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

\thankstext

e1e-mail: anpaliat@phys.uoa.gr \thankstexte2e-mail: bogadi.robert@gmail.com \thankstexte3e-mail: megandhreng@dut.ac.za

11institutetext: Institute of Systems Science, Durban University of Technology, Durban 4000, South Africa 22institutetext: Department of Mathematics, Faculty of Applied Sciences, Durban University of Technology, Durban 4000, South Africa

Lie symmetry approach to the time-dependent Karmarkar condition

Andronikos Paliathanasis \thanksrefaddr1,e1    Robert S. Bogadi \thanksrefaddr2,e2    Megandhren Govender \thanksrefaddr2,e3
(Received: date / Accepted: date)
Abstract

We obtain solutions of the time-dependent Einstein Field Equations which satisfy the Karmarkar condition via the method of Lie symmetries. Spherically symmetric spacetime metrics are used with metric functions set to impose conformal flatness, Weyl-free collapse and shear-free collapse. In particular, a solution was found which satisfies the heat-flux boundary condition of Santos, and a radiating stellar model was then obtained and investigated. Solutions obtained which do not allow for the application of the junction conditions at a boundary surface may lend themselves to cosmological models. This is a first attempt in generating solutions satisfying the Karmarkar condition via the method of Lie symmetries and our example of a radiating model highlights the viability of this method.

Keywords:
Karmarkar condition Lie symmetries Conformal flatness Weyl-free collapse Exact solutions
journal: Eur. Phys. J. C

1 Introduction

The gravitational collapse of stellar objects is of much interest in relativistic astrophysics, requiring the solution of time-dependent relativistic field equations. Gravitational collapse problems were pioneered by Oppenheimer and Synder opp , not long after General Relativity was formulated by Einstein. Initially, the Schwarzschild solution was used until the discovery of the Vaidya solution vaid which accommodates null-radiation in the exterior atmosphere due to energy being radiated away from the collapsing body of fluid. There have been numerous attempts at obtaining solutions of the Einstein field equations for describing a radiating body, simultaneously undergoing gravitational collapse, and these efforts typically employ boundary conditions, equations of state, initial static configurations and separation of variables bon ; pret ; bog . The boundary of a collapsing star divides the spacetime into two distinct regions, the interior, \mathcal{M}^{-}, and the exterior spacetime, +\mathcal{M}^{+}. The interior spacetime must match smoothly to the exterior spacetime in order to generate a physically viable model of a radiating star. Early attempts were made by Glass glas in which the Darmois and Lincherowitz matching conditions were utilised. Santos then established the appropriate boundary conditions for a spherically symmetric, shear-free, time-dependent metric which matches smoothly to the exterior Vaidya metric san . The Santos matching condition requires a non-vanishing pressure at the boundary for a star dissipating energy in the form of a radial heat flux. This is a necessary condition that ensures continuity of momentum flux across the boundary. The Santos junction conditions have been generalised to include shear nai ; thir1 , the cosmological constant as well as the electromagnetic field mah1 ; thir2 . Herrera and co-workers her1 ; her2 ; her3 have established important, fundamental results concerning matter distribution, stability of the shear-free condition, energy conditions and thermodynamic properties of gravitational collapse processes.
In developing models describing gravitational collapse, assumptions concerning the gravitational potentials and matter content of the gravitating body are often made. These have included acceleration-free and expansion-free collapse, Weyl-free collapse, anisotropic pressure configurations, the inclusion of shear and bulk viscosity and stipulations of equation of state mah2 ; gov ; red . Differential equations which arise, typically with respect to invariance of the junction conditions, lend themselves to the application of Lie symmetry methods and this can help to determine novel, exact solutions mso ; abe1 ; abe2 . In addition to restricting the matter content, conditions on the spacetime geometry are just as important. It is of interest to consider gravitational fields, typically represented by a Riemannian metric of four dimensions, to be immersed in a flat spacetime of higher dimension. Randall-Sundram and Anchordoqui-Bergliaffa re-established the conjecture that 4-dimensional spacetime might be embedded in higher dimensional flat space and much effort is made to achieve class one embedding mau1 . In general, an n-dimensional Riemannian spacetime is said to be of class p if it can be embedded into a flat space of dimension n + p. The Karmarkar condition kar is a necessary (but not sufficient) condition for a spacetime to be of class one pan . The derivation of the Karmarkar condition is purely geometric in nature providing relations among the components of the Riemann tensor. This in turn relates the metric potentials to one another and has thus assisted in model development. Of interest is to note that the Karmarkar condition together with pressure isotropy gives the interior Schwarzschild solution as the only bounded matter configuration. Recent attempts at modelling compact objects such as 4U 1538-52,PSR J1614-2230, Vela X-1 and Cen X-3 using the Karmarkar condition have produced models with favourable physical characteristics, consistent with observations bha ; mau2 . It is expected that the Karmarkar condition should also be favourable for time-dependent systems, perhaps improving stability which is an issue for shear-free spacetimes her2 .
Lie symmetry analysis is a very powerful tool for the study of nonlinear differential equations and we make use of this in generating solutions of differential equations obtained via Karmarkar’s condition. The steps that we follow in the analysis are: (i) we determine the Lie symmetries for each master partial differential equation for each model, (ii) the one-dimensional optimal system is determined in each case for the admitted Lie symmetries, (iii) we define similarity transformations from the Lie symmetries by using the Lie invariants which are used to reduce the master equation into an ordinary differential equation and (iv) exact closed-form solutions are then obtained. These steps have been applied before for various gravitational models with interesting results. Exact solutions describing charged radiation have been derived by applying Lie symmetries abe4 . Moreover, in abe2 Lie symmetries have been used to derive expanding and shearing models of radiating relativistic stars, while shear-free radiating stars were considered in abe3 . In GG , a new solution which describes a Euclidean star is derived by Lie symmetries and has the physical property of satisfying all the energy conditions and admitting a barotropic equation of state. For the field equations of the Schwarzschild model, conservation laws, invariant functions and differential operators derived by using the Lie symmetry analysis in ch5 ; ch6 . The Emden–Fowler equation which can describe gravitational spherically symmetric solutions was investigated via symmetry analysis in ch7 . For other applications of symmetry analysis in gravitational physics, we refer the reader to ch8 ; ch9 ; ch10 ; ch11 and the references therein.
The plan of this paper is as follows: In Section 2 we present the gravitational field equations for our analysis. In Section 3, the basic properties and definitions of Lie symmetries are given. The definition of the one-dimensional optimal system is discussed. The latter is necessary in order to perform a complete classification of the admitted similarity transformations. In Sections 4, 5 and 6 we solve the Karmarkar condition with respect to conformally flat, Weyl-free collapse and shear-free metrics respectively. A physical application is then given in Section 7 which describes radiating, Weyl-free collapse. In Section 8 we discuss the merits of the radiating stellar model obtained and in Section 9 we conclude on the suitability and novelty of our methods. Appendices A, B and C complete the presentation of this study where we present exact solutions for the three cases given in Sections 4, 5 and 6. These may be used for future studies.

2 Spherically symmetric spacetimes in relativity

The spherically symmetric line element is given by

ds2=A2(t,r)dt2+B2(t,r)dr2+Y2(t,r)dΩ2ds^{2}=-A^{2}\left(t,r\right)dt^{2}+B^{2}\left(t,r\right)dr^{2}+Y^{2}\left(t,r\right)d\Omega^{2} (1)

wheredΩ2~{}d\Omega^{2} is the line element of the two-sphere, that is,

dΩ2=dθ2+sin2θdϕ2.d\Omega^{2}=d\theta^{2}+\sin^{2}\theta d\phi^{2}. (2)

An energy-momentum tensor incorporating heat flux, qa=(0,q1,0,0)q^{a}=(0,q^{1},0,0), is used,

Tab\displaystyle T^{-}_{ab} =\displaystyle= (ρ+pt)uaub+ptgab+(prpt)χaχb+qaub+qbua\displaystyle(\rho+p_{t})u_{a}u_{b}+p_{t}g_{ab}+(p_{r}-p_{t})\chi_{a}\chi_{b}+q_{a}u_{b}+q_{b}u_{a}

where, ρ\rho is the energy density, prp_{r} the radial pressure, ptp_{t} the tangential pressure and qaq_{a} the heat flux vector. The timelike four-velocity of the fluid is uau_{a} and χa\chi_{a} is a spacelike unit four-velocity along the radial direction. These quantities must satisfy uaua=1u_{a}u^{a}=-1, uaqa=0u_{a}q^{a}=0, χaχa=1\chi_{a}\chi^{a}=1 and χaua=0\chi_{a}u^{a}=0. In co-moving coordinates we have

ua=A1δ0a,qa=qδ1a,χa=B1δ1au^{a}=A^{-1}\delta_{0}^{a}~{}~{},~{}~{}q^{a}=q\delta_{1}^{a}~{}~{},~{}~{}\chi^{a}=B^{-1}\delta_{1}^{a} (4)

The four-acceleration and expansion scalar are given by

wa=ua;bub,Θ=ua;aw_{a}=u_{a;b}u^{b}~{}~{},~{}~{}\Theta={u^{a}}_{;a} (5)

Einstein’s time-dependent field equations are then given by

ρ\displaystyle\rho =\displaystyle= 1A2(2B˙B+Y˙Y)Y˙Y\displaystyle\frac{1}{A^{2}}\left(2\frac{\dot{B}}{B}+\frac{\dot{Y}}{Y}\right)\frac{\dot{Y}}{Y} (6)
1B2[2Y′′Y+(YY)22BBYY(BY)2],\displaystyle-\frac{1}{B^{2}}\left[2\frac{Y^{\prime\prime}}{Y}+\left(\frac{Y^{\prime}}{Y}\right)^{2}-2\frac{B^{\prime}}{B}\frac{Y^{\prime}}{Y}-\left(\frac{B}{Y}\right)^{2}\right],
pr\displaystyle p_{r} =\displaystyle= 1A2[2Y¨Y(2A˙AY˙Y)Y˙Y]\displaystyle-\frac{1}{A^{2}}\left[2\frac{{\ddot{Y}}}{Y}-\left(2\frac{\dot{A}}{A}-\frac{\dot{Y}}{Y}\right)\frac{\dot{Y}}{Y}\right] (7)
+1B2(2AA+YY)YY1Y2,\displaystyle+\frac{1}{B^{2}}\left(2\frac{A^{\prime}}{A}+\frac{Y^{\prime}}{Y}\right)\frac{Y^{\prime}}{Y}-\frac{1}{Y^{2}},
pt\displaystyle p_{t} =\displaystyle= 1A2[B¨B+Y¨YA˙A(B˙B+Y˙Y)+B˙BY˙Y]\displaystyle-\frac{1}{A^{2}}\left[\frac{\ddot{B}}{B}+\frac{{\ddot{Y}}}{Y}-\frac{\dot{A}}{A}\left(\frac{\dot{B}}{B}+\frac{\dot{Y}}{Y}\right)+\frac{\dot{B}}{B}\frac{\dot{Y}}{Y}\right] (8)
+1B2[A′′A+Y′′YAABB+(AABB)YY],\displaystyle+\frac{1}{B^{2}}\left[\frac{A^{\prime\prime}}{A}+\frac{Y^{\prime\prime}}{Y}-\frac{A^{\prime}}{A}\frac{B^{\prime}}{B}+\left(\frac{A^{\prime}}{A}-\frac{B^{\prime}}{B}\right)\frac{Y^{\prime}}{Y}\right],
q\displaystyle q =\displaystyle= 2AB(Y˙YB˙BYYY˙YAA).\displaystyle\frac{2}{AB}\left(\frac{\dot{Y^{\prime}}}{Y}-\frac{\dot{B}}{B}\frac{Y^{\prime}}{Y}-\frac{\dot{Y}}{Y}\frac{A^{\prime}}{A}\right). (9)

In the above, ρ\rho, prp_{r}, ptp_{t} and qq are the energy density, radial pressure, tangential pressure and radial heat flux respectively.

2.1 The Karmarkar condition

The Karmarkar condition kar , which allows for the embedding of a four-dimensional spacetime into a five-dimensional pseudo-Euclidean space, is given in terms of the following relationship with respect to the components of the Riemann tensor, namely

10102323=1212303012201330,\mathcal{R}_{1010}\mathcal{R}_{2323}=\mathcal{R}_{1212}\mathcal{R}_{3030}-\mathcal{R}_{1220}\mathcal{R}_{1330}, (10)

where the notation (0,1,2,3)(0,1,2,3) represents the coordinates (t,r,θ,ϕ)(t,r,\theta,\phi). We then consider the metric (1) and calculate Karmarkar’s condition for a shearing, nonstatic spherically symmetric metric. The relevant nonzero Riemann tensor components are

1010\displaystyle\mathcal{R}_{1010} =\displaystyle= 1AB(B2B¨A+A2A′′BBAA2+B˙A˙B2),\displaystyle\frac{1}{AB}\biggl{(}-B^{2}\ddot{B}A+A^{2}{A}^{\prime\prime}B-{B}^{\prime}{A}^{\prime}A^{2}+\dot{B}\dot{A}B^{2}\biggr{)},
1212\displaystyle\mathcal{R}_{1212} =\displaystyle= YA2B(Y′′BA2+BYA2+B2B˙Y˙),\displaystyle\frac{Y}{A^{2}B}\biggl{(}-{Y}^{\prime\prime}BA^{2}+{B}^{\prime}{Y}^{\prime}A^{2}+B^{2}\dot{B}\dot{Y}\biggr{)},
1220\displaystyle\mathcal{R}_{1220} =\displaystyle= YAB(Y˙AB+B˙YA+AY˙B),\displaystyle\frac{-Y}{AB}\biggl{(}-{\dot{Y}}^{\prime}AB+\dot{B}{Y}^{\prime}A+{A}^{\prime}\dot{Y}B\biggr{)},
1330\displaystyle\mathcal{R}_{1330} =\displaystyle= Ysin2θAB(Y˙BA+B˙YA+AY˙B),\displaystyle\frac{-Y\sin^{2}{\theta}}{AB}\biggl{(}-{\dot{Y}}^{\prime}BA+\dot{B}{Y}^{\prime}A+{A}^{\prime}\dot{Y}B\biggr{)},
2323\displaystyle\mathcal{R}_{2323} =\displaystyle= Y2sin2θB2A2(B2A2Y2A2+Y˙2B2),\displaystyle\frac{Y^{2}\sin^{2}{\theta}}{B^{2}A^{2}}\biggl{(}B^{2}A^{2}-{{Y}^{\prime}}^{2}A^{2}+{\dot{Y}}^{2}B^{2}\biggr{)},
3030\displaystyle\mathcal{R}_{3030} =\displaystyle= Ysin2θAB2(Y¨B2A+A2AY+Y˙A˙B2).\displaystyle\frac{Y\sin^{2}{\theta}}{AB^{2}}\biggl{(}-\ddot{Y}B^{2}A+A^{2}{A}^{\prime}{Y}^{\prime}+\dot{Y}\dot{A}B^{2}\biggr{)}.

which results in the following expression for the Karmarkar condition:

0\displaystyle 0 =\displaystyle= AB(AYB˙+BAY˙ABY˙)2\displaystyle AB\bigg{(}A{Y}^{\prime}\dot{B}+B{A}^{\prime}\dot{Y}-AB\dot{Y}^{\prime}\bigg{)}^{2} (12)
+(A2B2A2Y2+B2Y˙2)\displaystyle+\bigg{(}A^{2}B^{2}-A^{2}{{Y}^{\prime}}^{2}+B^{2}{\dot{Y}}^{2}\bigg{)}
×(A2AB+A2BA′′+B2A˙B˙AB2B¨)\displaystyle\times\bigg{(}-A^{2}{A}^{\prime}{B}^{\prime}+A^{2}B{A}^{\prime\prime}+B^{2}\dot{A}\dot{B}-AB^{2}\ddot{B}\bigg{)}
(A2BYA2BY′′+B2B˙Y˙)\displaystyle-\bigg{(}A^{2}{B}^{\prime}{Y}^{\prime}-A^{2}B{Y}^{\prime\prime}+B^{2}\dot{B}\dot{Y}\bigg{)}
×(A2AY+B2A˙Y˙AB2Y¨).\displaystyle\times\bigg{(}A^{2}{A}^{\prime}{Y}^{\prime}+B^{2}\dot{A}\dot{Y}-AB^{2}\ddot{Y}\bigg{)}.

3 Lie symmetries of differential equations

In the context of geometry a differential equation (DE) may be considered as a function H=H(yi,uA,u,iA,u,ijA,)H=H(y^{i},u^{A},u_{,i}^{A},u_{,ij}^{A},...) in the jet-space B=B(yi,uA,u,iA,u,ijA,)B=B\left(y^{i},u^{A},u_{,i}^{A},u_{,ij}^{A},...\right), where {yi}\left\{y^{i}\right\} are the independent variables and uAu^{A} are the dependent variables, while comma means derivative with respect to the variable yiy^{i}, that is u,iA=uAyiu_{,i}^{A}=\frac{\partial u^{A}}{\partial y^{i}}.

Consider the infinitesimal one-parameter point transformation

x¯i\displaystyle\bar{x}^{i} =\displaystyle= xi+εξi(xk,uB),\displaystyle x^{i}+\varepsilon\xi^{i}(x^{k},u^{B})~{}, (13)
u¯A\displaystyle\bar{u}^{A} =\displaystyle= u¯A+εηA(xk,uB),\displaystyle\bar{u}^{A}+\varepsilon\eta^{A}(x^{k},u^{B})~{}, (14)

with generator

𝐗=ξi(xk,uB)xi+ηA(xk,uB)uA.\mathbf{X}=\xi^{i}(x^{k},u^{B})\partial_{x^{i}}+\eta^{A}(x^{k},u^{B})\partial_{u^{A}}~{}. (15)

The vector field 𝐗\mathbf{X} which defines the infinitesimal transformation (13)-(14) is called a Lie point symmetry of the DE HH if there exists a function κ\kappa such that the following condition holds Stephani ; Bluman ; ibra

𝐗[n](H)=κH,modH=0,\mathbf{X}^{[n]}(H)=\kappa H~{},~{}modH=0, (16)

where

𝐗[n]=𝐗+ηiAuiA+ηijAuijA+\mathbf{X}^{[n]}=\mathbf{X}+\eta_{i}^{A}\partial_{u_{i}^{A}}+\eta_{ij}^{A}\partial_{u_{ij}^{A}}+... (17)

is the nnth extension vector. Coefficient ηiA\eta_{i}^{A} of the first extension vector is defined as

ηiA=η,iA+u,iBη,BAξ,iju,jAu,iAu,jBξ,Bj,\eta_{i}^{A}=\eta_{,i}^{A}+u_{,i}^{B}\eta_{,B}^{A}-\xi_{,i}^{j}u_{,j}^{A}-u_{,i}^{A}u_{,j}^{B}\xi_{,B}^{j}~{}, (18)

coefficient ηijA\eta_{ij}^{A} of the second extension vector is given by the expression

ηijA\displaystyle\eta_{ij}^{A} =\displaystyle= η,ijA+2η,B(iAu,j)Bξ,ijku,kA+η,BCAu,iBu,jC2ξ,(i|B|kuj)Bu,kA\displaystyle\eta_{,ij}^{A}+2\eta_{,B(i}^{A}u_{,j)}^{B}-\xi_{,ij}^{k}u_{,k}^{A}+\eta_{,BC}^{A}u_{,i}^{B}u_{,j}^{C}-2\xi_{,(i|B|}^{k}u_{j)}^{B}u_{,k}^{A} (19)
ξ,BCku,iBu,jCu,kA+η,BAu,ijB2ξ,(jku,i)kA\displaystyle-\xi_{,BC}^{k}u_{,i}^{B}u_{,j}^{C}u_{,k}^{A}+\eta_{,B}^{A}u_{,ij}^{B}-2\xi_{,(j}^{k}u_{,i)k}^{A}
ξ,Bk(u,kAu,ijB+2u(,jBu,i)kA)\displaystyle-\xi_{,B}^{k}\left(u_{,k}^{A}u_{,ij}^{B}+2u_{(,j}^{B}u_{,i)k}^{A}\right)

while coefficient ηijjnA\eta_{ij...j_{n}}^{A} of the nnth extension vector is defined as

ηijjnA=Dηijjn1Auij..kADjnξk\eta_{ij...j_{n}}^{A}=D\eta_{ij...j_{n-1}}^{A}-u_{ij..k}^{A}D_{j_{n}}\xi^{k} (20)

The main application of Lie point symmetries of a DE is focused on the construction of invariant functions which can be used for the determination of invariant solutions also known as similarity solutions.

For the Lie point symmetry 𝐗\mathbf{X} of the differential equation HH we define the Lagrange system Stephani ; Bluman ; ibra as

dxiξi=duAηA=duiAη[i]A=duijAη[ij]A=\frac{dx^{i}}{\xi^{i}}=\frac{du^{A}}{\eta^{A}}=\frac{du_{i}^{A}}{\eta_{\left[i\right]}^{A}}=\frac{du_{ij}^{A}}{\eta_{\left[ij\right]}^{A}}=... (21)

whose solution provides the characteristic functions
W[0](yk,u),W[1](yk,u,ui)W^{\left[0\right]}\left(y^{k},u\right),~{}W^{\left[1\right]}\left(y^{k},u,u_{i}\right), W[2](yk,u,u,i,uij),W^{\left[2\right]}\left(y^{k},u,u_{,i},u_{ij}\right),...~{}.
The characteristic functions, can be applied to reduce the order of the DE (in the case of ordinary differential equations) or the number of the dependent variables (in the case of partial differential equations).

3.1 One-dimension optimal system

For a given differential equation HH which admit a Lie algebra GnG_{n} of dimension dimGn=n\dim G_{n}=n and elements {X1,X2,Xn},\left\{X_{1},~{}X_{2},~{}...~{}X_{n}\right\}, we consider the two generic vector fields olver

Z\displaystyle Z =i=1naiXi,W=\displaystyle=\sum\limits_{i=1}^{n}a_{i}X_{i}~{},\hskip 14.22636pt~{}W= i=1nbiXi,\displaystyle\sum\limits_{i=1}^{n}b_{i}X_{i}~{}, (22)

where ai,bia_{i},~{}b_{i} are constants.

The vector fields Z,WZ,~{}W are equivalent and leads to the same similarity transformation if and only if

𝐖=iAd(exp(εiXi))𝐙\mathbf{W}={\displaystyle\prod\limits_{i}}Ad\left(\exp\left(\varepsilon_{i}X_{i}\right)\right)\mathbf{Z} (23)

or

W=cZ,c=const.W=cZ~{},~{}c=const. (24)

where the operator Ad(exp(εXi))Ad\left(\exp\left(\varepsilon X_{i}\right)\right) is the Adjoint operator defined as olver

Ad(exp(εXi))Xj=Xjε[Xi,Xj]+12ε2[Xi,[Xi,Xj]]+Ad\left(\exp\left(\varepsilon X_{i}\right)\right)X_{j}=X_{j}-\varepsilon\left[X_{i},X_{j}\right]+\frac{1}{2}\varepsilon^{2}\left[X_{i},\left[X_{i},X_{j}\right]\right]+... (25)

which is used to determine the Adjoint representation. Hence, in order to perform a complete classification for the similarity solutions of a given differential equation we should determine all the one-dimensional independent symmetry vectors of the Lie algebra GnG_{n}.

In the following sections we consider special forms for the unknown metric functions such that there is only one unknown function, for that models we perform a detailed analysis of the Karmarkar condition by using the Lie’s theory. In particular we determine the Lie point symmetries and the one-dimensional optimal system for the Karmarkar condition, while we determine similarity solutions.

4 Model A: Conformally flat metric

Consider A(t,r)=B(t,r)A\left(t,r\right)=B\left(t,r\right) and Y(t,r)=rB(t,r)Y\left(t,r\right)=rB\left(t,r\right). In this case, the line element (1) is,

ds2=B2(t,r)(dt2+dr2+r2dΩ2).ds^{2}=B^{2}\left(t,r\right)\left(-dt^{2}+dr^{2}+r^{2}d\Omega^{2}\right). (26)

A spacetime with line element (26) is conformally flat and admits fifteen Conformal Killing vector fields (CKVs). For the conformally flat metric (26) where B(t,r)B\left(t,r\right) is the unique unknown function, the Karmarkar condition becomes,

0\displaystyle 0 =\displaystyle= 4r2B˙BB˙+B(r2B˙2+B2rB′′B+rB¨(BrB′′))\displaystyle-4r^{2}\dot{B}B^{\prime}\dot{B}^{\prime}+B\left(r^{2}\dot{B}^{\prime 2}+B^{\prime 2}-rB^{\prime\prime}B^{\prime}+r\ddot{B}\left(B^{\prime}-rB^{\prime\prime}\right)\right) (27)
+2rB2(rB¨+B)+2rB˙2(rB′′B)\displaystyle+2rB^{\prime 2}\left(r\ddot{B}+B^{\prime}\right)+2r\dot{B}^{2}\left(rB^{\prime\prime}-B^{\prime}\right)

We apply the Lie theory to equation (27) from which we obtain the Lie point symmetry vectors

X1=t,X2=BB,X3=B2B,X4=1rr,X_{1}=\partial_{t}~{},~{}X_{2}=B\partial_{B}~{},~{}X_{3}=B^{2}\partial_{B}~{},~{}X_{4}=\frac{1}{r}\partial_{r}~{},~{}
X5=tB2B,X6=(r2t2)B2B,X7=tt+rr.X_{5}=tB^{2}\partial_{B}~{},~{}X_{6}=\left(r^{2}-t^{2}\right)B^{2}\partial_{B}~{},~{}X_{7}=t\partial_{t}+r\partial_{r}.

The admitted Lie symmetries form a seven-dimensional Lie algebra GAG_{A}, i.e. dimGA=7\dim G_{A}=7, and the associated commutators are shown in Table 1. Moreover, in Table 2 we present the Adjoint representation for the elements of the Lie algebra GAG_{A}.

We continue by using the results in Tables 1 and 2 to derive the one-dimensional optimal system for the partial differential equation (27). Consider the generic symmetry vector

Z=α1X1+α2X2+α3X3+α4X4+α5X5+α6X6+α7X7Z=\alpha_{1}X_{1}+\alpha_{2}X_{2}+\alpha_{3}X_{3}+\alpha_{4}X_{4}+\alpha_{5}X_{5}+\alpha_{6}X_{6}+\alpha_{7}X_{7} (28)

From Table 2 we see that by applying the following adjoint representations

Z\displaystyle Z^{\prime} =\displaystyle= Ad(exp(ε2X2))Ad(exp(ε5X5))Ad(exp(ε6X6))×\displaystyle Ad\left(\exp\left(\varepsilon_{2}X_{2}\right)\right)Ad\left(\exp\left(\varepsilon_{5}X_{5}\right)\right)Ad\left(\exp\left(\varepsilon_{6}X_{6}\right)\right)\times (29)
Ad(exp(ε1X1))Z,\displaystyle Ad\left(\exp\left(\varepsilon_{1}X_{1}\right)\right)Z,

where for specific values of ε1,ε2,ε5\varepsilon_{1,}~{}\varepsilon_{2},~{}\varepsilon_{5} and ε6\varepsilon_{6} it follows

Z=α2X2+α4X4+α7X7.Z^{\prime}=\alpha_{2}^{\prime}X_{2}+\alpha_{4}^{\prime}X_{4}+\alpha_{7}^{\prime}X_{7}. (30)

Thus, the two vector fields ZZ^{\prime} and ZZ are equivalent and lead to the same similarity solution. Coefficient constants α2,α4\alpha_{2},~{}\alpha_{4} and α7\alpha_{7} are called relative invariants of the full adjoint action. Thus, in order to derive the relative invariants we solve the following system of partial differential equation olver

Δ(ϕ(αi))=Cijkαiαj\Delta\left(\phi\left(\alpha_{i}\right)\right)=C_{ij}^{k}\alpha^{i}\frac{\partial}{\partial\alpha_{j}} (31)

where CijkC_{ij}^{k} are the structure constants of the Lie algebra GAG_{A} as presented in 1. Hence, system (31) becomes

α5ϕα32α6ϕα5+α7ϕα1\displaystyle\alpha_{5}\frac{\partial\phi}{\partial\alpha_{3}}-2\alpha_{6}\frac{\partial\phi}{\partial\alpha_{5}}+\alpha_{7}\frac{\partial\phi}{\partial\alpha_{1}} =0,\displaystyle=0, (32)
α3ϕα3+α5ϕα5+α6ϕα6\displaystyle\alpha_{3}\frac{\partial\phi}{\partial\alpha_{3}}+\alpha_{5}\frac{\partial\phi}{\partial\alpha_{5}}+\alpha_{6}\frac{\partial\phi}{\partial\alpha_{6}} =0,\displaystyle=0, (33)
α2ϕα3\displaystyle-\alpha_{2}\frac{\partial\phi}{\partial\alpha_{3}} =\displaystyle= 0,\displaystyle 0, (34)
2α6ϕα3\displaystyle 2\alpha_{6}\frac{\partial\phi}{\partial\alpha_{3}} =\displaystyle= 0,\displaystyle 0, (35)
α1ϕα3(α2+α7)ϕα5\displaystyle-\alpha_{1}\frac{\partial\phi}{\partial\alpha_{3}}-\left(\alpha_{2}+\alpha_{7}\right)\frac{\partial\phi}{\partial\alpha_{5}} =\displaystyle= 0,\displaystyle 0, (36)
2α1ϕα5(α2+2α7)ϕα62α4ϕα3\displaystyle 2\alpha_{1}\frac{\partial\phi}{\partial\alpha_{5}}-\left(\alpha_{2}+2\alpha_{7}\right)\frac{\partial\phi}{\partial\alpha_{6}}-2\alpha_{4}\frac{\partial\phi}{\partial\alpha_{3}} =\displaystyle= 0,\displaystyle 0, (37)
α1ϕα1+α5ϕα5+α6ϕα6\displaystyle-\alpha_{1}\frac{\partial\phi}{\partial\alpha_{1}}+\alpha_{5}\frac{\partial\phi}{\partial\alpha_{5}}+\alpha_{6}\frac{\partial\phi}{\partial\alpha_{6}} =\displaystyle= 0,\displaystyle 0, (38)

from where it follows ϕ(αi)=ϕ(α2,α4,α7)\phi\left(\alpha_{i}\right)=\phi\left(\alpha_{2},\alpha_{4},\alpha_{7}\right), that is, the relative invariants are {α2,α4,α7}\left\{\alpha_{2},\alpha_{4},\alpha_{7}\right\}.

For α7=0\alpha_{7}=0, we find the equivalent symmetry vector

Z′′=α1′′X1+α2′′X2+α4′′X4.Z^{\prime\prime}=\alpha_{1}^{\prime\prime}X_{1}+\alpha_{2}^{\prime\prime}X_{2}+\alpha_{4}^{\prime\prime}X_{4}. (39)

On the other hand for α2=0\alpha_{2}=0, the equivalent symmetry vector is

Z′′′=α4′′′X4+α7′′′X7.Z^{\prime\prime\prime}=\alpha_{4}^{\prime\prime\prime}X_{4}+\alpha_{7}^{\prime\prime\prime}X_{7}. (40)

while when α4=0\alpha_{4}=0, the resulting equivalent symmetry vector is derived

Z′′′′=α2′′′′X2+α7′′′′X7.Z^{\prime\prime\prime\prime}=\alpha_{2}^{\prime\prime\prime\prime}X_{2}+\alpha_{7}^{\prime\prime\prime\prime}X_{7}. (41)

Similarly, the following equivalent symmetry vectors are obtained:

{α2=0,α4=0}Z¯=α¯7X7,\{\alpha_{2}=0,~{}\alpha_{4}=0\}~{}\Rightarrow~{}\bar{Z}=\bar{\alpha}_{7}X_{7}~{},
{α4=0,α7=0}Z¯=α¯1X1+α¯2X2,\{\alpha_{4}=0,~{}\alpha_{7}=0\}~{}\Rightarrow~{}\bar{Z}^{\prime}=\bar{\alpha}_{1}^{\prime}X_{1}+\bar{\alpha}_{2}^{\prime}X_{2}~{},
{α2=0,α7=0}Z¯′′=α¯1′′X1+α¯4′′X4+α¯5′′X5+α¯6′′X6,\{\alpha_{2}=0,~{}\alpha_{7}=0\}~{}\Rightarrow~{}\bar{Z}^{\prime\prime}=\bar{\alpha}_{1}^{\prime\prime}X_{1}+\bar{\alpha}_{4}^{\prime\prime}X_{4}+\bar{\alpha}_{5}^{\prime\prime}X_{5}+\bar{\alpha}_{6}^{\prime\prime}X_{6}~{},
{α2=0,α4=0,α7=0}Z¯′′′=α¯1′′′X1+α¯5′′′X5+α¯6′′′X6.\{\alpha_{2}=0,~{}\alpha_{4}=0,~{}\alpha_{7}=0\}~{}\Rightarrow~{}\bar{Z}^{\prime\prime\prime}=\bar{\alpha}_{1}^{\prime\prime\prime}X_{1}+\bar{\alpha}_{5}^{\prime\prime\prime}X_{5}+\bar{\alpha}_{6}^{\prime\prime\prime}X_{6}.

Thus, a one-dimensional optimal system is constructed from the one-dimensional Lie algebras:

{X1},{X2},{X3},{X4},{X5},{X6},{X7},\{X_{1}\},~{}\{X_{2}\},~{}\{X_{3}\},~{}\{X_{4}\},~{}\{X_{5}\},~{}\{X_{6}\},~{}\{X_{7}\},~{}
{X1+αX5},{X1+αX6},{X5+αX6},{X2+αX7},\{X_{1}+\alpha X_{5}\},~{}\{X_{1}+\alpha X_{6}\},~{}\{X_{5}+\alpha X_{6}\},~{}\{X_{2}+\alpha X_{7}\},~{}
{X5+αX7},{X1+αX4},{X4+αX5},{X4+αX6},\{X_{5}+\alpha X_{7}\},~{}\{X_{1}+\alpha X_{4}\},~{}\{X_{4}+\alpha X_{5}\},~{}\{X_{4}+\alpha X_{6}\},~{}
{X1+αX2+βX4},{X1+αX5+βX6},{X4+αX1+βX5},\{X_{1}+\alpha X_{2}+\beta X_{4}\},~{}\{X_{1}+\alpha X_{5}+\beta X_{6}\},~{}\{X_{4}+\alpha X_{1}+\beta X_{5}\},~{}
{X4+αX1+βX6},{X4+αX5+βX6},{X2+αX4+βX7}.\{X_{4}+\alpha X_{1}+\beta X_{6}\},~{}\{X_{4}+\alpha X_{5}+\beta X_{6}\},~{}\{X_{2}+\alpha X_{4}+\beta X_{7}\}.

We proceed by using the Lie symmetries to determine similarity transformations, in the following we present the application of Lie point symmetries which lead to exact solutions expressed in closed-form functions.

Table 1: Commutators of the admitted Lie point symmetries for the Karmarkar condition (26)
[,]\left[~{},~{}\right] 𝐗1\mathbf{X}_{1} 𝐗2\mathbf{X}_{2} 𝐗3\mathbf{X}_{3} 𝐗4\mathbf{X}_{4} 𝐗5\mathbf{X}_{5} 𝐗6\mathbf{X}_{6} 𝐗7\mathbf{X}_{7}
𝐗1\mathbf{X}_{1} 0 0 0 0 X3X_{3} 2X5-2X_{5} X1X_{1}
𝐗2\mathbf{X}_{2} 0 0 X3X_{3} 0 X5X_{5} X6X_{6} 0
𝐗3\mathbf{X}_{3} 0 X3-X_{3} 0 0 0 0 0
𝐗4\mathbf{X}_{4} 0 0 0 0 0 2X32X_{3} 0
𝐗5\mathbf{X}_{5} X3-X_{3} X5-X_{5} 0 0 0 0 X5-X_{5}
𝐗6\mathbf{X}_{6} 2X52X_{5} X6-X_{6} 0 2X3-2X_{3} 0 0 2X6-2X_{6}
𝐗7\mathbf{X}_{7} X1-X_{1} 0 0 0 X5X_{5} 2X62X_{6} 0
Table 2: Adjoint representation for the admitted Lie point symmetries for the Karmarkar condition (26)
Ad(e(ε𝐗i))𝐗jAd\left(e^{\left(\varepsilon\mathbf{X}_{i}\right)}\right)\mathbf{X}_{j} 𝐗1\mathbf{X}_{1} 𝐗2\mathbf{X}_{2} 𝐗3\mathbf{X}_{3} 𝐗4\mathbf{X}_{4} 𝐗5\mathbf{X}_{5} 𝐗6\mathbf{X}_{6} 𝐗7\mathbf{X}_{7}
𝐗1\mathbf{X}_{1} X1X_{1} X2X_{2} X3X_{3} X4X_{4} X5εX3X_{5}-\varepsilon X_{3} X5+2εX5ε2X3X_{5}+2\varepsilon X_{5}-\varepsilon^{2}X_{3} X7εX1X_{7}-\varepsilon X_{1}
𝐗2\mathbf{X}_{2} X1X_{1} X2X_{2} eεX3e^{-\varepsilon}X_{3} X4X_{4} eεX5e^{-\varepsilon}X_{5} eεX6e^{-\varepsilon}X_{6} X7X_{7}
𝐗3\mathbf{X}_{3} X1X_{1} X2+εX3X_{2}+\varepsilon X_{3} X3X_{3} X4X_{4} X5X_{5} X6X_{6} X7X_{7}
𝐗4\mathbf{X}_{4} X1X_{1} X2X_{2} X3X_{3} X4X_{4} X5X_{5} X62εX3X_{6}-2\varepsilon X_{3} X7X_{7}
𝐗5\mathbf{X}_{5} X1+εX3X_{1}+\varepsilon X_{3} X2+εX5X_{2}+\varepsilon X_{5} X3X_{3} X4X_{4} X5X_{5} X6X_{6} X7+εX5X_{7}+\varepsilon X_{5}
𝐗6\mathbf{X}_{6} X12εX5X_{1}-2\varepsilon X_{5} X2+εX6X_{2}+\varepsilon X_{6} X3X_{3} X4+2εX3X_{4}+2\varepsilon X_{3} X5X_{5} X6X_{6} X7+2εX6X_{7}+2\varepsilon X_{6}
𝐗7\mathbf{X}_{7} eεX1e^{\varepsilon}X_{1} X2X_{2} X3X_{3} X4X_{4} eεX5e^{-\varepsilon}X_{5} e2εX6e^{-2\varepsilon}X_{6} X7X_{7}

5 Model B: Weyl-free collapse

We proceed with our study on the Karmarkar condition by assuming Weyl-free collapse spacetimes, where

A(t,r)\displaystyle A\left(t,r\right) =\displaystyle= (1ω2r2)1/2B(t,r)\displaystyle\left(1-\omega^{2}r^{2}\right)^{1/2}B\left(t,r\right)
Y(t,r)\displaystyle Y\left(t,r\right) =\displaystyle= rB(t,r)\displaystyle rB\left(t,r\right) (42)

that is, the line element (1) becomes

ds2=B2(t,r)((1ω2r2)dt2+dr2+r2dΩ2)ds^{2}=B^{2}\left(t,r\right)\left(-\left(1-\omega^{2}r^{2}\right)dt^{2}+dr^{2}+r^{2}d\Omega^{2}\right) (43)

with ω0\omega\neq 0. Recall that in the limit where ω=0,\omega=0, the latter line element describes the conformally flat spacetime (26). The Karmarkar condition for the above line element is

0\displaystyle 0 =\displaystyle= rω2B2((r2ω2+1)B+r(r2ω21)B′′)\displaystyle r\omega^{2}B^{2}\left(\left(r^{2}\omega^{2}+1\right)B^{\prime}+r\left(r^{2}\omega^{2}-1\right)B^{\prime\prime}\right) (44)
+B(r2B˙((1r2ω2)B˙+2rω2B˙)+r(r2ω21)BB′′\displaystyle+B\bigg{(}r^{2}\dot{B}^{\prime}\left(\left(1-r^{2}\omega^{2}\right)\dot{B}^{\prime}+2r\omega^{2}\dot{B}\right)+r\left(r^{2}\omega^{2}-1\right)B^{\prime}B^{\prime\prime}
+r(r2ω21)(rB′′B)B¨+(1+r2ω2r4ω4)B2)\displaystyle+r\left(r^{2}\omega^{2}-1\right)\left(rB^{\prime\prime}-B^{\prime}\right)\ddot{B}+\left(1+r^{2}\omega^{2}-r^{4}\omega^{4}\right)B^{\prime 2}\bigg{)}
2r(B˙2((r2ω2+1)B+r(r2ω21)B′′)\displaystyle-2r\bigg{(}\dot{B}^{2}\left(\left(r^{2}\omega^{2}+1\right)B^{\prime}+r\left(r^{2}\omega^{2}-1\right)B^{\prime\prime}\right)
2r(r2ω21)BB˙B˙+(r2ω21)(rB¨+B)B2).\displaystyle-2r\left(r^{2}\omega^{2}-1\right)B^{\prime}\dot{B}^{\prime}\dot{B}+\left(r^{2}\omega^{2}-1\right)\left(r\ddot{B}+B^{\prime}\right)B^{\prime 2}\bigg{)}.

This admits a six-dimensional Lie algebra, GBG_{B}, consisted by the Lie symmetry vectors

Y1=t,Y2=BB,Y3=(1+ω2r2)B2B,Y_{1}=\partial_{t}~{},~{}Y_{2}=B\partial_{B}~{},~{}Y_{3}=\left(1+\omega^{2}r^{2}\right)B^{2}\partial_{B}~{},
Y4=e2ωt(ω2r21)B,Y5=e2ωt(ω2r21)B,Y_{4}=e^{-2\omega t}\left(\omega^{2}r^{2}-1\right)\partial_{B}~{},~{}Y_{5}=e^{2\omega t}\left(\omega^{2}r^{2}-1\right)\partial_{B}~{},~{}
Y6=(ω2r21)(2ω2BB(1+ω2r2)rr).Y_{6}=\left(\omega^{2}r^{2}-1\right)\left(2\omega^{2}B\partial_{B}-\frac{\left(1+\omega^{2}r^{2}\right)}{r}\partial_{r}\right).

In Table 3 we present the commutators of the Lie algebra GBG_{B} while in Table 4 the Adjoint representation is given, necessary for the derivation of the one-dimensional optimal system.

Hence, the system of the partial differential equations (31) which provides the relative invariants is simplified as

ϕα3=0,ϕα4=0,ϕα5=0,\frac{\partial\phi}{\partial\alpha_{3}}=0~{}~{},~{}~{}\frac{\partial\phi}{\partial\alpha_{4}}=0~{}~{},~{}~{}\frac{\partial\phi}{\partial\alpha_{5}}=0~{},~{} (45)

which means that ϕ(αI)=ϕ(α1,α2,α6)\phi\left(\alpha_{I}\right)=\phi\left(\alpha_{1},\alpha_{2},\alpha_{6}\right).

Assume the generic symmetry vector W=α1Y1+α2Y2+α3Y3+α4Y4+α5Y5+α6Y6W=\alpha_{1}Y_{1}+\alpha_{2}Y_{2}+\alpha_{3}Y_{3}+\alpha_{4}Y_{4}+\alpha_{5}Y_{5}+\alpha_{6}Y_{6}, then when α1α2α60\alpha_{1}\alpha_{2}\alpha_{6}\neq 0, the equivalent symmetry vector is

W=α1Y1+α2Y2+α6Y6.W^{\prime}=\alpha_{1}^{\prime}Y_{1}+\alpha_{2}^{\prime}Y_{2}+\alpha_{6}^{\prime}Y_{6}. (46)

For α6=0\alpha_{6}=0, the equivalent vector field is W′′=α1′′Y1+α2′′Y2,W^{\prime\prime}=\alpha_{1}^{\prime\prime}Y_{1}+\alpha_{2}^{\prime\prime}Y_{2},~{}for α4=0\alpha_{4}=0, it follows W′′′=α1′′′Y1+α3′′′Y3+α6′′′Y6,W^{\prime\prime\prime}=\alpha_{1}^{\prime\prime\prime}Y_{1}+\alpha_{3}^{\prime\prime\prime}Y_{3}+\alpha_{6}^{\prime\prime\prime}Y_{6},~{}while when α1=0\alpha_{1}=0 we have the equivalent symmetry vector W′′′′=α2′′′′Y2+α6′′′′Y6.~{}W^{\prime\prime\prime\prime}=\alpha_{2}^{\prime\prime\prime\prime}Y_{2}+\alpha_{6}^{\prime\prime\prime\prime}Y_{6}.~{}

Similarly, the following equivalent symmetry vectors are obtained:

{α1=0,α6=0}W¯=α¯2Y2,\{\alpha_{1}=0,~{}\alpha_{6}=0\}~{}\Rightarrow~{}\bar{W}=\bar{\alpha}_{2}Y_{2}~{},
{α1=0,α2=0}W¯=α¯3Y3+α¯6Y6,\{\alpha_{1}=0,~{}\alpha_{2}=0\}~{}\Rightarrow~{}\bar{W}^{\prime}=\bar{\alpha}_{3}^{\prime}Y_{3}+\bar{\alpha}_{6}^{\prime}Y_{6}~{},
{α2=0,α6=0}W¯′′=α¯1′′Y1+α¯3′′Y3,\{\alpha_{2}=0,~{}\alpha_{6}=0\}~{}\Rightarrow~{}\bar{W}^{\prime\prime}=\bar{\alpha}_{1}^{\prime\prime}Y_{1}+\bar{\alpha}_{3}^{\prime\prime}Y_{3}~{},
{α1=0,α2=0,α6=0}W¯′′′=α¯3′′′Y3+α¯4′′′Y4+α¯5′′′Y5.\{\alpha_{1}=0,~{}\alpha_{2}=0,~{}\alpha_{6}=0\}~{}\Rightarrow~{}\bar{W}^{\prime\prime\prime}=\bar{\alpha}_{3}^{\prime\prime\prime}Y_{3}+\bar{\alpha}_{4}^{\prime\prime\prime}Y_{4}+\bar{\alpha}_{5}^{\prime\prime\prime}Y_{5}.

We conclude that the one-dimensional optimal system consisted by the one-dimensional Lie algebras:

{Y1},{Y2},{Y3},{Y4},{Y5},{Y6},\{Y_{1}\}~{},~{}\{Y_{2}\}~{},~{}\{Y_{3}\}~{},~{}\{Y_{4}\}~{},~{}\{Y_{5}\}~{},~{}\{Y_{6}\}~{},
{Y1+αY2},{Y1+αY3},{Y1+αY6},{Y2+αY6},\{Y_{1}+\alpha Y_{2}\},~{}\{Y_{1}+\alpha Y_{3}\},~{}\{Y_{1}+\alpha Y_{6}\},~{}\{Y_{2}+\alpha Y_{6}\},~{}
{Y3+αY4},{Y3+αY5},{Y3+αY6},{Y4+αY5},\{Y_{3}+\alpha Y_{4}\},~{}\{Y_{3}+\alpha Y_{5}\},~{}\{Y_{3}+\alpha Y_{6}\},~{}\{Y_{4}+\alpha Y_{5}\},~{}
{Y1+αY3+αY6},{Y1+αY2+βY6},{Y3+αY4+βY5}.\{Y_{1}+\alpha Y_{3}+\alpha Y_{6}\},~{}\{Y_{1}+\alpha Y_{2}+\beta Y_{6}\},~{}\{Y_{3}+\alpha Y_{4}+\beta Y_{5}\}.

We proceed with the presentation of similarity solutions which are expressed by closed-form functions.

Table 3: Commutators of the admitted Lie point symmetries for the Karmarkar condition for the Weyl-free collapse spacetimes
[,]\left[~{},~{}\right] 𝐘1\mathbf{Y}_{1} 𝐘2\mathbf{Y}_{2} 𝐘3\mathbf{Y}_{3} 𝐘4\mathbf{Y}_{4} 𝐘5\mathbf{Y}_{5} 𝐘6\mathbf{Y}_{6}
𝐘1\mathbf{Y}_{1} 0 0 0 2ωY4-2\omega Y_{4} 2ωY52\omega Y_{5} 0
𝐘2\mathbf{Y}_{2} 0 0 Y3Y_{3} Y4Y_{4} Y5Y_{5} 0
𝐘3\mathbf{Y}_{3} 0 Y3-Y_{3} 0 0 0 0
𝐘4\mathbf{Y}_{4} 2ωY42\omega Y_{4} Y4-Y_{4} 0 0 0 4ω2Y44\omega^{2}Y_{4}
𝐘5\mathbf{Y}_{5} 2ωY5-2\omega Y_{5} Y5-Y_{5} 0 0 0 4ω2Y54\omega^{2}Y_{5}
𝐘6\mathbf{Y}_{6} 0 0 0 4ω2Y4-4\omega^{2}Y_{4} 4ω2Y5-4\omega^{2}Y_{5} 0
Table 4: Adjoint representation for the admitted Lie point symmetries for the Karmarkar condition for the Weyl-free collapse spacetimes
Ad(e(ε𝐘i))𝐘jAd\left(e^{\left(\varepsilon\mathbf{Y}_{i}\right)}\right)\mathbf{Y}_{j} 𝐘1\mathbf{Y}_{1} 𝐘2\mathbf{Y}_{2} 𝐘3\mathbf{Y}_{3} 𝐘4\mathbf{Y}_{4} 𝐘5\mathbf{Y}_{5} 𝐘6\mathbf{Y}_{6}
𝐘1\mathbf{Y}_{1} Y1Y_{1} Y2Y_{2} Y3Y_{3} e2ωεY4e^{2\omega\varepsilon}Y_{4} e2ωεY5e^{-2\omega\varepsilon}Y_{5} Y6Y_{6}
𝐘2\mathbf{Y}_{2} Y1Y_{1} Y2Y_{2} eεY3e^{-\varepsilon}Y_{3} eεY4e^{-\varepsilon}Y_{4} eεY5e^{-\varepsilon}Y_{5} Y6Y_{6}
𝐘3\mathbf{Y}_{3} Y1Y_{1} Y2+εY3Y_{2}+\varepsilon Y_{3} Y3Y_{3} Y4Y_{4} Y5Y_{5} Y6Y_{6}
𝐘4\mathbf{Y}_{4} Y12ωεY4Y_{1}-2\omega\varepsilon Y_{4} Y2+εY4Y_{2}+\varepsilon Y_{4} Y3Y_{3} Y4Y_{4} Y5Y_{5} Y64ω2εY4Y_{6}-4\omega^{2}\varepsilon Y_{4}
𝐘5\mathbf{Y}_{5} Y1+2ωεY5Y_{1}+2\omega\varepsilon Y_{5} Y2+εY5Y_{2}+\varepsilon Y_{5} Y3Y_{3} Y4Y_{4} Y5Y_{5} Y64ω2εY5Y_{6}-4\omega^{2}\varepsilon Y_{5}
𝐘6\mathbf{Y}_{6} Y1Y_{1} Y2Y_{2} Y3Y_{3} e4ω2εY4e^{4\omega^{2}\varepsilon}Y_{4} e4ω2εY5e^{4\omega^{2}\varepsilon}Y_{5} Y6Y_{6}

6 Model C: Shear-free collapse

Consider the shear-free collapse spacetime where Y(t,r)=rB(t,r)Y\left(t,r\right)=rB\left(t,r\right) and A(t,r)=B(t,r)NA\left(t,r\right)=B\left(t,r\right)^{-N} with N1N\neq-1. Hence, the line element (1) reads

ds2=B2(t,r)(B22N(t,r)dt2+dr2+r2dΩ2).ds^{2}=B^{2}\left(t,r\right)\left(-B^{-2-2N}\left(t,r\right)dt^{2}+dr^{2}+r^{2}d\Omega^{2}\right). (47)

The Karmarkar condition then gives

0\displaystyle 0 =\displaystyle= (2n2+3n+1)r2B˙2B2B2n+2n(n+1)r2B4\displaystyle\left(2n^{2}+3n+1\right)r^{2}\dot{B}^{2}B^{\prime 2}B^{2n+2}-n(n+1)r^{2}B^{\prime 4} (48)
2r(B˙2(nrB′′nB)(n1)rBB˙B˙rB¨B2)B2n+3\displaystyle-2r\left(\dot{B}^{2}\left(nrB^{\prime\prime}-nB^{\prime}\right)-(n-1)rB^{\prime}\dot{B}^{\prime}\dot{B}-r\ddot{B}B^{\prime 2}\right)B^{2n+3}
+r(rB˙2+B¨(BrB′′))B2n+4+nB(rB′′B)B2\displaystyle+r\left(r\dot{B}^{\prime 2}+\ddot{B}\left(B^{\prime}-rB^{\prime\prime}\right)\right)B^{2n+4}+nB^{\prime}\left(rB^{\prime\prime}-B^{\prime}\right)B^{2}
2n(n+2)rB3B.\displaystyle-2n(n+2)rB^{\prime 3}B.

The resulting Karmarkar condition admits a three dimensional Lie algebra, GCG_{C}, consisted by the Lie symmetry vectors

Z1=t,Z2=rr1N+1BB,Z3=tt+1N+1BB,Z_{1}=\partial_{t}~{},~{}Z_{2}=r\partial_{r}-\frac{1}{N+1}B\partial_{B}~{},~{}Z_{3}=t\partial_{t}+\frac{1}{N+1}B\partial_{B},~{} (49)

with commutators and Adjoint representation as given in Table 5.

Table 5: Commutators and Adjoint representation of the admitted Lie point symmetries for the Karmarkar condition for the Model C
[,]\left[~{},~{}\right] 𝐙1\mathbf{Z}_{1} 𝐙2\mathbf{Z}_{2} 𝐙3\mathbf{Z}_{3} Ad(e(ε𝐙i))𝐙jAd\left(e^{\left(\varepsilon\mathbf{Z}_{i}\right)}\right)\mathbf{Z}_{j} 𝐙1\mathbf{Z}_{1} 𝐙2\mathbf{Z}_{2} 𝐙3\mathbf{Z}_{3}
𝐙1\mathbf{Z}_{1} 0 0 Z1Z_{1} 𝐙1\mathbf{Z}_{1} Z1Z_{1} Z2Z_{2} Z3εZ1Z_{3}-\varepsilon Z_{1}
𝐙2\mathbf{Z}_{2} 0 0 0 𝐙2\mathbf{Z}_{2} Z1Z_{1} Z2Z_{2} Z3Z_{3}
𝐙3\mathbf{Z}_{3} Z1-Z_{1} 0 0 𝐙3\mathbf{Z}_{3} eεZ1e^{\varepsilon}Z_{1} Z2Z_{2} Z3Z_{3}

Easily we calculate that the relative invariants are α2,α3\alpha_{2},\alpha_{3}, from where we conclude that the one-dimensional optimal system consisted by the one-dimensional Lie algebras

{Z1},{Z2},{Z3},{Z1+αZ2},{Z2+αZ3}\left\{Z_{1}\right\}~{},~{}\left\{Z_{2}\right\}~{}\,,~{}\left\{Z_{3}\right\}~{},~{}\left\{Z_{1}+\alpha Z_{2}\right\}~{},~{}\left\{Z_{2}+\alpha Z_{3}\right\}

7 A radiating, Weyl-free model

By assuming Weyl-free collapse spacetimes where the line element is given by

ds2=B2(t,r)[(1ω2r2)dt2+dr2+r2dΩ2],ds^{2}=B^{2}\left(t,r\right)\left[-\left(1-\omega^{2}r^{2}\right)dt^{2}+dr^{2}+r^{2}d\Omega^{2}\right], (50)

application of the Lie symmetry vector X1+αX6X_{1}+\alpha X_{6} gives the similarity solution

B(t,r)=8αω2ω2r2+1(λ1e4zαω2λ0)1B\left(t,r\right)=-\frac{8\alpha\omega^{2}}{\omega^{2}r^{2}+1}\left(\lambda_{1}e^{-4z\alpha\omega^{2}}-\lambda_{0}\right)^{-1} (51)

where z=14αω2(ln(ω2r21ω2r2+1)+4αω2t)z=-\frac{1}{4\alpha\omega^{2}}\left(\ln\left(\frac{\omega^{2}r^{2}-1}{\omega^{2}r^{2}+1}\right)+4\alpha\omega^{2}t\right).

By setting λ0=0\lambda_{0}=0, we find that the solution satisfies the heat-flux boundary condition,

pr=qB|Σp_{r}=qB|_{\Sigma} (52)

and a simple radiating model may be generated. From the boundary condition constraint, an expression for α\alpha is obtained,

α=32+R2ω2Rω4ω1R2ω2\alpha=\frac{\sqrt{3}\sqrt{2+R^{2}\omega^{2}}-R\omega}{4\omega\sqrt{1-R^{2}\omega^{2}}} (53)

where R=rΣR=r_{\Sigma} is the co-moving boundary. We examine the mass function obtained by calculating,

m=r3BB˙22A2r2Br3B22Bm=\frac{r^{3}B\dot{B}^{2}}{2A^{2}}-r^{2}B^{\prime}-\frac{r^{3}B^{\prime 2}}{2B} (54)

and then search for a function ω(t)\omega(t) such that the mass is constant during the distant past and then decreases monotonically from some point in time as the stellar object undergoes non-adiabatic gravitational collapse.

During this investigation, a definitive form for the function ω(t)\omega(t) was not obtainable and an ad hoc approach was used in establishing an approximate relationship. This involved calculating values for ω(t)\omega(t) for t<tit<t_{i} such that the mass function remained constant. The function

ω(t)=abtc+dte\omega(t)=\frac{a}{bt-c}+\frac{d}{t-e} (55)

was found to be appropriate for approximating the data obtained.
We consider a mass of 3M3M_{\odot} prior to collapse (t<1000)(t<-1000) with a comoving boundary rΣ=3×1010cm/c=1.00sr_{\Sigma}=3\times 10^{10}cm/c=1.00s. The following parameters were then obtained for the function (55) by solving (54) for time tt, obtaining data for ω(t)\omega(t) at early times (t<1000)(t<-1000) and then fitting the data to the function (55):

(a,b,c)\displaystyle(a,b,c) \displaystyle\rightarrow (0.727479,0.191566,1399.22)\displaystyle(-0.727479,0.191566,1399.22)
(d,e)\displaystyle(d,e) \displaystyle\rightarrow (6.25399,136.291)\displaystyle(-6.25399,136.291) (56)

Now that ω(t)\omega(t) has been determined, the gravitational potentials are fully specified and a radiating model, based on the Karmarkar condition with vanishing Weyl stress, has been generated.

8 Discussion

The solutions obtained for the conformally flat metric (Model A) did not offer the possibility for closed systems satisfying the heat flux boundary condition. Pressure isotropy is however easily invoked which can assist in setting parameters. Cosmological models might then be possible, to be explored in future work.
A solution for the Weyl-free collapse metric (Model B) was found to satisfy the heat-flux boundary condition and this was exploited in the previous section. The resulting mass function is shown in Figure 1. The radial profile is typical of uniform density matter and most of the mass is radiated as the collapse proceeds. The energy density, pressure and heat flux are shown in Figures 2 - 4. The energy density is uniform and relatively low for a stellar collapse process. The pressure behaves in a similar manner. We note that the heat flux increases towards the surface boundary as the collapse proceeds, so that little heat is generated near the gravitating centre.

Refer to caption
Figure 1: Mass radial-temporal profile
Refer to caption
Figure 2: Energy density
Refer to caption
Figure 3: Pressure
Refer to caption
Figure 4: Heat flux

The shear-free solution (Model C) is yet to be explored and forms a basis for future work.

9 Conclusion

The use of Lie symmetries in obtaining solutions for conformally flat, Weyl-free and shear-free metrics which in addition, satisfy the Karmarkar condition has been shown. The solutions range from those that are almost trivial to those that are fairly complex. In searching for solutions that are consistent with the heat-flux boundary condition given according to Santos, a solution obtained for the Weyl-free collapse scenario was suitable. Other solutions from the other models did not offer a clear means in maintaining a time-independent heat-flux boundary condition. Some of these solutions may well be suited to cosmological models in which the energy density is related to the fourth power of the temperature. In particular, a solution was found with respect to Weyl-free collapse which allows for a radiating model to be constructed. We see from Figures 1 - 4 that the physical parameters behave in a physically viable manner. The mass decreases monotonically, in a way that is similar to other gravitational collapse models. We note that the mass does not display a simple linear time-dependence as shown in other studies nai2 . The solutions presented here compliment the recent solution obtained for the temporal evolution of a conformally flat interior, matched to a Vaidya exterior. This was obtained using Lie symmetry methods pal .

Appendix A Solutions for Model A

A.1 Static solution X1X_{1}

Consider the application of the Lie point symmetry vector X1X_{1} which leads to the the static solution B(t,r)=B(r).B\left(t,r\right)=B\left(r\right). Equation (27) becomes

BBr(BBr+2r(Br)2rBBrr)=0,BB_{r}\left(BB_{r}+2r\left(B_{r}\right)^{2}-rBB_{rr}\right)=0, (57)

which provides the solutions

B1(r)=λ0,B2(r)=1λ1r2+λ0.B_{1}\left(r\right)=\lambda_{0}~{},~{}~{}B_{2}\left(r\right)=-\frac{1}{\lambda_{1}r^{2}+\lambda_{0}}.

A.2 Similarity solution of X1+αX3X_{1}+\alpha X_{3}

The application of the Lie point symmetry vector X1+αX3X_{1}+\alpha X_{3} gives the exact solution

B(t,r)=(λ1r2+λ0αt)1.B\left(t,r\right)=\left(\lambda_{1}r^{2}+\lambda_{0}-\alpha t\right)^{-1}. (58)

A.3 Similarity solution of X1+αX4X_{1}+\alpha X_{4}

From the Lie point symmetry X1+αX4X_{1}+\alpha X_{4} we find the exact solutions

B(t,r)=(4α2λ1exp(r22αt4α2)+λ0)1.B\left(t,r\right)=\left(4\alpha^{2}\lambda_{1}\exp\left(-\frac{r^{2}-2\alpha t}{4\alpha^{2}}\right)+\lambda_{0}\right)^{-1}. (59)

A.4 Scaling solution X7X_{7}

Reduction with the Lie symmetry vector X7X_{7} leads to the scaling solution

B(t,r)\displaystyle B\left(t,r\right) =\displaystyle= (λ1exp((8u)1)u34×\displaystyle-\bigg{(}\lambda_{1}\int\frac{\exp\left(\left(8u\right)^{-1}\right)}{u^{\frac{3}{4}}}\times (60)
exp(1417u4+14u2+1u3)+λ0)1,\displaystyle\exp\left(-\frac{1}{4}\int\frac{\sqrt{17u^{4}+14u^{2}+1}}{u^{3}}\right)+\lambda_{0}\bigg{)}^{-1},

where u=r/tu=r/t.

A.5 Similarity solution of X1+αX4+βX3X_{1}+\alpha X_{4}+\beta X_{3}

The application of the Lie point symmetry X4+βX2X_{4}+\beta X_{2} provides the similarity solutions

B(t,r)=exp(2t2α+r22α)(λ1exp(2tα)λ2).B\left(t,r\right)=\frac{\exp\left(\frac{\sqrt{2}t}{2\sqrt{\alpha}}+\frac{r^{2}}{2\alpha}\right)}{\sqrt{\left(\lambda_{1}\exp\left(\frac{\sqrt{2}t}{\sqrt{\alpha}}\right)-\lambda_{2}\right)}}. (61)

A.6 Similarity solution of X1+αX5X_{1}+\alpha X_{5}

The exact solution which follows from the application of the symmetry vector X1+αX5X_{1}+\alpha X_{5} is

B(t,r)=2α(r2t2)1.B\left(t,r\right)=\frac{2}{\alpha}\left(r^{2}-t^{2}\right)^{-1}. (62)

A.7 Similarity solution of X1+αX5+βX6X_{1}+\alpha X_{5}+\beta X_{6}

Hence, the exact solution which follows from the application of the symmetry vector X1+αX5+βX6X_{1}+\alpha X_{5}+\beta X_{6} is expressed as follows,

B(t,r)\displaystyle B\left(t,r\right) =\displaystyle= 6(2βt36βr2t3αt2\displaystyle 6\bigg{(}2\beta t^{3}-6\beta r^{2}t-3\alpha t^{2} (63)
+(α24r2β28β2λ1)3212β2+αr22+λ0)1.\displaystyle+\frac{\left(\alpha^{2}-4r^{2}\beta^{2}-8\beta^{2}\lambda_{1}\right)^{\frac{3}{2}}}{12\beta^{2}}+\alpha\frac{r^{2}}{2}+\lambda_{0}\bigg{)}^{-1}.

A.8 Similarity solution of X4+αX5+βX6X_{4}+\alpha X_{5}+\beta X_{6}

From the application of the Lie symmetry vector X4+αX5+βX6X_{4}+\alpha X_{5}+\beta X_{6} it follows,

B(t,r)\displaystyle B\left(t,r\right) =\displaystyle= 6(2βt2r2βr42αtr223αt3\displaystyle 6\bigg{(}2\beta t^{2}r^{2}-\beta r^{4}-2\alpha tr^{2}-\frac{2}{3}\alpha t^{3} (64)
+β3t4+α2βt2+λ1ξ+λ0)1.\displaystyle+\frac{\beta}{3}t^{4}+\frac{\alpha^{2}}{\beta}t^{2}+\lambda_{1}\xi+\lambda_{0}\bigg{)}^{-1}.

Appendix B Solutions for Model B

B.1 Static solution Y1Y_{1}

The similarity solution which follows from the application of the Lie symmetry vector Y1Y_{1} is static and it is expressed as follows

B(t,r)=λ01+ω2r2,or,B(t,r)=1λ1r2+λ0B\left(t,r\right)=\frac{\lambda_{0}}{1+\omega^{2}r^{2}}~{},~{}or,~{}B\left(t,r\right)=\frac{1}{\lambda_{1}r^{2}+\lambda_{0}}

B.2 Similarity solution Y1+αY6Y_{1}+\alpha Y_{6}

Application of the Lie symmetry vector Y1+αY6Y_{1}+\alpha Y_{6} gives the similarity solution

B(t,r)\displaystyle B\left(t,r\right) =\displaystyle= 8αω2ω2r2+1(λ1e4zαω2λ0)1,or\displaystyle-\frac{8\alpha\omega^{2}}{\omega^{2}r^{2}+1}\left(\lambda_{1}e^{-4z\alpha\omega^{2}}-\lambda_{0}\right)^{-1}~{},~{}or
B(t,r)\displaystyle B\left(t,r\right) =\displaystyle= 2(4ω2α+1)ω2r2+1(λ1e(4ω2α+1)2azλ0)1,\displaystyle-\frac{2\left(4\omega^{2}\alpha+1\right)}{\omega^{2}r^{2}+1}\left(\lambda_{1}e^{-\frac{\left(4\omega^{2}\alpha+1\right)}{2a}z}-\lambda_{0}\right)^{-1}, (65)

where z=14αω2(ln(ω2r21ω2r2+1)+4αω2t)z=-\frac{1}{4\alpha\omega^{2}}\left(\ln\left(\frac{\omega^{2}r^{2}-1}{\omega^{2}r^{2}+1}\right)+4\alpha\omega^{2}t\right).

B.3 Similarity solution Y2+αY6Y_{2}+\alpha Y_{6}

Reduction with the Lie symmetry vector Y2+αY6Y_{2}+\alpha Y_{6} provides the similarity solution

B(t,r)=(ω2r21)14αω2(ω2r2+1)14αω21B¯(t),B\left(t,r\right)=\left(\omega^{2}r^{2}-1\right)^{-\frac{1}{4\alpha\omega^{2}}}\left(\omega^{2}r^{2}+1\right)^{\frac{1}{4\alpha\omega^{2}}-1}\bar{B}\left(t\right), (66)

where

B¯(t)\displaystyle\bar{B}\left(t\right) =\displaystyle= (λ1sin(2(2αω21)αt)\displaystyle\bigg{(}\lambda_{1}\sin{\left(\sqrt{\frac{2\left(2\alpha\omega^{2}-1\right)}{\alpha}}t\right)} (67)
λ2cos(2(2αω21)αt))12(2αω21)\displaystyle-\lambda_{2}\cos{\left(\sqrt{\frac{2\left(2\alpha\omega^{2}-1\right)}{\alpha}}t\right)}\bigg{)}^{\frac{1}{2\left(2\alpha\omega^{2}-1\right)}}

B.4 Similarity solution Y1+αY2+βY6Y_{1}+\alpha Y_{2}+\beta Y_{6}

Finally, reduction with respect to the Lie invariants of the vector field Y1+αY2+βY6Y_{1}+\alpha Y_{2}+\beta Y_{6} gives the similarity solution

B(t,r)=λ0(ω2r2+1)α4βω21(ω2r21)α4βω2e(α4βω2)ζB\left(t,r\right)=\lambda_{0}\frac{\left(\omega^{2}r^{2}+1\right)^{\frac{\alpha}{4\beta\omega^{2}}-1}}{\left(\omega^{2}r^{2}-1\right)^{\frac{\alpha}{4\beta\omega^{2}}}}e^{-\left(\alpha-4\beta\omega^{2}\right)\zeta} (68)

where ζ=14βω2(ln(ω2r21ω2r2+1)+4βω2t)\zeta=\frac{1}{4\beta\omega^{2}}\left(\ln\left(\frac{\omega^{2}r^{2}-1}{\omega^{2}r^{2}+1}\right)+4\beta\omega^{2}t\right).

Appendix C Solutions for Model C

C.1 Similarity solutions Z2Z_{2}

Reduction with respect to the symmetry vector leads to the similarity solution

B(t,r)=r1N+1(λ1(tt0))2N+14N2+3N+1.B\left(t,r\right)=r^{-\frac{1}{N+1}}\left(\lambda_{1}\left(t-t_{0}\right)\right)^{\frac{2N+1}{4N^{2}+3N+1}}. (69)

Acknowledgements

RB and MG acknowledge support from the office of the Deputy Vice-Chancellor for Research and Innovation at the Durban University of Technology.

Data Availability Statement

This manuscript has no associated data or the data will not be deposited. (Authors’ comment: All data was obtained using the formulae explicitly given in the article.)

References

  • (1) J. R. Oppenheimer, H. Snyder, Phys. Rev. 56, 455 (1939)
  • (2) P. C. Vaidya, Proc. Indian Acad. Sci. A 33, 264 (1951)
  • (3) W. B. Bonnor, A. K. G. de Oliveira, N. O. Santos, Phys. Rep. 181, 269 (1989)
  • (4) J. M. Z. Pretel, M. F. A. da Silva, Mon. Not. R. Astron. Soc. 495, 5027 (2020)
  • (5) R. S. Bogadi, M. Govender, S. Moyo, Eur. Phys. J. C 81, 922 (2021)
  • (6) E. N. Glass, Phys. Lett. A 867, 351 (1981)
  • (7) N. O. Santos, Mon. Not. R. Astron. Soc. 216, 403 (1985)
  • (8) N. F. Naidu, M. Govender, K. S. Govinder, Int. J. Mod. Phys. D 15, 1053 (2006)
  • (9) S. Thirukkanesh, S.S. Rajah, S.D. Maharaj, J. Math. Phys. 53, 032506 (2012)
  • (10) S. D. Maharaj, M. Govender, Pramana J. Phys. 54, 715 (2000)
  • (11) S. Thirukkanesh, S. Moopanar, M. Govender, Pramana J. Phys. 79, 223 (2012)
  • (12) L. Herrera, G. Le Denmat, N. O. Santos, Phys. Rev. D 79, 087505 (2009)
  • (13) L. Herrera, A. Di Prisco, N. O. Santos, Gen. Relativ. Grav. 42, 1585 (2010)
  • (14) L. Herrera, G. Le Denmat, N.O. Santos, Gen. Relativ. Grav. 44, 1143 (2012)
  • (15) S. D. Maharaj, M. Govender, Int. J. Mod. Phys. D 14, 667 (2005)
  • (16) M. Govender, R. S. Bogadi, D. B. Lortan, S. D. Maharaj, Int. J. Mod. Phys. D 25, 1650037 (2016)
  • (17) K. P. Reddy, M. Govender, W. Govender, S. D. Maharaj, Annals Phys. 429, 168458 (2021)
  • (18) A. M. Msomi, K. S. Govinder, S. D. Maharaj, Int. J. Theor. Phys. 51, 1290 (2012)
  • (19) G. Z. Abebe, K. S. Govinder, S. D. Maharaj, Int. J. Theor. Phys. 52, 3244 (2013)
  • (20) G. Z. Abebe, S. D. Maharaj, K. S. Govinder, Gen. Relativ. Grav. 46, 1650 (2014)
  • (21) S. K. Maurya, Y. K. Gupta, S. Ray, D. Deb, Eur. Phys. J. C 76, 693 (2016)
  • (22) K. R. Karmarkar, Proc. Indian Acad. Sci. A 27, 56 (1948)
  • (23) S. N. Pandey, S. P. Sharma, Gen. Relativ. Gravit. 14, 113 (1981)
  • (24) P. Bhar, Ksh. Newton Singh, F. Rahaman, N. Pant, S. Banerjee, Int. J. Mod. Phys. D. 26, 1750078 (2017)
  • (25) S. K. Maurya, S. D. Maharaj, Eur. Phys. J. C 77, 328 (2017)
  • (26) G. Z. Abebe, S. D. Maharaj, K. S. Govinder, Eur. Phys. J. C 75, 496 (2015)
  • (27) G. Z. Abebe, S. D. Maharaj, Eur. Phys. J. C 79, 849 (2019)
  • (28) K. S. Govinder, M. Govender, Gen. Relativ. Gravit. 44, 174 (2012)
  • (29) A. Paliathanasis and M. Tsamparlis, IJGMMP 11, 1450037 (2014)
  • (30) T. Christodoulakis, N. Dimakis, P.A. Terzis, G. Doulis, Th. Grammenos, E. Melas and A. Spanou, J. Geom. Phys. 71, 127 (2013)
  • (31) B. Muatjetjeja and C.M. Khalique, Acta Mathematica Scientia 32, 1959 (2012)
  • (32) E.D. Fackerell and D. Hartley, Gen. Relativ. Gravit. 32, 857 (2000)
  • (33) A. Paliathanasis, Mod. Phys. Lett. A 37, 2250119 (2022)
  • (34) M. Tsamparlis and A. Paliathanasis, Symmetry 10, 233 (2018)
  • (35) A. Janda, Acta Physica Polonica B 38, 3961 (2007)
  • (36) H. Stephani, Differential Equations: Their Solutions Using Symmetry, Cambridge University Press, New York, (1989)
  • (37) G. W. Bluman and S. Kumei, Symmetries of Differential Equations, Springer-Verlag, New York, (1989)
  • (38) N. H. Ibragimov, CRC Handbook of Lie Group Analysis of Differential Equations, Volume I: Symmetries, Exact Solutions, and Conservation Laws, CRS Press LLC, Florida (2000)
  • (39) P. J. Olver, Applications of Lie Groups to Differential Equations, Springer-Verlag, New York, (1993)
  • (40) N. F. Naidu, M. Govender, S. D. Maharaj, Eur. Phys. J. C 78, 48 (2018)
  • (41) A. Paliathanasis, M. Govender, G. Leon, Eur. Phys. J. C 81, 718 (2021)