SciELO - Scientific Electronic Library Online

 
vol.68 número4Non-perturbative field theoretical aspects of graphene and related systemsA dynamical dark energy solution to hubble tension in the light of the multimessenger era índice de autoresíndice de materiabúsqueda de artículos
Home Pagelista alfabética de revistas  

Servicios Personalizados

Revista

Articulo

Indicadores

Links relacionados

  • No hay artículos similaresSimilares en SciELO

Compartir


Revista mexicana de física

versión impresa ISSN 0035-001X

Rev. mex. fis. vol.68 no.4 México jul./ago. 2022  Epub 19-Mayo-2023

https://doi.org/10.31349/revmexfis.68.040701 

Research

Cosmological-static metric correspondence and Kruskal type solutions from symmetry transformations

J. A. Nieto1 

E. A. León1 

C. García-Quintero2 

1Facultad de Ciencias Físico-Matemáticas de la Universidad Autonóma de Sinaloa, 80010, Culiacán, Sinaloa, México.

2Department of Physics, The University of Texas at Dallas, Richardson, TX 75080, USA.


Abstract

We develop a formalism which provides a new view for the transformation of spherically symmetric metrics, regarding cosmological and Kruskal type metrics. Our analysis begins with some general relevant dynamically metrics in cosmology, and prove that they all can be transformed to a unique static form. We extend the formalism to obtain generalized Kruskal type coordinates in cosmology and black hole theory. This extended formalism provides a novel mechanism to obtain suitable coordinate charts associated with spherically symmetric metrics. In particular, we obtain explicitly new Kruskal type coordinates for extremal Reissner-Nordström and Schwarzschild-de-Sitter metrics, as well for an extension of the de-Sitter metric.

Keywords: Cosmology; black holes; symmetries

1. Introduction

The most illustrative analytical solutions to the field equations in general relativity, are those where the space has spherical symmetry. For instance, the Schwarzschild metric for black holes and the Friedmann-Lemaître-Robertson-Walker (FLRW) metric, that describes the behavior of the Universe at cosmological distances, have spherical symmetry (see Refs.1,2 and also3 for generalized models with extra dimensions). However, it is well known that the description of a non-trivial space-time cannot be complete with just one coordinate chart, and usually one has to consider several patches, where in each patch a distinct set of coordinates is valid2,4. In view of this, it can be useful to define new coordinates that cover more parts of the manifold. Then one sees that a transformation of space-time coordinates can fulfill two purposes: to reveal explicit symmetries of the space-time, as well as to extend the description of the space-time to regions that cannot be considered in the original setup. This allows, in cosmological and black hole models, to extend the description beyond the event horizons appearing in both cases5,6. In this article, we explore the relations between two possible forms of the metric: one where the coordinates associated with the space-time appears as dynamical, and other where the coordinates takes a static form.

The relation between dynamical type and static forms of the metric has been a topic of great interest in the literature7-12. For instance, the well known association of de-Sitter space with the FLRW metric has been used in a deeper analysis in general relativity7,8. However, interpretational problems between a static and a non-static representation of the same underlying space have been subject of debate13. In this work, one of the main ideas is to find, at a level of coordinate transformations, a link between spherically symmetric spaces (relevant to black hole theory) and cosmology. But more generally, we develop a general formalism based in coordinate transformation, that establishes a correspondence between static/non-static metrics. In particular, starting from time-dependent metrics we find the corresponding static metrics which turn out to be unique solutions.

We argue that our method can be extended to obtain Kruskal type coordinates in a number of scenarios in black-hole physics and cosmology. Specifically, assuming a general metric in the form

dS2=-f(r)dt2+dr2f(r)+r2dΩ2,

we transform it to coordinates where the metric takes a conformally flat form, at hypersurfaces with 2 = 0. We show that this leads to different possible mappings, including the Kruskal type transformations. From there, we discuss the resulting transformation for several spherically symmetric metrics, such as Schwarzschild, Reissner-Nordström, extremal Reissner-Norsdtröm, de-Sitter and Schwarzschild-de-Sitter. A relevant aspect of our approach is that we obtain three novel Kruskal transformations that highlights interesting features of Reissner-Norsdtröm and Schwarzschild-de-Sitter spaces, as well as a type of space described by a Generalized-de-Sitter metric. For all cases, the appropriate selection of integration constants assures two things: first, that the coordinates singularities can be removed; and second, that the different regions -for instance, interior and exterior of a black hole- can be distinguished in the Kruskal representation. We argue that this may shed some light on the underlying symmetries of a more general Kruskal formalism.

The rest of this work is divided as follows: in Sec. 2 we deal with the FLRW metric transformed to a ‘static’ type metric. We show the way this leads to a Friedmann equation with cosmological constant and zero matter density. The result is that for all spherical, hyperbolic and plane geometries, all converge to the same de-Sitter type metric. For sake of completeness, we also solve for the scale parameter α = α(T). In Sec. 3 we review a further generalization and find that the previous result of a Friedmann equation for vacuum is unique, as well as the general form for the FLRW metric. In Sec. 4 we show the way this procedure can be applied in general to spherically symmetric metrics; we find that there exists several possibilities for the solutions. One of this possibilities leads to Kruskal type coordinates, and in Sec. 5 we review some particular solutions for different static metrics. Finally, in Sec. 6 we make some final remarks.

2. From FLRW cosmology to static metrics

Consider the gravitational field equations with cosmological constant Λ:

Rμν-12gμν(R-2Λ)=8πGTμν. (1)

By assuming that the space is maximally symmetric with commoving coordinates (T, ρ, θ, Φ) describing a spherically symmetric space, one can solve (1) for the metric in the form

dS(1)2=-dT2+a2(T)(dρ21-kρ2+ρ2dΩ2), (2)

where 2 = d θ2 + sin2θ dΦ2 is the solid angle line element. Also, k can take the values 1, 0 or -1 denoting space-like slices at constant T corresponding to spherically, flat and hyperbolic topologies, respectively. Furthermore α(T) is the scale factor, whose evolution is obtained by assuming that the energy-momentum tensor takes the form

Tμν=(ρf+pf)uμuν+pfgμν. (3)

Here uμ is the four velocity, while ρƒ and ρƒ are the density of energy and pressure describing a perfect fluid. From there, one obtains the Friedmann equations

a˙2+ka2=8πG3ρf+Λ3, (4)

and

a¨a=-4πG3(ρf+3pf)+Λ3. (5)

Now consider the transformation from dS(1)2=gμνdxμdxν given in (2), to the static spherical symmetric form dS(2)2=γαβdx'αdx'β, namely

dS(2)2=-f(r)dt2+dr2f(r)+r2dΩ2. (6)

The solid angle is the same for both cases in such a way that the angular terms in (2) and (6) imply

r=aρ. (7)

From now on, we will denote partial derivatives respect to T with an overdot, while prime will mean partial derivatives respect to ρ, such as t˙=t/T and t’ = ∂t / ∂ρ. We start with the tensor transformation

gμν=yαxμyβxνγαβ. (8)

For g00, g11 and g01, after rearranging some terms, this leads to

t˙2=1f2(f+ρ2a˙2), (9)

t'2=a2f2(1-f1-kρ2), (10)

and

f4t˙2t'2=ρ2a2a˙2, (11)

respectively. Substituting (9) and (10) into (11) one obtains

f=1-ρ2(a˙2+k), (12)

where ƒ is considered to be a function of ρ and T.

By inserting (12) into (9), one obtains

t˙2=1-kρ2f2, (13)

while combining (10) and (12) one finds that

t'2=ρ2a2a˙2f2(1-kρ2). (14)

Now we use the fact that f'=-2ρ(α˙2+k) and f˙=-2ρ2a˙a¨ to take the partial derivatives of (13) respect to ρ and of (14) respect to T, in order to obtain ∂2 t/∂T∂ρ in both cases. After equating and performing some simplifications, the next relation appears:

&[(a˙2+k)-aa¨&][1+ρ2(a˙2-k)]=0. (15)

In general, the second factor is nonzero, since then it would imply that a is a function of ρ and this is incongruent with α = α(T); equivalently, the second factor equal to zero would imply, by (16), that ƒ can be put as a function of ρ only. It follows that only

a¨a=a˙2+ka2, (16)

holds. By noticing that d(a˙2+k)/dt=2a˙a¨, one can see that this equation is equivalent to

d(a˙2+k)dt=2a˙a(a˙2+k). (17)

As this can be expressed as a total derivative of logarithms, this leads to

a˙2+ka2=Γ, (18)

where Γ is an integration constant. Then, by (16) we have also the relation ä = Γa.

Observe that (18) reduces to the first Friedmann equation (4) for vacuum (ρƒ = 0) with cosmological constant Λ = 3Γ. This identification is validated by the comparison of ä = Γa with the second Friedmann equation in vacuum, namely ä/a = Λ / 3 in (5). It is interesting to note that the Friedmann equation emerges from a symmetry transformation, without invoking any dynamic equation such as the Einstein field equations.

Going back to (18), rewriting it as a˙=Γa2-k, we can solve for k = 0, 1 and -1. For the moment we take into account the cases Γ ≠ 0. It turns out that the equality would describe Minkowski space, as we shall see below.

For k = 0(Γ > 0), the result is a = e √Γt . For the closed topology where k = 1 (here also Γ > 0 is forced), a/(T) becomes a=(1/√Γ)cosh√Γ, where we have chosen T = 0 as the comoving time when a = 1/√Γ. Meanwhile, with k = -1, Γ can be either positive or negative. For Γ > 0, the solution is a = (a / √Γ)sinh /√ΓT. In this case, we have chosen the origin of time in such a way that a = 0 when T = 0. For k = -1 and Γ > 0, the solution corresponds to a = (1 / √|Γ|) sin√|Γ|T.

Concerning the function ƒ, we remember from (7) that a = r/ρ, that together with (12) and (17) imply that

f=1-Γr2. (19)

Summarizing this section, for Γ ≠ 0 we have the following solutions:

Curvature Metric
k = 0, Γ > 0 dS2=-dT2+e2ΓΤdρ2+ρ2dΩ2
k = 1, Γ > 0 dS2=-dT2+cos2ΓΤΓdρ21-ρ2+ρ2dΩ2
k = -1, Γ > 0 dS2=-dT2+sinh2ΓΤΓdρ21+ρ2+ρ2dΩ2
k = -1, Γ > 0 dS2=-dT2+sin2ΓΤΓdρ21+ρ2+ρ2dΩ2

The first solution is the usual de-Sitter space, while the second and third ones are the two types of Lanczos universe. The fourth solution is the only allowed solution with Γ > 0, and it corresponds to anti-de-Sitter space .

Finally, choosing Γ = 0 in (18) implies that the scale parameter obeys the equation a˙2+k=0, and the corresponding solutions for g11 in (1) are g11 = 1 for k = 0 and g11 = t2 / (1 + r2) for k = -1; this last solution is referred as Milne model. Both get mapped (in a trivial way) to Minkowski space. These two solutions, as well as the ones listed in (20), corresponds to the six possible transformations to the static form given in (2) (see Refs.12,13). The whole set of solutions share the corresponding static form of the metric

dS(2)2=-(1-Γr2)dt2+dr21-Γr2+r2dΩ2. (21)

As we shall see in the next section, this static form will be preserved even when generalizing the line element given in (2).

3. A further generalization.

Now, let us consider a more general form of the metric, but still assuming commoving time and radial symmetry. In this sense, the ansatz now reads as

dS(3)2=-dT2+a2(T)(b2(ρ)dρ2+ρ2dΩ2). (22)

If this metric is transformed to (6), then the relation (7), r = , is satisfied again. Even more, transformations (9) and (11) hold again. However, instead of (10) we have

t'2=a2f2(1-b2f). (23)

Substituting (9) and (23) into (11) leads after simplification to

f=1b2-ρ2a˙2. (24)

Insertion in (9) and (23) leads to the succinct expressions = 1/bƒ and t’ = ρaȧ b/ƒ. As in the previous section, we derive this relations with respect to ρ and with respect to T, respectively. By using ƒ = -2ρȧä and ƒ’ = -2(b’b -3 + ρȧ 2), and equating ∂ṫ/∂T with ∂t’/∂T, we see that after some algebra the next relation holds:

b'=ρb3(aa¨-a˙2). (25)

Since a = a(T) and b = b(ρ), (25) implies that

aa¨-a˙2=b'ρb3=κ, (26)

where κ is a constant. Note that the values of κ can be identified with those of κ (1, 0 or -1), by rescaling adequately the parameter a(T). With this identification, the previous relation for a(T) is just equation (16)1,2. Furthermore, integration of b’b -3 = κρ leads to

1b2=B-κρ2, (27)

with B another integration constant. Assuming local flatness at slices with T constant, B can be set equal to 1. Hence, even by considering a more general metric in our formalism, namely dS(3)2 in (22), a transformation to the static form given by (6) restricts the metric to the form given by (2), and with same solutions listed at the end of 2.

4. Extending the formalism to include Kruskal type coordinates

It turns out that the same formalism can lead to Kruskal type coordinates. Let us assume that the metric can take a form that is conformally flat in the space-time slices with dθ = dΦ = 0:

dS(4)2=N2(T,ρ)(-dT2+dρ2)+φ2(T,ρ)dΩ2. (28)

The transformation (8) of the metric components given in (28) to the static form (6), dS(2)2=-fdt2+f-1dr2+r2dΩ2, leads to

(tT)2=(tρ)2+fN-2 (29)

(rT)2=(rρ)2-f-1N-2, (30)

and

tTrT=tρrρ. (31)

Substitution of (29) and (30) into (31) gives, after clearing, the next relation:

N-2=(rρ)2f-(tρ)2f-1. (32)

This allows to simplify (29) and (30) as

tT=frρ (33)

and

rT=f-1tρ, (34)

respectively. The form of these two expressions suggests to take T(t, r) = θ(t) Φ(r) and ρ(t, r) = ξ = (t) Χ(r). Then those two relations leads to

1ξdΘdt=fΦdχdr=α, (35)

and

1Θdξdt=fχdΦdr=β, (36)

where α and β are constants. Both expressions imply various relations. First, from (35) we derive dθ/dt = αξ with respect to t and use (36). It results in

d2Θdt2-αβΘ=0. (37)

On the other hand, dividing (36) by (35) we have that αΦdΦ = βΧdΧ, giving the function relation

αΦ2-βχ2=σ, (38)

with σ another integration constant. There are several relevant possibilities for the product αβ in (37):

Case αβ = 0. Assuming that α = 0, then (35) implies that Θ is a constant, in such a way that ∂tT= 0; this in turn leads to ∂rρ = 0 due to (33), and also that Χ is constant. By (36), rescaling and shifting the origin of time, ξ can be set equal to t. Then, setting Θ = 1 and Χ = β-1, from the same relation (36), it results dT = dΦ = ƒ-1dr, while = dt. From (30) and (34) one learns that N2 = -ƒ/(∂tρ)2 = -ƒ. Inserting all this into the form of the metric (28) we obtain the same metric given in (6). A similar argument holds for the case β = 0. Thus, with αβ = 0 the transformation maps onto itself and N2 is proportional to ƒ, a reminiscence of what occurs with the use of tortoise coordinates, where dr* = (1 - rs/r)-1dr in such a way that -(1 - rs/r)dt2 + (1 - rs/r)-1dr2 transforms into -(1 - rs/r)(dt2 + dr*2), where rs is the Schwarzschild radius1,2.

Case αβ<0[/p]

By (37) we have that Θ ∝ sin(√|αβ|t + Φ0) and consequently -by (36)- that ξ ∝ cos(√|αβ|t + Φ0). We can fix the phase angle to zero in such a way that T = 0 coincides with t = 0. Also, without loss of generality we take α > 0 and ß < 0, that implies that σ > 0 in Eq. (38). Then the solutions are given by

Θ(t)=B1sin(-αβt), (39)

and

ξ(t)=B2cos(-αβt), (40)

where the relation B2=-β/αB1 holds in such a way that dΘ/dt=αξ in (36), is still satisfied.

From (36) we have that χ=(f/β)dΦ/dr, that together with (38) leads to

dΦdr=-βσ-αΦ2f, (41)

which can be integrated, yielding

Φ=σαsin(-αβdrf). (42)

Inserting this result in (38), we have that

χ=σ-βcos(-αβdrf). (43)

Remembering that T(t, r) = Θ(t)Φ(r) and ρ(t, r) = ξ(t)Χ(r), in this case we have

T(t,r)=sin(-αβdrf)sin(-αβt), (44)

and

ρ(t,r)=cos(-αβdrf)cos(-αβt). (45)

Here we have set B1 √σ/α = 1 (justified by rescaling coordinates). By using (45) in (32) for different ƒ in the metric, the factor N 2 appearing in (30) can be obtained. The result is

N2=f-αβ[cos(2-αβt)-cos(2-αβdrf)]. (46)

Case αß > 0

Equation (37) leads to Θ(t) = C1sinh(√αßt + Φ0). In order to get t = 0 when T = 0, we choose Φ0 = 0 and then we have the solutions

Θ(t)=C1sinh(αβt), (47)

and

ξ(t)=C2cosh(αβt), (48)

where C 2 = √ß/αC 1. Clearly, this two relations lead to αξ2 - ßΘ2 = ßC1 2, obtained also by using (35) and (36).

Now take into account that (38) implies χ=±α/βΦ, with σ = 0. Inserting this in (35), we obtain

1ΦdΦdr=±αβf-1. (49)

By integrating this expression, we see that

Φ(r)=Ae±αβdrf, (50)

and consequently (38) implies that

χ(r)=±Aαβe±αβdrf, (51)

with A constant. Thus, for αß > 0, the coordinates T(t, r) = Θ(t)Φ(r) and ρ(t, r) = ξ(t)Χ(r) are

T(t,r)=Φ(r)sinh(αβt), (52)

and

ρ(t,r)=Φ(r)cosh(αβt). (53)

Here, Φ(r) is given in (50) and we set A = C 1 = 1. Also, we have omitted a possible minus sign in ρ(t, r), since it just plays the role of an inversion of coordinates in the analysis of the regions considered.

The function N(T, ρ) can be obtained by inserting (33), (52) and (53) in (32). The result is:

N2=fΦ-2αβ. (54)

Observe that the function ƒ(r) determines all possible transformations, and the relations (52) and (53) determine Kruskal type coordinates for a given ƒ. In the next section we obtain the explicit form for several cases of interest.

5. Kruskal type solutions

For simplicity, we define γ = ±√αβ and proceed to obtain the Kruskal type solution for different cases, by changing ƒ in dS(2)2=-fdt2+f-1dr2+r2dΩ2.

5.1 Schwarzschild

In this case we have ƒ = 1 - rs /r, where rs = 2M is the Schwarzschild radius. This leads to dr/1-[rs/r]=r+rslnr/rs-1, for r > rs. Thus, substituting this result into (50), yields

ΦSchw=er2rsrrs-1, (55)

where we have set γ = 1/(2rs). Then (52) and (53) become

TSchw=er2rsrrs-1sinht2rs, (56)

and

ρSchw=er2rsrrs-1cosht2rs, (57)

respectively. Further, from (28) and (54) we find that the the metric is given by

dSSchw2=4rs3re-rrs(-dT2+dρ2)+r2dΩ2. (58)

We recognize in (56)-(58) the Kruskal transformation associated with the Schwarzschild metric14,15. As usual, the relation ρSchw2-TSchw2=er/rsr/rs-1 is useful to verify the properties of this space-time, and in particular to extend the analysis to the region r > rs (See1,2 and also16 and references therein for recent developments).

Moreover, it is worth mentioning that our formalism is more direct and general than the usually given in textbooks, since we just need to specify ƒ and then solve for Φ in order to obtain the full set of coordinate transformations for T, ρ and N(T ρ).

5.2. Reissner-Nordström

For the electric charged static black hole we have f=1-rs/r+Q2/r2=(r-r+)(r-r-)/r2, where r±=M±M2-Q2. The solution to f-1dr (again region with r>r +) is given by

drf=r+12M2-Q2[r+2ln(r-r+)-r-2ln(r-r-)]+const. (59)

By setting the integration constant to

r-2lnr--r+2lnr+2M2-Q2,

and following the same steps as in the Schwarzschild case, the Reissner-Nordström metric is transformed into the Kruskal form by means of

ΦR-N(r)=eγrrr+-1ζr+2rr--1ζr-2,

where ζ=γr+-r-. The corresponding coordinate transformations acquire the form

TR-N(t,r)=eγr(rr+-1)ζr+2(rr--1)ζr-2sinhγt,ρR-N(t,r)=eγr(rr+-1)ζr+2(rr--1)ζr-2coshγt. (60)

Then the Kruskal type solution for the charged static black hole is

dSR-N2=1γ2r2e-2γrrr+-12ζr-2+1rr--12ζr+2-1(-dT2+dρ2)+r2dΩ2. (61)

Thus, the different regions of this space can be visualized from

ρR-N2-TR-N2=e2γrrr+-12ζr+2rr--12ζr-2.

(For comparison set γ=r+-r-/r+2 and see for instance Refs.17 and 18).

5.3 Reissner-Nordström (Extremal)

The Reissner-Nordström extremal metric is of theoretical interest in several contexts. This solution is obtained from r±=M±M2-Q2 when M 2 = Q 2, which means that r ± = M. In such a case the metric takes the form dS2=-1-M/r2dt2+1-M/r-2dr2+r2dΩ2. So, our task is to determine the integral f-1dr=r2(r-M)-2dr. We obtain f-1dr=r-M2(r-M)+2Mln(r-M)+const and hence one may write the radial function Φ as ΦEx(r)=r/r+-1exp{r/r+-φ1r/r+-φ2/2r/r+-1}, where φ1 = (1 + √5)/2 and φ2 = (1 + √5)/2 . Also, we have set γ = (4M)-1 and the integration constant equal to -2MlnM. With this at hand, the coordinates (in the patch where r > r +) are given by

TExt,r=err+-φ1rr+-φ24rr+-1rr+-1sinht4r+,

ρEx(t,r)=e(rr+-φ1)(rr+-φ2)4(rr+-1)rr+-1cosht4r+ (62)

Now the metric would be specified by using N2=fΦ-2γ-2 , that yields

dSEx2=16r+4r2rr+-1e-rr+-φ1rr+-φ2rr+-1(-dT2+dρ2)+r2dΩ2. (63)

Also,

ρEx2-TEx2=(r/r+-1)exp(r/r+-φ1)(r/r+-φ2)&#091;2(r/r+-1)&#093;.

It is interesting to observe that the golden ratios φ1 and φ2 emerge in this extremal case (see Refs.19,20 and references therein).

5.4 De-Sitter space

In this case, for (17) we have f=1-Γr2, where Γ > 0 leads to de-Sitter space. Now we have f-1dr=Γ-1/2arctanh(Γr)+const. This leads to the radial function ΦdS(r)=(1-Γr)/(1+Γr)=(1-Γr)/1-Γr2 for r<1/Γ. We have chosen γ = -√Γ and the integration constant equal to zero. The Kruskal coordinates are then

TdS(t,r)=1-Γr1-Γr2sinh(Γt),ρdS(t,r)=1-Γr1-Γr2cosh(Γt). (64)

Thus, in this case the metric takes the simple form

dSdS2=1Γ(1-Γr)2(-dT2+dρ2)+r2dΩ2. (65)

The coordinates (64) yield ρdS2-TdS2=1-Γr/1+Γr. Note that the negative sign chosen in γ = -√Γ makes sense: with this selection the region 0 ≤ r ≤ 1/√Γ would be described, in the (T, p) system, by hyperbolas starting from ρdS2-TdS2=1 until reaching the asymptotes T = ±p. Observe that this is valid in the two quadrants where ρ > 0. On the other hand, γ > 0 would result in unbounded hyperbolas, since then ρdS2-TdS2 when r → 1/√Γ. Compare this solution with Refs.21-23.

5.5 Schwarzschild-de-Sitter metric

As we mentioned before, the solution for black hole with cosmological constant comes from taking the radial function as f=1-rs/r-Γr2. The three roots of the cubic equation r-rs-Γr3=0, namely (λ1, λ2, λ3), are real and distinct if M<1/33Γ (see 24 and references therein). In this case we set λ1 > λ2 > λ3, in such a way that the first root corresponds to the cosmological horizon r = λ1, the second root to the black hole event horizon r = λ2, and the third one is a negative (unphysical) root. It is worthwhile to mention that the two horizons at λ1 and λ2 coincide when M=33Γ-1. Then f-1dr=lnr-λ1σ1r-λ2σ2r-λ3σ3+const., where σi=λi/(1-3Γλi2), where i=1,2,3. By (50) we have

ΦS-dS(r)=(rλ1-1)γσ1(rλ2-1)γσ2(1-rλ3)γσ3. (66)

Here, we have chosen the integration constant equal to -σ1lnλ1-σ2lnλ2-σ3ln(-λ3). From (52) and (53), it results

TS-dS(t,r)=(rλ1-1)γσ1(rλ2-1)γσ2(1-rλ3)γσ3sinhγt,ρS-dS(t,r)=(rλ1-1)γσ1(rλ2-1)γσ2(1-rλ3)γσ3coshγt, (67)

where we have taken into account that λ3 < 0.

From (28) and (54) we find

dSS-dS2=Γλ1λ2λ3γ2rrλ1-11-2γσ1rλ2-11-γσ21-rλ31-2γσ3(-dT2+dρ2)+r2(T,ρ)dΩ2. (68)

5.6 A generalized de-Sitter metric

In Ref.25 a generalization of de-Sitter space is considered, where the metric is of the type (6), with ƒ = 1 - h 2 r 2 + q 4 r 4. Here two cosmological horizons arise, given by r±2=(h2±h4-4q4)/(2q4). In order to calculate Φ(r), we perform the integral

drf=1q4(r+2-r-2)1r+tanh-1rr+-1r-tan-1rr-. (69)

By choosing γ=-q4(r+2-r-2)r+, we have from (50) that

ΦGdS(r)=r+-rr++re-r+r-tan-1rr-, (70)

for the region with r < r -, where “GdS” stands for Generalized-de-Sitter space. The corresponding Kruskal type coordinates are given by

TGdS(t,r)=r+-rr++re-r+r-tan-1(rr-)sinh(γt),ρGdS(t,r)=r+-rr++re-r+r-tan-1(rr-)cosh(γt). (71)

Also, the relation

ρGdS2-TGdS2=r+-rr++re-2r+r-tan-1rr-

holds, confirming that the qualitative behavior of the space in these coordinates is very similar to that of the de-Sitter space analyzed before, in the patch where r < r -.

6. Final remarks

In this work we have analyzed the relationship between some spherical symmetric metrics for two cases: cosmological FLRW and Kruskal type metrics. In the first case, we have shown that, by imposing that the FLRW metric to be transformed into the form

dS2=-f(r)dt2+dr2f(r)+r2dΩ2, (72)

leads to ƒ = 1 - Γr 2 with solutions summarized at the end of 2. In fact, the only possibilities resulting from this symmetry transformation are the spaces known as: de-Sitter, anti-de-Sitter, Lanczos, Milne, and Minkowski. It is remarkable that as a by-product of the symmetry transformation, the Friedmann equation with cosmological constant emerges.

Next we moved in the reverse order: starting from the general metric (72), we applied a transformation to obtain a metric which is conformally flat in hypersurfaces with dθ = dΦ = 0. This led to two non-trivial possibilities, one in which the coordinates are proportional to sines and cosines; and a second solution in terms of hyperbolic trigonometric functions, that resembles the Kruskal solution for Schwarzschild space.

In 5 we used the method to explicitly obtain the coordinate (T, ρ) for several well known spaces: Schwarzschild, Reissner-Nordström, extremal Reissner-Norsdtröm, de-Sitter and Schwarzschild-de-Sitter. Here, the analysis was not exhaustive, in the sense that the main purpose was to show how the method of two-metric transformation correctly works. For instance, we only solved for the exterior regions in the case of black holes, and for the region inside the cosmological horizon in the de-Sitter case. Meanwhile, we found that choosing properly the integration constant γ in Φ(r)=Aexpγ[dr/f], given in (50), one preserves desirable properties of the Kruskal extended space. Specifically, in the extremal Reissner-Norsdström case we have chosen γ in such a way that the interior and the exterior regions of the black hole get uniquely represented in the Kruskal (T, ρ)-coordinates framework. Furthermore, in the de-Sitter case we favored a negative sign in γ, since this ensures that in the limit r → 1/√ Γ, the hyperbolas representing the space do not open to infinity. It is worthwhile to remark that we have obtained three novel Kruskal transformations: for the Reissner-Norsdtröm (where interestingly the golden ratio arises), Schwarzschild-de-Sitter space and the mentioned as Generalized-de-Sitter space.

Further interesting prospects of our work could emerge as follows. Clearly, one can find other Kruskal type solutions by considering other functions ƒ(r) in the metric (2) (for instance the space analyzed in Ref.26). In this sense, the formulation of Secs. 4 and 5 complements other works that consider the properties of static metrics in which g 11 = -1/g 00.

Another interesting aspect is that the formulation of this two-metric transformation can be generalized in a straightforward manner to higher dimensions. First, notice that the angular term r2 dΩ2 is a passive term in all the development. Then one may readily generalize it to higher dimensions. Of course, in this case one must modify also the function ƒ(r). For instance, in the Schwarzschild type metric in arbitrary D-dimensions one has

f=1-rsrD-3. (73)

There is at least one possible scenario in which such a generalization may have important and interesting consequences, namely black holes associated with parallelizable spheres. As it is known, the only parallelizable spheres are S1, S3, and S7, which corresponds to the existence of normed division algebras: real numbers, complex numbers, quaternions and octonions, respectively31-23. In this way, from the point of view of parallelizable spheres, the event horizon of black holes associated with the spheres S1, S3, and S7 seems even more interesting that the traditional S2-event/horizon.

Also, for further research it may be interesting to consider the connection between the transformations corresponding to negative and positive values of αβ in 4. Since for αβ < 0 the coordinate transformations are related to the trigonometric functions sine and cosine, while for αβ < 0 corresponds to hyperbolic trigonometric functions, one may expect a connection between these two scenarios. This may be analogue to the transformation between spheres and hyperbolas. In fact, one finds such example in complex variable, where the mapping ƒ = b(a2 z-2 - 1) transforms the complex variable z’ = u + iv to z = x + iy connecting the circumference u2 + v2 = b2 with the hyperbola x2 - y2 = a2. Moreover, we argue that this transformation admits a conformal mapping interpretation.

Acknowledgments

JAN and EAL recognize that this work was partially supported by PROFAPI-UAS, 2015. JAN would like to thank CUCEI in Universidad de Guadalajara, for hospitality during a stage of this work. CGQ acknowledge a Ph.D. fellowship by CONACyT.

References

1. C. W. Misner, K.S. Thorne, and J.A. Wheeler. Gravitation (W. H. Freeman and Company, San Francisco CA, 1970) pp. 616-617. [ Links ]

2. S. M. Carroll, Spacetime and geometry: An introduction to general relativity (Addison-Wesley, San Francisco CA, 2004) pp. 329-336. [ Links ]

3. E. A. León, J. A. Nieto, R. Nunez-Lopez, and A. Lipovka, Higher Dimensional Cosmology: Relations among the radii of two homogeneous spaces, Mod. Phys. Lett. A 26 (2011) 805, https://doi.org/10.1142/S0217732311035316. [ Links ]

4. B. Zink, E. Schnetter, and M. Tiglio, Multipatch methods in general relativistic astrophysics: Hydrodynamical flows on fixed backgrounds, Phys. Rev. D 77 (2008) 103015, https://doi.org/10.1103/PhysRevD.77.103015. [ Links ]

5. J. T. Giblin, D. Marolf, and R. Garvey, Spacetime Embedding Diagrams for Spherically Symmetric Black Holes, Gen. Rel. Grav. 36 (2004) 83, https://doi.org/10.1023/B:GERG.0000006695.17232.2e. [ Links ]

6. F. Melia, The apparent (gravitational) horizon in cosmology, Am. J. Phys. 86 (2018) 585, https://doi.org/10.1119/1.5045333. [ Links ]

7. R. C. Tolman, On the Astronomical Implications of the de Sitter Line Element for the Universe, Astrophys. J. 69 (1929) 245, https://doi.org/10.1086/143184. [ Links ]

8. C. Lanczos, Ein vereinfachendes Koordinatensystem für Einsteinschen Gravitationsgleichungen, Phys. Zeit. 23 (1922) 539. [ Links ]

9. R. L. Agacy, and W. H. McCrea, A Transformation of the De Sitter Metric and the law of Creation of Matter, MNRAS 123 (1962) 383, https://doi.org/10.1093/mnras/123.4.383. [ Links ]

10. M. G. J. van der Burg, Transformations Leaving the De Sitter Line-Element Invariant, MNRAS 131 (1966) 499, https://doi.org/10.1093/mnras/131.4.499. [ Links ]

11. P. S. Florides, The Robertson-Walker metrics expressible in static form, Gen. Rel. Grav. 12 (1980) 563, https://doi.org/10.1007/BF00756530. [ Links ]

12. A. Mitra, When can an “Expanding Universe” look “Static” and vice versa: A comprehensive study, Int. J. Mod. Phys D 24 (2015) 1550032, https://doi.org/10.1142/S0218271815500327. [ Links ]

13. F. Melia, Cosmological redshift in Friedmann-Robertson-Walker metrics with constant space-time curvature, MNRAS 422 (2012) 1418, https://doi.org/10.1111/j.1365-2966.2012.20714.x. [ Links ]

14. M. D. Kruskal, Maximal Extension of Schwarzschild Metric, Phys. Rev. 119 (1960) 1743, https://doi.org/10.1103/PhysRev.119.1743. [ Links ]

15. G. Szekeres, On the Singularities of a Riemannian Manifold, Publ. Math. Debrecen 7 (1960) 285, https://doi.org/10.1023/A:1020744914721. [ Links ]

16. C. Aviles-Niebla, P. A. Nieto-Marin, and J. A. Nieto, Towards exterior/interior correspondence of black-holes, Int. J. Geom. Meth. Mod. Phys. 17 (2020) 2050180, https://doi.org/10.1142/S0219887820501807. [ Links ]

17. J. C. Graves, and D. R. Brill, Oscillatory Character of Reissner-Nordström Metric for an Ideal Charged Wormhole, Phys. Rev. 120 (1960) 1507, https://doi.org/10.1103/PhysRev.120.1507. [ Links ]

18. J.-Q. Guo, and P. S. Joshi, Interior dynamics of neutral and charged black holes, Phys. Rev. D 92 (2015) 064013, https://doi.org/10.1103/PhysRevD.92.064013. [ Links ]

19. J. A. Nieto, E. A. Leon, and V. M. Villanueva, Higher-Dimensional Charged Black Holes as Constrained Systems, Int. J. Mod. Phys. D 22 (2013) 1350047, https://doi.org/10.1142/S0218271813500478. [ Links ]

20. N. Cruz, M. Olivares, and J. R. Villanueva, The golden ratio in Schwarzschild-Kottler black holes, Eur. Phys. J. C 77 (2017) 123, https://doi.org/10.1140/epjc/s10052-017-4670-7. [ Links ]

21. G. W. Gibbons, and S. W. Hawking, Cosmological event horizons, thermodynamics, and particle creation, Phys. Rev. D 15 (1977) 2738, https://doi.org/10.1103/PhysRevD.15.2738. [ Links ]

22. J. F. Beck, and A. Inomata, Kruskal-like representation of the de Sitter metric, J. Math. Phys. 25 (1984) 3039, https://doi.org/10.1063/1.526018. [ Links ]

23. W. Rindler, Analogies between Kruskal space and de Sitter space, Found. Phys. 15 (1985) 545. https://doi.org/10.1007/BF01882481. [ Links ]

24. S. Akcay, and R. Matzner, The Kerr-de Sitter universe, Class. Quant. Grav. 28 (2011) 085012, https://doi.org/10.1088/0264-9381/28/8/085012. [ Links ]

25. P. F. González-Díaz, Generalized de Sitter space, Phys. Rev. D 61 (2000) 024019, https://doi.org/10.1103/PhysRevD.61.024019. [ Links ]

26. J. Podolsky, The Structure of the Extreme Schwarzschild-de Sitter Space-time, Gen. Rel. Grav. 31 (1999) 1703, https://doi.org/10.1023/A:1026762116655. [ Links ]

27. M. Varadarajan, Kruskal coordinates as canonical variables for Schwarzschild black holes, Phys. Rev. D 63 (2001) 084007, https://doi.org/10.1103/PhysRevD.63.084007. [ Links ]

28. T. Jacobson, When is g tt g rr = −1?, Class. Quant. Grav. 24 (2007) 5717, https://doi.org/10.1088/0264-9381/24/22/N02. [ Links ]

29. K. A. Bronnikov, E. Elizalde, S. D. Odintsov, and O. B. Zaslavskii, Horizons versus singularities in spherically symmetric space-times, Phys. Rev. D 78 (2008) 064049, https://doi.org/10.1103/PhysRevD.78.064049. [ Links ]

30. K. Lake, Some notes on the Kruskal-Szekeres completion, Class. Quant. Grav. 27 (2010) 097001, https://doi.org/10.1088/0264-9381/27/9/097001. [ Links ]

31. F. Canfora, A. Giacomini, and R. Troncoso, Black holes, parallelizable horizons, and half-BPS states for the Einstein-Gauss-Bonnet theory in five dimensions, Phys. Rev. D 77 (2008) 024002, https://doi.org/10.1103/PhysRevD.77.024002. [ Links ]

32. J. A. Nieto, and L. N. Alejo-Armenta, Hurwitz theorem and parallelizable spheres from tensor analysis, Int. J. Mod. Phys. A 16 (2001) 4207, https://doi.org/10.1142/S0217751X01005213. [ Links ]

Appendix

A. Coordinate transformations from metric with scale parameter to its static form

For sake of completeness, we derive the explicit relation between the systems (t, r) and (T, ρ) listed in (16). First, we combine (8) and (10) to obtain

tρ=ρaa˙1-kρ2[1-ρ2(a˙2+k)]. (A.1)

Thus, the idea is to integrate (A.1) for the different cases.

Case k = 0

Considering a = e √Γ [see (16) for k = 0], (A.1) yields t/ρ=/ρ-2Γ-1ln(1-Γe2ΓTρ2). Hence, we can write

t=-12Γln(1-Γe2ΓTρ2)+g(T). (A.2)

Note that inserting (8) and (19) in (9) yields t/T=1/1-Γe2ΓTρ2. Comparing with ∂t/∂T obtained from (A.2), results in ġ = 1, and then

t=T-12Γln1-Γe2ΓTρ2 (A.3)

where we have chosen g(T) = T. This result, with e √Γρ, leads to10,13

T=12Γln(1-Γr2)+t, (A.4)

and

ρ=r1-Γr2e-Γt. (A.5)

Case k = 1

Now we have that a = cosh (√ΓT)/√Γ. Then (A.1) implies

tρ=ρsinh(ΓT)cosh(ΓT)Γ1-ρ2[1-ρ2cosh2(ΓT)]. (A.6)

Integration of this expression yields

t=1Γtanh-1[1-ρ2tanh(ΓT)]+g(T). (A.7)

Now, by using a and ƒ in (9) gives t/T=1-ρ2/1-ρ2cosh2(ΓT). Comparing with ∂t/∂T from (A.7) sets ġ = 0. By choosing g = 0, (A.7) is

tanh(Γt)tanh(ΓT)=1-ρ2. (A.8)

Meanwhile, as a = cosh (√ΓT)/√Γ, from (3) we have

r=ρcosh(ΓT)Γ. (A.9)

Note that (A.8) must be used for tanh(ΓT)>1-ρ2. But if tanh(ΓT)<1-ρ2, one needs to use the transformation given in11,13.

Case k = 1, Γ > 0

In this case a = sinh(√ΓT)/√Γ, and (A.1) leads to

tρ=ρsinh(ΓT)cosh(ΓT)Γ1+ρ2[1-ρ2sinh2(ΓT)]. (A.10)

Following the same steps as before, this implies

t=1Γtanh-1[1+ρ2tanh(ΓT)]+g(T). (A.11)

Again, we derive with respect to T, that sets ġ = 0 and then we obtain

tanh(Γt)=1+ρ2tanh(ΓT), (A.12)

where we chose g = 0. Furthermore, we also have that

r=ρsinh(ΓT)Γ.(A.13)

Here, we have used the scale parameter and (3). We note again that (A.12) is valid whenever tanhΓT>1/1+ρ2, while tanhΓT<1/1+ρ2 would lead to the transformation tanhΓttanhΓT=1/1+ρ2.

Case k = 1, Γ > 0

Here, a=sin(ΓT)/Γ. Then (A.1) yields

tρ=ρsin(|Γ|T)cos(|Γ|T)|Γ|1+ρ2[1+ρ2sin2(|Γ|T)]. (A.14)

We can integrate the previous equation respect to ρ to obtain

t=1|Γ|tan-1[1+ρ2tan(|Γ|T)]+g(T). (A.15)

Following the same steps as before, the coordinate transformation is

tan(Γt)=1+ρ2tan(ΓT). (A.16)

Additionally, in this case the coordinate r is

r=ρsin(|Γ|T)|Γ|, (A.17)

which completes all possible cases.

Received: December 16, 2021; Accepted: January 07, 2022

Creative Commons License This is an open-access article distributed under the terms of the Creative Commons Attribution License