This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Exact solutions of the Klein-Gordon equation in external electromagnetic fields on 3D de Sitter background

Alexey A. Magazev [email protected] Omsk State Technical University, Omsk, Russia Maria N. Boldyreva [email protected] Omsk State Technical University, Omsk, Russia
Abstract

In this study, we investigate the symmetry properties and the possibility of exact integration of the Klein–Gordon equation in the presence of an external electromagnetic field on 3D de Sitter background. We present an algorithm for constructing the first-order symmetry algebra and describe its structure in terms of Lie algebra extensions. Based on the well-known classification of the inequivalent subalgebras of the algebra 𝔰𝔬(1,3)\mathfrak{so}(1,3), we obtain the classification of the electromagnetic fields on dS3\mathrm{dS}_{3} admitting first-order symmetry algebras of the Klein–Gordon equation. Then, we select the integrable cases, and for each of them, we construct exact solutions, using the non-commutative integration method developed by Shapovalov and Shirokov. In Appendix, we present an original algebraic method for constructing the special local coordinates on de Sitter space, in which the basis vector fields for subalgebras of the algebra 𝔰𝔬(1,3)\mathfrak{so}(1,3) have the simplest form.

Introduction

One of the most important problems of modern theoretical and mathematical physics is the problem of finding exact solutions of relativistic wave equations in external fields. In quantum electrodynamics, for example, exact solutions of the Dirac and Klein–Gordon equations allow us to construct the so-called Furry picture. In the framework of this picture, the interaction with an external field is described exactly, and the interaction with a quantized photon field is taken into account using perturbation theory [1]. Exact solutions of the Dirac and Klein–Gordon equations are also extremely useful in studying various vacuum quantum effects in strong external fields, where the methods and approaches of the standard perturbation theory are not applicable (see, for instance, [2, 3]).

This study focuses on exact solutions of the Klein–Gordon equation in the presence of an external electromagnetic field on 3D de Sitter background dS3\mathrm{dS}_{3}. When writing this paper, we mainly followed two objectives. First of all, we would like to demonstrate the potential of a new powerful method for constructing exact solutions of relativistic wave equations called the method of noncommutative integration. This method was developed in the early 90s of the last century by Shapovalov and Shirokov [4]. In contrast to the well-known method of separation of variables, the application of which is associated with the study of commutative subalgebras of the first- and second-order symmetry operators (see, for instance, [5, 6, 7, 8]), the method of noncommutative integration allows us to construct exact solutions using only the algebra of the first-order symmetry operators, which is noncommutative in general [9, 10, 11]. Secondly, in the paper, we would like to give the exhaustive classification of electromagnetic fields on dS3\mathrm{dS}_{3} that admit the existence of a first-order symmetry algebra for the Klein-Gordon equation and, using this algebra, to construct the corresponding exact solutions in cases where this is possible.

Note that nn-dimensional de Sitter space dSn\mathrm{dS}_{n} is regularly in the center of the attention of specialists in mathematical and theoretical physics, which is mainly owing to two reasons. On the one side, from the modern viewpoint, it is considered that, in the four-dimensional case, this space describes the very early stages of the expansion of the universe quite accurately. On the other side, the space dSn\mathrm{dS}_{n} is a maximally symmetric Lorentzian manifold satisfying Einstein’s equations with a positive cosmological constant. In particular, the high symmetry of this space allows one to integrate relativistic wave equations and, as a consequence, to investigate in detail of various quantum effects on its background [2, 12, 13, 14]. The presence of an external electromagnetic field, however, can cause the symmetry breaking that makes the problem of classification of electromagnetic fields admitting exact solutions of relativistic wave equations relevant. We note that although the problem of integrating free relativistic equations on de Sitter background has long been solved (see, for instanse, [2, 15, 16]), exact solutions of these equations in external electromagnetic fields are still poorly systematized. At the same time, such exact solutions are extremely important, for instance, in connection with the study of the particle creation effect in an external electromagnetic field on a curved spacetime [17, 18, 19, 20].

The structure of this paper is as follows. In Section 1, we recall the necessary information about the algebra of the first-order symmetry operators for the Klein–Gordon equation. Based on the defining equations for these operators (see [21, 22]), we present an algorithm for constructing this algebra and describe its structure in terms of Lie algebra extensions. In Section 2, we outline some geometric and group properties of 3D de Sitter space as well as the well-known classification of all inequivalent subalgebras of the algebra 𝔰𝔬(1,3)\mathfrak{so}(1,3), which is the Lie algebra of the isometry group of dS3\mathrm{dS}_{3}. Using this classification, in Section 3, we obtain the list of all electromagnetic fields on dS3\mathrm{dS}_{3} that admit nontrivial first-order symmetry algebras of the Klein–Gordon equation. In the same section, we explicitly construct these algebras and describe their structures by writing down the corresponding commutation relations. In Section 4, we examine the integrability problem for the Klein–Gordon equation in external electromagnetic fields on dS3\mathrm{dS}_{3}. We emphasize that by integrability we mean that the Klein-Gordon equation can be constructively solved by the process of its reduction to an ordinary differential or algebraic equation (as it is understood in the theory of separation of variables). Using the condition of noncommutative integrability [4], we find all the integrable cases, and, for each of them, we reduce the original Klein–Gordon equation to an auxiliary ordinary differential equation. In addition, we express the general solution of the auxiliary equation in terms of known special functions where possible.

All local constructions are presented in special local coordinates on dS3\mathrm{dS}_{3}, in which the basis vector fields for subalgebras of 𝔰𝔬(1,3)\mathfrak{so}(1,3) have the simplest form. In Appendix, the original algebraic method for constructing such coordinates is described.

1 First-order Klein–Gordon symmetry operators

Let (M,g)(M,g) be a smooth nn-dimensional Lorentzian manifold of signature (+,,,)(+,-,\dots,-), x1,,xnx^{1},\dots,x^{n} are local coordinates on MM. We consider the covariant Klein–Gordon equation for a charged massive scalar field in the presence of an external electromagnetic field [2, 3]:

H^φ(gabDaDb+ζR+m2)φ=0.\hat{H}\varphi\equiv\left(g^{ab}D_{a}D_{b}+\zeta R+m^{2}\right)\varphi=0. (1)

Here, φ\varphi denotes the scalar field, gabg^{ab} are the contravariant components of the metric gg, RR is the Riemannian scalar curvature, mm and ee are the mass and charge of the field quanta, respectively. The generalized covariant derivatives DaD_{a} are defined as Daaie𝒜aD_{a}\equiv\nabla_{a}-ie\mathscr{A}_{a}, where a\nabla_{a} is the covariant derivative with respect to the coordinate vector field xa/xa\partial_{x^{a}}\equiv\partial/\partial x^{a}, and 𝒜a\mathscr{A}_{a} are the components of the electromagnetic potential 1-form 𝒜\mathscr{A}. The dimensionless parameter ζ\zeta can take two values: ζ=0\zeta=0 (the minimally coupled case) and ζ=(n1)/(4n)\zeta=(n-1)/(4n) (the conformally coupled case). In what follows we will use the Einstein summation convention unless stated otherwise. Latin indices are raised and lowered by the metric gg.

An operator X^\hat{X} is said to be a symmetry operator for Eq. (1) if the following condition is satisfied

[H^,X^]H^X^X^H^=Q^H^,[\hat{H},\hat{X}]\equiv\hat{H}\hat{X}-\hat{X}\hat{H}=\hat{Q}\hat{H}, (2)

where Q^\hat{Q} is an operator which depends on X^\hat{X} in general. It is known that the symmetry operators map solutions of Eq. (1) to other solutions, and the set of all symmetry operators 𝔐\mathfrak{M} forms a Lie algebra called the symmetry algebra of the Klein-Gordon equation [5].

The main object of our investigation is a subalgebra of 𝔐\mathfrak{M} generated by all the differential operators of the form

X^=Xa(x)Da+ieχ(x).\hat{X}=X^{a}(x)D_{a}+ie\chi(x). (3)

Here, Xa(x)X^{a}(x) and χ(x)\chi(x) are assumed to be smooth real functions on MM, possibly not everywhere defined. Obviously, the set of all such operators forms a certain subalgebra in 𝔐\mathfrak{M}, which we denote as 𝒢^\hat{\mathscr{G}}.

The conditions under which the operator (3) is a symmetry operator of the Klein–Gordon equation are well known (see, for instance, [21]). They can be obtained by substituting (3) into (2) and equating coefficients of DaDbD_{a}D_{b}, DaD_{a}, and 11 on both sides of the resulting relation. As a result, we obtain the system of equations

aXb+bXa=0,\nabla^{a}X^{b}+\nabla^{b}X^{a}=0, (4)
Xbabaχ=0,X^{b}\mathscr{F}_{ab}-\nabla_{a}\chi=0, (5)

where ab\mathscr{F}_{ab} are the components of the electromagnetic tensor defined by

ab=𝒜bxa𝒜axb.\mathscr{F}_{ab}=\frac{\partial\mathscr{A}_{b}}{\partial x^{a}}-\frac{\partial\mathscr{A}_{a}}{\partial x^{b}}.

Also, note that, for a massive scalar field, there should be Q^=0\hat{Q}=0.

The system of Eqs. (4) and (5) is a necessary and sufficient condition for X^\hat{X} to be a symmetry operator for the Klein–Gordon equation. In particular, Eq. (4) means that X=XaxaX=X^{a}\partial_{x^{a}} is a Killing vector of the Lorentzian manifold (M,g)(M,g). Killing vectors reflect the geometric symmetry of the spacetime and are the infinitesimal generators of its (local) isometry group Iso(M,g)\mathrm{Iso}(M,g). The set of all Killing vectors of (M,g)(M,g) forms a Lie algebra (with respect to the commutator of vector fields), which we denote as 𝔦𝔰𝔬(M,g)\mathfrak{iso}(M,g).

We note that not every Killing vector X=XaxaX=X^{a}\partial_{x^{a}} from 𝔦𝔰𝔬(M,g)\mathfrak{iso}(M,g) can be extended to a symmetry operator X^=XaDa+ieχ\hat{X}=X^{a}D_{a}+ie\chi of Eq. (1). Carter showed [21] that a scalar function χ\chi of the required form locally exists if and only if

X=0,\mathscr{L}_{X}\mathscr{F}=0, (6)

where X\mathscr{L}_{X} denotes the Lie derivative along the vector field XX. Indeed, if we rewrite Eq. (5) in terms of differential forms

dχ=iX,d\chi=-i_{X}\mathscr{F}, (7)

where iXabXadxbi_{X}\mathscr{F}\equiv\mathscr{F}_{ab}X^{a}dx^{b} is the interior product of the 2-form =12abdxadxb\mathscr{F}=\frac{1}{2}\,\mathscr{F}_{ab}\,dx^{a}\wedge dx^{b} and the vector field X=XaxaX=X^{a}\partial_{x^{a}}, then, it is easy to see that the local existence of the function χ\chi is equivalent to the statement that the 1-form iX-i_{X}\mathscr{F} is closed. This, in turn, is equivalent to Eq. (6), in view of the Cartan formula X=iXd+diX\mathscr{L}_{X}=i_{X}d+di_{X} and the condition d=0d\mathscr{F}=0.

Since the correspondence XXX\to\mathscr{L}_{X} is a homomorphism of Lie algebras, the set of all Killing vectors satisfying the condition (6) for a given electromagnetic field forms a subalgebra 𝒢\mathscr{G} of the Lie algebra 𝔦𝔰𝔬(M,g)\mathfrak{iso}(M,g). We will call 𝒢\mathscr{G} the admissible subalgebra.

Let XA=XAa(x)xaX_{A}=X_{A}^{a}(x)\partial_{x^{a}} be Killing vectors forming a basis of 𝒢\mathscr{G}, A=1,,dim𝒢A=1,\dots,\dim\mathscr{G}. Since 𝒢\mathscr{G} is a subalgebra, we have

[XA,XB]=CABCXC,A,B=1,,dim𝒢.[X_{A},X_{B}]=C_{AB}^{C}X_{C},\quad A,B=1,\dots,\dim\mathscr{G}. (8)

Here, the constants CABCC_{AB}^{C} are the structure constants of 𝒢\mathscr{G}. By the definition of the admissible subalgebra, for each vector field XAX_{A}, there exists a local function χA\chi_{A} satisfying the condition (7). Note that this function is not uniquely defined; instead of χA\chi_{A}, one can choose the function χA=χA+λA\chi_{A}^{\prime}=\chi_{A}+\lambda_{A}, where λA\lambda_{A} is an arbitrary constant. The meaning of this ambiguity will be explained below.

By construction, the first-order differential operators

X^A=XAa(x)Da+ieχA,A=1,,dim𝒢,\hat{X}_{A}=X_{A}^{a}(x)D_{a}+ie\chi_{A},\quad A=1,\dots,\dim\mathscr{G}, (9)

commute with the operator H^\hat{H}, i.e., X^A\hat{X}_{A} are symmetry operators for the Klein–Gordon equation (1). In the general case, however, the linear span of these operators does not form the whole algebra 𝒢^\hat{\mathscr{G}}. Indeed, using the commutation relations [Da,Db]=ieab[D_{a},D_{b}]=-ie\mathscr{F}_{ab} and (8), the commutator of X^A\hat{X}_{A} and X^B\hat{X}_{B} can be written as [11]

[X^A,X^B]=CABCX^C+𝐅ABX^0,[\hat{X}_{A},\hat{X}_{B}]=C_{AB}^{C}\hat{X}_{C}+\mathbf{F}_{AB}\hat{X}_{0}, (10)

where

𝐅AB(XA,XB)CABCχC.\mathbf{F}_{AB}\equiv\mathscr{F}(X_{A},X_{B})-C_{AB}^{C}\chi_{C}. (11)

Here, we have introduced the trivial symmetry operator X^0ie\hat{X}_{0}\equiv ie. In Ref. [11], it is shown that the quantities 𝐅AB\mathbf{F}_{AB} are constant on MM and satisfy the identity

CABD𝐅CD+CBCD𝐅AD+CCAD𝐅BD=0.C_{AB}^{D}\mathbf{F}_{CD}+C_{BC}^{D}\mathbf{F}_{AD}+C_{CA}^{D}\mathbf{F}_{BD}=0.

This allows us to interpret the quantities (11) as components of a cocycle of the admissible subalgebra 𝒢\mathscr{G} with values in the trivial 𝒢\mathscr{G}-module \mathbb{R} (for a review of basic results in Lie algebra cohomology and their applications in mathematical physics see [23, 24]). Moreover, from the commutation relations (10), it follows that the algebra 𝒢^\hat{\mathscr{G}} is a one-dimensional central extension of the Lie algebra 𝒢\mathscr{G} corresponding to the cocycle (11).

As we have already noted, the functions χA\chi_{A} are determined by Eq. (7) up to the addition of arbitrary constants. As can be seen from (9), the transformation χAχA=χA+λA\chi_{A}\to\chi_{A}^{\prime}=\chi_{A}+\lambda_{A} implies the transformation X^AX^A=X^A+λAX^0\hat{X}_{A}\to\hat{X}^{\prime}_{A}=\hat{X}_{A}+\lambda_{A}\hat{X}_{0}, which is a change of basis in the algebra 𝒢^\hat{\mathscr{G}}. In the general case, it leads to a change of the commutation relations (10), because the quantities (11) are transformed as follows:

𝐅AB𝐅AB=𝐅ABCABCλC.\mathbf{F}_{AB}\to\mathbf{F}^{\prime}_{AB}=\mathbf{F}_{AB}-C_{AB}^{C}\lambda_{C}. (12)
Remark 1.

In some cases, it may be possible to choose the constants λC\lambda_{C} in Eq. (12) such that all quantities 𝐅AB\mathbf{F}^{\prime}_{AB} vanish. Cocycles of this kind are called trivial cocycles or coboundaries. For example, any cocycle of a semisimple Lie algebra is trivial; this result is the content of Whitehead’s second lemma [25]. In the case of a trivial cocycle, the structure of the algebra 𝒢^\hat{\mathscr{G}} is especially simple, since in this situation the symmetry operators X^A\hat{X}_{A} can be chosen so that

[X^A,X^B]=CABCX^C.[\hat{X}_{A},\hat{X}_{B}]=C_{AB}^{C}\hat{X}_{C}.

This means that 𝒢^\hat{\mathscr{G}} is isomorphic to the direct sum of the algebras 𝒢\mathscr{G} and \mathbb{R}: 𝒢^𝒢\hat{\mathscr{G}}\simeq\mathscr{G}\oplus\mathbb{R}.

The above results allow us to list all electromagnetic fields admitting first-order symmetry operators of Eq. (1) for a given Lorentzian manifold (M,g)(M,g). It is preliminary convenient to introduce the following equivalence relation. Two electromagnetic fields with closed 2-forms \mathscr{F} and \mathscr{F}^{\prime} are called equivalent if they are connected by an isometry:

=τ,τIso(M,g).\mathscr{F}=\tau^{*}\mathscr{F}^{\prime},\quad\tau\in\mathrm{Iso}(M,g).

It is easy to show that the admissible subalgebras 𝒢\mathscr{G} and 𝒢\mathscr{G}^{\prime} corresponding to the equivalent electromagnetic fields \mathscr{F} and \mathscr{F}^{\prime} are conjugate in 𝔦𝔰𝔬(M,g)\mathfrak{iso}(M,g) by the transformation τ\tau: 𝒢=τ𝒢τ1\mathscr{G}^{\prime}=\tau\mathscr{G}\tau^{-1}. Conversely, if the subalgebras 𝒢\mathscr{G} and 𝒢\mathscr{G}^{\prime} are τ\tau-conjugate, then the spaces of closed 2-forms on MM invariant under the subalgebras 𝒢\mathscr{G} and 𝒢\mathscr{G}^{\prime} are connected by the transformation τ\tau^{*}. Thus, in order to describe fully the electromagnetic fields that admit a nontrivial first-order symmetry operator of Eq. (1), one may use the list of subalgebras of the algebra 𝔦𝔰𝔬(M,g)\mathfrak{iso}(M,g) up to conjugations. For each subalgebra 𝒢={XA}\mathscr{G}=\{X_{A}\} from this list, we can find the most general form of a closed 2-form \mathscr{F} such that111Note that the classes of electromagnetic fields corresponding non-conjugate subalgebras of 𝔦𝔰𝔬(M,g)\mathfrak{iso}(M,g) may have nonzero intersections.

XA=0,A=1,,dim𝒢,\mathscr{L}_{X_{A}}\mathscr{F}=0,\quad A=1,\dots,\dim\mathscr{G}, (13)

and then explicitly construct a basis of the Lie algebra 𝒢^\hat{\mathscr{G}} in accordance with the formula (9).

2 The algebra of Killing vectors of dS3\mathrm{dS}_{3} and its inequivalent subalgebras

De Sitter space dS3\mathrm{dS}_{3} is the three-dimensional one-sheeted hyperboloid in four-dimensional Minkowski space 1,3\mathbb{R}^{1,3}, described by the equation [2]

x02x12x22x32=α2.x_{0}^{2}-x_{1}^{2}-x_{2}^{2}-x_{3}^{2}=-\alpha^{2}. (14)

Here, α\alpha is a nonzero positive constant with units of length called the de Sitter radius. The metric of dS3\mathrm{dS}_{3} induced by the standard Minkowski metric ηij=diag(1,1,1,1)\eta_{ij}=\mathrm{diag}(1,-1,-1,-1) is non-degenerate and has the Lorentzian signature (+,,)(+,-,-). From the group-theoretic point of view, three-dimensional de Sitter space is the homogeneous space O(1,3)/O(1,2)O(1,3)/O(1,2), where O(1,n)O(1,n) denotes the pseudo-orthogonal group of all linear transformations that leave invariant a non-degenerate quadratic form of signature (1,n)(1,n). From the topological point of view, dS3\mathrm{dS}_{3} is the direct product of the real line and the two-dimensional sphere: dS3=×S2\mathrm{dS}_{3}=\mathbb{R}\times S^{2}. From now on, we assume that α=1\alpha=1. The scalar curvature of dS3\mathrm{dS}_{3}, in this case, is equal to R=6R=6.

The isometry group of de Sitter space dS3\mathrm{dS}_{3} is the Lorentz group O(1,3)O(1,3). The identity component SO+(1,3)SO^{+}(1,3) of this group, called the restricted Lorentz group, is generated by the six Killing vectors:

Jij=xixjxjxi,0i<j3.J_{ij}=x_{i}\,\frac{\partial}{\partial x^{j}}-x_{j}\frac{\partial}{\partial x^{i}},\quad 0\leq i<j\leq 3. (15)

(Raising and lowering the indices are performed by the metric ηij\eta_{ij}). The Lie algebra 𝔰𝔬(1,3)\mathfrak{so}(1,3) defined by these vector fields has the following commutation relations:

[Jij,Jkl]=ηjkJilηikJjl+ηilJjkηjlJik,i,j=0,1,2,3.[J_{ij},J_{kl}]=\eta_{jk}J_{il}-\eta_{ik}J_{jl}+\eta_{il}J_{jk}-\eta_{jl}J_{ik},\quad i,j=0,1,2,3.

We note that the vector fields J12J_{12}, J13J_{13}, and J23J_{23} correspond to ordinary spatial rotations, while the Killing vectors J01,J02J_{01},J_{02}, and J03J_{03} are associated with the Lorentz boosts.

As noted in the previous section, the first step in the classification of electromagnetic fields that admit first-order symmetry operators of Eq. (1) on dS3\mathrm{dS}_{3} is to list all inequivalent subalgebras of the algebra 𝔰𝔬(1,3)\mathfrak{so}(1,3). Every such subalgebra 𝒢𝔰𝔬(1,3)\mathscr{G}\subset\mathfrak{so}(1,3) is fixed by a set of nn linearly independent vector fields XA=i<jCAijJijX_{A}=\sum_{i<j}C_{A}^{ij}J_{ij}, which are some linear combinations of the Killing vectors (15).

All inequivalent proper subalgebras of the algebra 𝔰𝔬(1,3)\mathfrak{so}(1,3) are well known (see, for instance, [26, 27, 28]); their exhaustive list (up to conjugation) is given in Table 1. In this table, we label each inequivalent subalgebra 𝒢n,m\mathscr{G}_{n,m} by its dimension nn and its number mm in the list. If there is a one-parametric family of subalgebras, we denote this family as 𝒢n,ma\mathscr{G}^{a}_{n,m}, where aa is a parameter. The third column of the table shows the nonzero commutation relations between the basis vectors of the subalgebras.

Table 1: Inequivalent subalgebras of 𝔰𝔬(1,3)\mathfrak{so}(1,3).
Subalgebra Infinitesimal generators XAX_{A} Commutation relations
𝒢1,1\mathscr{G}_{1,1} X1=J03X_{1}=J_{03}
𝒢1,2\mathscr{G}_{1,2} X1=J12X_{1}=J_{12}
𝒢1,3a\mathscr{G}_{1,3}^{a} X1=J12+aJ03X_{1}=J_{12}+aJ_{03}, a>0a>0
𝒢1,4\mathscr{G}_{1,4} X1=J13J01X_{1}=J_{13}-J_{01}
𝒢2,1\mathscr{G}_{2,1} X1=J13J01X_{1}=J_{13}-J_{01}, X2=J23J02X_{2}=J_{23}-J_{02}
𝒢2,2\mathscr{G}_{2,2} X1=J12X_{1}=J_{12}, X2=J03X_{2}=J_{03}
𝒢2,3\mathscr{G}_{2,3} X1=J13J01X_{1}=J_{13}-J_{01}, X2=J03X_{2}=J_{03} [X1,X2]=X1[X_{1},X_{2}]=-X_{1}
𝒢3,1\mathscr{G}_{3,1} X1=J13J01X_{1}=J_{13}-J_{01}, X2=J23J02X_{2}=J_{23}-J_{02}, X3=J03X_{3}=J_{03} [X1,X3]=X1,[X2,X3]=X2[X_{1},X_{3}]=-X_{1},\ [X_{2},X_{3}]=-X_{2}
𝒢3,2\mathscr{G}_{3,2} X1=J13J01X_{1}=J_{13}-J_{01}, X2=J23J02X_{2}=J_{23}-J_{02}, X3=J12X_{3}=J_{12} [X1,X3]=X2,[X2,X3]=X1[X_{1},X_{3}]=X_{2},\ [X_{2},X_{3}]=-X_{1}
𝒢3,3a\mathscr{G}_{3,3}^{a} X1=J13J01X_{1}=J_{13}-J_{01}, X2=J23J02X_{2}=J_{23}-J_{02}, [X1,X3]=X2aX1[X_{1},X_{3}]=X_{2}-aX_{1},
X3=J12+aJ03X_{3}=J_{12}+aJ_{03}, a>0a>0 [X2,X3]=X1aX2[X_{2},X_{3}]=-X_{1}-aX_{2}
𝒢3,4\mathscr{G}_{3,4} X1=J12,X2=J13,X3=J23X_{1}=J_{12},\ X_{2}=J_{13},\ X_{3}=J_{23} [X1,X2]=X3,[X1,X3]=X2[X_{1},X_{2}]=X_{3},\ [X_{1},X_{3}]=-X_{2},
[X2,X3]=X1[X_{2},X_{3}]=X_{1}
𝒢3,5\mathscr{G}_{3,5} X1=J01,X2=J02,X3=J12X_{1}=J_{01},\ X_{2}=J_{02},\ X_{3}=J_{12} [X1,X2]=X3,[X1,X3]=X2[X_{1},X_{2}]=X_{3},\ [X_{1},X_{3}]=X_{2},
[X2,X3]=X1[X_{2},X_{3}]=-X_{1}
𝒢4,1\mathscr{G}_{4,1} X1=J13J01X_{1}=J_{13}-J_{01}, X2=J23J02,X3=J12X_{2}=J_{23}-J_{02},\ X_{3}=J_{12}, [X1,X3]=X2[X_{1},X_{3}]=X_{2}, [X1,X4]=X1[X_{1},X_{4}]=-X_{1},
X4=J03X_{4}=J_{03} [X2,X3]=X1[X_{2},X_{3}]=-X_{1}, [X2,X4]=X2[X_{2},X_{4}]=-X_{2}

Let us give some explanations for the subalgebras listed in Table 1.

The one-dimensional subalgebras 𝒢1,1\mathscr{G}_{1,1}, 𝒢1,2\mathscr{G}_{1,2}, and 𝒢1,4\mathscr{G}_{1,4} generate one-parameter subgroups of Lorentz boosts SO(1,1)SO(1,1), rotations SO(2)SO(2), and null rotations SO(0,1)SO(0,1), respectively. The family of one-dimensional subalgebras 𝒢1,3a\mathscr{G}^{a}_{1,3} parametrized by the positive parameter aa generates the family of one-parameter loxodromic transformations.

The subalgebra 𝒢2,1\mathscr{G}_{2,1} is associated with a two-dimensional Abelian group of transformations consisting of zero rotations. The subalgebra 𝒢2,2\mathscr{G}_{2,2} corresponds to a two-dimensional Abelian transformation group consisting of Lorentz boosts (in the x3x^{3} direction), rotations in the plane x1x2x^{1}x^{2}, and their combinations. The subalgebra 𝒢2,3\mathscr{G}_{2,3} generates a non-Abelian two-dimensional group isomorphic to the affine group A(1)A(1).

The subalgebra 𝒢3,1\mathscr{G}_{3,1} is the Lie algebra of Bianchi type V; the corresponding group is isomorphic to the group of Euclidean homotheties Hom(2)\mathrm{Hom}(2). The subalgebra 𝒢3,2\mathscr{G}_{3,2} is the Lie algebra of Bianchi type VII0; it is isomorphic to the Lie algebra of the euclidean group E(2)E(2). The subalgebra 𝒢3,3a\mathscr{G}_{3,3}^{a}, where a>0a>0, is of Bianchi type VIIa. The subalgebra 𝒢3,4\mathscr{G}_{3,4} is of Bianchi type IX; the corresponding transformation group is the rotation group SO(3)SO(3). Finally, the subalgebra 𝒢3,5\mathscr{G}_{3,5} of Bianchi type VIII is isomorphic to the Lie algebra of the group SL(2,)SL(2,\mathbb{R}).

There is only one (up to conjugation) four-dimensional subalgebra 𝒢4,1\mathscr{G}_{4,1}; it is isomorphic to the Lie algebra of the group of Euclidean similitudes Sim(2)\mathrm{Sim}(2).

There are no five-dimensional subalgebras of the Lie algebra 𝔰𝔬(1,3)\mathfrak{so}(1,3).

3 Electromagnetic fields admitting nontrivial first-order symmetry algebras on dS3\mathrm{dS}_{3}

For subsequent purposes, we need to fix a local coordinate system in the space dS3\mathrm{dS}_{3} and rewrite the basis vector fields XAX_{A} for each of the subalgebras 𝒢n,m𝔰𝔬(1,3)\mathscr{G}_{n,m}\subset\mathfrak{so}(1,3) in these coordinates. Since our problem is to find a closed 2-form \mathscr{F} satisfying Eq. (13), it is reasonable to use a coordinate system in which the vector fields XAX_{A} have the simplest form. In this sense, local coordinates that “rectify” the integral submanifolds of the system of vector fields X1,,XnX_{1},\dots,X_{n} are most suitable. Let us recall the rigorous definition of such coordinates in the context of the situation under consideration.

Let 𝒢n,m\mathscr{G}_{n,m} be a subalgebra from Table 1, X1,,XnX_{1},\dots,X_{n} are its basis vector fields. Denote by rr the dimension of the space spanned by the vectors X1|x,,Xn|xX_{1}|_{x},\dots,X_{n}|_{x} for a point xdS3x\in\mathrm{dS}_{3} in general position, i. e.

rsupxdS3rankXAa(x).r\equiv\sup\limits_{x\in\mathrm{dS}_{3}}\mathrm{rank}\,\|X_{A}^{a}(x)\|. (16)

If x0dS3x_{0}\in\mathrm{dS}_{3} is a point in general position, then, according to the well-known Frobenius theorem, there are local coordinates (q,u)=(q1,,qr(q,u)=(q^{1},\dots,q^{r}, u1,,u3r)u^{1},\dots,u^{3-r}) near of x0dS3x_{0}\in\mathrm{dS}_{3} such that the integral submanifolds of {X1,,Xn}\{X_{1},\dots,X_{n}\} intersect this coordinate chart along the “slices” u1=c1,,u3r=c3ru^{1}=c_{1},\dots,u^{3-r}=c_{3-r}, where c1,,c3rc_{1},\dots,c_{3-r} are arbitrary constants (see Ref. [29]). In this case, q1,,qrq^{1},\dots,q^{r} are regarded as local coordinates on the “slices”, and the vector fields XAX_{A} being tangent to the “slices” are written as

XA=a=1rXAa(q,u)qa,A=1,,n.X_{A}=\sum\limits_{a=1}^{r}X_{A}^{a}(q,u)\frac{\partial}{\partial q^{a}},\quad A=1,\dots,n.

In turn, the coordinates u1,,u3ru^{1},\dots,u^{3-r} are local invariants of the transformation group Gn,mG_{n,m} generated by the Lie algebra 𝒢n,m\mathscr{G}_{n,m}. We call the coordinates (q,u)=(q1,,qr(q,u)=(q^{1},\dots,q^{r}, u1,,u3r)u^{1},\dots,u^{3-r}) the rectifying coordinates associated with the subalegbras 𝒢n,m\mathscr{G}_{n,m}.

Let us now consider each of the subalgebras from Table 1. In each case, it is not hard to construct local rectifying coordinates associated with 𝒢n,m\mathscr{G}_{n,m} and then to find the most general form of the closed 2-form \mathscr{F} satisfying Eq. (13). The results of these calculations are summarized in Table 2, in which for each subalgebra 𝒢n,m\mathscr{G}_{n,m} we indicate the number rr calculated by Eq. (16), the vector fields XAX_{A}, and the corresponding invariant closed 2-form \mathscr{F} in the rectifying coordinates. In this table, f,f1,f2f,f_{1},f_{2} denote arbitrary smooth functions of their arguments, and μ,μ1,μ2\mu,\mu_{1},\mu_{2} are arbitrary constants. The functions xi=xi(q1,,qr,u1,,u3r)x^{i}=x^{i}(q^{1},\dots,q^{r},u^{1},\dots,u^{3-r}) defining the rectifying local coordinates associated with the subalgebras 𝒢n,m\mathscr{G}_{n,m} as well as a constructive algebraic method for their construction are given in Appendix A.

Table 2: Basis vector fields XA𝒢n,mX_{A}\in\mathscr{G}_{n,m} and the corresponding closed invariant 2-forms \mathscr{F} in rectifying local coordinates (q,u)(q,u).
Subalgebra rr Infinitesimal generators XAX_{A} Closed 2-form \mathscr{F}, such that
XA=0\mathscr{L}_{X_{A}}\mathscr{F}=0 for all A=1,,nA=1,\dots,n
𝒢1,1\mathscr{G}_{1,1} 1 X1=q1X_{1}=\partial_{q^{1}} dq1df1(u1,u2)+f2(u1,u2)du1du2dq^{1}\wedge df_{1}(u^{1},u^{2})+f_{2}(u^{1},u^{2})du^{1}\wedge du^{2}
𝒢1,2\mathscr{G}_{1,2}
𝒢1,3a\mathscr{G}_{1,3}^{a}
𝒢1,4\mathscr{G}_{1,4}
𝒢2,1\mathscr{G}_{2,1} 2 X1=q1,X2=q2X_{1}=\partial_{q^{1}},\ X_{2}=\partial_{q^{2}} μdq1dq2+f1(u1)dq1du1+\mu\,dq^{1}\wedge dq^{2}+f_{1}(u^{1})dq^{1}\wedge du^{1}+\hfill
𝒢2,2\mathscr{G}_{2,2} +f2(u1)dq2du1\hfill+f_{2}(u^{1})dq^{2}\wedge du^{1}
𝒢2,3\mathscr{G}_{2,3} X1=q1,X2=q1q1+q2X_{1}=\partial_{q^{1}},\quad X_{2}=-q^{1}\partial_{q^{1}}+\partial_{q^{2}} exp(q2)dq1(f1(u1)dq2+df1(u1))+\exp(q^{2})dq^{1}\wedge\left(f_{1}(u^{1})dq^{2}+df_{1}(u^{1})\right)+\hfill
+f2(u1)dq2du1\hfill+f_{2}(u^{1})dq^{2}\wedge du^{1}
𝒢3,1\mathscr{G}_{3,1} 3 X1=q1,X2=q2,X_{1}=\partial_{q^{1}},\ X_{2}=\partial_{q^{2}}, exp(q3)(μ1dq1+μ2dq2)dq3\exp(q^{3})(\mu_{1}dq^{1}+\mu_{2}dq^{2})\wedge dq^{3}
X3=q1q1q2q2+q3X_{3}=-q^{1}\partial_{q^{1}}-q^{2}\partial_{q^{2}}+\partial_{q^{3}}
𝒢3,2\mathscr{G}_{3,2} 2 X1=q1,X2=q2,X3=q2q1+q1q2X_{1}=\partial_{q^{1}},\ X_{2}=\partial_{q^{2}},\ X_{3}=-q^{2}\partial_{q^{1}}+q^{1}\partial_{q^{2}} μdq1dq2\mu\,dq^{1}\wedge dq^{2}
𝒢3,3a\mathscr{G}_{3,3}^{a} 3 X1=q1,X2=q2,X_{1}=\partial_{q^{1}},\ X_{2}=\partial_{q^{2}}, exp(aq3)(μ1cos(q3)+μ2sin(q3))dq1dq3+\exp(aq^{3})\left(\mu_{1}\cos(q^{3})+\mu_{2}\sin(q^{3})\right)dq^{1}\wedge dq^{3}+\hfill
X3=(aq1+q2)q1+(q1aq2)q2+q3X_{3}=-(aq^{1}+q^{2})\partial_{q^{1}}+(q^{1}-aq^{2})\partial_{q^{2}}+\partial_{q^{3}} +exp(aq3)(μ1sin(q3)μ2cos(q3))dq2dq3\hfill+\exp(aq^{3})\left(\mu_{1}\sin(q^{3})-\mu_{2}\cos(q^{3})\right)dq^{2}\wedge dq^{3}
𝒢3,4\mathscr{G}_{3,4} 2 X1=q1,X_{1}=\partial_{q^{1}}, μcos(q2)dq1dq2\mu\cos(q^{2})dq^{1}\wedge dq^{2}
X2=sin(q1)tan(q2)q1+cos(q1)q2,X_{2}=\sin(q^{1})\tan(q^{2})\partial_{q^{1}}+\cos(q^{1})\partial_{q^{2}},
X3=cos(q1)tan(q2)q1sin(q1)q2X_{3}=\cos(q^{1})\tan(q^{2})\partial_{q^{1}}-\sin(q^{1})\partial_{q^{2}}
𝒢3,5\mathscr{G}_{3,5} 2 X1=q1,X_{1}=\partial_{q^{1}},
X2=sinh(q1)tan(q2)q1+cosh(q1)q2,X_{2}=\sinh(q^{1})\tan(q^{2})\partial_{q^{1}}+\cosh(q^{1})\partial_{q^{2}},
X3=cosh(q1)tan(q2)q1+sinh(q1)q2X_{3}=\cosh(q^{1})\tan(q^{2})\partial_{q^{1}}+\sinh(q^{1})\partial_{q^{2}}
𝒢4,1\mathscr{G}_{4,1} 3 X1=q1,X2=q2,X3=q2q1+q1q2X_{1}=\partial_{q^{1}},\ X_{2}=\partial_{q^{2}},\ X_{3}=-q^{2}\partial_{q^{1}}+q^{1}\partial_{q^{2}}, 0
X4=q1q1q2q2+q3X_{4}=-q^{1}\partial_{q^{1}}-q^{2}\partial_{q^{2}}+\partial_{q^{3}}

By construction, the electromagnetic fields listed in Table 2 admit nontrivial first-order symmetry algebras of the Klein–Gordon eqiation (1). To construct bases of these algebras explicitly, we again consider each subalgebra 𝒢n,m\mathscr{G}_{n,m} separately. For each vector fields XA𝒢n,mX_{A}\in\mathscr{G}_{n,m} from Table 2, we find the function χA\chi_{A}, that is, a solution of Eq. (7), and we then write out the explicit form of the associated symmetry operator X^A\hat{X}_{A} according to Eq. (9). Having constructed the basis {X^0=ie,X^1,,X^n}\{\hat{X}_{0}=ie,\hat{X}_{1},\dots,\hat{X}_{n}\} of the symmetry algebra 𝒢^n,m\hat{\mathscr{G}}_{n,m}, we list the nonzero commutation relations between its basis elements.

Symmetry algebras 𝒢^1,1\hat{\mathscr{G}}_{1,1}, 𝒢^1,2\hat{\mathscr{G}}_{1,2}, 𝒢^1,3a\hat{\mathscr{G}}_{1,3}^{a} and 𝒢^1,4\hat{\mathscr{G}}_{1,4}

X^1=D1ief1(u1,u2).\hat{X}_{1}=D_{1}-ief_{1}(u^{1},u^{2}).

Symmetry algebras 𝒢^2,1\hat{\mathscr{G}}_{2,1} and 𝒢^2,2\hat{\mathscr{G}}_{2,2}

X^1=D1ie(12μq2+f1(u1)𝑑u1),X^2=D2+ie(12μq1+f2(u1)𝑑u1),\hat{X}_{1}=D_{1}-ie\left(\frac{1}{2}\,\mu q^{2}+\int f_{1}(u^{1})du^{1}\right),\quad\hat{X}_{2}=D_{2}+ie\left(\frac{1}{2}\,\mu q^{1}+\int f_{2}(u^{1})du^{1}\right),
[X^1,X^2]=μX^0.[\hat{X}_{1},\hat{X}_{2}]=\mu\hat{X}_{0}.

Symmetry algebra 𝒢^2,3\hat{\mathscr{G}}_{2,3}

X^1=D1,X^2=q1D1+D2ief2(u1)𝑑u1,\hat{X}_{1}=D_{1},\quad\hat{X}_{2}=-q^{1}D_{1}+D_{2}-ie\int f_{2}(u^{1})du^{1},
[X^1,X^2]=X^1.[\hat{X}_{1},\hat{X}_{2}]=-\hat{X}_{1}.

Symmetry algebra 𝒢^3,1\hat{\mathscr{G}}_{3,1}

X^1=D1ieμ1exp(q3),X^2=D2ieμ2exp(q3),X^3=q1D1q2D2+ieexp(q3)(μ1q1+μ2q2),\hat{X}_{1}=D_{1}-ie\mu_{1}\exp(q^{3}),\quad\hat{X}_{2}=D_{2}-ie\mu_{2}\exp(q^{3}),\\ \hat{X}_{3}=-q^{1}D_{1}-q^{2}D_{2}+ie\exp(q^{3})\left(\mu_{1}q^{1}+\mu_{2}q^{2}\right), (17)
[X^1,X^2]=0,[X^1,X^3]=X^1,[X^2,X^3]=X^2.[\hat{X}_{1},\hat{X}_{2}]=0,\quad[\hat{X}_{1},\hat{X}_{3}]=-\hat{X}_{1},\quad[\hat{X}_{2},\hat{X}_{3}]=-\hat{X}_{2}. (18)

Symmetry algebra 𝒢^3,2\hat{\mathscr{G}}_{3,2}

X^1=D1ieμq2,X^2=D2+ieμq1,X^3=q2D1+q1D2+i2eμ((q1)2+(q2)2),\hat{X}_{1}=D_{1}-ie\mu q^{2},\quad\hat{X}_{2}=D_{2}+ie\mu q^{1},\quad\hat{X}_{3}=-q^{2}D_{1}+q^{1}D_{2}+\frac{i}{2}\,e\mu\,\left((q^{1})^{2}+(q^{2})^{2}\right), (19)
[X^1,X^2]=μX^0,[X^1,X^3]=X^2,[X^2,X^3]=X^1.[\hat{X}_{1},\hat{X}_{2}]=\mu\hat{X}_{0},\quad[\hat{X}_{1},\hat{X}_{3}]=\hat{X}_{2},\quad[\hat{X}_{2},\hat{X}_{3}]=-\hat{X}_{1}. (20)

Symmetry algebra 𝒢^3,3a\hat{\mathscr{G}}_{3,3}^{a}

X^1=D1+ieexp(aq3)(μ2aμ11+a2cos(q3)μ1+aμ21+a2sin(q3)),\hat{X}_{1}=D_{1}+ie\exp(aq^{3})\left(\frac{\mu_{2}-a\mu_{1}}{1+a^{2}}\,\cos(q^{3})-\frac{\mu_{1}+a\mu_{2}}{1+a^{2}}\,\sin(q^{3})\right), (21)
X^2=D2+ieexp(aq3)(μ2aμ11+a2sin(q3)+μ1+aμ21+a2cos(q3)),\hat{X}_{2}=D_{2}+ie\exp(aq^{3})\left(\frac{\mu_{2}-a\mu_{1}}{1+a^{2}}\,\sin(q^{3})+\frac{\mu_{1}+a\mu_{2}}{1+a^{2}}\,\cos(q^{3})\right), (22)
X^3=(aq1+q2)D1+(q1aq2)D2+D3++ieexp(aq3)(μ1q1μ2q2)cos(q3)+ieexp(aq3)(μ1q2+μ2q1)cos(q3),\hat{X}_{3}=-\left(aq^{1}+q^{2}\right)D_{1}+\left(q^{1}-aq^{2}\right)D_{2}+D_{3}+\\ +ie\exp(aq^{3})\left(\mu_{1}q^{1}-\mu_{2}q^{2}\right)\cos(q^{3})+ie\exp(aq^{3})\left(\mu_{1}q^{2}+\mu_{2}q^{1}\right)\cos(q^{3}), (23)
[X^1,X^2]=0,[X^1,X^3]=X^2aX^1,[X^2,X^3]=X^1aX^2.[\hat{X}_{1},\hat{X}_{2}]=0,\quad[\hat{X}_{1},\hat{X}_{3}]=\hat{X}_{2}-a\hat{X}_{1},\quad[\hat{X}_{2},\hat{X}_{3}]=-\hat{X}_{1}-a\hat{X}_{2}. (24)

Symmetry algebra 𝒢^3,4\hat{\mathscr{G}}_{3,4}

X^1=D1ieμsin(q2),\hat{X}_{1}=D_{1}-ie\mu\sin(q^{2}), (25)
X^2=sin(q1)tan(q2)D1+cos(q1)D2+ieμsin(q1)cos(q2),\hat{X}_{2}=\sin(q^{1})\tan(q^{2})D_{1}+\cos(q^{1})D_{2}+ie\mu\sin(q^{1})\cos(q^{2}), (26)
X^3=cos(q1)tan(q2)D1sin(q1)D2+ieμcos(q1)cos(q2),\hat{X}_{3}=\cos(q^{1})\tan(q^{2})D_{1}-\sin(q^{1})D_{2}+ie\mu\cos(q^{1})\cos(q^{2}), (27)
[X^1,X^2]=X^3,[X^2,X^3]=X^1,[X^3,X^1]=X^2.[\hat{X}_{1},\hat{X}_{2}]=\hat{X}_{3},\quad[\hat{X}_{2},\hat{X}_{3}]=\hat{X}_{1},\quad[\hat{X}_{3},\hat{X}_{1}]=\hat{X}_{2}. (28)

Symmetry algebra 𝒢^3,5\hat{\mathscr{G}}_{3,5}

X^1=D1ieμsin(q2),\hat{X}_{1}=D_{1}-ie\mu\sin(q^{2}), (29)
X^2=sinh(q1)tan(q2)D1+cosh(q1)D2+ieμsinh(q1)cos(q2),\hat{X}_{2}=\sinh(q^{1})\tan(q^{2})D_{1}+\cosh(q^{1})D_{2}+ie\mu\sinh(q^{1})\cos(q^{2}), (30)
X^3=cosh(q1)tan(q2)D1+sinh(q1)D2+ieμcosh(q1)cos(q2),\hat{X}_{3}=\cosh(q^{1})\tan(q^{2})D_{1}+\sinh(q^{1})D_{2}+ie\mu\cosh(q^{1})\cos(q^{2}), (31)
[X^1,X^2]=X^3,[X^1,X^3]=X^2,[X^2,X^3]=X^1.[\hat{X}_{1},\hat{X}_{2}]=\hat{X}_{3},\quad[\hat{X}_{1},\hat{X}_{3}]=\hat{X}_{2},\quad[\hat{X}_{2},\hat{X}_{3}]=\hat{X}_{1}. (32)

Symmetry algebra 𝒢^4,1\hat{\mathscr{G}}_{4,1}

X^1=D1,X^2=D2,X^3=q2D1+q1D2,X^4=q1D1q2D2+D3,\hat{X}_{1}=D_{1},\quad\hat{X}_{2}=D_{2},\quad\hat{X}_{3}=-q^{2}D_{1}+q^{1}D_{2},\quad\hat{X}_{4}=-q^{1}D_{1}-q^{2}D_{2}+D_{3},
[X^1,X^2]=0,[X^1,X^3]=X^2,[X^1,X^4]=X^1,[X^2,X^3]=X^1,[X^2,X^4]=X^2.[\hat{X}_{1},\hat{X}_{2}]=0,\quad[\hat{X}_{1},\hat{X}_{3}]=\hat{X}_{2},\quad[\hat{X}_{1},\hat{X}_{4}]=-\hat{X}_{1},\quad[\hat{X}_{2},\hat{X}_{3}]=-\hat{X}_{1},\quad[\hat{X}_{2},\hat{X}_{4}]=-\hat{X}_{2}.

4 Noncommutative integrability of the Klein–Gordon equation on dS3\mathrm{dS}_{3}

In this study, integrability of the Klein–Gordon equation (1) is understood in the sense of the following definition [10, 4].

Definition 1.

The Klein-Gordon equation (1) is integrable if we can reduce the construction of a basis of its solutions to solving an ordinary differential equation.

Suppose that Eq. (1) admits some (n+1)(n+1)-dimensional first-order symmetry algebra 𝒢^\hat{\mathscr{G}} with the basis

X^0=ie,X^1=X1a(x)Da+ieχ1(x),,X^n=Xna(x)Da+ieχn(x).\hat{X}_{0}=ie,\ \hat{X}_{1}=X_{1}^{a}(x)D_{a}+ie\chi_{1}(x),\ \dots,\ \hat{X}_{n}=X_{n}^{a}(x)D_{a}+ie\chi_{n}(x). (33)

As shown in Section 1, this algebra is isomorphic to a one-dimensional central extension of the nn-dimensional Lie algebra 𝒢={XA=XAa(x)a,A=1,,n}\mathscr{G}=\{X_{A}=X_{A}^{a}(x)\partial_{a},\ A=1,\dots,n\} preserving the 2-form \mathscr{F} of the electromagnetic field. The symmetry operators (33) satisfy the commutation relations

[X^A,X^B]=CABCX^C+𝐅ABX^0,[\hat{X}_{A},\hat{X}_{B}]=C_{AB}^{C}\hat{X}_{C}+\mathbf{F}_{AB}\hat{X}_{0},

where CABCC_{AB}^{C} are the structure constants of the Lie algebra 𝒢\mathscr{G}, 𝐅AB\mathbf{F}_{AB} are the components of a cocycle of 𝒢\mathscr{G} with values in the trivial 𝒢\mathscr{G}-module \mathbb{R}. In order to avoid some technical difficulties, we assume that the infinitesimal generators XAX_{A} of the algebra 𝒢\mathscr{G} are algebraically independent, that is, there is no symmetrized polynomial PP in nn non-commuting variables such that P(X1,,Xn)0P(X_{1},\dots,X_{n})\equiv 0.

Let us describe a method for reducing the Klein–Gordon equation (1) to a differential equation including fewer independent variables than the original one. For this purpose, we adapt the method of noncommutative integration of linear partial differential equations suggested by Shapovalov and Shirokov (see [4, 9, 10]).

The basic ingredient for the noncommutative integration method is the so-called λ\lambda-representation of the Lie symmetry algebra 𝒢^\hat{\mathscr{G}}, that is, its operator-irreducible representation realized by an ind𝒢^\operatorname{\mathrm{ind}}\hat{\mathscr{G}}-parametric family of first-order operators acting in the space of partially holomorphic functions of (dim𝒢^ind𝒢^)/2(\dim\hat{\mathscr{G}}-\operatorname{\mathrm{ind}}\hat{\mathscr{G}})/2 variables. Here, the non-negative integer ind𝒢^\operatorname{\mathrm{ind}}\hat{\mathscr{G}}, called the index of the Lie algebra 𝒢^\hat{\mathscr{G}}, is defined as the dimension of regular orbits of the corresponding coadjoint representation [30].

In the case of 𝒢^\hat{\mathscr{G}}, the Lie algebra is realized by the operators (33); the λ\lambda-representation of 𝒢^\hat{\mathscr{G}} is defined by the operators

^0=ie,^A=aAμ(λ)λμ+bA(λ,J),A=1,,n,\hat{\ell}_{0}=-ie,\quad\hat{\ell}_{A}=a_{A}^{\mu}(\lambda)\,\frac{\partial}{\partial\lambda^{\mu}}+b_{A}(\lambda,J),\quad A=1,\dots,n,

which satisfy the commutation relations

[^A,^B]=CABC^C+𝐅AB^0,A,B=1,,n,[\hat{\ell}_{A},\hat{\ell}_{B}]=C_{AB}^{C}\hat{\ell}_{C}+\mathbf{F}_{AB}\hat{\ell}_{0},\quad A,B=1,\dots,n,

where λ=(λ1,,λs)s\lambda=(\lambda^{1},\dots,\lambda^{s})\in\mathbb{C}^{s}, s=(dim𝒢^ind𝒢^)/2s=\left(\dim\hat{\mathscr{G}}-\operatorname{\mathrm{ind}}\hat{\mathscr{G}}\right)/2, J=(J1,,Jl)lJ=(J_{1},\dots,J_{l})\in\mathbb{R}^{l}, l=ind𝒢^1l=\operatorname{\mathrm{ind}}\hat{\mathscr{G}}-1. The index ind𝒢^\operatorname{\mathrm{ind}}\hat{\mathscr{G}}, in this case, is calculated by the formula

ind𝒢^=dim𝒢^supf𝒢^rankCABCfC+𝐅ABf0,\operatorname{\mathrm{ind}}\hat{\mathscr{G}}=\dim\hat{\mathscr{G}}-\sup\limits_{f\in\hat{\mathscr{G}}^{*}}\operatorname{\mathrm{rank}}\|C_{AB}^{C}f_{C}+\mathbf{F}_{AB}f_{0}\|, (34)

where 𝒢^\hat{\mathscr{G}}^{*} is the dual space to the Lie algebra 𝒢^\hat{\mathscr{G}}. The operator irreducibility means that all Casimir invariants of the λ\lambda-representation are multiples of the identity operator. For convenience, we require that the operators ^A\hat{\ell}_{A} be skew-symmetric with respect to some measure dμ(λ)d\mu(\lambda). We note that there is an efficient computational algorithm for constructing λ\lambda-representations based on the fact that these representations can be obtained as a quantization result of the Lie–Poisson bracket in Darboux coordinates [31].

Now, we consider the system of equations

X^A(x,x)φJ(x,λ)=^A(λ,λ,J)φJ(x,λ),A=1,,n,\hat{X}_{A}(x,\partial_{x})\varphi_{J}(x,\lambda)=-\hat{\ell}_{A}(\lambda,\partial_{\lambda},J)\varphi_{J}(x,\lambda),\quad A=1,\dots,n, (35)

where φJ(x,λ)\varphi_{J}(x,\lambda) is a function of the variables x=(x1,,xm)x=(x^{1},\dots,x^{m}) and λ=(λ1,,λs)\lambda=(\lambda^{1},\dots,\lambda^{s}) depending on the real parameters J=(J1,,Jl)J=(J_{1},\dots,J_{l}). We note that this system of equations is compatible because the sets of the operators {X^A}\{\hat{X}_{A}\} and {^A}\{\hat{\ell}_{A}\} form representations of the same Lie algebra 𝒢^\hat{\mathscr{G}}. Since X^A\hat{X}_{A} are symmetry operators of the Klein–Gordon equation (1), the space of all solutions of the system (35) is invariant under the operator H^\hat{H}. Solving Eq. (35) by the method of characteristics, we have the general solution in the form

φJ(x,λ)=eRJ(x,λ)ΦJ(v1(x,λ),,vm~(x,λ)),\varphi_{J}(x,\lambda)=e^{R_{J}(x,\lambda)}\Phi_{J}(v^{1}(x,\lambda),\dots,v^{\tilde{m}}(x,\lambda)), (36)

where RJ(x,λ)R_{J}(x,\lambda) is some function, and ΦJ(v)=ΦJ(v1,,vm~)\Phi_{J}(v)=\Phi_{J}(v^{1},\dots,v^{\tilde{m}}) is an arbitrary function of m~\tilde{m} variables v1=v1(x,λ),,vm~=vm~(x,λ)v^{1}=v^{1}(x,\lambda),\dots,v^{\tilde{m}}=v^{\tilde{m}}(x,\lambda) that are characteristics of the system (35).

As we have already noted, the space of the solutions of Eq. (35) is invariant under H^\hat{H}. Substituting (36) in the Klein–Gordon equation (1) and multiplying the result by the factor eRJ(x,λ)e^{-R_{J}(x,\lambda)}, we arrive at the differential equation for the unknown function ΦJ(v)\Phi_{J}(v):

H~^(v,v,J)ΦJ(v)=0,\hat{\tilde{H}}(v,\partial_{v},J)\Phi_{J}(v)=0, (37)

where H~^(v,v,J)\hat{\tilde{H}}(v,\partial_{v},J) is a second-order linear differential operator. The differential equation (37) is called the reduced equation. It is clear that the reduced equation involves fewer independent variables than the original Klein–Gordon equation (1). Thus, we have reduced Eq. (1) for φ\varphi as a function of (q,u)(q,u) to the ordinary differential equation (37) for ΦJ\Phi_{J} as a function of vv.

Let us now calculate the number m~\tilde{m} of independent variables in the reduced equation, which is equal to the number of functionally independent characteristics of the system of equations (35). Due to the algebraic independence of the generators XA=XAa(x)xaX_{A}=X_{A}^{a}(x)\partial_{x^{a}}, this number is obviously equal to

m~=m+12(dim𝒢^ind𝒢^)(dim𝒢^1)=m12(dim𝒢^+ind𝒢^)+1.\tilde{m}=m+\frac{1}{2}\left(\dim\hat{\mathscr{G}}-\operatorname{\mathrm{ind}}\hat{\mathscr{G}}\right)-\left(\dim\hat{\mathscr{G}}-1\right)=m-\frac{1}{2}\left(\dim\hat{\mathscr{G}}+\operatorname{\mathrm{ind}}\hat{\mathscr{G}}\right)+1.

If m~1\tilde{m}\leq 1, then the reduced equation (37) is an algebraic or ordinary differential equation. Thus, if the symmetry algebra 𝒢^\hat{\mathscr{G}} satisfies the condition

dim𝒢^+ind𝒢^2m,\dim\hat{\mathscr{G}}+\operatorname{\mathrm{ind}}\hat{\mathscr{G}}\geq 2m, (38)

then the Klein–Gordon equation (1) is integrable in the sense of Definition 1.

If the number of independent variables in the reduced equation is equal to m~=1\tilde{m}=1, then, substituting a solution of Eq. (37) into (36), we obtain a family of solutions of Eq. (1) depending on s=dim𝒢^ms=\dim\hat{\mathscr{G}}-m parameters λ=(λ1,,λs)\lambda=(\lambda^{1},\dots,\lambda^{s}) and l=2m1dim𝒢^l=2m-1-\dim\hat{\mathscr{G}} parameters J=(J1,,Jl)J=(J_{1},\dots,J_{l}). This family can be chosen as a basis of the solution space of the Klein–Gordon equation.

If m~=0\tilde{m}=0, the reduced equation (37) is purely algebraic and can be considered as the restriction imposed on the parameters J=(J1,,Jl)J=(J_{1},\dots,J_{l}). In this case, the number ss of variables λ\lambda is s=dim𝒢^m1s=\dim\hat{\mathscr{G}}-m-1, whereas the number of functionally independent parameters JJ is equal to l1=2mdim𝒢^l-1=2m-\dim\hat{\mathscr{G}}. As a result, the set of the functions φJ(x,λ)=CeRJ(x,λ)\varphi_{J}(x,\lambda)=C\cdot e^{R_{J}(x,\lambda)}, where CC is a constant, forms a basis of the solution space of the Klein–Gordon equation.

Based on the above theory, let us select all the integrable cases for the Klein–Gordon equation (1) on 3D de Sitter space. In order to do this, we must verify the condition (38) for each of the symmetry algebras 𝒢^n,k\hat{\mathscr{G}}_{n,k} given in Section 3. Since m=dimdS3=3m=\dim\mathrm{dS}_{3}=3, this condition takes the form

dim𝒢^+ind𝒢^6.\dim\hat{\mathscr{G}}+\operatorname{\mathrm{ind}}\hat{\mathscr{G}}\geq 6. (39)

The results of this verification are summarized in Table 3, which also contains the dimensions and the indices of algebras 𝒢^n,m\hat{\mathscr{G}}_{n,m} and the numbers ss, ll, and m~\tilde{m} that are the numbers of variables λ\lambda, JJ, and vv, respectively. Table 3 shows that the Klein–Gordon equation on dS3\mathrm{dS}_{3} is integrable only for electromagnetic fields invariant under three- and four-dimensional subalgebras of the Lie algebra 𝔰𝔬(1,3)\mathfrak{so}(1,3).

Table 3: Verification of the condition (39) for symmetry algebras 𝒢^n,m\hat{\mathscr{G}}_{n,m}.
Symmetry algebra 𝒢^\hat{\mathscr{G}} dim𝒢^\dim\hat{\mathscr{G}} ind𝒢^\operatorname{\mathrm{ind}}\hat{\mathscr{G}} ss ll m~\tilde{m} Is condition (39) true?
𝒢^1,1\hat{\mathscr{G}}_{1,1}, 𝒢^1,2\hat{\mathscr{G}}_{1,2}, 𝒢^1,3a\hat{\mathscr{G}}_{1,3}^{a}, 𝒢^1,4\hat{\mathscr{G}}_{1,4} 2 2 0 1 2 No
𝒢^2,1\hat{\mathscr{G}}_{2,1}, 𝒢^2,2\hat{\mathscr{G}}_{2,2}, 𝒢^2,3\hat{\mathscr{G}}_{2,3} 3 1 1 0 2 No
𝒢^3,1\hat{\mathscr{G}}_{3,1}, 𝒢^3,2\hat{\mathscr{G}}_{3,2}, 𝒢^3,3a\hat{\mathscr{G}}_{3,3}^{a}, 𝒢^3,4\hat{\mathscr{G}}_{3,4}, 𝒢^3,5\hat{\mathscr{G}}_{3,5} 4 2 1 1 1 Yes
𝒢^4,1\hat{\mathscr{G}}_{4,1} 5 3 1 3 0 Yes

Now, we demonstrate how the Klein–Gordon equation on 3D de Sitter space can be reduced to an ordinary differential equation in the integrable cases. In order to do this, we use the non-commutative integration method described above. Moreover, we construct solutions of Eq. (1) in terms of special functions where possible. The case of the subalgebra 𝒢4,1\mathscr{G}_{4,1} corresponds to the absence of an electromagnetic field and is not considered here as trivial.

Since we aim to demonstrate the possibility of integrating the Klein–Gordon equation, here, we restrict ourselves only to the local construction of its exact solutions. The global aspect related to investigating the behavior of the solutions on the whole space dS3\mathrm{dS}_{3} and the choice of appropriate function space will not be discussed in this study.

4.1 Case 𝒢3,1\mathscr{G}_{3,1}

In the local coordinates q=(q1,q2,q3)3q=(q^{1},q^{2},q^{3})\in\mathbb{R}^{3}, determined by (A.9), (A.10), the metric of dS3\mathrm{dS}_{3} has the form

ds2=exp(2q3)[(dq1)2(dq2)2]+(dq3)2.ds^{2}=-\exp(2q^{3})\left[(dq^{1})^{2}-(dq^{2})^{2}\right]+(dq^{3})^{2}. (40)

We note that this coordinate chart covers only the “half” of the hyperboloid (14) corresponding the condition x3>0x^{3}>0. In the local coordinates qaq^{a}, the infinitesimal generators of the subalgebra 𝒢3,1\mathscr{G}_{3,1} have the form

X1=q1,X2=q2,X3=q1q1q2q2+q3,X_{1}=\frac{\partial}{\partial q^{1}},\quad X_{2}=\frac{\partial}{\partial q^{2}},\quad X_{3}=-q^{1}\frac{\partial}{\partial q^{1}}-q^{2}\frac{\partial}{\partial q^{2}}+\frac{\partial}{\partial q^{3}},

and the closed 2-form \mathscr{F} invariant under these vector fields is written as (see Table 2):

=exp(q3)(μ1dq1+μ2dq2)dq3.\mathscr{F}=\exp(q^{3})(\mu_{1}dq^{1}+\mu_{2}dq^{2})\wedge dq^{3}. (41)

Here, μ1\mu_{1} and μ2\mu_{2} are arbitrary constants.

We choose the electromagnetic potential corresponding to (41) in the form

𝒜=exp(q3)(μ1dq1+μ2dq2).\mathscr{A}=\exp(q^{3})\left(\mu_{1}dq^{1}+\mu_{2}dq^{2}\right). (42)

The Klein–Gordon equation (1) for the metric (40) and the electromagnetic potential (42) has the following form

H^φ=[exp(2q3)(2(q1)22(q2)2)+2(q3)2+2q32ieexp(q3)(μ1q1+μ2q2)q33ieexp(q3)(μ1q1+μ2q2)e2exp(2q3)(μ1q1+μ2q2)2+6ζ+m2]φ=0.\hat{H}\varphi=\left[-\exp\left(-2q^{3}\right)\left(\frac{\partial^{2}}{\partial(q^{1})^{2}}-\frac{\partial^{2}}{\partial(q^{2})^{2}}\right)+\frac{\partial^{2}}{\partial(q^{3})^{2}}+2\,\frac{\partial}{\partial q^{3}}-\right.\\ \left.-2ie\exp(q^{3})\left(\mu_{1}q^{1}+\mu_{2}q^{2}\right)\frac{\partial}{\partial q^{3}}-3ie\exp(q^{3})(\mu_{1}q^{1}+\mu_{2}q^{2})-\right.\\ \left.-e^{2}\exp(2q^{3})\left(\mu_{1}q^{1}+\mu_{2}q^{2}\right)^{2}+6\zeta+m^{2}\right]\varphi=0. (43)

The symmetry operators for Eq. (43) are given by the formulas (17). Taking into account the electromagnetic potential (42), these operators are written in the explicit form as

X^1=q1ieμ1exp(q3),X^2=q2ieμ2exp(q3),X^3=q1q1q2q2+q3.\hat{X}_{1}=\frac{\partial}{\partial q^{1}}-ie\mu_{1}\exp(q^{3}),\quad\hat{X}_{2}=\frac{\partial}{\partial q^{2}}-ie\mu_{2}\exp(q^{3}),\quad\hat{X}_{3}=-q^{1}\frac{\partial}{\partial q^{1}}-q^{2}\frac{\partial}{\partial q^{2}}+\frac{\partial}{\partial q^{3}}.

Together with the trivial operator X^0=ie\hat{X}_{0}=ie, these operators define the symmetry algebra 𝒢^3,1\hat{\mathscr{G}}_{3,1}, and it follows from the commutation relations (18) that this algebra is isomorphic to the direct sum 𝒢3,1\mathscr{G}_{3,1}\oplus\mathbb{R} (a trivial central extension). Using (34), it is easy to verify that ind𝒢^3,1=2\operatorname{\mathrm{ind}}\hat{\mathscr{G}}_{3,1}=2.

The λ\lambda-representation of the algebra 𝒢^3,1\hat{\mathscr{G}}_{3,1} acts in the space of functions of one real variable λ\lambda and depend on one real parameter JJ:

^0=ie,^1=iJλ,^2=iλ,^3=λλ+12.\hat{\ell}_{0}=-ie,\quad\hat{\ell}_{1}=iJ\lambda,\quad\hat{\ell}_{2}=i\lambda,\quad\hat{\ell}_{3}=\lambda\,\frac{\partial}{\partial\lambda}+\frac{1}{2}.

It is easy to see that these operators are skew-symmetric under the usual Lebesgue measure dλd\lambda on \mathbb{R}. Solving the system of equations (35), we obtain

φJ(q1,q2,q3,λ)=ΦJ(v)exp[iλ(Jq1+q2)12q3+ieexp(q3)(μ1q1+μ2q2)],\varphi_{J}\left(q^{1},q^{2},q^{3},\lambda\right)=\Phi_{J}\left(v\right)\exp\left[-i\lambda(Jq^{1}+q^{2})-\frac{1}{2}\,q^{3}+ie\exp(q^{3})\left(\mu_{1}q^{1}+\mu_{2}q^{2}\right)\right], (44)

where v=λexp(q3)v=\lambda\exp(-q^{3}). Substituting the obtained solution into (43), after some algebra, we obtain an ordinary differential equation for an unknown function ΦJ(v)\Phi_{J}(v):

v2ΦJ′′(v)+[(J2+1)v22e(Jμ1+μ2)v+m2+6ζ+e2(μ12+μ22)34]ΦJ(v)=0.v^{2}\Phi_{J}^{\prime\prime}(v)+\left[(J^{2}+1)v^{2}-2e(J\mu_{1}+\mu_{2})v+m^{2}+6\zeta+e^{2}(\mu_{1}^{2}+\mu_{2}^{2})-\frac{3}{4}\right]\Phi_{J}(v)=0. (45)

Thus, the family of the functions (44), parameterized by the parameters JJ and λ\lambda, forms a basis of the solution space for the Klein-Gordon equation (43) if the function ΦJ(v)\Phi_{J}(v) is a solution of the reduced equation (45).

If instead of the variable vv, we introduce the new complex variable z=2iJ2+1vz=2i\sqrt{J^{2}+1}\,v, Eq. (45) is reduced to Whittaker’s equation

Φ~J′′(z)+[14+αz+14β2z2]Φ~J(z)=0,\tilde{\Phi}_{J}^{\prime\prime}(z)+\left[-\frac{1}{4}+\frac{\alpha}{z}+\frac{\frac{1}{4}-\beta^{2}}{z^{2}}\right]\tilde{\Phi}_{J}(z)=0,

where Φ~J(z)=ΦJ(2iJ2+1v)\tilde{\Phi}_{J}(z)=\Phi_{J}(2i\sqrt{J^{2}+1}\,v), α=ie(Jμ1+μ2)/J2+1\alpha=-ie\left(J\mu_{1}+\mu_{2}\right)/\sqrt{J^{2}+1}, β2=1m26ζe(μ12+μ22)\beta^{2}=1-m^{2}-6\zeta-e\left(\mu_{1}^{2}+\mu_{2}^{2}\right). This equation has the regular singular point z=0z=0 and the irregular singular point z=z=\infty. Its two linearly independent solutions Mα,β(z)M_{\alpha,\beta}(z) and Wα,β(z)W_{\alpha,\beta}(z), called the Whittaker functions, can be expressed in terms of confluent hypergeometric functions [32]:

Mα,β(z)=exp(z2)z12+βM(12+βα,1+2β,z),M_{\alpha,\beta}(z)=\exp\left(-\frac{z}{2}\right)z^{\frac{1}{2}+\beta}\,M\left(\frac{1}{2}+\beta-\alpha,1+2\beta,z\right),
Wα,β(z)=exp(z2)z12+βU(12+βα,1+2β,z).W_{\alpha,\beta}(z)=\exp\left(-\frac{z}{2}\right)z^{\frac{1}{2}+\beta}U\left(\frac{1}{2}+\beta-\alpha,1+2\beta,z\right).

4.2 Case 𝒢3,2\mathscr{G}_{3,2}

Using the local coordinates (q,u)=(q1,q2,u1)3(q,u)=(q^{1},q^{2},u^{1})\in\mathbb{R}^{3} defined by (A.11), (A.12), for the metric of dS3\mathrm{dS}_{3} we have

ds2=exp(2u1)[(dq1)2(dq2)2]+(du1)2.ds^{2}=-\exp(-2u^{1})\left[(dq^{1})^{2}-(dq^{2})^{2}\right]+(du^{1})^{2}. (46)

As in the previous case, this coordinate chart covers only the “half” of the space dS3\mathrm{dS}_{3} corresponding to the positive values of the coordinate x3x^{3} in 1,3\mathbb{R}^{1,3}. In the coordinates (q1,q2,u1)(q^{1},q^{2},u^{1}), the basic vector fields XA𝒢3,2X_{A}\in\mathscr{G}_{3,2} and the corresponding electromagnetic field \mathscr{F} have the form (see Table 2):

X1=q1,X2=q2,X3=q2q1+q1q2,X_{1}=\frac{\partial}{\partial q^{1}},\quad X_{2}=\frac{\partial}{\partial q^{2}},\quad X_{3}=-q^{2}\frac{\partial}{\partial q^{1}}+q^{1}\frac{\partial}{\partial q^{2}},
=μdq1dq2.\mathscr{F}=\mu\,dq^{1}\wedge dq^{2}.

Here, μ\mu is a constant.

Choosing the electromagnetic field potential in the form

𝒜=12μ(q1dq2q2dq1),\mathscr{A}=\frac{1}{2}\,\mu\left(q^{1}dq^{2}-q^{2}dq^{1}\right), (47)

we obtain that the Klein–Gordon equation for the given electromagnetic field and the metric (46) is

H^φ=[exp(2u1)(2(q1)22(q2)2)+2(u1)22u1ieμexp(2u1)(q2q1q1q2)++14e2μ2exp(2u1)((q1)2+(q2)2)+6ζ+m2]φ=0.\hat{H}\varphi=\left[-\exp(2u^{1})\left(\frac{\partial^{2}}{\partial(q^{1})^{2}}-\frac{\partial^{2}}{\partial(q^{2})^{2}}\right)+\frac{\partial^{2}}{\partial(u^{1})^{2}}-2\,\frac{\partial}{\partial u^{1}}-ie\mu\exp(2u^{1})\left(q^{2}\frac{\partial}{\partial q^{1}}-q^{1}\frac{\partial}{\partial q^{2}}\right)+\right.\\ \left.+\frac{1}{4}\,e^{2}\mu^{2}\exp(2u^{1})\left((q^{1})^{2}+(q^{2})^{2}\right)+6\zeta+m^{2}\right]\varphi=0. (48)

The first-order symmetries of this equation are given by the operators (19), which, taking into account (47), can be written as

X^1=q112ieμq2,X^2=q2+12ieμq1,X^3=q2q1+q1q2.\hat{X}_{1}=\frac{\partial}{\partial q^{1}}-\frac{1}{2}\,ie\mu q^{2},\quad\hat{X}_{2}=\frac{\partial}{\partial q^{2}}+\frac{1}{2}\,ie\mu q^{1},\quad\hat{X}_{3}=-q^{2}\frac{\partial}{\partial q^{1}}+q^{1}\frac{\partial}{\partial q^{2}}.

Together with the trivial operator X^0=ie\hat{X}_{0}=ie, these operators form the symmetry algebra 𝒢^3,2\hat{\mathscr{G}}_{3,2} isomorphic to a one-dimensional central extension of the algebra 𝒢3,2\mathscr{G}_{3,2}. From the commutation relations (20), it follows that this extension is indecomposable for μ0\mu\neq 0.

The λ\lambda-representation of the algebra 𝒢^3,2\hat{\mathscr{G}}_{3,2} acts in the space of holomorphic functions on the complex plane \mathbb{C}:

^0=ie,^1=iλ12ieμλ,^2=λ12eμλ,^3=iλλiJ.\hat{\ell}_{0}=-ie,\quad\hat{\ell}_{1}=i\,\frac{\partial}{\partial\lambda}-\frac{1}{2}\,ie\mu\lambda,\quad\hat{\ell}_{2}=-\frac{\partial}{\partial\lambda}-\frac{1}{2}\,e\mu\lambda,\quad\hat{\ell}_{3}=i\lambda\,\frac{\partial}{\partial\lambda}-iJ. (49)

Here, λ\lambda\in\mathbb{C}, JJ\in\mathbb{R}. The operators (49) are skew-symmetric under the Gaussian measure μ(λ)=exp(12e|λ|2)\mu(\lambda)=\exp\left(-\frac{1}{2}\,e|\lambda|^{2}\right) on \mathbb{C}.

The general solution of the system of equations (35), in this case, has the form

φJ(q1,q2,u1,λ)=ΦJ(v)[q1+i(q2+λ)]Jexp[12eμλ(iq1+q2)+14eμ((q1)2+(q2)2)],\varphi_{J}\left(q^{1},q^{2},u^{1},\lambda\right)=\Phi_{J}(v)\left[q^{1}+i\left(q^{2}+\lambda\right)\right]^{J}\exp\left[\frac{1}{2}\,e\mu\lambda\left(iq^{1}+q^{2}\right)+\frac{1}{4}\,e\mu\left((q^{1})^{2}+(q^{2})^{2}\right)\right], (50)

where v=u1v=u^{1}. Substituting this function into Eq. (48), after some algebraic calculations, we obtain the following ordinary differential equation for the unknown function ΦJ(v)\Phi_{J}(v):

ΦJ′′(v)2ΦJ(v)+[6ζ+m2eμ(2J+1)exp(2v)]ΦJ(v)=0.\Phi_{J}^{\prime\prime}(v)-2\Phi_{J}^{\prime}(v)+\left[6\zeta+m^{2}-e\mu\left(2J+1\right)\exp(2v)\right]\Phi_{J}(v)=0. (51)

Two linearly independent solutions of this equation can be expressed in terms of Bessel functions of the first and second kind, respectively:

ΦJ(1)(v)=exp(v)Jα(iexp(v)eμ(2J+1)),ΦJ(2)(v)=exp(v)Yα(iexp(v)eμ(2J+1)).\Phi_{J}^{(1)}(v)=\exp(v)J_{\alpha}\left(i\exp(v)\sqrt{e\mu(2J+1)}\right),\quad\Phi_{J}^{(2)}(v)=\exp(v)Y_{\alpha}\left(i\exp(v)\sqrt{e\mu(2J+1)}\right).

Here, α=1m26ζ\alpha=\sqrt{1-m^{2}-6\zeta}.

Thus, the set of the functions (50), where ΦJ(u)\Phi_{J}(u) is a solution of the reduced equation (51), can be regarded as a basis of the solution space for Eq. (48). This basis is parameterized by the numbers λ\lambda and JJ.

4.3 Case 𝒢3,3a\mathscr{G}_{3,3}^{a}

In this case, the rectifying local coordinates q=(q1,q2,q3)3q=(q^{1},q^{2},q^{3})\in\mathbb{R}^{3} on dS3\mathrm{dS}_{3} are given by (A.13), (A.14). In these coordinates, the metric ds2ds^{2}, the generators XA𝒢3,3aX_{A}\in\mathscr{G}_{3,3}^{a}, and the electromagnetic field \mathscr{F} invariant under the subalgebra 𝒢3,3a\mathscr{G}_{3,3^{a}} have the form:

ds2=exp(2aq3)[(dq1)2(dq2)2]+a2(dq3)2;ds^{2}=-\exp(2aq^{3})\left[(dq^{1})^{2}-(dq^{2})^{2}\right]+a^{2}(dq^{3})^{2};
X1=q1,X2=q2,X3=(aq1+q2)q1+(q1aq2)q2+q3;X_{1}=\frac{\partial}{\partial q^{1}},\quad X_{2}=\frac{\partial}{\partial q^{2}},\quad X_{3}=-\left(aq^{1}+q^{2}\right)\frac{\partial}{\partial q^{1}}+\left(q^{1}-aq^{2}\right)\frac{\partial}{\partial q^{2}}+\frac{\partial}{\partial q^{3}};
=exp(aq3)[μ1cos(q3)+μ2sin(q3)]dq1dq3+exp(aq3)[μ1sin(q3)μ2cos(q3)]dq2dq3.\mathscr{F}=\exp(aq^{3})\left[\mu_{1}\cos(q^{3})+\mu_{2}\sin(q^{3})\right]dq^{1}\wedge dq^{3}+\exp(aq^{3})\left[\mu_{1}\sin(q^{3})-\mu_{2}\cos(q^{3})\right]dq^{2}\wedge dq^{3}.

Here, μ1\mu_{1}, μ2\mu_{2}, and a>0a>0 are arbitrary constants.

Let us fix the electromagnetic field potential in the form

𝒜=μ1exp(aq3)[q1cos(q3)+q2sin(q3)]dq3+μ2exp(aq3)[q1sin(q3)q2cos(q3)]dq3.\mathscr{A}=\mu_{1}\exp(aq^{3})\left[q^{1}\cos(q^{3})+q^{2}\sin(q^{3})\right]dq^{3}+\mu_{2}\exp(aq^{3})\left[q^{1}\sin(q^{3})-q^{2}\cos(q^{3})\right]dq^{3}. (52)

Then, the corresponding Klein–Gordon equation can be written as

H^φ={exp(2aq3)(2(q1)22(q2)2)+1a22(q3)2+2aq32iea2exp(aq3)[(μ1q1μ2q2)cos(q3)+(μ2q1+μ1q2)sin(q3)]q3iea2exp(aq3)[(3aq1+q2)μ1+(q13aq2)μ2]cos(q3)iea2exp(aq3)[(3aq2q1)μ1+(3aq1+q2)μ2]sin(q3)e2a2exp(2aq3)[(q1μ1q2μ2)cos(q3)(q1μ2+q2μ1)sin(q3)]+6ζ+m2}φ=0.\hat{H}\varphi=\left\{-\exp(-2aq^{3})\left(\frac{\partial^{2}}{\partial(q^{1})^{2}}-\frac{\partial^{2}}{\partial(q^{2})^{2}}\right)+\frac{1}{a^{2}}\,\frac{\partial^{2}}{\partial(q^{3})^{2}}+\frac{2}{a}\,\frac{\partial}{\partial q^{3}}-\right.\\ \left.-\frac{2ie}{a^{2}}\,\exp(aq^{3})\left[\left(\mu_{1}q^{1}-\mu_{2}q^{2}\right)\cos(q^{3})+\left(\mu_{2}q^{1}+\mu_{1}q^{2}\right)\sin(q^{3})\right]\frac{\partial}{\partial q^{3}}-\right.\\ \left.-\frac{ie}{a^{2}}\,\exp(aq^{3})\left[(3aq^{1}+q^{2})\mu_{1}+(q^{1}-3aq^{2})\mu_{2}\right]\cos(q^{3})-\right.\\ \left.-\frac{ie}{a^{2}}\,\exp(aq^{3})\left[(3aq^{2}-q^{1})\mu_{1}+(3aq^{1}+q^{2})\mu_{2}\right]\sin(q^{3})-\right.\\ \left.-\frac{e^{2}}{a^{2}}\,\exp(2aq^{3})\left[\left(q^{1}\mu_{1}-q^{2}\mu_{2}\right)\cos(q^{3})-\left(q^{1}\mu_{2}+q^{2}\mu_{1}\right)\sin(q^{3})\right]+6\zeta+m^{2}\right\}\varphi=0. (53)

The symmetry algebra 𝒢^3,3a\hat{\mathscr{G}}^{a}_{3,3} of this equation is generated by the trivial operator X^0=ie\hat{X}_{0}=ie and the operators (21)–(23), which in view of Eq. (52) can be written in the form

X^1=q1ieexp(aq3)1+a2[(aμ1μ2)cos(q3)+(μ1+aμ2)sin(q3)],\hat{X}_{1}=\frac{\partial}{\partial q^{1}}-\frac{ie\exp(aq^{3})}{1+a^{2}}\left[\left(a\mu_{1}-\mu_{2}\right)\cos(q^{3})+\left(\mu_{1}+a\mu_{2}\right)\sin(q^{3})\right],
X^2=q2+ieexp(aq3)1+a2[(μ1+aμ2)cos(q3)(aμ1μ2)sin(q3)],\hat{X}_{2}=\frac{\partial}{\partial q^{2}}+\frac{ie\exp(aq^{3})}{1+a^{2}}\left[\left(\mu_{1}+a\mu_{2}\right)\cos(q^{3})-\left(a\mu_{1}-\mu_{2}\right)\sin(q^{3})\right],
X^3=(aq1+q2)q1+(q1aq2)q2+q3.\hat{X}_{3}=-(aq^{1}+q^{2})\frac{\partial}{\partial q^{1}}+(q^{1}-aq^{2})\frac{\partial}{\partial q^{2}}+\frac{\partial}{\partial q^{3}}.

As follows from the commutation relations (24), 𝒢^3,3a\hat{\mathscr{G}}^{a}_{3,3} is isomorphic to the direct sum of the algebra 𝒢3,3\mathscr{G}_{3,3} and the one-dimensional center X^0\langle\hat{X}_{0}\rangle\simeq\mathbb{R}.

The λ\lambda-representation of the Lie algebra 𝒢^3,3a\hat{\mathscr{G}}^{a}_{3,3} acts in the space of functions of one real variable λ\lambda:

^1=iJexp(aλ)cosλ,^2=iJexp(aλ)sinλ,^3=λ.\hat{\ell}_{1}=iJ\exp(a\lambda)\cos\lambda,\quad\hat{\ell}_{2}=iJ\exp(a\lambda)\sin\lambda,\quad\hat{\ell}_{3}=\frac{\partial}{\partial\lambda}.

The general solution of the system of equations (35) is given by

φJ(q1,q2,q3,λ)=ΦJ(v)exp[ieexp(aq3)(aμ1μ21+a2q1μ1+aμ21+a2q2)cos(q3)++ieexp(aq3)(μ1+aμ21+a2q1+aμ1μ21+a2q2)sin(q3)iJexp(aλ)(q1cos(λ)+q2sin(λ))],\varphi_{J}(q^{1},q^{2},q^{3},\lambda)=\Phi_{J}(v)\exp\left[ie\exp(aq^{3})\left(\frac{a\mu_{1}-\mu_{2}}{1+a^{2}}\,q^{1}-\frac{\mu_{1}+a\mu_{2}}{1+a^{2}}\,q^{2}\right)\cos(q^{3})+\right.\\ \left.+ie\exp(aq^{3})\left(\frac{\mu_{1}+a\mu_{2}}{1+a^{2}}\,q^{1}+\frac{a\mu_{1}-\mu_{2}}{1+a^{2}}\,q^{2}\right)\sin(q^{3})-iJ\exp(a\lambda)\left(q^{1}\cos(\lambda)+q^{2}\sin(\lambda)\right)\right],

where v=q3λv=q^{3}-\lambda. Substituting it into the Klein-Gordon equation (53), we obtain the reduced equation for the unknown function ΦJ(v)\Phi_{J}(v):

ΦJ′′(v)+2aΦJ(v)+[2ea2Jexp(av)1+a2(aμ1μ2)cos(v)2ea2Jexp(av)1+a2(μ1+aμ2)sin(v)++a2J2exp(2av)+a2(m2+6ζ)+e2a2(μ12+μ22)1+a2]ΦJ(v)=0.\Phi_{J}^{\prime\prime}(v)+2a\Phi_{J}^{\prime}(v)+\left[-\frac{2ea^{2}J\exp(-av)}{1+a^{2}}\left(a\mu_{1}-\mu_{2}\right)\cos(v)-\frac{2ea^{2}J\exp(-av)}{1+a^{2}}\left(\mu_{1}+a\mu_{2}\right)\sin(v)+\right.\\ \left.+a^{2}J^{2}\exp(-2av)+a^{2}(m^{2}+6\zeta)+\frac{e^{2}a^{2}(\mu_{1}^{2}+\mu_{2}^{2})}{1+a^{2}}\right]\Phi_{J}(v)=0. (54)

In general, the solutions of this ordinary differential equation are apparently not expressed in terms of known special functions. For specific parameter values, (54) can be studied numerically or using asymptotic methods.

4.4 Case 𝒢3,4\mathscr{G}_{3,4}

In the local coordinates (q1,q2,u1)(q^{1},q^{2},u^{1}) that are given by (A.15), the metric of dS3\mathrm{dS}_{3} is written as

ds2=cosh2(u1)cos2(q2)d(q1)2cosh2(u1)d(q2)2+d(u1)2.ds^{2}=-\cosh^{2}(u^{1})\cos^{2}(q^{2})d(q^{1})^{2}-\cosh^{2}(u^{1})d(q^{2})^{2}+d(u^{1})^{2}.

We note that these coordinates cover de Sitter space almost everywhere. In particular, q1,q2q^{1},q^{2} are spherical coordinates on the spacelike surfaces x0=constx^{0}=\mathrm{const}.

In the coordinates (q1,q2,u1)(q^{1},q^{2},u^{1}), the infinitesimal generators XA𝒢3,4X_{A}\in\mathscr{G}_{3,4} and the invariant closed 2-form \mathscr{F} have the form (see Table 2):

X1=q1,X2=sin(q1)tan(q2)q1+cos(q1)q2,X3=cos(q1)tan(q2)q1sin(q1)q2;X_{1}=\frac{\partial}{\partial q^{1}},\quad X_{2}=\sin(q^{1})\tan(q^{2})\,\frac{\partial}{\partial q^{1}}+\cos(q^{1})\,\frac{\partial}{\partial q^{2}},\quad X_{3}=\cos(q^{1})\tan(q^{2})\,\frac{\partial}{\partial q^{1}}-\sin(q^{1})\,\frac{\partial}{\partial q^{2}};
=μcos(q2)dq1dq2,\mathscr{F}=\mu\cos(q^{2})dq^{1}\wedge dq^{2}, (55)

where μ\mu is a constant. If we choose the electromagnetic field potential in the form

𝒜=μsin(q2)dq1,\mathscr{A}=-\mu\sin(q^{2})dq^{1}, (56)

then the Klein-Gordon equation for this case is written as follows:

H^φ={2(u1)2+2tanh(u1)u11cosh2(u1)(1cos2(q2)2(q1)2+2(q2)2tan(q2)q2)2ieμtan(q2)cosh2(u1)cos(q2)q1+e2μ2tan2(q2)cosh2(u1)+6ζ+m2}φ=0.\hat{H}\varphi=\left\{\frac{\partial^{2}}{\partial(u^{1})^{2}}+2\tanh(u^{1})\,\frac{\partial}{\partial u^{1}}-\frac{1}{\cosh^{2}(u^{1})}\left(\frac{1}{\cos^{2}(q^{2})}\,\frac{\partial^{2}}{\partial(q^{1})^{2}}+\frac{\partial^{2}}{\partial(q^{2})^{2}}-\tan(q^{2})\,\frac{\partial}{\partial q^{2}}\right)-\right.\\ \left.-\frac{2ie\mu\tan(q^{2})}{\cosh^{2}(u^{1})\cos(q^{2})}\,\frac{\partial}{\partial q^{1}}+\frac{e^{2}\mu^{2}\tan^{2}(q^{2})}{\cosh^{2}(u^{1})}+6\zeta+m^{2}\right\}\varphi=0. (57)

The first-order symmetry algebra of this equation is generated by the operators (25)–(27), which in view of (56) are explicitly given by

X^1=q1,\hat{X}_{1}=\frac{\partial}{\partial q^{1}},\quad
X^2=sin(q1)tan(q2)q1+cos(q1)q2+ieμsin(q1)cos(q2),\hat{X}_{2}=\sin(q^{1})\tan(q^{2})\frac{\partial}{\partial q^{1}}+\cos(q^{1})\frac{\partial}{\partial q^{2}}+\frac{ie\mu\sin(q^{1})}{\cos(q^{2})},
X^3=cos(q1)tan(q2)q1sin(q1)q2+ieμcos(q1)cos(q2).\hat{X}_{3}=\cos(q^{1})\tan(q^{2})\frac{\partial}{\partial q^{1}}-\sin(q^{1})\frac{\partial}{\partial q^{2}}+\frac{ie\mu\cos(q^{1})}{\cos(q^{2})}.

Together with the trivial operator X^0=ie\hat{X}_{0}=ie, these operators form the Lie algebra 𝒢^3,4\hat{\mathscr{G}}_{3,4} isomorphic to the direct sum 𝒢3,4\mathscr{G}_{3,4}\oplus\mathbb{R} (it can be easily seen from (28) and, in fact, follows from the semisimplicity of the Lie algebra 𝒢3,4𝔰𝔬(3)\mathscr{G}_{3,4}\simeq\mathfrak{so}(3)).

We realize the λ\lambda-representation of the Lie algebra 𝒢3,4\mathscr{G}_{3,4} by the following family of operators

^1=iλλ+iJ,^2=i2(1λ2)λ+iJλ,^3=12(1+λ2)λ+Jλ.\hat{\ell}_{1}=-i\lambda\,\frac{\partial}{\partial\lambda}+iJ,\quad\hat{\ell}_{2}=\frac{i}{2}\left(1-\lambda^{2}\right)\frac{\partial}{\partial\lambda}+iJ\lambda,\quad\hat{\ell}_{3}=-\frac{1}{2}\left(1+\lambda^{2}\right)\frac{\partial}{\partial\lambda}+J\lambda. (58)

Here λ\lambda\in\mathbb{C}, J(0,+)J\in(0,+\infty). The measure under which these operators are skew-symmetric is determined by the equality dμ(λ)=(1+|λ|2)2(J+1)dλdλ¯d\mu(\lambda)=(1+|\lambda|^{2})^{2(J+1)}d\lambda\wedge d\bar{\lambda}. Solving the system of equations (35), we find

φJ(q1,q2,u1,λ)=ΦJ(v)[(λ2exp(iq1)+exp(iq1))cos(q2)2iλsin(q2)]J××[(iλexp(iq1)cos(q2)+sin(q2)+1)(sin(q2)1)(iλexp(iq1)cos(q2)+sin(q2)1)cos(q2)]eμ,\varphi_{J}(q^{1},q^{2},u^{1},\lambda)=\Phi_{J}(v)\left[(\lambda^{2}\exp(iq^{1})+\exp(-iq^{1}))\cos(q^{2})-2i\lambda\sin(q^{2})\right]^{J}\times\\ \times\left[\frac{(i\lambda\exp(iq^{1})\cos(q^{2})+\sin(q^{2})+1)(\sin(q^{2})-1)}{(i\lambda\exp(iq^{1})\cos(q^{2})+\sin(q^{2})-1)\cos(q^{2})}\right]^{e\mu}, (59)

where v=u1v=u^{1}. Substituting this function into the Klein–Gordon equation (57), we obtain the ordinary differential equation for the unknown function ΦJ(u)\Phi_{J}(u):

ΦJ′′(v)+2tanh(v)ΦJ(v)+(m2+6ζ+J(J+1)e2μ2cosh2(v))ΦJ(v)=0.\Phi_{J}^{\prime\prime}(v)+2\tanh(v)\Phi^{\prime}_{J}(v)+\left(m^{2}+6\zeta+\frac{J(J+1)-e^{2}\mu^{2}}{\cosh^{2}(v)}\right)\Phi_{J}(v)=0. (60)

Two linearly independent solutions of this equation are expressed in terms of associated Legendre functions of the first and second kind:

ΦJ(1)(v)=1cosh(v)Pνσ(tanh(v)),ΦJ(2)(v)=1cosh(v)Qνσ(tanh(v)),\Phi_{J}^{(1)}(v)=\frac{1}{\sqrt{\cosh(v)}}\,P_{\nu}^{\sigma}(\tanh(v)),\quad\Phi_{J}^{(2)}(v)=\frac{1}{\sqrt{\cosh(v)}}\,Q_{\nu}^{\sigma}(\tanh(v)),

where the parameters ν\nu and σ\sigma are defined as

ν=(J+12)2e2μ212,σ=1m26ζ.\nu=\sqrt{\left(J+\frac{1}{2}\right)^{2}-e^{2}\mu^{2}}-\frac{1}{2},\quad\sigma=\sqrt{1-m^{2}-6\zeta}.

Thus, the family of functions (59) parameterized by the numbers JJ and λ\lambda forms a basis of the solution space of Eq. (57) if the function ΦJ(v)\Phi_{J}(v) is a solution to the ordinary differential equation (60).

4.5 Case 𝒢3,5\mathscr{G}_{3,5}

In the local coordinates (A.17) and (A.18), the metric of dS3\mathrm{dS}_{3} is written as

ds2=sin2(u1)cos2(q2)d(q1)2+sin2(u1)d(q2)2+d(u1)2.ds^{2}=-\sin^{2}(u^{1})\cos^{2}(q^{2})d(q^{1})^{2}+\sin^{2}(u^{1})d(q^{2})^{2}+d(u^{1})^{2}.

The basic vector fields XA𝒢3,5X_{A}\in\mathscr{G}_{3,5} have the form (see Table 2)

X1=q1,X2=sinh(q1)tan(q2)q1+cosh(q1)q2,X3=cosh(q1)tan(q2)q1+sinh(q1)q2,X_{1}=\frac{\partial}{\partial q^{1}},\quad X_{2}=\sinh(q^{1})\tan(q^{2})\frac{\partial}{\partial q^{1}}+\cosh(q^{1})\frac{\partial}{\partial q^{2}},\quad X_{3}=\cosh(q^{1})\tan(q^{2})\frac{\partial}{\partial q^{1}}+\sinh(q^{1})\frac{\partial}{\partial q^{2}},

and the corresponding closed invariant 2-form \mathscr{F} is given by the expression (55).

Choosing the electromagnetic field potential in the form (56), we obtain the following Klein-Gordon equation:

H^φ=[2(u1)2+2cot(u1)u11sin2(u1)(1cos2(q2)2(q1)22(q2)2+tan(q2)q2)2ieμtan(q2)sin2(u1)cos(q2)q1+e2μ2tan2(q2)sin2(u1)+6ζ+m2]φ=0.\hat{H}\varphi=\left[\frac{\partial^{2}}{\partial(u^{1})^{2}}+2\cot(u^{1})\,\frac{\partial}{\partial u^{1}}-\frac{1}{\sin^{2}(u^{1})}\left(\frac{1}{\cos^{2}(q^{2})}\,\frac{\partial^{2}}{\partial(q^{1})^{2}}-\frac{\partial^{2}}{\partial(q^{2})^{2}}+\tan(q^{2})\,\frac{\partial}{\partial q^{2}}\right)-\right.\\ \left.-\frac{2ie\mu\tan(q^{2})}{\sin^{2}(u^{1})\cos(q^{2})}\,\frac{\partial}{\partial q^{1}}+\frac{e^{2}\mu^{2}\tan^{2}(q^{2})}{\sin^{2}(u^{1})}+6\zeta+m^{2}\right]\varphi=0. (61)

The symmetry operators (29)–(31), for the selected electromagnetic potential having the form

X^1=q1,X^2=sinh(q1)tan(q2)q1+cosh(q1)q2+ieμsinh(q1)cos(q2),\hat{X}_{1}=\frac{\partial}{\partial q^{1}},\quad\hat{X}_{2}=\sinh(q^{1})\tan(q^{2})\,\frac{\partial}{\partial q^{1}}+\cosh(q^{1})\frac{\partial}{\partial q^{2}}+\frac{ie\mu\sinh(q^{1})}{\cos(q^{2})},
X^3=cosh(q1)tan(q2)q1+sinh(q1)q2+ieμcosh(q1)cos(q2),\quad\hat{X}_{3}=\cosh(q^{1})\tan(q^{2})\,\frac{\partial}{\partial q^{1}}+\sinh(q^{1})\,\frac{\partial}{\partial q^{2}}+\frac{ie\mu\cosh(q^{1})}{\cos(q^{2})},

together with the trivial symmetry operator X^0=ie\hat{X}_{0}=ie form the Lie algebra 𝒢^3,5\hat{\mathscr{G}}_{3,5} isomorphic to the direct sum 𝒢3,5\mathscr{G}_{3,5}\oplus\mathbb{R}.

It was shown in [31] that the Lie algebra 𝔰𝔬(1,2)\mathfrak{so}(1,2) has the two different λ\lambda-representations corresponding to discrete and continuous series of representations of the group SO(1,2)SO(1,2). As an example, we choose the representation corresponding to a continuous series (the second case is analyzed similarly and simply leads to a different basis for solutions to the Klein - Gordon equation):

^1=λλ+iJ+12,^2=12(λ2+1)λ+(iJ+12)λ,^3=12(λ21)λ+(iJ+12)λ.\hat{\ell}_{1}=\lambda\,\frac{\partial}{\partial\lambda}+iJ+\frac{1}{2},\quad\hat{\ell}_{2}=\frac{1}{2}\left(\lambda^{2}+1\right)\frac{\partial}{\partial\lambda}+\left(iJ+\frac{1}{2}\right)\lambda,\quad\hat{\ell}_{3}=\frac{1}{2}\left(\lambda^{2}-1\right)\frac{\partial}{\partial\lambda}+\left(iJ+\frac{1}{2}\right)\lambda.

Here, λ\lambda\in\mathbb{R}, J[0,+)J\in[0,+\infty). These operators are skew-symmetric under the Lebesgue measure dλd\lambda on \mathbb{R}.

It is easy to obtain the general solution of the system of equations (35) in this case:

φJ(q1,q2,u1,λ)=ΦJ(v)[2λsin(q2)+(exp(q1)λ2exp(q1))cos(q2)]iJ12××[λexp(q1)cos(q2)+1sin(q2)cos(q2)λexp(q1)(1sin(q2))]ieμ.\varphi_{J}(q^{1},q^{2},u^{1},\lambda)=\Phi_{J}(v)\left[2\lambda\sin(q^{2})+\left(\exp(q^{1})-\lambda^{2}\exp(-q^{1})\right)\cos(q^{2})\right]^{-iJ-\frac{1}{2}}\times\\ \times\left[\frac{\lambda\exp(-q^{1})\cos(q^{2})+1-\sin(q^{2})}{\cos(q^{2})-\lambda\exp(-q^{1})\left(1-\sin(q^{2})\right)}\right]^{ie\mu}.

Substituting it into (61), after a series of algebraic transformations, we obtain the ordinary differential equation for the unknown function ΦJ(v)\Phi_{J}(v):

ΦJ′′(v)+2cot(v)ΦJ(v)+[m2+6ζ+J2e2μ2+14sin2(v)]ΦJ(v)=0.\Phi^{\prime\prime}_{J}(v)+2\cot(v)\Phi^{\prime}_{J}(v)+\left[m^{2}+6\zeta+\frac{J^{2}-e^{2}\mu^{2}+\frac{1}{4}}{\sin^{2}(v)}\right]\Phi_{J}(v)=0.

The two linearly independent solutions of this equation are expressed in terms of associated Legendre functions of the first and second kind:

ΦJ(1)(v)=1sin(v)Pνσ(cos(v)),ΦJ(2)(v)=1sin(v)Qνσ(cos(v)),\Phi_{J}^{(1)}(v)=\frac{1}{\sqrt{\sin(v)}}\,P_{\nu}^{\sigma}(\cos(v)),\quad\Phi_{J}^{(2)}(v)=\frac{1}{\sqrt{\sin(v)}}\,Q_{\nu}^{\sigma}(\cos(v)),

where the parameters ν\nu and σ\sigma are defined as

ν=1+m2+6ζ12,σ=e2μ2J2.\nu=\sqrt{1+m^{2}+6\zeta}-\frac{1}{2},\quad\sigma=\sqrt{e^{2}\mu^{2}-J^{2}}.

Conclusion

Employing the defining equations for the first-order symmetry operators of the Klein–Gordon equation in an external electromagnetic field [21, 22], we have given an algorithm for constructing these operators and described the structure of the corresponding Lie algebra 𝒢^\hat{\mathscr{G}} in terms of Lie algebra extensions. In particular, we have shown that 𝒢^\hat{\mathscr{G}} is a one-dimensional central extension of the subalgebra 𝒢\mathscr{G} of the Killing vector fields that preserve the electromagnetic field tensor. These results allowed us to suggest a symmetry-based approach to the classification of electromagnetic fields that admit nontrivial first-order symmetry algebras of the Klein–Gordon equation on 3D de Sitter background. Within the approach, we have obtained the list of all such electromagnetic fields on dS3\mathrm{dS}_{3} (see Table 2) based on the well-known classification of inequivalent subalgebras of the Lie algebra 𝔰𝔬(1,3)\mathfrak{so}(1,3). Also, we have explicitly constructed the corresponding first-order symmetry algebras 𝒢^\hat{\mathscr{G}}. We emphasize that all computations were carried out in local coordinates in which the basis vector fields of subalgebras 𝒢𝔰𝔬(1,3)\mathscr{G}\in\mathfrak{so}(1,3) have the simplest form. An original algebraic method for constructing such coordinates is presented in Appendix.

In the second part of the paper, we have briefly described the method of noncommutative integration for linear partial differential equations developed by Shapovalov and Shirokov [4]. Unlike the method of separation of variables, the noncommutative integration method more efficiently exploits the first-order symmetry algebra and allows one to integrate the equation without involving second-order symmetry operators. Applying the integrability condition (38) to the constructed first-order symmetry algebras 𝒢^\hat{\mathscr{G}} on dS3\mathrm{dS}_{3}, we selected the classes of electromagnetic fields admitting the noncommutative integrability of the Klein–Gordon equation (see Table 3). All such electromagnetic fields correspond to subalgebras 𝒢3,1\mathscr{G}_{3,1}, 𝒢3,2\mathscr{G}_{3,2}, 𝒢3,3a\mathscr{G}_{3,3}^{a}, 𝒢3,4\mathscr{G}_{3,4}, 𝒢3,5\mathscr{G}_{3,5}, and 𝒢4,1\mathscr{G}_{4,1} of the Lie algebra 𝔰𝔬(1,3)\mathfrak{so}(1,3) (see Table 1). As the four-dimensional subalgebra 𝒢4,1\mathscr{G}_{4,1} corresponds to the electromagnetic field with ab=0\mathscr{F}_{ab}=0, the Klein-Gordon equation in this electromagnetic field is equivalent to the free Klein–Gordon equation, and we, therefore, did not consider this case. In contrast, the cases of three-dimensional subalgebras lead to nontrivial electromagnetic fields, and we examined them in detail. In each case, we reduced the original Klein–Gordon equation (1) to an auxiliary ordinary differential equation and expressed its general solution in terms of special functions where possible.

Acknowledgments

The reported study was funded by RFBR, project number 19-32-90200. Dr. S. V. Danilova is gratefully acknowledged for careful reading of the manuscript.

Appendix: An algebraic method for constructing rectifying coordinates

Let GG be an nn-dimensional Lie transformation group effectively acting on a smooth mm-dimensional manifold MM on the left. The orbit of a point x0Mx_{0}\in M is a subset 𝒪x0\mathcal{O}_{x_{0}} of MM such that 𝒪x0={xM:x=gx0,gG}\mathcal{O}_{x_{0}}=\{x\in M\colon x=gx_{0},\ g\in G\}. Every orbit 𝒪x0\mathcal{O}_{x_{0}} is diffeomorphic to the quotient space G/Gx0G/G_{x_{0}}, where Gx0={gG:gx0=x0}G_{x_{0}}=\{g\in G\colon gx_{0}=x_{0}\} is the stability group of x0x_{0}, also called the isotropy subgroup of x0x_{0}.

It is known that the orbits of the group GG are immersed submanifolds of MM. A group action is called semi-regular if all its orbits have the same dimension. The action is called regular if, in addition, each point xMx\in M has a neighborhood whose intersection with each of the orbits is a pathwise connected subset. It is clear that an action of GG on MM is semi-regular if and only if the dimension of GxG_{x} does not depend on xMx\in M.

Remark A.1.

In general, Lie group actions are not regular or even semi-regular. However, if an action of GG on MM is not semi-regular, then we may restrict our attention to the induced semi-regular action of GG on M~=MS\tilde{M}=M\setminus S, where SS is the union of all singular orbits of GG. Obviously, SS forms a set of measure zero in MM.

Manifolds with regular or semi-regular Lie group actions are locally arranged especially simply. The following fundamental result is a direct consequence of the famous Frobenius’ theorem.

Theorem A.1.

Let the orbits of a semi-regular action of GG on MM have dimension rr. Then, for any point x0Mx_{0}\in M, there exist rectifying local coordinates (q,u)=(q1,,qr,u1,,umr)(q,u)=(q^{1},\dots,q^{r},u^{1},\dots,u^{m-r}) near x0x_{0} such that every orbit intersects the given coordinate chart in the surfaces u1=c1,,umr=cmru^{1}=c_{1},\dots,u^{m-r}=c_{m-r}, where c1,,cmrc_{1},\dots,c_{m-r} are arbitrary constants. If, moreover, the action of GG on MM is regular, then the coordinate chart can be chosen so that each orbit intersects it in at most one such surface.

The proof and detailed discussion of this theorem can be found in Olver’s book [33]. As a consequence of Theorem A.1, the decomposition of MM into the orbits of a semi-regular action forms a foliation of MM.

In practice, rectifying coordinates (q,u)(q,u) are usually constructed using the infinitesimal technique.

Let e1,,ene_{1},\dots,e_{n} be a basis in the Lie algebra 𝒢\mathscr{G} of the Lie group GG. Each basic element eA𝒢e_{A}\in\mathscr{G} can be associated with an infinitesimal generator XAX_{A} of the action GG on MM according to the formula

XA(x)=ddtExp(teA)x|t=0,A=1,,n.X_{A}(x)=\frac{d}{dt}\operatorname{\mathrm{Exp}}(te_{A})x\,\Big{|}_{t=0},\quad A=1,\dots,n.

Here, xMx\in M, Exp:𝒢G\operatorname{\mathrm{Exp}}:\mathscr{G}\to G is the exponential map. These infinitesimal generators satisfy the commutation relations

[XA,XB]=CABCXC,A,B=1,,n,[X_{A},X_{B}]=C_{AB}^{C}X_{C},\quad A,B=1,\dots,n, (A.1)

where the constants CABCC_{AB}^{C} are the structure constants of the Lie algebra 𝒢\mathscr{G}. Thus, the correspondence eAXAe_{A}\to X_{A} extended by linearity to the whole algebra 𝒢\mathscr{G} is a Lie algebra homomorphism. In the case of an effective action, this homomorphism is an isomorphism. Since the action of the group GG on MM is supposed to be effective, we will identify the Lie algebra 𝒢\mathscr{G} with the algebra of infinitesimal generators XAX_{A}.

The infinitesimal generators X1,,XnX_{1},\dots,X_{n} at the point x0Mx_{0}\in M form a basis of the tangent space Tx0𝒪x0T_{x_{0}}\mathcal{O}_{x_{0}} to the orbit 𝒪x0\mathcal{O}_{x_{0}} at x0x_{0}. If we introduce the notation

rdimspan(X1,,Xn)|x0,r\equiv\dim\mathrm{span}\,(X_{1},\dots,X_{n})|_{x_{0}}, (A.2)

this means that dim𝒪x0=r\dim\mathcal{O}_{x_{0}}=r or, equivalently, dimGx0=nr\dim G_{x_{0}}=n-r. Obviously, if the group GG acts on MM semi-regularly, then rr does not depend on the choice of x0x_{0}. We also note that the isotropy subgroup Gx0G_{x_{0}} is generated by those infinitesimal generators from 𝒢\mathscr{G} that vanish at x0x_{0}; the corresponding subalgebra 𝒢x0𝒢\mathscr{G}_{x_{0}}\subset\mathscr{G} is called the isotropy subalgebra of the point x0x_{0}:

𝒢x0={cAeA𝒢:cAXAa(x0)=0,cA}.\mathscr{G}_{x_{0}}=\{c^{A}e_{A}\in\mathscr{G}\colon c^{A}X_{A}^{a}(x_{0})=0,\ c^{A}\in\mathbb{R}\}.

Now, we describe a general scheme for constructing rectifying coordinates (q,u)(q,u) near a point x0Mx_{0}\in M. Let us first make a suitable linear change of coordinates in MM and, if necessary, a change of basis in 𝒢\mathscr{G}, so that

Xa(x0)=xa,a=1,,r;Xα(x0)=0,α=r+1,,n.X_{a}(x_{0})=\frac{\partial}{\partial x^{a}},\quad a=1,\dots,r;\quad X_{\alpha}(x_{0})=0,\quad\alpha=r+1,\dots,n. (A.3)

From (A.3), it follows that the generators Xr+1,,XnX_{r+1},\dots,X_{n} form a basis of the isotropy algebra 𝒢x0\mathscr{G}_{x_{0}}. We consider an (mr)(m-r)-dimensional submanifold Sx0MS_{x_{0}}\subset M formed by the points of the form s(u)=(x01,s(u)=(x_{0}^{1},\dots, x0r,x0r+1+u1,,x0m+umr)x_{0}^{r},x_{0}^{r+1}+u^{1},\dots,x_{0}^{m}+u^{m-r}), where u=(u1,,umr)mru=(u^{1},\dots,u^{m-r})\in\mathbb{R}^{m-r}. By continuity, the orbits of the group GG transversally intersect the submanifold Sx0S_{x_{0}} in a neighborhood of x0x_{0}. Thus, each point xx of the neighborhood can be uniquely written as x=gs(u)x=gs(u), where s(u)Sx0s(u)\in S_{x_{0}} and gGg\in G.

By virtue of (A.3), the basis vectors e1,,ere_{1},\dots,e_{r} span the subspace 𝒢\mathscr{M}\subset\mathscr{G} that is the complement of the isotropy subalgebra 𝒢x0\mathscr{G}_{x_{0}} in 𝒢\mathscr{G}. Since the points s(u)s(u) smoothly depend on the parameters u=(u1,,umr)u=(u^{1},\dots,u^{m-r}) and s(0)=x0s(0)=x_{0}, the direct sum decomposition 𝒢=𝒢s(u)\mathscr{G}=\mathscr{M}\oplus\mathscr{G}_{s(u)} is correct for all small values of uu. This allows us to write the group element gGg\in G in the canonical coordinates of the second kind as

g=a=1rExp(qaea)α=r+1nExp(hαeα(u)),(no summation over a and α),g=\prod_{a=1}^{r}\operatorname{\mathrm{Exp}}(q^{a}e_{a})\cdot\prod_{\alpha=r+1}^{n}\operatorname{\mathrm{Exp}}(h^{\alpha}e_{\alpha}(u)),\quad\text{(no summation over $a$ and $\alpha$)},

where er+1(u),,en(u)e_{r+1}(u),\dots,e_{n}(u) is a basis of the isotropy subalgebra 𝒢s(u)\mathscr{G}_{s(u)}. Taking into account that the transformation corresponding to the group element Exp(hαeα(u))\operatorname{\mathrm{Exp}}(h^{\alpha}e_{\alpha}(u)) leaves the point s(u)s(u) fixed, the decomposition x=gs(u)x=gs(u) can be rewritten in the form:

x=Exp(q1e1)Exp(qrer)s(u).x=\operatorname{\mathrm{Exp}}(q^{1}e_{1})\dots\operatorname{\mathrm{Exp}}(q^{r}e_{r})s(u). (A.4)

The formula (A.4) gives a local diffeomorphism between (x1,,xm)(x^{1},\dots,x^{m}) and (q1,,qr,u1,,umr)(q^{1},\dots,q^{r},u^{1},\dots,u^{m-r}) that determines the desired transformation to rectifying coordinates.

Remark A.2.

Note that in the coordinate chart constructed according to (A.4), the vector field X1X_{1} is automatically “rectifed”, i.e., X1=q1X_{1}=\partial_{q^{1}}.

In the general case, the explicit calculation of the transformation xExp(teA)xx\mapsto\operatorname{\mathrm{Exp}}(te_{A})x is reduced to finding the integral trajectory of the vector field XAX_{A} passing through the point xx. This, in turn, leads to a system of autonomous first-order differential equations. However, if the Lie group GG acts linearly, one can avoid the procedure of integrating differential equations and perform the required calculations, using only linear algebra tools. Indeed, for a linear action of GG on MM the components of the corresponding infinitesimal generators can be chosen to be linear and homogeneous functions of xax^{a}:

XA=ρAabxaxb,A=1,,n,X_{A}=-\rho_{Aa}^{b}x^{a}\partial_{x^{b}},\quad A=1,\dots,n,

where ρAab\rho_{Aa}^{b} are constants. In this case, from the commutation relations (A.1), it follows that the matrices ρAρAab\rho_{A}\equiv\|\rho_{Aa}^{b}\| satisfy the relations

[ρA,ρB]=CABCρC,A,B=1,,n,[\rho_{A},\rho_{B}]=C_{AB}^{C}\rho_{C},\quad A,B=1,\dots,n,

which mean that the mapping eAρAe_{A}\mapsto\rho_{A} defines a representation of the Lie algebra 𝒢\mathscr{G}.

The linearity of the generators XAX_{A} with respect to coordinates allows us to write the action of the group element Exp(teA)\operatorname{\mathrm{Exp}}(te_{A}) on the point xMx\in M in the form etρAxe^{-t\rho_{A}}x, where eY=k=0Yk/k!e^{Y}=\sum_{k=0}^{\infty}Y^{k}/k! denotes the matrix exponential of a matrix YY. Thus, the transformation x=x(q,u)x=x(q,u) to rectifying coordinates can be made by the formula (A.4) with Exp(qaea)\operatorname{\mathrm{Exp}}(q^{a}e_{a}) replaced by eqaρae^{-q^{a}\rho_{a}}. We note that the calculation of matrix exponentials reduces to purely algebraic manipulations and can be carried out using specialized mathematical software (Maple, Wolfram Mathematica, etc.).

Note that the algebraic construction of rectifying coordinates can also be performed for a non-linear action of the Lie group GG if it is induced by a linear semi-regular action of GG on some extended manifold. Indeed, let the Lie group GG act linearly and semi-regularly on a manifold M~\tilde{M} and let MM~M\subset\tilde{M} be a submanifold of the form

M={xM~:Jν(x)=0,ν=1,,m~m},M=\{x\in\tilde{M}\colon J_{\nu}(x)=0,\quad\nu=1,\dots,\tilde{m}-m\},

where {Jν(x)}\{J_{\nu}(x)\} is the set of functionally independent invariants of the action of GG on M~\tilde{M}, m=dimMm=\dim M, m~=dimM~\tilde{m}=\dim\tilde{M}. Let (q,u~)=(q1,,qr,u~1,,u~m~r)(q,\tilde{u})=(q^{1},\dots,q^{r},\tilde{u}^{1},\dots,\tilde{u}^{\tilde{m}-r}) be rectifying local coordinates in a neighborhood of x0Mx_{0}\in M. Since the functions Jν(x)J_{\nu}(x) are invariants of the action of GG, we have Jν(x(q,u~))=κν(u~)J_{\nu}(x(q,\tilde{u}))=\kappa_{\nu}(\tilde{u}), that is, in the rectifying coordinates, JνJ_{\nu} depend only on the variables u~ν\tilde{u}^{\nu}, ν=1,,m~r\nu=1,\dots,\tilde{m}-r. Equating functions κν(u~)\kappa_{\nu}(\tilde{u}) to zero and taking into account their functional independence, we find u~ν=u~ν(u1,,umr)\tilde{u}^{\nu}=\tilde{u}^{\nu}(u^{1},\dots,u^{m-r}), where {u1,,umr}\{u^{1},\dots,u^{m-r}\} is some collection of new parameters. The variables (q1,,qr,u1,,umr)(q^{1},\dots,q^{r},u^{1},\dots,u^{m-r}) obtained in this way can be taken to be rectifying local coordinates on the invariant submanifold MM.

Let us give an example. Consider the standard linear action of the group G=SO(1,2)G=SO(1,2) on Minkowski space 1,3\mathbb{R}^{1,3} generated by the infinitesimal generators:

X1=x0x1x1x0,X2=x0x2x2x0,X3=x2x1+x1x2.X_{1}=-x^{0}\frac{\partial}{\partial x^{1}}-x^{1}\frac{\partial}{\partial x^{0}},\quad X_{2}=-x^{0}\frac{\partial}{\partial x^{2}}-x^{2}\frac{\partial}{\partial x^{0}},\quad X_{3}=-x^{2}\frac{\partial}{\partial x^{1}}+x^{1}\frac{\partial}{\partial x^{2}}. (A.5)

The commutation relations between these generators are

[X1,X2]=X3,[X2,X3]=X1,[X1,X3]=X2.[X_{1},X_{2}]=X_{3},\quad[X_{2},X_{3}]=-X_{1},\quad[X_{1},X_{3}]=X_{2}.

The corresponding Lie algebra 𝒢\mathscr{G} is isomorphic to the semisimple algebra 𝔰𝔬(1,2)\mathfrak{so}(1,2) and is a subalgebra of the Lie algebra 𝔰𝔬(1,3)\mathfrak{so}(1,3) (see the subalgebra 𝒢3,5\mathscr{G}_{3,5} in Table 1). We note that the orbits of GG in 1,3\mathbb{R}^{1,3} are two-dimensional, except the singular orbits that are points of the line L={x1,3:x1=x2=x0=0}L=\{x\in\mathbb{R}^{1,3}\colon x^{1}=x^{2}=x^{0}=0\}. Thus, the action is semi-regular (and even regular) on the open set M~=1,3L\tilde{M}=\mathbb{R}^{1,3}\setminus L, and, in accordance with (A.2), we have r=2r=2.

The function J(x)=(x1)2+(x2)2+(x3)2(x0)2J(x)=(x^{1})^{2}+(x^{2})^{2}+(x^{3})^{2}-(x^{0})^{2} is an invariant of the action since XAJ=0X_{A}J=0 for all A=1,2,3A=1,2,3 . This means that the space dS3={x1,3:J(x)=1}\mathrm{dS}_{3}=\{x\in\mathbb{R}^{1,3}\colon J(x)=1\} regarded as the surface in 1,3\mathbb{R}^{1,3} is invariant under the group GG. Almost all orbits of the corresponding induced action are one-dimensional submanifolds in dS3\mathrm{dS}_{3}, except two singular orbits that are points P±=(0,0,±1,0)P_{\pm}=(0,0,\pm 1,0). Therefore, we have a semi-regular action of the group GG on the open set M=dS3{P±}M=\mathrm{dS}_{3}\setminus\{P_{\pm}\}.

Let us fix the point x0=(1,0,0,0)Mx_{0}=(1,0,0,0)\in M and construct rectifying coordinates (q1,q2,u)(q^{1},q^{2},u) in some neighborhood of this point. As the first step, we choose the new basis X1=X1,X2=X3,X3=X2X_{1}^{\prime}=X_{1},\ X_{2}^{\prime}=X_{3},\ X_{3}^{\prime}=X_{2} in the Lie algebra 𝒢\mathscr{G} and make the change of coordinates in M~\tilde{M}: y1=x0,y2=x2,y3=x3,y0=x1y^{1}=-x^{0},\ y^{2}=x^{2},\ y^{3}=x^{3},\ y^{0}=x^{1}. Then, in these coordinates, we have

X1=y0y1+y1y0,X2=y0y2y2y0,X3=y2y1+y1y2.X^{\prime}_{1}=y^{0}\frac{\partial}{\partial y^{1}}+y^{1}\frac{\partial}{\partial y^{0}},\quad X^{\prime}_{2}=y^{0}\frac{\partial}{\partial y^{2}}-y^{2}\frac{\partial}{\partial y^{0}},\quad X^{\prime}_{3}=y^{2}\frac{\partial}{\partial y^{1}}+y^{1}\frac{\partial}{\partial y^{2}}. (A.6)

At the point x0x_{0} whose new coordinates are y1=y2=y3=0y^{1}=y^{2}=y^{3}=0, y0=1y^{0}=1, the infinitesimal generators (A.6) have the form X1(x0)=y1X^{\prime}_{1}(x_{0})=\partial_{y^{1}}, X2(x0)=y2X^{\prime}_{2}(x_{0})=\partial_{y^{2}}, X3(x0)=0X^{\prime}_{3}(x_{0})=0.

The vector fields XA=ρAijyiyjX^{\prime}_{A}=-\rho_{Ai}^{j}y^{i}\partial_{y^{j}} being infinitesimal generators of the linear action of GG on M~\tilde{M} determine the matrices ρA=ρAij\rho_{A}=\|\rho_{Ai}^{j}\| forming a finite-dimensional representation of the Lie algebra 𝒢\mathscr{G} in the space 1,3\mathbb{R}^{1,3}:

ρ1=(0001000000001000),ρ2=(0000000100000100),ρ3=(0100100000000000).\rho_{1}=\left(\begin{array}[]{cccc}0&0&0&-1\\ 0&0&0&0\\ 0&0&0&0\\ -1&0&0&0\\ \end{array}\right),\quad\rho_{2}=\left(\begin{array}[]{cccc}0&0&0&0\\ 0&0&0&-1\\ 0&0&0&0\\ 0&1&0&0\\ \end{array}\right),\quad\rho_{3}=\left(\begin{array}[]{cccc}0&-1&0&0\\ -1&0&0&0\\ 0&0&0&0\\ 0&0&0&0\\ \end{array}\right).

According to the formula (A.4), the functions yi(q,u~)y^{i}(q,\tilde{u}) defining rectifying coordinates in a neighborhood of x0x_{0} are obtained by computing the expression y=eq1ρ1eq2ρ2s(u~)y=e^{-q^{1}\rho_{1}}e^{-q^{2}\rho_{2}}s(\tilde{u}), where s(u~)=(0,0,u~1,u~21)s(\tilde{u})=(0,0,\tilde{u}^{1},\tilde{u}^{2}-1). Carrying out some calculations and returning to the original coordinates xix^{i}, we find

x1=(u~2+1)coshq1cosq2,x2=(u~2+1)sinq2,x3=u~1,x0=(u~2+1)sinhq1cosq2.x^{1}=\left(\tilde{u}^{2}+1\right)\cosh q^{1}\cos q^{2},\quad x^{2}=\left(\tilde{u}^{2}+1\right)\sin q^{2},\quad x^{3}=\tilde{u}^{1},\quad x^{0}=-\left(\tilde{u}^{2}+1\right)\sinh q^{1}\cos q^{2}. (A.7)

The formulas (A.7) define the transformation to rectifying coordinates on M~\tilde{M} in a neighborhood of the point x0x_{0}.

To construct the rectifying local coordinates (q1,q2,u)(q^{1},q^{2},u) on de Sitter space dS3\mathrm{dS}_{3}, we substitute (A.7) into the invariant J(x)J(x):

κ(u~)J(x(q,u~))=(u~1)2+(u~2+1)2.\kappa(\tilde{u})\equiv J(x(q,\tilde{u}))=(\tilde{u}^{1})^{2}+\left(\tilde{u}^{2}+1\right)^{2}.

The equation κ(u~)=1\kappa(\tilde{u})=1 has the parametric family of solutions u~1=cosu\tilde{u}^{1}=\cos u, u~2=sinu1\tilde{u}^{2}=\sin u-1, where 0u<2π0\leq u<2\pi. Substituting these solutions into (A.7), we arrive at the equalities

x1=coshq1cosq2sinu,x2=sinq2sinu,x3=cosu,x0=sinhq1cosq2sinu,x^{1}=\cosh q^{1}\cos q^{2}\sin u,\quad x^{2}=\sin q^{2}\sin u,\quad x^{3}=\cos u,\quad x^{0}=-\sinh q^{1}\cos q^{2}\sin u, (A.8)

which give the desired transformation to rectifying coordinates. In this chart, the coordinates of the point x0x_{0} are q1=q2=0q^{1}=q^{2}=0, u=π/2u=\pi/2, and

rankxiqa,xiu|x0=3\mathrm{rank}\,\left\|\frac{\partial x^{i}}{\partial q^{a}},\frac{\partial x^{i}}{\partial u}\right\|\Big{|}_{x_{0}}=3

at the point x0x_{0}. It means that the variables q1q^{1}, q2q^{2}, and uu can indeed be considered as local coordinates on dS3\mathrm{dS}_{3} near x0x_{0}. The domain of the coordinate chart is defined by the inequalities

<q1<+,π2<q2<π2,0<u<π.-\infty<q^{1}<+\infty,\quad-\frac{\pi}{2}<q^{2}<\frac{\pi}{2},\quad 0<u<\pi.

It can be verified by direct calculation that the infinitesimal generators (A.5) in the coordinates (q1,q2,u)(q^{1},q^{2},u) take the form

X1=q1,X2=sinh(q1)tan(q2)q1+cosh(q1)q2,X3=cosh(q1)tan(q2)q1+sinh(q1)q2.X_{1}=\frac{\partial}{\partial q^{1}},\quad X_{2}=\sinh(q^{1})\tan(q^{2})\frac{\partial}{\partial q^{1}}+\cosh(q^{1})\frac{\partial}{\partial q^{2}},\quad X_{3}=\cosh(q^{1})\tan(q^{2})\frac{\partial}{\partial q^{1}}+\sinh(q^{1})\frac{\partial}{\partial q^{2}}.

Let us now list rectifying local coordinates (q1,,qr,u1,,u3r)(q^{1},\dots,q^{r},u^{1},\dots,u^{3-r}) on de Sitter space dS31,3\mathrm{dS}_{3}\subset\mathbb{R}^{1,3} for each subalgebra 𝒢n,m\mathscr{G}_{n,m} in Table 1.

Subalgebra 𝒢1,1\mathscr{G}_{1,1} (r=1r=1):

x1=cos(u2),x2=cos(u1)sin(u2),x3=sin(u1)sin(u2)cosh(q1),x0=sin(u1)sin(u2)sinh(q1).x^{1}=\cos(u^{2}),\quad x^{2}=\cos(u^{1})\sin(u^{2}),\quad x^{3}=\sin(u^{1})\sin(u^{2})\cosh(q^{1}),\quad x^{0}=-\sin(u^{1})\sin(u^{2})\sinh(q^{1}).

Subalgebra 𝒢1,2\mathscr{G}_{1,2} (r=1r=1):

x1=cosh(u1)cos(u2)sin(q1),x2=cosh(u1)cos(u2)cos(q1),x3=cosh(u1)sin(u2),x^{1}=-\cosh(u^{1})\cos(u^{2})\sin(q^{1}),\quad x^{2}=\cosh(u^{1})\cos(u^{2})\cos(q^{1}),\quad x^{3}=\cosh(u^{1})\sin(u^{2}),
x0=sinh(u1).x^{0}=\sinh(u^{1}).

Subalgebra 𝒢1,3a\mathscr{G}_{1,3}^{a} (r=1r=1):

x1=cosh(u1)cos(u2)sin(q1),x2=cosh(u1)cos(u2)cos(q1),x^{1}=-\cosh(u^{1})\cos(u^{2})\sin(q^{1}),\quad x^{2}=\cosh(u^{1})\cos(u^{2})\cos(q^{1}),\quad
x3=cosh(u1)sin(u2)cosh(aq1)sinh(u1)sinh(aq1),x^{3}=\cosh(u^{1})\sin(u^{2})\cosh(aq^{1})-\sinh(u^{1})\sinh(aq^{1}),
x0=cosh(u1)sin(u2)sinh(aq1)+sinh(u1)cosh(aq1).x^{0}=-\cosh(u^{1})\sin(u^{2})\sinh(aq^{1})+\sinh(u^{1})\cosh(aq^{1}).

Subalgebra 𝒢1,4\mathscr{G}_{1,4} (r=1r=1):

x1=q1(cosh(u1)sin(u2)sinh(u1)),x2=cosh(u1)cos(u2),x^{1}=q^{1}\left(\cosh(u^{1})\sin(u^{2})-\sinh(u^{1})\right),\quad x^{2}=\cosh(u^{1})\cos(u^{2}),\quad
x3=12(q1)2(sinh(u1)cosh(u1)sin(u2))+cosh(u1)sin(u2),x^{3}=\frac{1}{2}\,(q^{1})^{2}\left(\sinh(u^{1})-\cosh(u^{1})\sin(u^{2})\right)+\cosh(u^{1})\sin(u^{2}),
x0=12(q1)2(sinh(u1)cosh(u1)sin(u2))+sinh(u1).x^{0}=\frac{1}{2}\,(q^{1})^{2}\left(\sinh(u^{1})-\cosh(u^{1})\sin(u^{2})\right)+\sinh(u^{1}).

Subalgebra 𝒢2,1\mathscr{G}_{2,1} (r=2r=2):

x1=q1exp(u1),x2=q2exp(u1),x^{1}=q^{1}\exp(-u^{1}),\quad x^{2}=q^{2}\exp(-u^{1}),\
x3=cosh(u)12exp(u1)((q1)2+(q2)2),x0=sinh(u)12exp(u1)((q1)2+(q2)2).x^{3}=\cosh(u)-\frac{1}{2}\,\exp(-u^{1})\left((q^{1})^{2}+(q^{2})^{2}\right),\quad x^{0}=\sinh(u)-\frac{1}{2}\,\exp(-u^{1})\left((q^{1})^{2}+(q^{2})^{2}\right).

Subalgebra 𝒢2,2\mathscr{G}_{2,2} (r=2r=2):

x1=cos(u1)cos(q1),x2=cos(u1)sin(q1),x3=sin(u1)cosh(q2),x0=sin(u1)sinh(q2).x^{1}=\cos(u^{1})\cos(q^{1}),\quad x^{2}=\cos(u^{1})\sin(q^{1}),\quad x^{3}=\sin(u^{1})\cosh(q^{2}),\quad x^{0}=-\sin(u^{1})\sinh(q^{2}).

Subalgebra 𝒢2,3\mathscr{G}_{2,3} (r=2r=2):

x1=q1exp(q2)cos(u1),x2=sin(u1),x^{1}=q^{1}\exp(q^{2})\cos(u^{1}),\quad x^{2}=\sin(u^{1}),
x3=cos(u1)(cosh(q2)12(q1)2exp(q2)),x0=cos(u1)(sinh(q2)+12(q1)2exp(q2)).x^{3}=\cos(u^{1})\left(\cosh(q^{2})-\frac{1}{2}\,(q^{1})^{2}\exp(q^{2})\right),\quad x^{0}=-\cos(u^{1})\left(\sinh(q^{2})+\frac{1}{2}\,(q^{1})^{2}\exp(q^{2})\right).

Subalgebra 𝒢3,1\mathscr{G}_{3,1} (r=3r=3):

x1=q1exp(q3),x2=q2exp(q3),x^{1}=q^{1}\exp(q^{3}),\ x^{2}=q^{2}\exp(q^{3}), (A.9)
x3=cosh(q3)12exp(q3)((q1)2+(q2)2),x0=sinh(q3)12exp(q3)((q1)2+(q2)2).x^{3}=\cosh(q^{3})-\frac{1}{2}\,\exp(q^{3})\left((q^{1})^{2}+(q^{2})^{2}\right),\quad x^{0}=-\sinh(q^{3})-\frac{1}{2}\,\exp(q^{3})\left((q^{1})^{2}+(q^{2})^{2}\right). (A.10)

Subalgebra 𝒢3,2\mathscr{G}_{3,2} (r=2r=2):

x1=q1exp(u1),x2=q2exp(u1),x^{1}=q^{1}\exp(-u^{1}),\quad x^{2}=q^{2}\exp(-u^{1}), (A.11)
x3=cosh(u)12exp(u)((q1)2+(q2)2),x0=sinh(u)12exp(u)((q1)2+(q2)2).x^{3}=\cosh(u)-\frac{1}{2}\,\exp(-u)\left((q^{1})^{2}+(q^{2})^{2}\right),\quad x^{0}=\sinh(u)-\frac{1}{2}\,\exp(-u)\left((q^{1})^{2}+(q^{2})^{2}\right). (A.12)

Subalgebra 𝒢3,3a\mathscr{G}_{3,3}^{a} (r=3r=3):

x1=q1exp(aq3),x2=q2exp(aq3),x^{1}=q^{1}\exp(aq^{3}),\quad x^{2}=q^{2}\exp(aq^{3}), (A.13)
x3=cosh(aq3)12exp(aq3)((q1)2+(q2)2),x0=sinh(aq3)12exp(aq3)((q1)2+(q2)2).x^{3}=\cosh(aq^{3})-\frac{1}{2}\,\exp(aq^{3})\left((q^{1})^{2}+(q^{2})^{2}\right),\ x^{0}=-\sinh(aq^{3})-\frac{1}{2}\,\exp(aq^{3})\left((q^{1})^{2}+(q^{2})^{2}\right). (A.14)

Subalgebra 𝒢3,4\mathscr{G}_{3,4} (r=2r=2):

x1=cosh(u1)sin(q1)cos(q2),x2=cosh(u1)cos(q1)cos(q2),x3=cosh(u1)sin(q2),x^{1}=-\cosh(u^{1})\sin(q^{1})\cos(q^{2}),\quad x^{2}=\cosh(u^{1})\cos(q^{1})\cos(q^{2}),\quad x^{3}=\cosh(u^{1})\sin(q^{2}), (A.15)
x0=sinh(u1).x^{0}=\sinh(u^{1}). (A.16)

Subalgebra 𝒢3,5\mathscr{G}_{3,5} (r=2r=2):

x1=sin(u1)cosh(q1)cos(q2),x2=sin(u1)sin(q2),x3=cos(u1),x^{1}=\sin(u^{1})\cosh(q^{1})\cos(q^{2}),\quad x^{2}=\sin(u^{1})\sin(q^{2}),\quad x^{3}=\cos(u^{1}), (A.17)
x0=sin(u1)sinh(q1)cos(q2).x^{0}=-\sin(u^{1})\sinh(q^{1})\cos(q^{2}). (A.18)

Subalgebra 𝒢4,1\mathscr{G}_{4,1} (r=3r=3):

x1=q1exp(q3),x2=q2exp(q3),x^{1}=q^{1}\exp(q^{3}),\quad x^{2}=q^{2}\exp(q^{3}),\quad
x3=cosh(q3)12exp(q3)((q1)2+(q2)2),x0=sinh(q3)12exp(q3)((q1)2+(q2)2).x^{3}=\cosh(q^{3})-\frac{1}{2}\,\exp(q^{3})\left((q^{1})^{2}+(q^{2})^{2}\right),\ x^{0}=-\sinh(q^{3})-\frac{1}{2}\,\exp(q^{3})\left((q^{1})^{2}+(q^{2})^{2}\right).

References

  • [1] Furry W 1951 Physical Review 81 115
  • [2] Birrell N and Davies P 1984 Quantum fields in curved space 7 (Cambridge university press)
  • [3] Grib A A, Mostepanenko V and Mamayev S 1994 Vacuum quantum effects in strong fields (Fridmann Lab.)
  • [4] Shapovalov A V and Shirokov I 1995 Theoretical and Mathematical Physics 104 921–934
  • [5] Miller Jr W 1977
  • [6] Woodhouse N 1975 Communications in Mathematical Physics 44 9–38
  • [7] Shapovalov V 1979 Siberian Mathematical Journal 20 790–800
  • [8] Benenti S 2016 Symmetry, Integrability and Geometry: Methods and Applications 12 013
  • [9] Bagrov V G, Baldiotti M C, Gitman D M and Shirokov I V 2002 Journal of Mathematical Physics 43 2284–2305
  • [10] Klishevich V V 2001 Classical and Quantum Gravity 18 3735
  • [11] Magazev A A 2012 Theoretical and Mathematical Physics 173 1654–1667
  • [12] López-Ortega A 2004 General Relativity and Gravitation 36 1299–1319
  • [13] Allen B 1985 Physical Review D 32 3136–3149
  • [14] Yagdjian K and Galstian A 2009 Communications in mathematical physics 285 293–344
  • [15] Otchik V S 1985 Classical and Quantum Gravity 2 539–543
  • [16] Polarski D 1989 Classical and Quantum Gravity 893–900
  • [17] Garriga J 1994 Physical Review D 49 6343–6346
  • [18] Villalba V M 1995 Physical Review D 52 3742–3745
  • [19] Moradi S 2009 Modern Physics Letters A 24 1129–1136
  • [20] Bavarsad E, Kim S P, Stahl C and Xue S S 2018 Physical Review D 97
  • [21] Carter B 1977 Physical Review D 16 3395
  • [22] Van Holten J 2007 Physical Review D 75 025027
  • [23] de Azcárraga J A and Izquierdo J M 1998 Lie groups, Lie algebras, cohomology and some applications in physics (Cambridge University Press)
  • [24] de Azcárraga J A, Izquierdo J M and Bueno J 1998 arXiv preprint physics/9803046
  • [25] Jacobson N 1979 Lie algebras 10 (Courier Corporation)
  • [26] Fris I and Winternitz P 1964 Invariant expansions of relativistic amplitudes and subgroups of the proper Lorentz group Tech. rep. Joint Inst. of Nuclear Research, Dubna, USSR Lab. of Theoretical Physics
  • [27] Barannik A, Barannik L and Fushchich V 1991 Subgroup analysis of Galilean and Poincare groups and reduction of nonlinear equations. (Naukova Dumka, Kiev)
  • [28] Hall G S 2004 Symmetries and curvature structure in general relativity (World Scientific)
  • [29] Olver P J 2000 Applications of Lie groups to differential equations vol 107 (Springer Science & Business Media)
  • [30] Panyushev D I 2003 The index of a Lie algebra, the centralizer of a nilpotent element, and the normalizer of the centralizer Mathematical Proceedings of the Cambridge Philosophical Society vol 134 (Cambridge University Press) pp 41–59
  • [31] Shirokov I V 2000 Theoretical and Mathematical Physics 123 754–767
  • [32] Abramowitz M, Stegun I A and Romer R H 1988 Handbook of mathematical functions with formulas, graphs, and mathematical tables
  • [33] Olver P J 1995 Equivalence, invariants and symmetry (Cambridge University Press)