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

Regular frames for spherically symmetric black holes revisited

A.  V. Toporensky Sternberg Astronomical Institute, Lomonosov Moscow State University Kazan Federal University, Kremlevskaya 18, Kazan 420008, Russia [email protected]    O. B. Zaslavskii Department of physics and Technology, Kharkov V.N. Karazin National University, 4 Svoboda Square, Kharkov 61022, Ukraine
Kazan Federal University, Kremlevskaya 18, Kazan 420008, Russia
[email protected]
Abstract

We consider a space-time of a spherically symmetric black hole with one simple horizon. As a standard coordinate frame fails in its vicinity, this requires continuation across the horizon and constructing frames which are regular there. Up to now, several standard frames of such a kind are known. It was shown in literature before, how some of them can be united in one picture as different limits of a general scheme. However, some types of frames (the Kruskal-Szekeres and Lemaître ones) and transformations to them from the original one remained completely disjoint. We show that the Kruskal-Szekeres and Lemaître frames stem from the same root. Overall, our approach in some sense completes the procedure and gives the most general scheme. We relate the parameter of transformation e0e_{0} to the specific energy of fiducial observers and show that in the limit e00e_{0}\rightarrow 0 a homogeneous metric under the horizon can be obtained by a smooth limiting transition.

frame, black  hole, coordinate transformations
pacs:
04.70.Bw, 97.60.Lf

I Introduction

The Schwarzschild black hole schw is the core object of general relativity, the properties of its space-time play a crucial role in understanding space-time of black holes. The standard coordinate system in which the Schwarzschild metric is written uses so-called curvature (Schwarzschildean) coordinates and fails on the event horizon.  To repair this drawback, there are several ”standard” transformations and corresponding coordinate systems - such as Eddington-Filkestein coordinates (EF), Kruskal-Szekeres ones, Novikov systems, Gullstrand-Painlevé (GP) and Lemaître coordinates. All these coordinates and methods of their constructions look very different. Meanwhile, it turned out that all of these transformations (or at least their part) can be united, if one introduces an additional parameter in the coordinate transformation. This parameter e0e_{0} has the meaning of the energy per unit mass for a reference (fiducial) observer. In this sense, particle dynamics is encoded in the typical transformations to the regular frame. Then, taking the limit e0e_{0}\rightarrow\infty, one can recover some familiar coordinate systems and metrics mart , finch , jose . In this sense, previous metrics are contained as different limiting cases of some more general one.

The approach developed in mart , jose does not include the transformation to the synchronous frame. Meanwhile, this frame simplifies the whole picture and thus plays an important role. The construction of such a frame for the Schwarzschild metric was done by Lemaître lem . Quite recently, this was generalized to metrics other than the Schwarzschild one. To this end, two different procedures were suggested, bron , 3 .

In spite of the fact that some frames were combined in a single general construction, the whole picture remains quite intricate and even the ways of particular unifications also look very different. Unifying particular approaches and metrics, we can separate the whole set of possible transformations to two kinds. The first one (A) envolves the parameter e0e_{0} having the meaning of energy per unit mass of fiducial observers. This includes such systems in which fiducial observers (characterized by a constant value of a spatial coordinate) move not freely. The bright example is the Kruskal-Szekeres system. The second class (B) contains the transition to synchronous systems. For example, this concerns the Novikov system nov63 or more general Bronnikov - Dymnikova - Galactionov (BDG) system bron . In appearance, classes A and B look completely separated, derived from different requirements and seem to be complementary to each other. However, we show that, as a matter of fact, there is deep and very simple connection between both classes. We also consider the limit quite different from mart , finch , jose where it was implied that e0>1e_{0}>1. This is the limit e00e_{0}\rightarrow 0. Then, another synchronous system typical of the Kantwoski-Sachs (KS) metric komp , ks appears explicitly.

The most general regular frame is the one suggested by Fomin fom . Wrongly, this paper was not noticed in due course and remained poorly cited. We use the approach of Fomin to show that all other ones can be obtained from it.

Our consideration applies to a class of metrics more general than the Schwarzschild one. It includes the Schwarzschild-de Sitter, Reissner-Nordström metrics, etc. Further generalization is straightforward. We use the geometric system of units in which fundamental constants G=c=1G=c=1.

II Generalized Gullstrand-Painlevé frame

Let us consider the metric

ds2=fdt2+dr2f+r2dω2,ds^{2}=-fdt^{2}+\frac{dr^{2}}{f}+r^{2}d\omega^{2}\text{,} (1)

where dω2=dθ2+dϕ2sin2θd\omega^{2}=d\theta^{2}+d\phi^{2}\sin^{2}\theta. It represents the spherically symmetric solution of the Einstein equations, provided the components of the stress-energy tensor obey the relation Trr=T00T_{r}^{r}=T_{0}^{0}. Let r=r+r=r_{+} correspond to the event horizon, so f(r+)=0.f(r_{+})=0. The original frame fails in the vicinity of the horizon. To repair the situation, one can introduce a new time variable

dt~=e0dt+e0drfVe0dt+P0drf,d\tilde{t}=e_{0}dt+e_{0}\frac{dr}{f}V\equiv e_{0}dt+P_{0}\frac{dr}{f}\text{,} (2)

where initially we consider e0e_{0} as a positive function of coordinates rr and tt,

P0=e0V=e02f,P_{0}=e_{0}V=\sqrt{e_{0}^{2}-f}, (3)
V1fe02.V\equiv\sqrt{1-\frac{f}{e_{0}^{2}}}\text{.} (4)

After substitution in (1), one obtains

ds2=fe02dt~2+2dt~dre0V+dr2e02+r2dω2.ds^{2}=-\frac{f}{e_{0}^{2}}d\tilde{t}^{2}+\frac{2d\tilde{t}dr}{e_{0}}V+\frac{dr^{2}}{e_{0}^{2}}+r^{2}d\omega^{2}\text{.} (5)

It can be also written in the form

ds2=dt~2+1e02(dr+Ve0dt~)2+r2dω2.ds^{2}=-d\tilde{t}^{2}+\frac{1}{e_{0}^{2}}(dr+Ve_{0}d\tilde{t})^{2}+r^{2}d\omega^{2}\text{.} (6)

Now,

gt~t~=1grt~=Ve0grr=f.g^{\tilde{t}\tilde{t}}=-1\text{, }g^{r\tilde{t}}=Ve_{0}\text{, }g^{rr}=f. (7)

As dt~d\tilde{t} should be a total differential, the integrability conditions have to be fulfilled:

(e0),r=e0(e0),tP0f.\left(e_{0}\right)_{,r}=\frac{e_{0}\left(e_{0}\right)_{,t}}{P_{0}f}\text{.} (8)

Eq. (5) corresponds to eq. (3.19) of finch . Let e0=conste_{0}=const. If e0=1e_{0}=1, we arrive at the Gullstrand-Painlevé frame gul , p , generalized to an arbitrary ff 3 . If f=12mrf=1-\frac{2m}{r} in (5), we return to the Schwarzschild version of the GP system lake , mart , jose .

For the GP system the cross term with dt~drd\tilde{t}dr defines a coordinate flow velocity VV which has a direct physical meaning - it is the 3-velocity of a free falling particle with the unit energy with respect to the static coordinate system (see below). From (5) we can see that this direct correspondence is lost since for e01e_{0}\neq 1, the additional factor (1/e0)(1/e_{0}) appears. The reason is, however, rather clear – since the intervals of physical distance dldl in t~=const\tilde{t}=const sections are connected with the interval of coordinate distance drdr via dl=dr/e0dl=dr/e_{0}, the physical 3-velocity of the generalized GP system with respect to the stationary system is still equal to VV, as it should be. So, we still can think of the coordinate system with e0=conste_{0}=const as realized by free falling particles with the energy e0e_{0}.

This can be confirmed by direct simple calculations. We can choose tetrads attached to an static observer. Then, in coordinates (t,r,θ,ϕ)(t,r,\theta,\phi)

h(0)μ=f(1,0,0,0),h_{(0)\mu}=\sqrt{f}(-1,0,0,0), (9)
h(1)μ=(0,1f,0,0).h_{\left(1\right)\mu}=(0,\frac{1}{\sqrt{f}},0,0). (10)

Let us introduce the three-velocity in a standard way 72

V(i)=h(i)μuμh(0)μuμ,V^{(i)}=-\frac{h_{(i)\mu}u^{\mu}}{h_{(0)\mu}u^{\mu}}\text{,} (11)

where uμu^{\mu} is the four-velocity. Eq. (11) is valid in general. Now, we will apply it to motion of a particle moving freely with a constant specific energy e0e_{0}. For pure radial motion,

dtdτ=e0f,\frac{dt}{d\tau}=\frac{e_{0}}{f}\text{,} (12)
drdτ=P0,\frac{dr}{d\tau}=-P_{0}\text{,} (13)

where τ\tau is the proper time along the trajectory, P0P_{0} is given by (3). Then, uμ=(e0f,P0,0,0)u^{\mu}=(\frac{e_{0}}{f},-P_{0},0,0). By substitution into (11), we obtain that

V(1)=V.V^{(1)}=V\text{.} (14)

Thus VV has the meaning of a velocity measured by a static observer, P0=e0VP_{0}=e_{0}V being the corresponding momentum. Then, e0e_{0} can be thought of as the energy of some effective particle moving in the given background,

e0=f1V2P0=Vf1V2.e_{0}=\frac{\sqrt{f}}{\sqrt{1-V^{2}}}\text{, }P_{0}=\frac{V\sqrt{f}}{\sqrt{1-V^{2}}}\text{.} (15)

If a particle moves not freely, e0e_{0} ceases to be an integral of motion and depends on time. However, as stressed in Sec. 3 of our , even in this case equations of motion retain their validity with e0=e0(t)e_{0}=e_{0}(t). Moreover, we can admit formally dependence of spatial coordinates as well in e0e_{0} that enters transformation (2). Although e0e_{0} is not an integral of motion in this case, formulas (15) show that this can be still considered as a pure local Lorentz transformation. Below we will see how this helps in constructing regular frames.

The case of e0=1e_{0}=1 is a special one since spatial sections dt~=0d\tilde{t}=0 are flat. In this case spatial intervals are simply differences r2r1r_{2}-r_{1}. When e01e_{0}\neq 1, the factor 1/e021/e_{0}^{2} before dr2dr^{2} makes them non-flat. For e01e_{0}\geq 1, the proper distance between points with fixed r2r_{2} and r1r_{1} measured along the hypersurface t~=const\tilde{t}=const is less than in the case when e0=1e_{0}=1. This is some reminiscent of the Lorentz contraction in special relativity (SR). Indeed, in SR (f=1(f=1) the proper distance along the hypersurface t=constt=const (e0=1(e_{0}=1, V=0)V=0) is equal to Δr=r2r1\Delta r=r_{2}-r_{1}. If another observer passes by him with velocity VV and the specific energy e0=11V2>1e_{0}=\frac{1}{\sqrt{1-V^{2}}}>1, the proper distance measured in its own frame (t~=const)(\tilde{t}=const) equals Δl=Δre0=Δr1V2<Δr\Delta l=\frac{\Delta r}{e_{0}}=\Delta r\sqrt{1-V^{2}}<\Delta r.  However, if e0<1e_{0}<1, the corresponding situation has no analogue in SR since along the surface t~=const\tilde{t}=const the proper distance Δl>Δr\Delta l>\Delta r. This is due to the fact that the space-time is curved since in the flat one such observers are absent. Meanwhile, for any fixed rr there is a lower bound on possible e0e_{0} which is equal to f(r)f(r). Therefore, among all states with different e0e_{0} and fixed dr,dr, the maximum value of the proper distance dl=dr/e0dl=dr/e_{0} is achieved for a minimum value of e0=f(r)e_{0}=\sqrt{f(r)} that coincides with the distance in the static frame. This is natural since an observer with minimal possible e0e_{0} for a given rr has zero flow velocity (4) and thus coincides with a stationary observer. In this sense, there is some analogy with SR again since the minimum of the proper distance is achieved in the rest frame.

If we rescale time according to

t~=t^e0\tilde{t}=\hat{t}e_{0} (16)

with e0=const>0e_{0}=const>0,

ds2=fdt^2+2dt^drV+dr2e02+r2dω2.ds^{2}=-fd\hat{t}^{2}+2d\hat{t}drV+\frac{dr^{2}}{e_{0}^{2}}+r^{2}d\omega^{2}. (17)

It follows from (2), (16) that

t^=tdrfV=t𝑑rV,\hat{t}=t-\int\frac{dr}{f}V=t-\int dr^{\ast}V\text{,} (18)

where

r=drfr^{\ast}=\int\frac{dr}{f} (19)

is the tortoise coordinate.

If f=12Mrf=1-\frac{2M}{r} and we write e0=1p2e_{0}=\frac{1}{p^{2}}, we return to the coordinates of Ref. mart - see eq. (3.5) there. It is worth noting that eq. (5) is valid even if e0e_{0} depends on both coordinates. Meanwhile, transformation from (5) to (17) implies that e0=conste_{0}=const.

If one takes the limit e0e_{0}\rightarrow\infty in (17), the metric in the EF coordinates is reproduced:

ds2=fdt^2+2dt^dr+r2dω2.ds^{2}=-fd\hat{t}^{2}+2d\hat{t}dr+r^{2}d\omega^{2}\text{.} (20)

In doing so, V1V\rightarrow 1,

t^=tr=u\hat{t}=t-r^{\ast}=u (21)

is the EF coordinate.

It is worth noting that the transformation (2) somewhat generalizes that in jose . However, in contrast to jose , we do not use the double GP coordinates and obtain the limit e0e_{0}\rightarrow\infty directly from (5) after rescaling the time coordinate t~t^\tilde{t}\rightarrow\hat{t}. If one starts with t^\hat{t} from the very beginning, the limiting transition from mart is reproduced easily with their TT equal to t^\hat{t}. Meanwhile, as the transformation used in mart does not include the parameter e0e_{0}, it is relatively restricted in that it is unable to describe the diversity of different approaches.

III Diagonal metric

We can consider transformation that can be interpreted as modification of the approach developed by Fomin fom . For completeness, we present here the main features of this approach, though in contrast to the original paper, we use our notations with the parameters e0e_{0} and P0P_{0}.

Let in new coordinates TT, ρ\rho the metric be regular and diagonal,

ds2=F(T,ρ)dT2+G(T,ρ)dρ2+r2(ρ,T)dω2.ds^{2}=-F(T,\rho)dT^{2}+G(T,\rho)d\rho^{2}+r^{2}(\rho,T)d\omega^{2}\text{.} (22)

We perform transformations according to which

dt=e0fFdTGP0fdρ,dt=\frac{e_{0}}{f}\sqrt{F}dT-\frac{\sqrt{G}P_{0}}{f}d\rho, (23)
dr=e0GdρP0FdT,dr=e_{0}\sqrt{G}d\rho-P_{0}\sqrt{F}dT\text{,} (24)

where P0P_{0} is given by eq. (3).

The inverse transformation reads

dT=1F(dte0+drfP0),dT=\frac{1}{\sqrt{F}}(dte_{0}+\frac{dr}{f}P_{0})\text{,} (25)
dρ=1G(dtP0+dre0f).d\rho=\frac{1}{\sqrt{G}}\left(dtP_{0}+\frac{dre_{0}}{f}\right). (26)

It is easy to check that in new coordinates the metric is indeed diagonal. If we put in (25) F=1F=1 and, instead of ρ\rho, will use our previous coordinate rr, this would correspond to the transformation (2) that leads to the GP metric (5).

It follows from (23) - (26) that

r,Tr,ρGF=V,\frac{r_{,T}}{r_{,\rho}}\sqrt{\frac{G}{F}}=-V\text{,} (27)
F=t,T2f(1V2)=r,T2(1V2)fV2,F=t_{,T}^{2}f(1-V^{2})=\frac{r_{,T}^{2}(1-V^{2})}{fV^{2}}\text{,} (28)
G=t,ρ2f(1V2)V2=r,ρ2(1V2)f.G=\frac{t_{,\rho}^{2}f(1-V^{2})}{V^{2}}=\frac{r_{,\rho}^{2}(1-V^{2})}{f}\text{.} (29)

Eqs. (27), (28) correspond to eqs. (9), (10) of fom .

To ensure that the left hand side of (23) and (24) is the total differential, the integrability conditions should be satisfied:

(e0fF),ρ=(GP0f),T\left(\frac{e_{0}}{f}\sqrt{F}\right)_{,\rho}=-\left(\frac{\sqrt{G}P_{0}}{f}\right)_{,T} (30)
(e0G),T=(P0F),ρ\left(e_{0}\sqrt{G}\right)_{,T}=-\left(P_{0}\sqrt{F}\right)_{,\rho} (31)

Here,

f,ρ=e0Gf(r),f_{,\rho}=e_{0}\sqrt{G}f^{\prime}(r)\text{,} (32)
f,T=P0Ff(r)f_{,T}=-P_{0}\sqrt{F}f^{\prime}(r) (33)

After substitution into (30) we get

FGf(r)+(e0F),ρ+(P0G),T=0-\sqrt{FG}f^{\prime}(r)+\left(e_{0}\sqrt{F}\right)_{,\rho}+\left(P_{0}\sqrt{G}\right)_{,T}=0 (34)

If F=1F=1, (25) coincides with (2). In general, e0=e0(ρe_{0}=e_{0}(\rho, T)T). Then, it cannot be interpreted as a conserved energy, although one can define the quantity VV formally according to (15).

If we assume that e0e_{0} is finite (at least, finite near the horizon), it follows from (15) the universal behavior of V:V:

V=1fe021f2e021κe02(rr+),V=\sqrt{1-\frac{f}{e_{0}^{2}}}\approx 1-\frac{f}{2e_{0}^{2}}\approx 1-\frac{\kappa}{e_{0}^{2}}(r-r_{+})\text{,} (35)

where we took into account that

f2κ(rr+),f\approx 2\kappa(r-r_{+}), (36)

κ\kappa is the surface gravity that agrees with eqs. (21), (22) of fom .

It is instructive to analyze the example suggested by Fomin for the Schwarzschild metric, when V=tanht2r+V=\tanh\frac{t}{2r\,+}. In this case,

e0=1r+rcosht2r+.e_{0}=\sqrt{1-\frac{r_{+}}{r}}\cosh\frac{t}{2r_{+}}\text{.} (37)

We see that for a fixed r>r+r>r_{+}, limte0=\lim_{t\rightarrow\infty}e_{0}=\infty. If, instead, we fix tt, then limrr+e0=0\lim_{r\rightarrow r_{+}}e_{0}=0. For our purposes, it is important that e0e_{0} remain finite and nonzero for an observer falling in a black hole. Then, we consider tt and rr as related by equations of motion. For a geodesic observer with some ee,

drdt=fee2f,\frac{dr}{dt}=\frac{f}{e}\sqrt{e^{2}-f}\text{,} (38)

whence near the horizon of the Schwarzschild black hole

tr+ln(rr+r+).\frac{t}{r_{+}}\approx\ln(\frac{r-r_{+}}{r_{+}})\text{.} (39)

As a result,

e01.e_{0}\approx 1\text{.} (40)

Then, the transformations (23), (24) acquire the meaning of the local Lorentz transformations and are equivalent to eqs. (7) of fom . It follows from (23), (24) that

tr=P0fe0,\frac{t^{\prime}}{r^{\prime}}=-\frac{P_{0}}{fe_{0}}\text{,} (41)

where prime denotes derivative with respect to ρ\rho. Eq. (41) corresponds to eq. (17) of bron . It can be also rewritten in the form

tr=Vf.\frac{t^{\prime}}{r^{\prime}}=-\frac{V}{f}\text{.} (42)

IV Synchronous system and relation to BDG

Now, we will consider a particular case when e0e_{0} does not depend on TT. Then, it follows from (34) with (30) with (32), (33) that

de0dρF+e0(F),ρ12FGf(r)+P0(G),T=0.\frac{de_{0}}{d\rho}\sqrt{F}+e_{0}\left(\sqrt{F}\right)_{,\rho}-\frac{1}{2}\sqrt{FG}f^{\prime}(r)+P_{0}\left(\sqrt{G}\right)_{,T}=0. (43)

And, (31) gives us

e0(G),T+P0(F),ρ+1P0Fe0de0dρ12e0P0FGf(r)=0.e_{0}\left(\sqrt{G}\right)_{,T}+P_{0}\left(\sqrt{F}\right)_{,\rho}+\frac{1}{P_{0}}\sqrt{F}e_{0}\frac{de_{0}}{d\rho}-\frac{1}{2}\frac{e_{0}}{P_{0}}\sqrt{FG}f^{\prime}(r)=0. (44)

From these two equations we obtain that F,ρ=0F_{,\rho}=0, so eqs. (43) and (44) are equivalent. If F=F(T)F=F(T), we can always rescale time to achieve F=1F=1. Then, the frame becomes synchronous. The function GG should obey the equation

P0(G),T=(fG2de0dρ),P_{0}\left(\sqrt{G}\right)_{,T}=(\frac{f^{\prime}\sqrt{G}}{2}-\frac{de_{0}}{d\rho})\text{,} (45)

whence

G=P0e0μ(ρ,r).\sqrt{G}=\frac{P_{0}}{e_{0}}\mu(\rho,r). (46)

It follows from (3) and (33) that

(G),T=f2e0.\left(\sqrt{G}\right)_{,T}=\frac{f^{\prime}}{2e_{0}}\text{.} (47)

After substitution in (45) we obtain

P02μ,T=e0e0.P_{0}^{2}\mu_{,T}=-e_{0}^{\prime}e_{0}\text{.} (48)

Then, it is easy to find the solution with the help of the ansatz

μ=z(ρ)+e0e0η(r),\mu=z(\rho)+e_{0}e_{0}^{\prime}\eta(r)\text{,} (49)

where η=P02\eta^{\prime}=P_{0}^{-2}, whence

η=rdr¯P03(r¯)\eta=\int^{r}\frac{d\bar{r}}{P_{0}^{3}(\bar{r})} (50)

for a given ρ\rho. It follows from (46) that

G=P0e0[z(ρ)+e0e0η(r)],\sqrt{G}=\frac{P_{0}}{e_{0}}[z(\rho)+e_{0}e_{0}^{\prime}\eta(r)], (51)

so

ds2=dT2+(P0e0)2[z(ρ)+e0e0η(r)]2dρ2+r2dω2.ds^{2}=-dT^{2}+\left(\frac{P_{0}}{e_{0}}\right)^{2}[z(\rho)+e_{0}e_{0}^{\prime}\eta(r)]^{2}d\rho^{2}+r^{2}d\omega^{2}\text{.} (52)

It can be written also in the form

ds2=dT2+(r,ρe0)2dρ2+r2dω2ds^{2}=-dT^{2}+\left(\frac{r_{,\rho}}{e_{0}}\right)^{2}d\rho^{2}+r^{2}d\omega^{2} (53)

where we used (24).

This exactly corresponds (in our notations) to eqs. (19), (20) of bron .

Thus we obtained the synchronous form of the metric from the local Lorentz transformation following the approach of fom . Meanwhile, it was found in bron due to analysis of equations of motion of geodesics with different energies.

If the requirement (e0T)ρ=0\left(\frac{\partial e_{0}}{\partial T}\right)_{\rho}=0 is relaxed, the metric function depends, in general, on both TT and ρ\rho. Then, world lines of fiducial observers with ρ=const\rho=const and θ=const\theta=const, ϕ=const\phi=const are not geodesics. Indeed, in this case we have for the nonzero component of the four-acceleration aμa^{\mu}:

aρ=F,ρ2FG,a^{\rho}=\frac{F_{,\rho}}{2FG}\text{,} (54)
a2=aμaμ=(F,ρ)24GF2.a^{2}=a_{\mu}a^{\mu}=\frac{\left(F_{,\rho}\right)^{2}}{4GF^{2}}. (55)

As it is assumed, by construction, that FF and GG are finite and nonzero on the horizon, acceleration aa remains finite there.

If, instead of TT and ρ\rho, one uses TT and rr, the generalization of the GP frame can be obtained. Indeed, it follows from (23), (24) that

ds2=fFe02dT22P0e02FdTdr+Fdr2e02.ds^{2}=-\frac{fF}{e_{0}^{2}}dT^{2}-2\frac{P_{0}}{e_{0}^{2}}FdTdr+\frac{Fdr^{2}}{e_{0}^{2}}. (56)

Obviously, metric (56) is regular in the vicinity of the horizon. If e0=conste_{0}=const, F=1F=1, we return to (17) after change TTT\rightarrow-T.

V Double GP coordinates

In Ref. jose , two coordinates t~\tilde{t} and τ\tau were used for constructing a regular Schwarzschild metric. These coordinates represent advanced and retarded GP coordinates. In this Section, we extend the corresponding procedure considering more general metrics (1). It is quite straightforward but, bearing in mind that corresponding formulas can be useful in further applications, we list them explicitly. Let us introduce the coordinate τ\tau according to

dτ=e0dte0drfV.d\tau=e_{0}dt-e_{0}\frac{dr}{f}V\text{.} (57)

Then,

ds2=f4P02e02[f(dt~2+dτ2)]2(2e02f)dτdt~]+r2dω2.ds^{2}=\frac{f}{4P_{0}^{2}e_{0}^{2}}[f(d\tilde{t}^{2}+d\tau^{2})]-2(2e_{0}^{2}-f)d\tau d\tilde{t}]+r^{2}d\omega^{2}. (58)

This metric is still deficient near the horizon. To repair this, one can introduce Kruskal-type variables t~\tilde{t}^{\prime} and τ\tau^{\prime}. Let, for simplicity, e0e_{0} be constant. Then,

t~=t0exp(κt~e0)=t0exp(κt+κχe0),\tilde{t}^{\prime}=t_{0}\exp(\frac{\kappa\tilde{t}}{e_{0}})=t_{0}\exp(\kappa t+\kappa\frac{\chi}{e_{0}})\text{,} (59)
τ=t0exp(κt~e0)=t0exp(κt+κχe0),\tau^{\prime}=-t_{0}\exp(-\frac{\kappa\tilde{t}}{e_{0}})=-t_{0}\exp(-\kappa t+\kappa\frac{\chi}{e_{0}})\text{,} (60)

where t0t_{0} is some constant,

χ=dre02ff=𝑑re02f=e0𝑑rV,\chi=\int\frac{dr\sqrt{e_{0}^{2}-f}}{f}=\int dr^{\ast}\sqrt{e_{0}^{2}-f}=e_{0}\int dr^{\ast}V\text{,} (61)

κ\kappa is the surface gravity, rr^{\ast} is defined in (19). Then, one can check that in variables t~\tilde{t}^{\prime}, τ\tau^{\prime} the metric takes the form

ds2=f4P02e02κ2[f(dt~2t~2+dτ2τ2)]+2(2e02f)dττdt~t~]+r2dω2.ds^{2}=\frac{f}{4P_{0}^{2}e_{0}^{2}\kappa^{2}}[f(\frac{d\tilde{t}^{\prime 2}}{\tilde{t}^{\prime 2}}+\frac{d\tau^{\prime 2}}{\tau^{\prime 2}})]+2(2e_{0}^{2}-f)\frac{d\tau^{\prime}}{\tau^{\prime}}\frac{d\tilde{t}^{\prime}}{\tilde{t}^{\prime}}]+r^{2}d\omega^{2}\text{.} (62)

Near the horizon,

f2κ(rr+)χ=e02κln(rr+)+χreg,f\approx 2\kappa(r-r_{+})\text{, }\chi=\frac{e_{0}}{2\kappa}\ln(r-r_{+})+\chi_{reg}\text{,} (63)

where χreg\chi_{reg} is regular near r+r_{+}. Then,

r12κln(rr+),r^{\ast}\approx\frac{1}{2\kappa}\ln(r-r_{+})\text{,} (64)
t~t0expκv,\tilde{t}^{\prime}\approx t_{0}\exp\kappa v\text{,} (65)
τt0exp(κu),\tau^{\prime}\approx-t_{0}\exp(-\kappa u)\text{,} (66)

where

u=trv=t+r.u=t-r^{\ast}\text{, }v=t+r^{\ast}\text{.} (67)

Then, near the horizon, taking into account (35), we have

f2κ(t~τt02).f\approx-2\kappa\left(\frac{\tilde{t}^{\prime}\tau^{\prime}}{t_{0}^{2}}\right)\text{.} (68)

As a result, the metric (62) is regular near the horizon. If t0=Mt_{0}=M and f=12Mrf=1-\frac{2M}{r}, we return to the Schwarzschild case considered in jose . The whole space-time splits to four regions, similarly to the Kruskal metric in the Schwarzschild case. Transformations (59), (60) correspond to the quadrant I in jose and can be adjusted to other quadrants. We will not dwell upon on this.

If e0=1e_{0}=1, we return to the standard transformations that bring the metric into the Kruskal form. Now, we can also perform a limiting transition  e0e_{0}\rightarrow\infty and observe that

u=lime0τe0=drf=tr,u=\lim_{e_{0}\rightarrow\infty}\frac{\tau}{e_{0}}=-\int\frac{dr}{f}=t-r^{\ast}\text{,} (69)
v=lime0t~e0=t+r.v=\lim_{e_{0}\rightarrow\infty}\frac{\tilde{t}}{e_{0}}=t+r^{\ast}\text{.} (70)

In this limit,

ds2=fdudv+r2dω2,ds^{2}=-fdudv+r^{2}d\omega^{2}\text{,} (71)

so uu and vv have the meaning of the Eddington-Filkenstein coordinates. In doing so, t~\tilde{t}^{\prime} and τ\tau^{\prime} have the meaning of standard Kruskal coordinates.

It is worth noting  an important scale property of coordinates t~\tilde{t}^{\prime} and τ\tau^{\prime}. One can compare two limits: (i) e0e_{0}\rightarrow\infty for any fixed rr+r\geq r_{+} and (ii) the horizon limit rr+r\rightarrow r_{+} for any fixed e0e_{0}. In both limits these coordinates behave in the same manner. We see that the value e0e_{0} does not have a crucial influence on the coordinate frame, the metric remains regular on the horizon.

VI Some examples

In this section we present some examples how different metrics, initially discovered using totally different approaches can be incorporated into the general scheme described in the present paper.

First of all, we can note that Eq. (53) is generalization of the Lemaître - Tolman - Bondi solution (LBT) of Einstein equations valid for dust. To see this, is sufficient to write

e02=1+h(ρ)e_{0}^{2}=1+h(\rho)\text{. } (72)

Then, it corresponds to eq. (103.6) of LL , where we used hh instead of ff in LL and ρ\rho instead of RR. Meanwhile, we would like to stress that the metric (53) is more general and, in particular, its origination can have nothing to do with dust.

From the other hand, (53)  can be considered as generalization of the Novikov frame nov63 used for the description of the Schwarzschild metric, if we identify e02(ρ)=R21+R2e_{0}^{2}(\rho)=\frac{R^{\ast 2}}{1+R^{\ast 2}} in eq. (31.12a) of mtw .

Another interesting example appears if we put e0=1e_{0}=1 and

f(r)=1H2r2.f(r)=1-H^{2}r^{2}. (73)

It is convenient to rescale ρ\rho in such a way that z(ρ)=1.z(\rho)=1.

It is seen from (3) that

P0=Hr.P_{0}=Hr. (74)

Then, eq. (51) gives us

G=Hrz(ρ).\sqrt{G}=Hrz(\rho)\text{.} (75)

It is convenient to take z(ρ)=1z(\rho)=1. It follows from (24) that

r,ρ=Hrr_{,\rho}=Hr (76)

and it follows from (31) that

r,T=r,ρ.r_{,T}=-r_{,\rho}\text{.} (77)

As a result, we can write

r=r0exp(HρHT),r=r_{0}\exp(H\rho-HT)\text{,} (78)

where r0r_{0} is a constant. We see that the expression for rr is factorized into a product of a function of ρ\rho and a function of TT. This means that by appropriate redefinition of ρ\rho in the form χ=exp(Hρ)\chi=\exp(H\rho) we can kill all the dependence rr upon ρ\rho and obtain a metric with the dependence upon TT only. It is convenient also to choose r0=H1r_{0}=H^{-1} and make redefinition T~=T\tilde{T}=-T. Then,

ds2=dT2+exp(2HT)H2(dχ2+χ2dω2),ds^{2}=-dT^{2}+\frac{\exp(2HT)}{H^{2}}(d\chi^{2}+\chi^{2}d\omega^{2})\text{,} (79)

where we omitted tilde. This is nothing else than the standard Friedmann form of the de Sitter flat metric. It is interesting that allowing for non-constant e0e_{0} it is possible to get also positively and negatively curved de Sitter solutions, see bron .

As for GP form of the metric, it has the form

ds2=(1H2r2)dt2+2Hrdrdt+dr2ds^{2}=-(1-H^{2}r^{2})dt^{2}+2Hrdrdt+dr^{2} (80)

from which we can extract the Hubble law for the velocity of the flow V=HrV=Hr. It is known that this form is valid not only for de Sitter solution, but for an arbitrary Friedmann cosmology Faraoni .

Note that the fact that the resulting diagonal metric (79) appears to be a homogeneous one explicitly is connected with a particular form of the function ff in eq. (3) and particular value e0=1e_{0}=1 which leads to factorizable expression for rr. Meanwhile, in the next section we will see that there exists another family of homogeneous metrics existing for an arbitrary function ff.

VII The limit e00e_{0}\to 0

Let us consider the limiting transition e00e_{0}\rightarrow 0. It cannot be done in the metric (5) directly. In this limit, the axis rr and TT become collinear since in (24) the term with dρd\rho drops out. As a result, these coordinates fail to be suitable for constructing a regular frame. Also, the limit under discussion cannot be taken in the form of metric (52), (53). Formally, the proper distance between two arbitrary points with different values of their radial coordinate rr grows like 1/e01/e_{0} and the metric becomes degenerate.

However, for a synchronous metric the limit e0=0e_{0}=0 is allowed. To make a meaningful result, we need to rescale the spatial coordinate according to ρ=e0ρ~\rho=e_{0}\tilde{\rho} and take the limit under discussion only afterwards. Then, it follows from (24) with e0=0e_{0}=0, F=1F=1 that

T=rdr¯g(r¯),T=-\int^{r}\frac{d\bar{r}}{\sqrt{g(\bar{r})}}, (81)

where g=f>0g=-f>0. Thus this transformation is legitimate under the horizon only. It brings the metric in the form

ds2=dT2+g(r(T))dρ~2+r2(T)dω2.ds^{2}=-dT^{2}+g(r(T))d\tilde{\rho}^{2}+r^{2}(T)d\omega^{2}\text{.} (82)

Schematically, the timelike geodesics with e0=0e_{0}=0 are depicted on Fig. 1 where a relevant part of the Kruskal diagram is depicted.

Refer to caption
Figure 1: Timelike geodesics for e0=0e_{0}=0.

It is worth nothing that in synchronous form (82) of the metric the variable ρ~\tilde{\rho} is always a spatial one, and TT is always a temporal one. However, some peculiarities of the e0=0e_{0}=0 case lead to peculiar properties of the corresponding synchronous frame. It can be easily seen that the metric now depends on the temporal coordinate only, becoming an homogeneous one. This is however not surprising, since, as the T=constT=const hypersurface coincides now with the r=constr=const hypersurface, and any spatial dependence in a spherically symmetric metric is in fact the rr-dependence. Therefore, it is clear that the hypersurface T=constT=const in the e0=0e_{0}=0 case has no spatial dependence at all. The central singularity is not present in the any nonsingular T=constT=const plane, and instead, is present in the observer’s future.

This form of the metric can also be obtained directly from (1) if one interchanges the role of coordinates rr and tt and makes the coefficient gt~t~=1g_{\tilde{t}\tilde{t}}=-1 by rescaling the time coordinate. This is just the form, first introduced by Novikov - see nov61 and eqs. 2.4.8 and 2.4.9 in fn . It can be considered as particular case of the cosmological Kantowski-Sachs metric.

The cosmological interpretation of this metrics gives a non-formal explanation of a curious fact about time needed to reach a singularity from a horizon. Indeed, the coordinate time before cosmological singularity ΔT=r+\Delta T=r_{+} obviously does not depend on a particular motion of an observer. As for the proper time from the horizon crossing to singularity hitting, it differs from ΔT\Delta T by a Lotentz factor originating from the relative motion of the object in question with respect to the e0=0e_{0}=0 frame. As it is known from the SR, the Lorentz factor can only make the proper time shorter, so ΔT\Delta T is the maximum possible proper time from a horizon crossing to a singularity hitting, and it is achieved if the observer moves along the geodesic with e0=0e_{0}=0 – see our for detail, where other formal and informal treatments of this question have been given.

Returning to the GP metric (5), we can note that despite the original GP coordinate system has no smooth e00e_{0}\rightarrow 0 limit, we can easily write down another coordinate system with a smooth limit at e0=0e_{0}=0. Indeed, if instead of t~\tilde{t} and rr, one uses coordinates t~\tilde{t} and tt, then, after substitution of (2) into (5), we obtain the  metric in the form

ds2=gP02dt~2+g2dt2P02+2ge0P02dtdt~+r2(t~)dω2,ds^{2}=-\frac{g}{P_{0}^{2}}d\tilde{t}^{2}+\frac{g^{2}dt^{2}}{P_{0}^{2}}+\frac{2ge_{0}}{P_{0}^{2}}dtd\tilde{t}+r^{2}(\tilde{t})d\omega^{2}\text{,} (83)

g=f>0g=-f>0 under the horizon It can be rewritten in the form

ds2=dt~2+g2P02(dt+dt~e0g)2+r2(t~)dω2.ds^{2}=-d\tilde{t}^{2}+\frac{g^{2}}{P_{0}^{2}}(dt+\frac{d\tilde{t}e_{0}}{g})^{2}+r^{2}(\tilde{t})d\omega^{2}. (84)

As under the horizon the coordinate tt is space-like, the metric is expressed through one space-like and one time-like coordinate (in contrast to the original GP which has two time-like coordinates under the horizon). The non-diagonal term defines a coordinate ”flow velocity” e0g-\frac{e_{0}}{g} which can be interpreted as a velocity with respect to e0=0e_{0}=0 frame. Indeed, in the e0=0e_{0}=0 limit it vanishes. It is known that the 3-velocity with respect to e0=0e_{0}=0 frame of a radially falling particle with the energy ee is equal to e/P-e/P (see eq. (97) in we ). We get this value from the coordinate velocity if we remember that physical distance interval dldl is connected with the interval of the space-like coordinate dtdt through dl=(g/P)dtdl=(g/P)dt.

So that, this metric, in some sense dual to GP, has better behavior under the horizon than the original GP and allows a smooth transition to e0=0e_{0}=0 limit.

VIII Summary

Thus we established the connection between two kinds of approaches, both of them being connected with the particle dynamics through the parameter e0.e_{0}\,. In this sense, we revealed the meaning of main coordinate transformations from the original metric. Outside the horizon, some results are known but we extended corresponding interpretation, having considered the region inside the horizon.

Previous papers showed how to unify separate metrics and transformations. We made a next step and showed how one can unify the whole classes of unifying transformation. Namely, if the parameter of coordinate transformation e0=e0(ρ)e_{0}=e_{0}(\rho), the Fomin metric (22) turns into the BDG frame. It is worth noting that the metric (22) is more general than the BDG one in that the coordinate lines of observers with ρ=const\rho=const are not necessarily geodesics.

It is also shown that, when e00e_{0}\rightarrow 0, (22) turns smoothly into the metric considered by Novikov nov61 . To the best of our knowledge, existence of this limit was not considered before in the context of black hole metric under the horizon. Thus the coordinates frame such as the Kruskal-Szekeres, homogeneous Kantowski-Sacks metric inside the horizon and Lemaître ones, which look so differently, are now united as elements of a whole picture.

By contrary, the generalized GP metric has no smooth  e00e_{0}\rightarrow 0 limit. In a sense, we proposed a metric which can be considered as dual to GP. This new form of metric has a good behavior under the horizon, in particular, it is regular for e0=0e_{0}=0.

It is of interest to try extension of the approach under discussion to the rotating case. Especially interesting in this context is the possibility to buid a general approach that would unite the coordinate transformations to regular frames with the the Janis-Newman algorithm jan that relates static solutions and rotating metrics. Also, it is of interest to generalize the approach under discussion to higher dimensions. All this requires separate treatment.

IX Acknowledgement

This paper has been supported by the Kazan Federal University Strategic Academic Leadership Program. AT has been supported by the Interdisciplinary Scientific and Educational School of Moscow University in Fundamental and Applied Space Research.

References

  • (1) K. Schwarzschild, “Uber das Gravitationsfeld eines Massenpunktes nach der Einsteinschen Theorie”, Sitzungsberichte der Königlich Preussischen Akademie Wissenschaften 1916, 189 (1916).
  • (2) K. Martel and E. Poisson, Regular coordinate systems for Schwarzschild and other spherical spacetimes, Am. J. Phys. 69 (2001) 476-480, [arXiv:gr-qc/0001069].
  • (3) T. K. Finch, Coordinate families for the Schwarzschild geometry based on radial timelike geodesics, [arXiv:1211.4337].
  • (4) J. P. S. Lemos, D. L. F. G. Silva, Maximal extension of the Schwarzschild metric: From Painlev́-Gullstrand to Kruskal-Szekeres, [arXiv:2005.14211].
  • (5) L’Univers en expansion, Ann. Soc. Sci. Braxelles, Ser. A, 1933, v. 53, p. 51.
  • (6) K. Bronnikov, I. Dymnikova, E. Galaktionov, Multi-horizon spherically symmetric spacetimes with several scales of vacuum energy, Class. Quant. Grav. 29, 095025 (2012)[arXiv:1204.0534].
  • (7) A. Toporensky, O. Zaslavskii, S. Popov, Unified approach to redshift in cosmological /black hole spacetimes and synchronous frame, Eur. J. Phys. 39, 015601 (2018), [arXiv:1704.08308].
  • (8) I. D. Novikov, Doctoral dissertation, Shternberg Astronomical Institute, Moscow.
  • (9) A.S. Kompaneets and A.S. Chernov, Solution of the Gravitation Equations for a Homogeneous Anisotropic Model, Zh. Eksp. Teor. Fiz. 47, 1939 (1964); Sov. Phys. JETP 20, 1303 (1965).
  • (10) R. Kantowski and R.K. Sachs, Some Spatially Homogeneous Anisotropic Relativistic Cosmological Models, J. Math. Phys. 7, 443 (1966).
  • (11) P. I. Fomin. Coordinate transformations that eliminate singularities on the gravitational radius in the Schwarzschild metric. Sov. Physics JETP, 27 (1968) 483.
  • (12) A. Gullstrand, Allgemeine Lösung des statischen Einkörperproblems in der Einsteinschen Gravitationstheorie, Arkiv. Mat. Astron. Fys. 16 (1922) 1.
  • (13) P. Painlevé, La mecanique classique et la theorie de la relativité, C. R. Acad. Sci. (Paris) 173 (1923) 677.
  • (14) Lake, K. A class of quasi-stationary regular line elements for the Schwarzschild geometry. [arXiv:gr-qc/9407005].
  • (15) J. M. Bardeen, W. H. Press, and S. A. Teukolsky, Rotating black holes: locally nonrotating frames, energy extraction, and scalar synchrotron radiation, Astrophys. J. 178, 347 (1972).
  • (16) A. V. Toporensky and O. B. Zaslavskii, On strategies of motion under the black hole horizon, Int. J. of Mod. Phys. D (2020) 2030003 [arXiv:1905.02150].
  • (17) Gautreau, R., Hoffmann, B. The Schwarzschild radial coordinate as a measure of proper distance. Phys. Rev. D 17, 2552–2555 (1978).
  • (18) M. R. Francis and A. Kosowsky, Geodesics in the generalized Schwarzschild solution, Am. J. Phys. 72, 1204 (2004), [arXiv:gr-qc/0311038].
  • (19) Frolov V P and Novikov I D 1998 Physics of Black Holes (Dordrecht: Kluwer Academic Publishers).
  • (20) Landau, L.D.; Lifshitz, E.M. The Classical Theory of Fields; Pergamon Press: Oxford, 1983.
  • (21) I. D. Novikov, Note on the space-time metric inside the Schwarzchild singular sphere, Soviet Astronomy, 5 (1961) 423 (Astron. Zh. 38, 564 (1961)).
  • (22) M. Kruskal, “Maximal extension of Schwarzschild metric”, Phys. Rev. 119, 1743 (1959).
  • (23) G. Szekeres, “On the singularities of a Riemannian manifold”, Publicationes Mathematicae Debrecen 7, 285 (1959).
  • (24) C.W. Misner, K. S. Thorne, and J. A. Wheeler, Gravitation (Freeman, San Francisco, 1973).
  • (25) A. Toporensky, O. Zaslavskii, Flow and peculiar velocities for generic motion in spherically symmetric black holes, Gravitation and Cosmology, 27, 126 (2021).
  • (26) G. Vachon, R. Vanderwee and V. Faraoni, Revisiting geodesic observers in cosmology, EPJC, 81:820 (2021), [arXiv:2108.01782].
  • (27) E. T. Newman and A. I. Janis, Note on the Kerr Spinning-Particle Metric, J. Math. Phys. 6, 915 (1965).