}_{0}}}$). The Poincare Lie group is then a semi direct product of the direct product of the Lorentz group on the abelian group of space-time translations.

5.2.3. Compliance-Energy Lie algebras: Para-Poincare

When the mass m tends to infinity, the structure of the de Sitter Lie algebras given by (70), (71) and (72) becomes

Equation (79)

Equation (80)

Equation (81)

and defines the Para-Poincare Lie algebra charcterized by the compliance C and the energy E0 related to the curvature $\kappa =\tfrac{1}{\sqrt{{{CE}}_{0}}}$ (the radius $r=\sqrt{{{CE}}_{0}}$). The Para-Poincare Lie group is then a semi direct product of the Para-Lorentz group on the abelian group of impulses-time translations.

5.3. One finite parameter Lie algebras: Galilei, Para-Galilei and Carroll

When two of the three parameters become infinite, the structure of the de Sitter Lie algebras gives rise to a Lie algebra characterized by the remaining one. In one dimension of space (case of ${{ \mathcal O }}_{\pm }(3)$), these algebras are the nilpotent ones.

5.3.1. Energy Lie algebra: Carroll

The structure

Equation (82)

Equation (83)

Equation (84)

defining the Carroll Lie algebra is obtained from that of the Poincare Lie algebra defined by (76), (77) and (78) when the mass m tends to infinity or from that of the Para-Poincare Lie algebras defined (88), (86) and (90) when the compliance tends to infinity. The Carroll Lie algebras characterized by a finite energy E0.

5.3.2. Compliance Lie algebra: Para-Galilei

The structure

Equation (85)

Equation (86)

Equation (87)

defining the Para-Galilei Lie algebras is obtained from that of the Newton-Hooke Lie algebras defined by (73), (74) and (75) when the mass m tends to infinity or from that of the Para-Poincare Lie algebras defined (88), (86) and (90) when the energy tends to infinity. The Para-Galilei Lie algebras characterized by a finite compliance C.

5.3.3. Mass Lie algebra: Galilei

The structure

Equation (88)

Equation (89)

Equation (90)

defining the Galilei Lie algebra is obtained from that of the Newton-Hooke Lie algebras defined by (73), (74) and (75) when the compliance C tends to infinity or from that of the Poincare Lie algebras defined (76), (77) and (78)when the energy tends to infinity. The Galilei Lie algebras characterized by a finite mass m.

We can say that the mass m is galilean, the compliance C is para-galilean and the energy E0 is carrollian.

5.4. Zero parameters Lie algebra: Static

The zero parameters Lie algebra is the static Lie algebra which is obtained from the Galilei Lie algebra as the mass tends to infinity, from the Carroll Lie algebra as the energy tends to infinity or from the Para-Galilei Lie algebra as the compliance tends to infinity.Its structure is

Equation (91)

Equation (92)

Equation (93)

It is the abelian Lie algebra in the case of the dimension one of space (case of ${{ \mathcal O }}_{\pm }(3)$).

5.5. Summarizing cube

All the Lie brackets defining these Lie algebras have in common the Lie brackets

the others are summarized in the table 2 while the limiting process is given by the figure 2 where the horizontal arrows represent the contractions as the mass $m\to \infty $ (static limit), the vertical arrows represent the contractions as the energy ${E}_{0}\to \infty $(Newtonian limit) and the oblique arrows represent the contractions as the compliance $C\to \infty $ (flat limit). If we use coordinates $\left(\tfrac{1}{m},\tfrac{1}{C},\tfrac{1}{{E}_{0}}\right)$, the kinematical Lie algebras constitute then the cube below (figure 2). The table 3 give a comparison of kinematical Lie algebras distribution obtained through the dynamical contraction process above with that obtained by the kinematical contraction process as given by MacRae [8]. The relative (absolute) time groups in the kinematical contraction process correspond to the fine (infinite) energy groups in the dynamical contraction process. Similarly the relative (absolute) space correspond to the finite (infinite) mass while the cosmological (local) groups correspond to the finite (infinite) compliance.

Table 2.  The kinematical Lie algebras in term of the mass, the compliance and the energy.

Lie symbol Lie algebra name $[{Q}_{i},H]$ $[{Q}_{i},{Q}_{j}]$ $[{Q}_{i},{P}_{j}]$ $[{P}_{i},{P}_{j}]$ $[{P}_{i},H]$
${{dS}}_{\pm }$ ${de}\,{Sitter}$ $\tfrac{1}{m}{P}_{i}$ $-\tfrac{1}{{{mE}}_{0}}{J}_{k}{\epsilon }_{{ij}}^{k}$ $\tfrac{1}{{E}_{0}}H{\delta }_{{ij}}$ $\pm \tfrac{1}{{{CE}}_{0}}{J}_{k}{\epsilon }_{{ij}}^{k}$ $\pm \tfrac{1}{C}{Q}_{i}$
P Poincare $\tfrac{1}{m}{P}_{i}$ $-\tfrac{1}{{{mE}}_{0}}{J}_{k}{\epsilon }_{{ij}}^{k}$ $\tfrac{1}{{E}_{0}}H{\delta }_{{ij}}$ 0 0
${{NH}}_{\pm }$ NewtonHooke $\tfrac{1}{m}{P}_{i}$ 0 0 0 $\pm \tfrac{1}{C}{Q}_{i}$
P± ${Para}\,{Poincare}$ 0 0 $\tfrac{1}{{E}_{0}}H{\delta }_{{ij}}$ $\pm \tfrac{1}{{{CE}}_{0}}{J}_{k}{\epsilon }_{{ij}}^{k}$ $\pm \tfrac{1}{C}{Q}_{i}$
G Galilei $\tfrac{1}{m}{P}_{i}$ 0 0 0 0
G± ParaGalilei 0 0 0 0 $\pm \tfrac{1}{C}{Q}_{i}$
C Carroll 0   $\tfrac{1}{{E}_{0}}H{\delta }_{{ij}}$ 0 0
S Static 0 0 0 0 0

Table 3.  Distribution of kinematical Lie algebras through the kinematical versus dynamical contractions.

Kinematical process Dynamical process Kinematical Lie algebras
Relative time groups finite energy groups ${{dS}}_{\pm },P,{P}_{\pm },C$
Absolute time groups infinite energy groups ${{NH}}_{\pm },G,{G}_{\pm },S$
Relative space groups finite mass groups ${{dS}}_{\pm },{{NH}}_{\pm },P,G$
Absolute space infinite mass groups ${P}_{\pm },{G}_{\pm },C,S$
Cosmological groups finite compliance groups ${{dS}}_{\pm },{{NH}}_{\pm },{P}_{\pm },{G}_{\pm }$
Local groups infinite compliance groups $P,G,C,S$

6. A glance at the physics associated to the kinematical Lie algebras

Let us have a look at the physics associated to the kinematical Lie algebras in function of the three dynamical parameters:compliance C, mass m and energy E0.

6.1. Poisson brackets

We know that the Poisson bracket of two functions defined on the dual ${{ \mathcal G }}^{* }$ of any Lie algebra ${ \mathcal G }$ is defined by

Equation (94)

where ai are the coordinates on ${{ \mathcal G }}^{* }$ and ${K}_{{ij}}(a)=-{a}_{k}{C}_{{ij}}^{k}$ are the matrix elements of the Kirillov form.

Let the general element of the dual of a kinematical Lie algebra be ${j}_{k}{J}^{k* }+{q}_{k}{Q}^{k* }+{\pi }_{k}{P}^{k* }+{{EH}}^{* }$ where jk are the components of the angular momentum conjugated to the angle θk, πk are the components of the linear momentum conjugate to the space translation xk, qk are the components of the position conjugated to the impulse pk while E is an energy conjugated to the time translation t.

From (70), (71) and (72) follows that the Kirillov matrix associated to the de Sitter groups is

Equation (95)

From the Lie brackets (see previous section) defining the kinematical Lie algebras in function of m,E0 and C follows then that the kinematical Poisson-Lie algebras are defined by the common Poisson brackets

Equation (96)

and the other Poisson brackets given by the table 4.

Table 4.  The kinematical Poisson-Lie algebras.

Lie symbol Lie algebra name $\left\{E,{q}_{i}\right\}$ $\left\{{q}_{i},{q}_{j}\right\}$ $\left\{{\pi }_{i},{q}_{j}\right\}$ $\left\{{\pi }_{i},{\pi }_{j}\right\}$ $\left\{E,{\pi }_{i}\right\}$
${{dS}}_{\pm }$ ${de}\,{Sitter}$ $\tfrac{1}{m}{\pi }_{i}$ $\tfrac{1}{{{mE}}_{0}}{j}_{k}{\epsilon }_{{ij}}^{k}$ $\tfrac{E}{{E}_{0}}{\delta }_{{ij}}$ $\mp \tfrac{1}{{{CE}}_{0}}{j}_{k}{\epsilon }_{{ij}}^{k}$ $\pm \tfrac{1}{C}{q}_{i}$
${{NH}}_{\pm }$ NewtonHooke $\tfrac{1}{m}{\pi }_{i}$ 0 0 0 $\pm \tfrac{1}{C}{q}_{i}$
P Poincare $\tfrac{1}{m}{\pi }_{i}$ $\tfrac{1}{{{mE}}_{0}}{j}_{k}{\epsilon }_{{ij}}^{k}$ $\tfrac{E}{{E}_{0}}{\delta }_{{ij}}$ 0 0
G Galilei $\tfrac{1}{m}{\pi }_{i}$ 0 0 0 0
P± ${Para}\,{Poincare}$ 0 0 $\tfrac{E}{{E}_{0}}{\delta }_{{ij}}$ $\mp \tfrac{1}{{{CE}}_{0}}{j}_{k}{\epsilon }_{{ij}}^{k}$ $\pm \tfrac{1}{C}{q}_{i}$
G± ParaGalilei 0 0 0 0 $\pm \tfrac{1}{C}{q}_{i}$
C Carroll 0   $\tfrac{E}{{E}_{0}}{\delta }_{{ij}}$ 0 0
S Static 0 0 0 0 0

6.2. Equations of Change

We rewrite (94) as $\left\{f,g\right\}={X}_{f}(g)$ where the vector field Xf is defined by

Equation (97)

and verifies $[{X}_{f},{X}_{g}]={X}_{\left\{f,g\right\}}$. It is known that the mapping ρ defined by ρ(f) = Xf is a realization of the Lie algebra ${ \mathcal G }$.

Use of the structure of de Sitter Lie algebras $d{{ \mathcal S }}_{\pm }$ defined by the Lie brackets (70), (71) and (72) permits to verify that the de Sitter is the realized by the vector fields

Equation (98)

Equation (99)

Equation (100)

Equation (101)

The realizations of the other kinematical Lie algebras are obtained from (98) to (101) through the dynamical contraction process defined in the previous section. No need to make them explicit here.

If Xf is the generating function of the one parameter diffeomorphism ${{\rm{\Phi }}}_{s}:{{ \mathcal G }}^{* }\to {{ \mathcal G }}^{* }$ and if Xg is the generating function of the one parameter diffeomorphism ${{\rm{\Phi }}}_{\lambda }:{{ \mathcal G }}^{* }\to {{ \mathcal G }}^{* }$, then the equation of change of the function g with respect to s is $\tfrac{{dg}}{{ds}}={X}_{f}(g)$ while the equation of change of f with respect λ is $\tfrac{{df}}{d\lambda }={X}_{g}(f)$, $\tfrac{{dg}}{{ds}}=-\tfrac{{df}}{d\lambda }$.

For each parameter s, let define Fs as ${F}_{s}=\left\{f\in {{ \mathcal G }}^{* }:\tfrac{{df}}{{ds}}=0\right\}$ and let ${V}_{s}={{ \mathcal G }}^{* }/{F}_{s}$ be the variables submanifold of ${{ \mathcal G }}^{* }$. The equations of change describe how the coordinates on Vs change with respect the ad hoc parameter.

Note that as ${{\rm{\Phi }}}_{s}\circ {{\rm{\Phi }}}_{t}={{\rm{\Phi }}}_{s+t}$, the change parameters must be additive. It is true for the longitudinal angle. We show in the appendix that it is also true for the momentum parameter pi, the space translation xi and the time translation parameter t appearing in the dS± Lie algebra element $X={J}_{k}{\theta }^{k}+{Q}_{k}{p}^{k}+{P}_{k}{x}^{k}+{Ht}$. We use the one spatial Poincaré Para-Poincaré and Newton-Hooke, Lie algebras to respectively associate a non additive boost to momentum, a non additive force to space translation and to time translation a non additive dampinglike coefficient.

In the maximal case of the de Sitter case, it follows from tt follows (98) to (101) that the equations of change with

  • the angle θi are
    Equation (102)
    meaning that ji, qi, πi and E are constant with respect θi.It follows that ${F}_{{\theta }^{i}}=\left\{{j}_{i},{q}_{i},{\pi }_{i},E\right\}$ is quadridimensional while ${V}_{{\theta }^{i}}=\left\{{j}_{k},{q}_{k},{\pi }_{k},\,k\ne i\right\}$ is 6-dimensional, a direct sum ${V}_{{\theta }^{i}}={V}_{{\theta }^{i}}({j}_{k})\oplus {V}_{{\theta }^{i}}({q}_{k})\oplus {V}_{{\theta }^{i}}({\pi }_{k}),k\ne i$ of 2-dimensional irreducible subspaces under the operator $\tfrac{d}{d{\theta }^{i}}$.
  • the impulse pi are
    Equation (103)
    meaning that ji, qi are constant with respect pi. We then have that ${F}_{{p}^{i}}=\left\{{j}_{i},{q}_{i}\right\}$ is 2 dimensional and that ${V}_{{p}^{i}}=\left\{{j}_{k},{q}_{k},{\pi }_{l},E\right\},k\ne i$ is 8-dimensional a direct sum ${V}_{{p}^{i}}={V}_{{p}^{i}}({j}_{k},{q}_{k})\oplus {V}_{{p}^{i}}({\pi }_{l},E),k\ne i$, two 4 dimensional irreducible subspaces $\tfrac{d}{{{dp}}^{i}}$.
  • the space translation xi are
    Equation (104)
    meaning that ji, πi are constant with respect xi. We then have that ${F}_{{x}^{i}}=\left\{{j}_{i},{\pi }_{i}\right\}$ is 2 dimensional and that ${V}_{{x}^{i}}=\left\{{j}_{k},{\pi }_{k},{q}_{l},E\right\},k\ne i$ is 8-dimensional a direct sum ${V}_{{x}^{i}}={V}_{{x}^{i}}({j}_{k},{\pi }_{k})\oplus {V}_{{x}^{i}}({q}_{l},E),k\ne i$, two 4 dimensional irreducible subspaces $\tfrac{d}{{{dx}}^{i}}$.
  • the time translation t are
    Equation (105)
    meaning that ji, E are constant with respect t. We then have that ${F}_{t}=\left\{{j}_{i},E\right\}$ is 4 dimensional and that ${V}_{t}=\left\{{q}_{i},{\pi }_{i}\right\}$, a 6 dimensional irreducible space $\tfrac{d}{{dt}}$.

The equations of change corresponding to all kinematical Lie algebra are given in the table 5. In this table we illustrate the equations change with respect the longitude angle φ = θ3, the momentum up p = p3, the altitude z = x3 and the time t. Also the greek indices take the value 1 and 2 while the latin indices take the values 1, 2 and 3. The equations of the form $\tfrac{{df}}{{ds}}=0$ do not appear in the table. Moreover the first column of the table contains the dimensions of Vs, the second one indicates the first order differential equations, the third one the second order differential equation when available, finally the last one shows the corresponding kinematical Lie algebras paired as mother-daughter in the parental relations (see figure 2) up-down for the variations with respect time (energy E0 is absent in the equations), front-backward for the variations with respect the altitude z (mass m is absent in the equations) and right-left for the variations with respect the momentum up p (compliance C is absent in the equations). We also notice that ${V}_{\varphi }={V}_{\varphi }({j}_{\mu })\oplus {V}_{\varphi }({q}_{\mu })\oplus {V}_{\varphi }({\pi }_{\mu })$ (i.e. $6=2+2+2$ as sum of dimensions) under the differential rotation operator $\tfrac{d}{d\varphi }$. Note that ${V}_{\varphi }({j}_{\mu })$ means that the components of j form an irreducible entity under the differential operator $\tfrac{d}{d\varphi }$. Also Vt is irreducible under the differential time operator $\tfrac{d}{{dt}}$ in the de Sitter and Newton-Hooke case, that ${V}_{z}={V}_{z}({j}_{\mu },{\pi }_{\mu })\oplus {V}_{z}(q,E)$ (i.e. $6=4+2$) is irreducible under the differential altitude operator $\tfrac{d}{{dz}}$ in the de Sitter and Para-Poincaré cases, and finally that ${V}_{p}={V}_{p}({j}_{\mu },{q}_{\mu })\oplus {V}_{p}(\pi ,E)$ (i.e.6 = 4 + 2) is irreducible under the differential momentum up operator $\tfrac{d}{{dp}}$ in the de Sitter and Poincaré cases. The reader can also verify that the three dimensional manifolds are irreducible under the operator $\tfrac{d}{{dt}}$. They are direct sums (i.e. 3 = 2 + 1) under the operators $\tfrac{d}{{dz}}$ and $\tfrac{d}{{dp}}$. Finally the two dimensional ones are irreducible and correspond to the pairs containing the static Lie algebra as a daughter. Note also that all the ten coordinates on ${{ \mathcal G }}^{* }$ are constant in the Carroll and static Lie algebras cases.

Table 5.  Equations of change.

${{dimV}}_{s}$ 1st order DE 2nd order DE Lie algebras
6 $\tfrac{d{\alpha }_{\mu }}{d\varphi }={\alpha }_{\nu }{\epsilon }_{\mu }^{\nu };\,{\alpha }_{\mu }={j}_{\mu },{q}_{\mu },{\pi }_{\mu }$ $\tfrac{{d}^{2}{\alpha }_{\mu }}{d{\varphi }^{2}}=-{\alpha }_{\mu }$ All
6 $\tfrac{d}{{dt}}({q}_{i},{\pi }_{i})=\left(\tfrac{{\pi }_{i}}{m},\pm \tfrac{{q}_{i}}{C}\right)$ $\tfrac{{d}^{2}f}{{{dt}}^{2}}=\pm \tfrac{f}{{mC}}$ ${{dS}}_{\pm },{{NH}}_{\pm }$
3 $\tfrac{{{dq}}_{i}}{{dt}}=\tfrac{{\pi }_{i}}{m}$,   (cst velocity) No one P, G
3 $\tfrac{d{\pi }_{i}}{{dt}}=\pm \tfrac{{q}_{i}}{C}$,  (cst force) No one P± ,G±
6 $\tfrac{d}{{dz}}({j}_{\mu },{\pi }_{\mu })=({\pi }_{\nu },\mp \tfrac{{j}_{\nu }}{{{CE}}_{0}}){\epsilon }_{\mu }^{\nu },\tfrac{d}{{dz}}(q,E)=\left(\tfrac{E}{{E}_{0}},\mp \tfrac{q}{C}\right)$ $\tfrac{{d}^{2}f}{{{dz}}^{2}}=\mp \tfrac{f}{{{CE}}_{0}}$ ${{dS}}_{\pm }$ ,P±
3 $\tfrac{{{dj}}_{\mu }}{{dz}}={\pi }_{\nu }{\epsilon }_{\mu }^{\nu }$ (cst pseudo-momentum), $\tfrac{{dE}}{{dz}}=\mp \tfrac{q}{C}$ (cst force) No one ${{NH}}_{\pm }$ ,G±
3 $\tfrac{{{dj}}_{\mu }}{{dz}}={\pi }_{\nu }{\epsilon }_{\mu }^{\nu }$ (cst pseudo-momentum),$\tfrac{{dq}}{{dz}}=\tfrac{E}{{E}_{0}}$ (cst number) No one $P,C$
2 $\tfrac{{{dj}}_{\mu }}{{dz}}={\pi }_{\nu }{\epsilon }_{\mu }^{\nu }$ (cst pseudo-momentum) No one G, S
6 $\tfrac{d}{{dp}}({j}_{\mu },{q}_{\mu })=({q}_{\nu },\tfrac{{j}_{\nu }}{{{mE}}_{0}}){\epsilon }_{\mu }^{\nu },\tfrac{d}{{dp}}(\pi ,E)=-\left(\tfrac{E}{{E}_{0}},\tfrac{\pi }{m}\right)$ $\tfrac{{d}^{2}f}{{{dp}}^{2}}=-\tfrac{f}{{{mE}}_{0}}$ ${{dS}}_{\pm }$ ,P
3 $\tfrac{{{dj}}_{\mu }}{{dp}}={q}_{\nu }{\epsilon }_{\mu }^{\nu }$ (cst pseudo-position), $\tfrac{{dE}}{{dp}}=-\tfrac{\pi }{m}$ (cst velocity) No one ${{NH}}_{\pm }$, G
3 $\tfrac{{{dj}}_{\mu }}{{dp}}={q}_{\nu }{\epsilon }_{\mu }^{\nu }$ (pseudo-position),$\tfrac{d\pi }{{dp}}=-\tfrac{E}{{E}_{0}}$ (cst number) No one P, C
2 $\tfrac{{{dj}}_{\mu }}{{dp}}={q}_{\nu }{\epsilon }_{\mu }^{\nu }$, (pseudo-position) No one G± , S

7. Conclusion

In this paper we have shown how to obtain straightfully all kinematical Lie algebras from the de Sitter Lie algebras through contraction using dynamical parameters mass m, compliance C and energy E0. We consider that it is the best kinematical Lie algebras contraction process. We had also a little glance at the physics associated to each kinematical Lie algebra. We noticed that ${V}_{\varphi }={V}_{\varphi }({j}_{\mu })\oplus {V}_{\varphi }({q}_{\mu })\oplus {V}_{\varphi }({\pi }_{\mu })$ (i.e. 6 = 2 + 2 + 2) under the differential rotation operator $\tfrac{d}{d\varphi }$, where ${V}_{s}({j}_{\mu })$ means the components of j form an irreducible entity under the differential operator $\tfrac{d}{{ds}}$. Also Vt is irreducible under the differential time operator $\tfrac{d}{{dt}}$ in the de Sitter and Newton-Hooke case, that ${V}_{z}={V}_{z}({j}_{\mu },{\pi }_{\mu })\oplus {V}_{z}(q,E)$ (i.e. 6 = 4 + 2) under the differential altitude operator $\tfrac{d}{{dz}}$ in the de Sitter and Para-Poincaré cases and finally that ${V}_{p}={V}_{p}({j}_{\mu },{q}_{\mu })\oplus {V}_{p}(\pi ,E)$ (i.e.6 = 4 + 2) under the differential momentum operator $\tfrac{d}{{dp}}$ in the de Sitter and Poincaré cases.

We notice from table 4 that positions do not commute in the de Sitter and Poincaré cases, that linear momenta do not commute in de Sitter and Para-Poicare cases and that the uncertainty—like relation $\left\{{\pi }_{i},{q}_{j}\right\}=\tfrac{E}{{E}_{0}}{\delta }_{{ij}}$ occurs in finite energy (relative time) groups.

Appendix

This appendix serves to show how non additive parameters such as Lorentz boost are obtained from additive ones such as momentum. We use the brackets of table 2.

A.1. From momentum to boost

The Poincaré Lie algebra in one space dimension is defined by the Lie brackets

Equation (106)

where Q genrates momenta, P generates space translations while P generates time translations. We verify that $\exp ({x}_{0}^{{\prime} }P+{t}_{0}^{{\prime} }H)={{Ad}}_{\exp ({pQ}+{xP}+{tH})}(\exp ({x}_{0}P+{t}_{0}H))$ gives the Poincare space-time transformations

Equation (107)

where p is an additive momentum.

If we define the boost by $v=\sqrt{\tfrac{{E}_{0}}{m}}\tanh \left(\tfrac{p}{\sqrt{{{mE}}_{0}}}\right)$, then we recover the corresponding non additive boosts composition law $v^{\prime\prime} =\tfrac{v+v^{\prime} }{1+\tfrac{{mvv}^{\prime} }{{E}_{0}}}$. It is the usual Lorentz one when ${E}_{0}={{mc}}^{2}$. Similary if slowness is defined by $s=\sqrt{\tfrac{m}{{E}_{0}}}\tanh \left(\tfrac{p}{\sqrt{{{mE}}_{0}}}\right)$, then the slowness composition law is $s^{\prime\prime} =\tfrac{s+s^{\prime} }{1+\tfrac{{E}_{0}{vv}^{\prime} }{m}}$.

The limits of (107) are given in the table

Equation (108)

where $v=\tfrac{p}{m}$ is a Galilean boost and $s=\tfrac{p}{{E}_{0}}$ is a Carrollian slowness.

A.2. From space translation to force

The Lie algebra of one spatial Para-Poincaré Lie algebra P± is defined by the Lie brackets

Equation (109)

We verify that $\exp ({p}_{0}^{{\prime} }Q+{t}_{0}^{{\prime} }H)={{Ad}}_{\exp ({pQ}+{xP}+{tH})}(\exp ({p}_{0}Q+{t}_{0}H))$ gives the Para-Poincare momentum-time transformations

Equation (110)

in the ${P}_{-}$ case and

Equation (111)

in the ${P}_{+}$ case. The additive parameter x is non compact (compact) in the ${P}_{-}({P}_{+})$ case.

If $f=\sqrt{\tfrac{{E}_{0}}{C}}\tanh \left(\tfrac{x}{\sqrt{{{CE}}_{0}}}\right)$ $\left(f=\sqrt{\tfrac{{E}_{0}}{C}}\tan \left(\tfrac{x}{\sqrt{{{CE}}_{0}}}\right)\right)$ is a force for ${P}_{-}$ (${P}_{+}$) while $\phi =\sqrt{\tfrac{C}{{E}_{0}}}\tanh \left(\tfrac{x}{\sqrt{{{CE}}_{0}}}\right)$ $\left(\phi =\sqrt{\tfrac{C}{{E}_{0}}}\tan \left(\tfrac{x}{\sqrt{{{CE}}_{0}}}\right)\right)$ is an inverse of force for ${P}_{-}$ (${P}_{+}$), then we get the non additive composition laws $f^{\prime\prime} =\tfrac{f+f^{\prime} }{1\mp \tfrac{{Cff}^{\prime} }{{E}_{0}}}$ and $\phi ^{\prime\prime} =\tfrac{\phi +\phi ^{\prime} }{1\mp \tfrac{{E}_{0}\phi \phi ^{\prime} }{C}}$ for the P± case. Moreover we have that

Equation (112)

where $f=\pm \tfrac{x}{C}$ is a force for the Para-Galilei case P± an force and $\phi =-\tfrac{x}{{E}_{0}}$ is a Carrollian inverse of force.

A.3. From time translation to dampinglike coefficient

The Lie algebra of one spatial Newton-Hooke Lie algebra ${{NH}}_{\pm }$ is defined by the Lie brackets

Equation (113)

We verify that $\exp ({p}_{0}^{{\prime} }Q+{x}_{0}^{{\prime} }P)={{Ad}}_{\exp ({pQ}+{xP}+{tH})}(\exp ({p}_{0}Q+{x}_{0}P))$ gives the Newton-Hooke momentum-space transformations

Equation (114)

in the NH+ case and

Equation (115)

in the ${{NH}}_{-}$ case. The additive parameter t is non compact (compact) in the ${{NH}}_{+}({{NH}}_{-})$ case. Moreover the dampinglike coefficient $b=\sqrt{\tfrac{m}{C}}\tanh \left(\tfrac{t}{\sqrt{{mC}}}\right)$ ($b=\sqrt{\tfrac{m}{C}}\tan \left(\tfrac{t}{\sqrt{{mC}}}\right)$) for the NH+(${{NH}}_{-}$) whose the dimension is MT−1 and $\beta =\sqrt{\tfrac{C}{m}}\tanh \left(\tfrac{t}{\sqrt{{mC}}}\right)$ and an inverse of a dampinglike $\left(\beta =\sqrt{\tfrac{C}{m}}\tan \left(\tfrac{t}{\sqrt{{mC}}}\right)\right)$ for the NH+$\left({{NH}}_{-}\right)$ whose the dimension is ${M}^{-1}T$ satisfy the non additive composition laws $b^{\prime\prime} =\tfrac{b+b^{\prime} }{1\pm \tfrac{{Cbb}^{\prime} }{m}}$ and $\beta ^{\prime\prime} =\tfrac{\beta +\beta ^{\prime} }{1\pm \tfrac{m\beta \beta ^{\prime} }{C}}$. for the ${{NH}}_{\pm }$ cases.

We finally have

Equation (116)

Please wait… references are loading.
10.1088/2399-6528/ab4683