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

Charged black-bounce spacetimes

Edgardo Franzin IDa,b,c    Stefano Liberati IDa,b,c    Jacopo Mazza IDa,b,c    Alex Simpson IDd    and Matt Visser IDd
Abstract

Given the recent development of rotating black-bounce–Kerr spacetimes, for both theoretical and observational purposes it becomes interesting to see whether it might be possible to construct black-bounce variants of the entire Kerr–Newman family. Specifically, herein we shall consider black-bounce–Reissner–Nordström and black-bounce–Kerr–Newman spacetimes as particularly simple and clean everywhere-regular black hole “mimickers” that deviate from the Kerr–Newman family in a precisely controlled and minimal manner, and smoothly interpolate between regular black holes and traversable wormholes. While observationally the electric charges on astrophysical black holes are likely to be extremely low, |Q|/m1|Q|/m\ll 1, introducing any non-zero electric charge has a significant theoretical impact. In particular, we verify the existence of a Killing tensor (and associated Carter-like constant) but without the full Killing tower of principal tensor and Killing–Yano tensor, also we discuss how, assuming general relativity, the black-bounce–Kerr–Newman solution requires an interesting, non-trivial matter/energy content.

1 Introduction

Given the demonstrated existence of the black-bounce–Schwarzschild [1, 2, 3, 4, 5, 6], and the black-bounce–Kerr [7, 8, 9] geometries, it is intuitive to suspect that analogous black-bounce variants of both the Reissner–Nordström (RN) and Kerr–Newman (KN) spacetimes will exist [10]; and that they would be amenable to reasonably tractable general relativistic analyses.

Specifically, in ref. [7], three of the current authors used the Newman–Janis procedure [11, 12, 13] to transmute the original black-bounce–Schwarzschild spacetime [1, 2, 3, 4] into an axisymmetric rotating version. Herein, we shall propose that the procedure by which one obtains a “black-bounce” variant from a known pre-existing solution in either axisymmetry or spherical symmetry can be simplified, and in doing so we advocate for two new candidate spacetimes: a spherically symmetric black-bounce with electrical charge (black-bounce–Reissner–Nordström), as well as the axisymmetric rotating equivalent (black-bounce–Kerr–Newman).

Given any spherically symmetric or axisymmetric geometry equipped with some metric gμνg_{\mu\nu} which possesses a curvature singularity at r=0r=0, the proposed procedure is explicitly designed to transmute said geometry into a globally regular candidate spacetime whilst retaining the manifest symmetries. In standard t,r,θ,ϕt,r,\theta,\phi curvature coordinates, the procedure is simply as follows:

  • Leave the object dr{\mathrm{d}}r in the line element undisturbed.

  • Whenever the metric components gμνg_{\mu\nu} have an explicit rr-dependence, replace the rr-coordinate by r2+2\sqrt{r^{2}+\ell^{2}}, where \ell is some length scale (typically associated with the Planck length).

Leaving the object dr{\mathrm{d}}r unchanged implies that the rr-coordinate still performs an identical role to the curvature rr-coordinate in terms of the spatial slicings of the spacetime, and ensures that we are not simply making a coordinate transformation. As we shall shortly see, replacing rr2+2r\rightarrow\sqrt{r^{2}+\ell^{2}} has the advantage of ‘smoothing’ the geometry into something which is globally regular.

Apart from regularity, these spacetimes exhibit the interesting feature that they are a 1-parameter class of geometries smoothly interpolating between standard general relativity electrovac black holes and traversable wormholes [14, 15, 16, 17, 18]. In particular, the classical energy conditions will have non-trivial behaviour in these spacetimes [16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28]. Furthermore, since these geometries can be fine-tuned to be arbitrarily close to the usual Kerr family, they can serve as black hole “mimickers” potentially of interest to the observational community [29, 30, 31, 32, 33, 34, 35, 36, 37, 38, 39, 40].

2 Black-bounce–Reissner–Nordström geometry

To begin with, let us consider the Reissner–Nordström solution to the electrovac Einstein equations of general relativity expressed in terms of standard t,r,θ,ϕt,r,\theta,\phi curvature coordinates and geometrodynamic units (we shall use everywhere a (,+,+,+)(-,+,+,+) signature)

ds2=fRN(r)dt2+dr2fRN(r)+r2dΩ22,fRN(r)=12mr+Q2r2.{\mathrm{d}}s^{2}=-f_{\text{RN}}(r){\mathrm{d}}t^{2}+\frac{{\mathrm{d}}r^{2}}{f_{\text{RN}}(r)}+r^{2}{\mathrm{d}}\Omega^{2}_{2}\ ,\quad f_{\text{RN}}(r)=1-\frac{2m}{r}+\frac{Q^{2}}{r^{2}}\ . (2.1)

We now perform our “regularising” procedure, replacing rr2+2r\rightarrow\sqrt{r^{2}+\ell^{2}} in the metric. The new candidate spacetime is described by the following line element

ds2=f(r)dt2+dr2f(r)+(r2+2)dΩ22,f(r)=12mr2+2+Q2r2+2.{\mathrm{d}}s^{2}=-f(r){\mathrm{d}}t^{2}+\frac{{\mathrm{d}}r^{2}}{f(r)}+\left(r^{2}+\ell^{2}\right){\mathrm{d}}\Omega^{2}_{2}\ ,\quad f(r)=1-\frac{2m}{\sqrt{r^{2}+\ell^{2}}}+\frac{Q^{2}}{r^{2}+\ell^{2}}\ . (2.2)

One can immediately see that the natural domains for the angular and temporal coordinates are unaffected by the regularisation procedure. In contrast the natural domain of the rr coordinate expands from r[0,+)r\in[0,+\infty) to r(,+)r\in(-\infty,+\infty). Asymptotic flatness is preserved, as are the manifest spherical and time translation symmetries. Given the diagonal metric environment, it is trivial to establish the following tetrad

(et^)μ\displaystyle\big{(}e_{\hat{t}}\big{)}^{\mu} =1|f(r)|(1,0,0,0),(er^)μ=|f(r)|(0,1,0,0),\displaystyle=\frac{1}{\sqrt{|f(r)|}}\left(1,0,0,0\right),\quad\big{(}e_{\hat{r}}\big{)}^{\mu}=\sqrt{|f(r)|}\left(0,1,0,0\right),
(eθ^)μ\displaystyle\big{(}e_{\hat{\theta}}\big{)}^{\mu} =1r2+2(0,0,1,0),(eϕ^)μ=1r2+2sinθ(0,0,0,1),\displaystyle=\frac{1}{\sqrt{r^{2}+\ell^{2}}}\left(0,0,1,0\right),\quad\big{(}e_{\hat{\phi}}\big{)}^{\mu}=\frac{1}{\sqrt{r^{2}+\ell^{2}}\,\sin\theta}\left(0,0,0,1\right), (2.3)

and straightforward to verify that this is indeed a solution of gμνeμ^eν^μ=νημ^ν^g_{\mu\nu}e_{\hat{\mu}}{}^{\mu}e_{\hat{\nu}}{}^{\nu}=\eta_{\hat{\mu}\hat{\nu}}. (Note however that the 1-1 in the Minkowski metric corresponds to the timelike direction, and is therefore in the t^t^\hat{t}\hat{t} position when f(r)>0f(r)>0, and in the r^r^\hat{r}\hat{r} position when instead f(r)<0f(r)<0). The analysis that follows is, when appropriate, performed with respect to this orthonormal basis.

Furthermore, note that as 0\ell\to 0 one recovers the standard Reissner–Nordström geometry, while when m0m\to 0 and Q0Q\to 0 one recovers the standard Morris–Thorne wormhole [14, 15, 16, 41]:

ds2=dt2+dr2+(r2+2)dΩ22.{\mathrm{d}}s^{2}=-{\mathrm{d}}t^{2}+{\mathrm{d}}r^{2}+(r^{2}+\ell^{2})\,{\mathrm{d}}\Omega^{2}_{2}\ . (2.4)

2.1 Kretschmann scalar

Our first task consists in showing that our spacetime is indeed everywhere regular. Conveniently, given that our spacetime is static, examination of the Kretschmann scalar K=RμνρσRμνρσK=R^{\mu\nu\rho\sigma}\,R_{\mu\nu\rho\sigma} will be sufficient to accomplish this task [4].

Indeed, a simple computation shows that the Kretschmann scalar is quartic in QQ and given by

K\displaystyle K =4(r2+2)6{(34102r2+14r4)Q4\displaystyle=\frac{4}{(r^{2}+\ell^{2})^{6}}\Bigg{\{}\left(3\ell^{4}-10\ell^{2}r^{2}+14r^{4}\right)Q^{4} (2.5)
+r2+2[22(223r2)r2+22m(54112r2+12r4)]Q2\displaystyle+\sqrt{r^{2}+\ell^{2}}\left[2\ell^{2}(2\ell^{2}-3r^{2})\sqrt{r^{2}+\ell^{2}}-2m(5\ell^{4}-11\ell^{2}r^{2}+12r^{4})\right]Q^{2}
+38+(9m2+6r2)6+3r2(r2m2)48m2(4r4)r2+2+12m2r6}.\displaystyle+3\ell^{8}+(9m^{2}+6r^{2})\ell^{6}+3r^{2}(r^{2}-m^{2})\ell^{4}-8m\ell^{2}(\ell^{4}-r^{4})\sqrt{r^{2}+\ell^{2}}+12m^{2}r^{6}\Bigg{\}}\ .

In the limit as r0r\rightarrow 0 we have the finite result

limr0K=128Q4+8(25m)7Q2+4(328m+9m2)6.\lim_{r\rightarrow 0}K=\frac{12}{\ell^{8}}\,Q^{4}+\frac{8(2\ell-5m)}{\ell^{7}}\,Q^{2}+\frac{4\left(3\ell^{2}-8m\ell+9m^{2}\right)}{\ell^{6}}\ . (2.6)

So in this manifestly static situation we are then guaranteed that all of the orthonormal curvature tensor components are automatically finite [4] (nonetheless for completeness we provide in Appendix A expressions for other curvature invariants). We may then conclude that the geometry is indeed globally regular.

2.2 Curvature tensors

In order to complete our investigation on curvature, we display here the explicit forms for the various non-zero curvature components. This is mostly conveniently done in the above introduced orthonormal basis. The orthonormal components of the Riemann curvature tensor are given by

Rt^r^t^r^\displaystyle R^{\hat{t}\hat{r}}{}_{\hat{t}\hat{r}} =23r2(r2+2)3Q2m(22r2)(r2+2)5/2,\displaystyle=\frac{\ell^{2}-3r^{2}}{(r^{2}+\ell^{2})^{3}}Q^{2}-\frac{m(\ell^{2}-2r^{2})}{(r^{2}+\ell^{2})^{5/2}}\ , (2.7a)
Rt^θ^=t^θ^Rt^ϕ^t^ϕ^\displaystyle R^{\hat{t}\hat{\theta}}{}_{\hat{t}\hat{\theta}}=R^{\hat{t}\hat{\phi}}{}_{\hat{t}\hat{\phi}} =r2(r2+2)3Q2mr2(r2+2)5/2,\displaystyle=\frac{r^{2}}{(r^{2}+\ell^{2})^{3}}Q^{2}-\frac{mr^{2}}{(r^{2}+\ell^{2})^{5/2}}\ , (2.7b)
Rr^θ^=r^θ^Rr^ϕ^r^ϕ^\displaystyle R^{\hat{r}\hat{\theta}}{}_{\hat{r}\hat{\theta}}=R^{\hat{r}\hat{\phi}}{}_{\hat{r}\hat{\phi}} =r22(r2+2)3Q2+m(22r2)2r2+2(r2+2)5/2,\displaystyle=\frac{r^{2}-\ell^{2}}{(r^{2}+\ell^{2})^{3}}Q^{2}+\frac{m(2\ell^{2}-r^{2})-\ell^{2}\sqrt{r^{2}+\ell^{2}}}{(r^{2}+\ell^{2})^{5/2}}\ , (2.7c)
Rθ^ϕ^θ^ϕ^\displaystyle R^{\hat{\theta}\hat{\phi}}{}_{\hat{\theta}\hat{\phi}} =r2(r2+2)3Q2+2mr2+2r2+2(r2+2)5/2,\displaystyle=-\frac{r^{2}}{(r^{2}+\ell^{2})^{3}}Q^{2}+\frac{2mr^{2}+\ell^{2}\sqrt{r^{2}+\ell^{2}}}{(r^{2}+\ell^{2})^{5/2}}\ , (2.7d)

and in the limit as r0r\rightarrow 0 we find

Rt^r^t^r^Q2m4,Rt^θ^t^θ^0,Rr^θ^r^θ^Q2+22m4,Rθ^ϕ^θ^ϕ^12.R^{\hat{t}\hat{r}}{}_{\hat{t}\hat{r}}\rightarrow\frac{Q^{2}-m\ell}{\ell^{4}}\ ,\quad R^{\hat{t}\hat{\theta}}{}_{\hat{t}\hat{\theta}}\rightarrow 0\ ,\quad R^{\hat{r}\hat{\theta}}{}_{\hat{r}\hat{\theta}}\rightarrow-\frac{Q^{2}+\ell^{2}-2m\ell}{\ell^{4}}\ ,\quad R^{\hat{\theta}\hat{\phi}}{}_{\hat{\theta}\hat{\phi}}\rightarrow\frac{1}{\ell^{2}}\ . (2.8)

For the orthonormal components of the Ricci tensor we find

Rt^t^\displaystyle R^{\hat{t}}{}_{\hat{t}} =2r2(r2+2)3Q2m2(r2+2)5/2,\displaystyle=\frac{\ell^{2}-r^{2}}{(r^{2}+\ell^{2})^{3}}Q^{2}-\frac{m\ell^{2}}{(r^{2}+\ell^{2})^{5/2}}\ , (2.9a)
Rr^r^\displaystyle R^{\hat{r}}{}_{\hat{r}} =Q2(r2+2)2+2(3m2r2+2)(r2+2)5/2,\displaystyle=-\frac{Q^{2}}{(r^{2}+\ell^{2})^{2}}+\frac{\ell^{2}(3m-2\sqrt{r^{2}+\ell^{2}})}{(r^{2}+\ell^{2})^{5/2}}\ , (2.9b)
Rθ^θ^\displaystyle R^{\hat{\theta}}{}_{\hat{\theta}} =Rϕ^=ϕ^r22(r2+2)3Q2+2m2(r2+2)5/2,\displaystyle=R^{\hat{\phi}}{}_{\hat{\phi}}=\frac{r^{2}-\ell^{2}}{(r^{2}+\ell^{2})^{3}}Q^{2}+\frac{2m\ell^{2}}{(r^{2}+\ell^{2})^{5/2}}\ , (2.9c)

and in the limit as r0r\rightarrow 0

Rt^t^Q2m4;Rr^r^3m22Q24;Rθ^=θ^Rϕ^ϕ^2mQ24.R^{\hat{t}}{}_{\hat{t}}\rightarrow\frac{Q^{2}-m\ell}{\ell^{4}}\ ;\quad R^{\hat{r}}{}_{\hat{r}}\rightarrow\frac{3m\ell-2\ell^{2}-Q^{2}}{\ell^{4}}\ ;\quad R^{\hat{\theta}}{}_{\hat{\theta}}=R^{\hat{\phi}}{}_{\hat{\phi}}\rightarrow\frac{2m\ell-Q^{2}}{\ell^{4}}\ . (2.10)

Finally, for the Weyl tensor we obtain

Ct^r^t^r^\displaystyle C^{\hat{t}\hat{r}}{}_{\hat{t}\hat{r}} =2Ct^θ^=t^θ^2Ct^ϕ^=t^ϕ^2Cr^θ^=r^θ^2Cr^ϕ^=r^ϕ^Cθ^ϕ^θ^ϕ^\displaystyle=-2C^{\hat{t}\hat{\theta}}{}_{\hat{t}\hat{\theta}}=-2C^{\hat{t}\hat{\phi}}{}_{\hat{t}\hat{\phi}}=-2C^{\hat{r}\hat{\theta}}{}_{\hat{r}\hat{\theta}}=-2C^{\hat{r}\hat{\phi}}{}_{\hat{r}\hat{\phi}}=C^{\hat{\theta}\hat{\phi}}{}_{\hat{\theta}\hat{\phi}}
=2(23r2)3(r2+2)3Q2+22r2+23m(22r2)3(r2+2)5/2,\displaystyle=\frac{2(\ell^{2}-3r^{2})}{3(r^{2}+\ell^{2})^{3}}Q^{2}+\frac{2\ell^{2}\sqrt{r^{2}+\ell^{2}}-3m(\ell^{2}-2r^{2})}{3(r^{2}+\ell^{2})^{5/2}}\ , (2.11)

and in the limit as r0r\rightarrow 0

Ct^r^t^r^2Q2+223m34.C^{\hat{t}\hat{r}}{}_{\hat{t}\hat{r}}\rightarrow\frac{2Q^{2}+2\ell^{2}-3m\ell}{3\ell^{4}}\ . (2.12)

Now that the regularity of spacetime is out of question, we shall proceed by discussing its geometrical properties, i.e. horizons and characteristic orbits.

2.3 Horizons and surface gravity

In view of the diagonal metric environment, horizon locations are characterised by

gtt\displaystyle g_{tt} =0rH=S1(m+S2m2Q2)22.\displaystyle=0\quad\Longrightarrow\quad r_{H}=S_{1}\sqrt{\left(m+S_{2}\sqrt{m^{2}-Q^{2}}\right)^{2}-\ell^{2}}\ . (2.13)

Here S1,S2=±1S_{1},S_{2}=\pm 1, and choice of sign for S1S_{1} dictates which universe we are in, whilst the choice of sign on S2S_{2} corresponds to an outer/inner horizon respectively. For horizons to exist we need both |Q|m|Q|\leq m and m±m2Q2\ell\leq m\pm\sqrt{m^{2}-Q^{2}}. The case |Q|=m|Q|=m while m\ell\leq m leads to extremal horizons at rH=±m22r_{H}=\pm\sqrt{m^{2}-\ell^{2}}. When |Q|>m|Q|>m there are no horizons, and the geometry is that of a traversable wormhole. Finally when |Q|m|Q|\leq m but \ell is large enough, >m±m2Q2\ell>m\pm\sqrt{m^{2}-Q^{2}}, first the inner horizons vanish and then the outer horizons vanish.

The structure of the maximally-extended spacetime can be visualised with the aid of the Penrose diagrams in figures 1 and 2 for the two qualitatively different regular black hole cases. Both diagrams will be relevant for the rotating generalisation of the next section, too; they are analogous to those of [7] (both) and [1] (the first one).

Refer to caption
Figure 1: Penrose diagram for a regular black hole with only outer horizons, corresponding to mm2Q2<<m+m2Q2m-\sqrt{m^{2}-Q^{2}}<\ell<m+\sqrt{m^{2}-Q^{2}}. The lower (upper) portion of the diagram corresponds to the r>0r>0 (r<0r<0) universe; the diagram continues indefinitely above and below the portion shown by repetition of this fundamental block. Here, r+r_{+} is rHr_{H} with S2=+1S_{2}=+1; the sign in front of it is S1S_{1}.
Refer to caption
Figure 2: Penrose diagram for a regular black hole with outer and inner horizons, corresponding to <mm2Q2\ell<m-\sqrt{m^{2}-Q^{2}}. Vertical lines of the same colour are identified, as the right-hand (left-hand) part of the diagram represents the r>0r>0 (r<0r<0) universe; the diagram continues indefinitely above and below the portion shown. Here r+r_{+} (resp. rr_{-}) is rHr_{H} with S2=+1S_{2}=+1 (1-1); the sign in front of it is S1S_{1}.

It is straightforward to calculate the surface gravity at the event horizon in our universe for the black-bounce–Reissner–Nordström spacetime. Keeping in mind that we are still working in curvature coordinates, the surface gravity κH\kappa_{H} reduces to [42]

κH=limrrH12rgttgttgrr.\kappa_{H}=\lim_{r\rightarrow r_{H}}\frac{1}{2}\frac{\partial_{r}g_{tt}}{\sqrt{-g_{tt}g_{rr}}}\ . (2.14)

For our metric we have gtt=1/grrg_{tt}=-1/g_{rr} so this simplifies to

κH=12rgtt|rH=(m+m2Q2)22m2Q2(m+m2Q2)3=κHRNrH2rH2+2,\displaystyle\kappa_{H}=\left.\frac{1}{2}\partial_{r}g_{tt}\right|_{r_{H}}=\frac{\sqrt{(m+\sqrt{m^{2}-Q^{2}})^{2}-\ell^{2}}\;\sqrt{m^{2}-Q^{2}}}{(m+\sqrt{m^{2}-Q^{2}})^{3}}=\kappa_{H}^{\text{RN}}\sqrt{\frac{r_{H}^{2}}{r_{H}^{2}+\ell^{2}}}\ , (2.15)

where κHRN\kappa_{H}^{\text{RN}} is the surface gravity of a Reissner–Nordström black hole of the same mass and charge, and as usual the associated Hawking temperature is given by kBTH=2πκHk_{B}T_{H}=\frac{\hbar}{2\pi}\kappa_{H} [42]. (This gives the usual Reissner–Nordström result as 0\ell\to 0.)

2.4 Innermost stable circular orbit and photon sphere

Let us briefly examine the coordinate locations of the notable orbits for our candidate spacetime, the innermost stable circular orbit (ISCO) and photon sphere [43, 44, 45]. Firstly we define the object

ϵ={1massive particle, i.e. timelike worldline0massless particle, i.e. null worldline.\epsilon=\begin{cases}-1&\quad\mbox{massive particle, \emph{i.e.}\ timelike worldline}\\ 0&\quad\mbox{massless particle, \emph{i.e.}\ null worldline}.\end{cases} (2.16)

Considering the affinely parameterised tangent vector to the worldline of a massive or massless particle, and fixing θ=π/2\theta=\pi/2 in view of spherical symmetry, we obtain the reduced equatorial problem

ds2dλ2=gtt(dtdλ)2+grr(drdλ)2+(r2+2)(dϕdλ)2=ϵ.\frac{{\mathrm{d}}s^{2}}{{\mathrm{d}}\lambda^{2}}=-g_{tt}\left(\frac{{\mathrm{d}}t}{{\mathrm{d}}\lambda}\right)^{2}+g_{rr}\left(\frac{{\mathrm{d}}r}{{\mathrm{d}}\lambda}\right)^{2}+(r^{2}+\ell^{2})\left(\frac{{\mathrm{d}}\phi}{{\mathrm{d}}\lambda}\right)^{2}=\epsilon\ . (2.17)

The Killing symmetries yield the following expressions for the conserved energy EE, and angular momentum per unit mass LL

(12mr2+2+Q2r2+2)(dtdλ)=E;(r2+2)(dϕdλ)=L,\left(1-\frac{2m}{\sqrt{r^{2}+\ell^{2}}}+\frac{Q^{2}}{r^{2}+\ell^{2}}\right)\left(\frac{{\mathrm{d}}t}{{\mathrm{d}}\lambda}\right)=E\ ;\quad(r^{2}+\ell^{2})\left(\frac{{\mathrm{d}}\phi}{{\mathrm{d}}\lambda}\right)=L\ , (2.18)

yielding the “effective potential” for geodesic orbits

Vϵ(r)=(12mr2+2+Q2r2+2)[ϵ+L2r2+2].V_{\epsilon}(r)=\left(1-\frac{2m}{\sqrt{r^{2}+\ell^{2}}}+\frac{Q^{2}}{r^{2}+\ell^{2}}\right)\left[-\epsilon+\frac{L^{2}}{r^{2}+\ell^{2}}\right]\ . (2.19)

Null orbits

For the massless case of null orbits, e.g. photon orbits, set ϵ=0\epsilon=0 and solve V0(r)=0V_{0}^{\prime}(r)=0 for the location of the “photon sphere”. We have

V0(r)\displaystyle V_{0}^{\prime}(r) =2L2r[3mr2+2](r2+2)5/24L2r(r2+2)3Q2,\displaystyle=\frac{2L^{2}r\left[3m-\sqrt{r^{2}+\ell^{2}}\right]}{(r^{2}+\ell^{2})^{5/2}}-\frac{4L^{2}r}{(r^{2}+\ell^{2})^{3}}Q^{2}\ , (2.20)

yielding the analytic location for the photon sphere in our universe outside horizons:

rγ=m2(9m+39m28Q2)2Q22.r_{\gamma}=\sqrt{\frac{m}{2}\left(9m+3\sqrt{9m^{2}-8Q^{2}}\right)-2Q^{2}-\ell^{2}}\ . (2.21)

In the limit as Q,0Q,\ell\rightarrow 0 we reproduce the standard Schwarzschild result, rγ=3mr_{\gamma}=3m, as expected.

Timelike orbits

For the massive case of timelike orbits, the ISCO is found via setting V1(r)=0V_{-1}^{\prime}(r)=0. We have

V1(r)\displaystyle V_{-1}^{\prime}(r) =2r[m(r2+2)+L2(3mr2+2)](r2+2)5/22r(2L2+r2+2)(r2+2)3Q2.\displaystyle=\frac{2r\left[m(r^{2}+\ell^{2})+L^{2}(3m-\sqrt{r^{2}+\ell^{2}})\right]}{(r^{2}+\ell^{2})^{5/2}}-\frac{2r(2L^{2}+r^{2}+\ell^{2})}{(r^{2}+\ell^{2})^{3}}Q^{2}\ . (2.22)

Equating V1(r)=0V_{-1}^{\prime}(r)=0 and solving for rr is not analytically feasible. We may make life easier via the change of variables z=r2+2z=\sqrt{r^{2}+\ell^{2}}, giving

V1(z)=2z22z6{(2L2+z2)Q2+z[L2(z3m)mz2]}.V_{-1}^{\prime}(z)=-\frac{2\sqrt{z^{2}-\ell^{2}}}{z^{6}}\left\{(2L^{2}+z^{2})Q^{2}+z\left[L^{2}(z-3m)-mz^{2}\right]\right\}\ . (2.23)

Assuming some fixed orbit at some rcr_{c}, hence fixing the corresponding zc=rc2+2z_{c}=\sqrt{r_{c}^{2}+\ell^{2}}, we may rearrange to find the required angular momentum per unit mass LcL_{c} as a function of zcz_{c} and the metric parameters. It follows that the ISCO will be located at the coordinate location where LcL_{c} is minimised [43, 44, 45]. We find

Lc\displaystyle L_{c} =zcmzcQ2zc23mzc+2Q2;Lczc=6m2zc2mzc(9Q2+zc2)+4Q42mzcQ2(zc23mzc+2Q2)3/2.\displaystyle=\frac{z_{c}\sqrt{mz_{c}-Q^{2}}}{\sqrt{z_{c}^{2}-3mz_{c}+2Q^{2}}}\ ;\quad\Longrightarrow\quad\frac{\partial L_{c}}{\partial z_{c}}=-\frac{6m^{2}z_{c}^{2}-mz_{c}(9Q^{2}+z_{c}^{2})+4Q^{4}}{2\sqrt{mz_{c}-Q^{2}}(z_{c}^{2}-3mz_{c}+2Q^{2})^{3/2}}\ . (2.24)

Equating Lc/zc=0\partial L_{c}/\partial z_{c}=0 and solving for zcz_{c}, rearranging for rcr_{c}, and discounting complex roots leaves the analytic ISCO location for timelike particles in our universe:

rc=9m4Q46m2Q2(A2+2Am2+4m4)+(A2+2Am2+4m4)2A22m2mAr_{c}=\frac{\sqrt{9m^{4}Q^{4}-6m^{2}Q^{2}\left(A^{2}+2Am^{2}+4m^{4}\right)+\left(A^{2}+2Am^{2}+4m^{4}\right)^{2}-A^{2}\ell^{2}m^{2}}}{mA} (2.25)

where now

A\displaystyle A =[2m2Q4+m2(±B9m2)Q2+8m6]13;\displaystyle=\left[2m^{2}Q^{4}+m^{2}\left(\pm B-9m^{2}\right)Q^{2}+8m^{6}\right]^{\frac{1}{3}}\ ;
B\displaystyle B =4Q49m2Q2+5m4.\displaystyle=\sqrt{4Q^{4}-9m^{2}Q^{2}+5m^{4}}\ . (2.26)

It is easily verified that in the limit as Q,0Q,\ell\rightarrow 0, rc6mr_{c}\rightarrow 6m, as expected for Schwarzschild.

2.5 Stress-energy tensor

Starting from a given geometry, the discussion of the associated stress-energy tensor necessarily requires us to fix the geometrodynamics. In the following we shall assume that this is everywhere described by general relativity. While this might seem a crude assumption for geometries associated to a possible regularisation of singularities by quantum gravity, we do expect it — once the regular spacetime settles down in an equilibrium state after the collapse — to be a good approximation everywhere, if the final regularisation scale \ell is much larger than the Planck scale, or at least sufficiently far away from the core region otherwise.

Given the above assumption, the determination of the stress-energy tensor associated to our spacetime is easily accomplished by computing the non-zero components of the Einstein tensor. This leads to the following decomposition for the total stress-energy tensor Tμ^ν^T^{\hat{\mu}}{}_{\hat{\nu}} valid outside the outer horizon (and inside the inner horizon):

18πGμ^=ν^Tμ^=ν^[Tbb]μ^+ν^[TQ]μ^=ν^diag(ε,pr,pt,pt).\frac{1}{8\pi}\,G^{\hat{\mu}}{}_{\hat{\nu}}=T^{\hat{\mu}}{}_{\hat{\nu}}=\left[T_{\text{bb}}\right]^{\hat{\mu}}{}_{\hat{\nu}}+\left[T_{Q}\right]^{\hat{\mu}}{}_{\hat{\nu}}=\operatorname{diag}\left(-\varepsilon,p_{r},p_{t},p_{t}\right)\ . (2.27)

In contrast, between the inner and outer horizons we have:

18πGμ^=ν^Tμ^=ν^[Tbb]μ^+ν^[TQ]μ^=ν^diag(pr,ε,pt,pt).\frac{1}{8\pi}\,G^{\hat{\mu}}{}_{\hat{\nu}}=T^{\hat{\mu}}{}_{\hat{\nu}}=\left[T_{\text{bb}}\right]^{\hat{\mu}}{}_{\hat{\nu}}+\left[T_{Q}\right]^{\hat{\mu}}{}_{\hat{\nu}}=\operatorname{diag}\left(p_{r},-\varepsilon,p_{t},p_{t}\right)\ . (2.28)

Here [Tbb]μ^ν^\left[T_{\text{bb}}\right]^{\hat{\mu}}{}_{\hat{\nu}} is the stress-energy tensor for the original electrically neutral black-bounce spacetime from ref. [1], and [TQ]μ^ν^\left[T_{Q}\right]^{\hat{\mu}}{}_{\hat{\nu}} is the charge-dependent contribution to the stress-energy. Examination of the non-zero components of the Einstein tensor, outside the outer horizon (and inside the inner horizon), yields the following:

ε\displaystyle\varepsilon =Q2(r222)8π(r2+2)3+2(4mr2+2)8π(r2+2)5/2,\displaystyle=\frac{Q^{2}(r^{2}-2\ell^{2})}{8\pi(r^{2}+\ell^{2})^{3}}+\frac{\ell^{2}\left(4m-\sqrt{r^{2}+\ell^{2}}\right)}{8\pi(r^{2}+\ell^{2})^{5/2}}\ , (2.29a)
pr\displaystyle-p_{r} =Q2r28π(r2+2)3+28π(r2+2)2,\displaystyle=\frac{Q^{2}r^{2}}{8\pi(r^{2}+\ell^{2})^{3}}+\frac{\ell^{2}}{8\pi(r^{2}+\ell^{2})^{2}}\ , (2.29b)
pt\displaystyle p_{t} =Q2r28π(r2+2)3+2(r2+2m)8π(r2+2)5/2.\displaystyle=\frac{Q^{2}r^{2}}{8\pi(r^{2}+\ell^{2})^{3}}+\frac{\ell^{2}\left(\sqrt{r^{2}+\ell^{2}}-m\right)}{8\pi(r^{2}+\ell^{2})^{5/2}}\ . (2.29c)

For the radial null energy condition (NEC) [16, 17, 18, 19, 20, 24, 23, 22, 21, 25, 26], outside (inside) the outer (inner) horizon, we have:

ε+pr=2[r2+22mr2+2+Q2]4π(r2+2)3=2f(r)4π(r2+2)2.\displaystyle\varepsilon+p_{r}=-\frac{\ell^{2}\left[r^{2}+\ell^{2}-2m\sqrt{r^{2}+\ell^{2}}+Q^{2}\right]}{4\pi(r^{2}+\ell^{2})^{3}}=-\frac{\ell^{2}f(r)}{4\pi(r^{2}+\ell^{2})^{2}}\ . (2.30)

It is then clear that outside the outer horizon (or inside the inner one), where f(r)>0f(r)>0, we will have violation of the radial NEC. At the horizons, where f(r)=0f(r)=0, we always have (ε+pr)|H=0\left.\left(\varepsilon+p_{r}\right)\right|_{H}=0. (This on-horizon marginal satisfaction of the NEC is a quite generic phenomenon [46, 42, 47, 48].)

We can trivially extract the explicit form for [TQ]μ^ν^\left[T_{Q}\right]^{\hat{\mu}}{}_{\hat{\nu}}:

[TQ]μ^ν^\displaystyle\left[T_{Q}\right]^{\hat{\mu}}{}_{\hat{\nu}} =Q2r28π(r2+2)3diag(22r21,1,1,1)\displaystyle=\frac{Q^{2}r^{2}}{8\pi(r^{2}+\ell^{2})^{3}}\operatorname{diag}\left(\frac{2\ell^{2}}{r^{2}}-1,-1,1,1\right)
=Q2r28π(r2+2)3[diag(1,1,1,1)+diag(22r2,0,0,0)].\displaystyle=\frac{Q^{2}r^{2}}{8\pi(r^{2}+\ell^{2})^{3}}\left[\operatorname{diag}\left(-1,-1,1,1\right)+\operatorname{diag}\left(\frac{2\ell^{2}}{r^{2}},0,0,0\right)\right]\ . (2.31)

In the current situation, the first term above can be interpreted as the usual Maxwell stress-energy tensor [49]

[TMaxwell]μ^=ν^14π[Fμ^Fα^α^ν^14δμ^F2ν^],\left[T_{\text{Maxwell}}\right]^{\hat{\mu}}{}_{\hat{\nu}}=\frac{1}{4\pi}\left[-F^{\hat{\mu}}{}_{\hat{\alpha}}F^{\hat{\alpha}}{}_{\hat{\nu}}-\frac{1}{4}\delta^{\hat{\mu}}{}_{\hat{\nu}}F^{2}\right]\ , (2.32)

while the second term can be interpreted as the stress-energy of “charged dust”, with the density of the dust involving both the bounce parameter \ell and the total charge QQ. Overall we have

[TQ]μ^=ν^[TMaxwell]μ^+ν^ΞVμ^Vν^.\left[T_{Q}\right]^{\hat{\mu}}{}_{\hat{\nu}}=\left[T_{\text{Maxwell}}\right]^{\hat{\mu}}{}_{\hat{\nu}}+\Xi\;V^{\hat{\mu}}V_{\hat{\nu}}\ . (2.33)

The vector Vμ^V^{\hat{\mu}} is the normalised unit timelike eigenvector of the stress-energy, which in the current situation reduces to the normalised time-translation Killing vector, while the dust density Ξ\Xi has to be determined. We obtain

[TQ]t^=t^εem=[TMaxwell]t^t^Ξ=18πE2Ξ.\left[T_{Q}\right]^{\hat{t}}{}_{\hat{t}}=-\varepsilon_{\text{em}}=\left[T_{\text{Maxwell}}\right]^{\hat{t}}{}_{\hat{t}}-\Xi=-\frac{1}{8\pi}E^{2}-\Xi\ . (2.34)

Comparing with eq. (2.5), we find for the electric field strength EE

E\displaystyle E =Qr(r2+2)3/2=ERN[r3(r2+2)3/2],\displaystyle=\frac{Qr}{(r^{2}+\ell^{2})^{3/2}}=E_{\text{RN}}\left[\frac{r^{3}}{(r^{2}+\ell^{2})^{3/2}}\right]\ , (2.35)

where ERNE_{\text{RN}} is the electric field strength of a Reissner–Nordström black hole. For the density of the dust Ξ\Xi we find

Ξ=14πQ22(r2+2)3.\Xi=-\frac{1}{4\pi}\frac{Q^{2}\ell^{2}}{(r^{2}+\ell^{2})^{3}}\ . (2.36)

All told, we have the following form for the electromagnetic stress-energy tensor for our regularised Reissner–Nordström spacetime

[TQ]μ^=ν^14π[Fμ^Fα^α^ν^14δμ^F2ν^]14πQ22(r2+2)3Vμ^Vν^.\left[T_{Q}\right]^{\hat{\mu}}{}_{\hat{\nu}}=\frac{1}{4\pi}\left[-F^{\hat{\mu}}{}_{\hat{\alpha}}F^{\hat{\alpha}}{}_{\hat{\nu}}-\frac{1}{4}\delta^{\hat{\mu}}{}_{\hat{\nu}}F^{2}\right]-\frac{1}{4\pi}\frac{Q^{2}\ell^{2}}{(r^{2}+\ell^{2})^{3}}V^{\hat{\mu}}V_{\hat{\nu}}\ . (2.37)

Finally, the electromagnetic potential is easily extracted via integrating eq. (2.35), and in view of asymptotic flatness we may set the constant of integration to zero, yielding

Aμ=(Φem(r),0,0,0)=Qr2+2(1,0,0,0).A_{\mu}=(\Phi_{\text{em}}(r),0,0,0)=-\frac{Q}{\sqrt{r^{2}+\ell^{2}}}\;(1,0,0,0). (2.38)

Note that this really is simply the electromagnetic potential from standard Reissner–Nordström spacetime, Q/r-Q/r, under the map rr2+2r\to\sqrt{r^{2}+\ell^{2}}.

It is easy to verify that the electromagnetic field-strength tensor Fμν=μAννAμF_{\mu\nu}=\nabla_{\mu}A_{\nu}-\nabla_{\nu}A_{\mu} satisfies F[μν,σ]=0F_{[\mu\nu,\sigma]}=0. The inhomogeneous Maxwell equation is, using z=r2+2z=\sqrt{r^{2}+\ell^{2}}:

μ^Fμ^ν^\displaystyle\nabla^{\hat{\mu}}F_{\hat{\mu}\hat{\nu}} =Q2z6z22mz+Q2(1,0,0,0).\displaystyle={\frac{Q\ell^{2}}{z^{6}}\;\sqrt{z^{2}-2mz+Q^{2}}}\;\left(1,0,0,0\right). (2.39)

(We warn the reader that the situation will get somewhat messier once we add rotation.)

3 Black-bounce–Kerr–Newman geometry

As expected, the black-bounce–Kerr–Newman geometry is qualitatively more complicated (but also physically richer) than the black-bounce–Reissner–Nordström one. We start from Kerr–Newman geometry in standard Boyer–Lindquist coordinates [50, 51]

dsKN2=ΔKNρKN2(asin2θdϕdt)2+sin2θρKN2[(r2+a2)dϕadt]2+ρKN2ΔKNdr2+ρKN2dθ2,{\mathrm{d}}s^{2}_{\text{KN}}=-\frac{\Delta_{\text{KN}}}{\rho^{2}_{\text{KN}}}(a\sin^{2}\theta{\mathrm{d}}\phi-{\mathrm{d}}t)^{2}+\frac{\sin^{2}\theta}{\rho^{2}_{\text{KN}}}\left[(r^{2}+a^{2}){\mathrm{d}}\phi-a{\mathrm{d}}t\right]^{2}+\frac{\rho^{2}_{\text{KN}}}{\Delta_{\text{KN}}}{\mathrm{d}}r^{2}+\rho^{2}_{\text{KN}}{\mathrm{d}}\theta^{2}\ , (3.1)

where

ρKN2=r2+a2cos2θ;ΔKN=r2+a22mr+Q2.\rho^{2}_{\text{KN}}=r^{2}+a^{2}\cos^{2}\theta\ ;\qquad\Delta_{\text{KN}}=r^{2}+a^{2}-2mr+Q^{2}\ . (3.2)

Applying our procedure, our new candidate spacetime is given by the line element

ds2=Δρ2(asin2θdϕdt)2+sin2θρ2[(r2+2+a2)dϕadt]2+ρ2Δdr2+ρ2dθ2,{\mathrm{d}}s^{2}=-\frac{\Delta}{\rho^{2}}(a\sin^{2}\theta{\mathrm{d}}\phi-{\mathrm{d}}t)^{2}+\frac{\sin^{2}\theta}{\rho^{2}}\left[(r^{2}+\ell^{2}+a^{2}){\mathrm{d}}\phi-a{\mathrm{d}}t\right]^{2}+\frac{\rho^{2}}{\Delta}{\mathrm{d}}r^{2}+\rho^{2}{\mathrm{d}}\theta^{2}\ , (3.3)

where now ρ2\rho^{2} and Δ\Delta are modified:

ρ2=r2+2+a2cos2θ;Δ=r2+2+a22mr2+2+Q2.\rho^{2}=r^{2}+\ell^{2}+a^{2}\cos^{2}\theta\ ;\qquad\Delta=r^{2}+\ell^{2}+a^{2}-2m\sqrt{r^{2}+\ell^{2}}+Q^{2}\ . (3.4)

The natural domains of the angular and temporal coordinates are unaffected, while the radial coordinate again extends from the positive half line to the entire real line, and both manifest axisymmetry and asymptotic flatness are preserved. This spacetime is now stationary but not static, hence we must examine more than just the Kretschmann scalar to draw any conclusion as to regularity.

3.1 Curvature tensors

Again using z:=r2+2z:=\sqrt{r^{2}+\ell^{2}}, where this shorthand now stands for the equatorial value of the parameter ρ\rho, the Ricci scalar is given by

R=22m(ρ42z4)+Q2z3+z3(ρ22Δ)ρ6z3.R=2\ell^{2}\frac{m\left(\rho^{4}-2z^{4}\right)+Q^{2}z^{3}+z^{3}\left(\rho^{2}-2\Delta\right)}{\rho^{6}z^{3}}. (3.5)

It is clearly finite in the limit r0r\to 0, i.e. zz\to\ell. So are the Kretschmann scalar and the invariants RμνRμνR^{\mu\nu}R_{\mu\nu} and CμνρσCμνρσC^{\mu\nu\rho\sigma}C_{\mu\nu\rho\sigma}. The only potentially dangerous behaviour arises from the denominators, which, in the r0r\to 0 limit, take the form

12(2+a2cos2θ)6.\frac{1}{\ell^{2}(\ell^{2}+a^{2}\cos^{2}\theta)^{6}}. (3.6)

As long as 0\ell\neq 0, therefore, these quantities are never infinite.

We now turn our attention to the Einstein and Ricci tensors. We note that the (mixed) Einstein tensor GνμG^{\mu}_{\ \nu} can be diagonalised over the real numbers: its four eigenvectors {eμ^}μ=t,r,θ,ϕ\{e_{\hat{\mu}}\}_{\mu=t,r,\theta,\phi} form a globally defined tetrad [52] and have explicit Boyer–Lindquist components

(et^)μ\displaystyle\big{(}e_{\hat{t}}\big{)}^{\mu} =1ρ2|Δ|(r2+2+a2,0,0,a),(er^)μ=|Δ|ρ2(0,1,0,0),\displaystyle=\frac{1}{\sqrt{\rho^{2}\absolutevalue{\Delta}}}\left(r^{2}+\ell^{2}+a^{2},0,0,a\right),\quad\big{(}e_{\hat{r}}\big{)}^{\mu}=\sqrt{\frac{\absolutevalue{\Delta}}{\rho^{2}}}\left(0,1,0,0\right),
(eθ^)μ\displaystyle\big{(}e_{\hat{\theta}}\big{)}^{\mu} =1ρ2(0,0,1,0),(eϕ^)μ=1sinθρ2(asin2θ,0,0,1).\displaystyle=\frac{1}{\sqrt{\rho^{2}}}\left(0,0,1,0\right),\quad\big{(}e_{\hat{\phi}}\big{)}^{\mu}=\frac{1}{\sin\theta\sqrt{\rho^{2}}}\left(a\sin^{2}\theta,0,0,1\right). (3.7)

Eigenvectors are defined up to multiplicative, possibly dimensionful constants. This choice of normalisation ensures that the tetrad {eμ^}\{e_{\hat{\mu}}\} is orthonormal and reduces to eq. (2.3) in the limit a0a\to 0. We will therefore use the tetrad (3.7), along with the coordinate basis, to express components of tensors.

In particular, the components of the Einstein tensor are

Gt^t^\displaystyle G_{\hat{t}\hat{t}} =signΔ2[2m(z2ρ2)+z(ρ22Δ)]+Q2z(ρ22)ρ6z,\displaystyle=\operatorname{sign}\Delta\frac{\ell^{2}\left[2m\left(z^{2}-\rho^{2}\right)+z\left(\rho^{2}-2\Delta\right)\right]+Q^{2}z\left(\rho^{2}-\ell^{2}\right)}{\rho^{6}z}\ , (3.8a)
Gr^r^\displaystyle G_{\hat{r}\hat{r}} =signΔ2[2m(z2ρ2)+ρ2z]+Q2z(ρ22)ρ6z,\displaystyle=-\operatorname{sign}\Delta\frac{\ell^{2}\left[2m\left(z^{2}-\rho^{2}\right)+\rho^{2}z\right]+Q^{2}z\left(\rho^{2}-\ell^{2}\right)}{\rho^{6}z}\ , (3.8b)
Gθ^θ^\displaystyle G_{\hat{\theta}\hat{\theta}} =2[m(ρ42ρ2z2+2z4)+ρ2z3]+Q2z3(ρ22)ρ6z3,\displaystyle=\frac{\ell^{2}\left[m\left(-\rho^{4}-2\rho^{2}z^{2}+2z^{4}\right)+\rho^{2}z^{3}\right]+Q^{2}z^{3}\left(\rho^{2}-\ell^{2}\right)}{\rho^{6}z^{3}}\ , (3.8c)
Gϕ^ϕ^\displaystyle G_{\hat{\phi}\hat{\phi}} =2[m(ρ42ρ2z2+6z4)+z3(2Δρ2)]+Q2z3(ρ232)ρ6z3.\displaystyle=\frac{\ell^{2}\left[m\left(-\rho^{4}-2\rho^{2}z^{2}+6z^{4}\right)+z^{3}\left(2\Delta-\rho^{2}\right)\right]+Q^{2}z^{3}\left(\rho^{2}-3\ell^{2}\right)}{\rho^{6}z^{3}}\ . (3.8d)

We note that

Gt^t^+Gr^r^=signΔ22Δρ6G_{\hat{t}\hat{t}}+G_{\hat{r}\hat{r}}=-\operatorname{sign}\Delta\;\frac{2\ell^{2}\Delta}{\rho^{6}} (3.9)

which is well-behaved at Δ=0\Delta=0.

The Ricci tensor is clearly diagonal, too:

Rt^t^\displaystyle R_{\hat{t}\hat{t}} =signΔm2(ρ42ρ2z2+4z4)+Q2z3(ρ222)ρ6z3,\displaystyle=\operatorname{sign}\Delta\frac{m\ell^{2}\left(-\rho^{4}-2\rho^{2}z^{2}+4z^{4}\right)+Q^{2}z^{3}\left(\rho^{2}-2\ell^{2}\right)}{\rho^{6}z^{3}}\ , (3.10a)
Rr^r^\displaystyle R_{\hat{r}\hat{r}} =signΔ2(m(ρ4+2ρ2z24z4)2Δz3)+Q2z3(22ρ2)ρ6z3,\displaystyle=\operatorname{sign}\Delta\frac{\ell^{2}\left(m\left(\rho^{4}+2\rho^{2}z^{2}-4z^{4}\right)-2\Delta z^{3}\right)+Q^{2}z^{3}\left(2\ell^{2}-\rho^{2}\right)}{\rho^{6}z^{3}}\ , (3.10b)
Rθ^θ^\displaystyle R_{\hat{\theta}\hat{\theta}} =Q2ρ422ρ6[Δ+ρ2(mz)z],\displaystyle=\frac{Q^{2}}{\rho^{4}}-\frac{2\ell^{2}}{\rho^{6}}\left[\Delta+\frac{\rho^{2}(m-z)}{z}\right], (3.10c)
Rϕ^ϕ^\displaystyle R_{\hat{\phi}\hat{\phi}} =2m2(2z2ρ2)+Q2z(ρ222)ρ6z.\displaystyle=\frac{2m\ell^{2}\left(2z^{2}-\rho^{2}\right)+Q^{2}z\left(\rho^{2}-2\ell^{2}\right)}{\rho^{6}z}\ . (3.10d)

Similarly, we note that

Rt^t^+Rr^r^=signΔ22Δρ6R_{\hat{t}\hat{t}}+R_{\hat{r}\hat{r}}=-\operatorname{sign}\Delta\;\frac{2\ell^{2}\Delta}{\rho^{6}} (3.11)

which is well-behaved at Δ=0\Delta=0.

From these expressions one immediately notices that the curvature tensors are rational polynomials in the variable z=r2+2z=\sqrt{r^{2}+\ell^{2}}, which is strictly positive, and their denominators never vanish. The same is true for the Riemann and Weyl tensors; we thus conclude that the spacetime is free of curvature singularities.

3.2 Horizons, surface gravity and ergosurfaces

Horizons are now associated to the roots of Δ\Delta:

rH=S1(m+S2m2Q2a2)22,r_{H}=S_{1}\sqrt{\left(m+S_{2}\sqrt{m^{2}-Q^{2}-a^{2}}\right)^{2}-\ell^{2}}, (3.12)

where S1S_{1} and S2S_{2} are defined as in section 2.3. The spacetime structures corresponding to mm2a2Q2<<m+m2a2Q2m-\sqrt{m^{2}-a^{2}-Q^{2}}<\ell<m+\sqrt{m^{2}-a^{2}-Q^{2}} and <mm2a2Q2\ell<m-\sqrt{m^{2}-a^{2}-Q^{2}} are analogous to their non-spinning counterparts and are represented by the diagrams in figures 1 and 2.

In the Kerr–Newman geometry, one demands Q2+a2m2Q^{2}+a^{2}\leq m^{2} to avoid the possibility of naked singularities. In our case, we need not worry about this eventuality and may consider arbitrary values of spin and charge. Thus, if Q2+a2>m2Q^{2}+a^{2}>m^{2} or >m+S2m2Q2a2\ell>m+S_{2}\sqrt{m^{2}-Q^{2}-a^{2}}, the spacetime has no horizon.

If horizons are present, their surface gravity is given by

κS2:=12ddr(Δr2+2+a2)|rH=κS2KNrH2rH2+2,\kappa_{S_{2}}:=\frac{1}{2}\derivative{r}\evaluated{\bigg{(}\frac{\Delta}{r^{2}+\ell^{2}+a^{2}}\bigg{)}}_{r_{H}}=\kappa^{\text{KN}}_{S_{2}}\sqrt{\frac{r_{H}^{2}}{r_{H}^{2}+\ell^{2}}}, (3.13)

where κS2KN\kappa^{\text{KN}}_{S_{2}} is the surface gravity relative to the inner, when S2=1S_{2}=-1, or outer, when S2=+1S_{2}=+1, horizon of a Kerr–Newman black hole with mass mm, spin aa and charge QQ.

The ergosurface is determined by gtt=0g_{tt}=0, which is a quadratic equation in rr. The roots are given by:

rerg=S1(m+S2m2Q2a2cos2θ)22,r_{\text{erg}}=S_{1}\sqrt{\left(m+S_{2}\sqrt{m^{2}-Q^{2}-a^{2}\cos^{2}\theta}\right)^{2}-\ell^{2}}, (3.14)

where S1,S2S_{1},S_{2} are as before.

3.3 Geodesics and equatorial orbits

Consider a test particle with mass μ\mu, energy EE, component of angular momentum (per unit mass) along the rotation axis LzL_{z} and zero electric charge. Its trajectory xμ(τ)x^{\mu}(\tau) is governed by the following set of first-order differential equations (see e.g. ref. [53]):

ρ2dtdτ\displaystyle\rho^{2}\derivative{t}{\tau} =a(LzaEsin2θ)+(r2+2)+a2Δ[E(r2+2+a2)Lza],\displaystyle=a(L_{z}-aE\sin^{2}\theta)+\frac{(r^{2}+\ell^{2})+a^{2}}{\Delta}[E(r^{2}+\ell^{2}+a^{2})-L_{z}a], (3.15a)
ρ2drdτ\displaystyle\rho^{2}\derivative{r}{\tau} =±,\displaystyle=\pm\sqrt{\mathcal{R}}, (3.15b)
ρ2dθdτ\displaystyle\rho^{2}\derivative{\theta}{\tau} =±Θ,\displaystyle=\pm\sqrt{\Theta}, (3.15c)
ρ2dϕdτ\displaystyle\rho^{2}\derivative{\phi}{\tau} =Lzsin2θaE+aΔ[E(r2+2+a2)Lza],\displaystyle=\frac{L_{z}}{\sin^{2}\theta}-aE+\frac{a}{\Delta}[E(r^{2}+\ell^{2}+a^{2})-L_{z}a], (3.15d)

where

\displaystyle\mathcal{R} =[E(r2+2+a2)Lza]2Δ[μ2(r2+2)+(LzaE)2+𝒬],\displaystyle=[E(r^{2}+\ell^{2}+a^{2})-L_{z}a]^{2}-\Delta[\mu^{2}(r^{2}+\ell^{2})+(L_{z}-aE)^{2}+\mathcal{Q}], (3.16)
Θ\displaystyle\Theta =𝒬cos2θ[a2(μ2E2)+Lz2sin2θ],\displaystyle=\mathcal{Q}-\cos^{2}\theta\bigg{[}a^{2}(\mu^{2}-E^{2})+\frac{L_{z}^{2}}{\sin^{2}\theta}\bigg{]}, (3.17)

and 𝒬\mathcal{Q} is a generalised Carter constant associated to the existence of a Killing tensor discussed in section 3.4 below.

In view of the existence of the Killing tensor, there exist orbits that lie entirely on the equatorial plane θ=π/2\theta=\pi/2. Exploiting the conserved quantities, their motion is effectively one-dimensional and governed by the effective potential \mathcal{R}: circular orbits, in particular, are given by

=0andddr=0;\mathcal{R}=0\qquad\text{and}\qquad\derivative{\mathcal{R}}{r}=0; (3.18)

when, in addition,

d2dr2>0,\derivative[2]{\mathcal{R}}{r}>0, (3.19)

the orbits are stable.

Solutions to (3.18) can be easily found by exploiting known results on the Kerr–Newman geometry [51, 50, 54, 55]. Indeed, writing (3.16) in terms of z:=r2+2z:=\sqrt{r^{2}+\ell^{2}}, one immediately recognises the textbook result for a Kerr–Newman spacetime in which the Boyer–Lindquist radius has been given the uncommon name zz. Moreover,

ddr=dzdrddz,\derivative{\mathcal{R}}{r}=\derivative{z}{r}\;\derivative{\mathcal{R}}{z}, (3.20)

so

ddz=0ddr=0.\derivative{\mathcal{R}}{z}=0\quad\Longrightarrow\quad\derivative{\mathcal{R}}{r}=0. (3.21)

Furthermore, at the critical point

d2dr2=(dzdr)2d2dz2.\frac{{\mathrm{d}}^{2}\mathcal{R}}{{\mathrm{d}}r^{2}}=\left(\frac{{\mathrm{d}}z}{{\mathrm{d}}r}\right)^{2}\;\frac{{\mathrm{d}}^{2}\mathcal{R}}{{\mathrm{d}}z^{2}}. (3.22)

So stability (or lack thereof) is unaffected by the substitution rzr\longleftrightarrow z. Therefore, suppose z0z_{0} is such that

(z0)=0andd(z0)dz=0;\mathcal{R}(z_{0})=0\qquad\text{and}\qquad\derivative{\mathcal{R}(z_{0})}{z}=0; (3.23)

that is, suppose the Kerr–Newman spacetime has a circular orbit at radius z=z0z=z_{0}, then the black–bounce-Kerr–Newman spacetime has a circular orbit at r=r0:=z022r=r_{0}:=\sqrt{z_{0}^{2}-\ell^{2}}. Clearly, this mapping is allowed only if z0z_{0}\geq\ell.

Non-circular and non-equatorial orbits, instead, require a more thorough analysis.

3.4 Killing tensor and non-existence of the Killing tower

The existence of the generalised Carter constant 𝒬\mathcal{Q} introduced in the previous section is guaranteed by the fact that the tensor

Kμν=ρ2(lμnν+lνnμ)+(r2+2)gμνK_{\mu\nu}=\rho^{2}\left(l_{\mu}n_{\nu}+l_{\nu}n_{\mu}\right)+\left(r^{2}+\ell^{2}\right)g_{\mu\nu} (3.24)

is a Killing tensor; it is easy to explicitly check that K(μν;λ)=0K_{(\mu\nu;\lambda)}=0. Here

lμ=(r2+2+a2Δ,1,0,aΔ)andnμ=12ρ2(r2+2+a2,Δ,0,a)l^{\mu}=\left(\frac{r^{2}+\ell^{2}+a^{2}}{\Delta},1,0,\frac{a}{\Delta}\right)\quad\text{and}\quad n^{\mu}=\frac{1}{2\rho^{2}}\left(r^{2}+\ell^{2}+a^{2},-\Delta,0,a\right) (3.25)

are a pair of geodesic null vectors belonging to a generalised Kinnersley tetrad — see ref. [7].

Based on proposition 1.3 in [56], it has recently been established [57] that when one defines the Carter operator 𝒦Φ=μ(KμννΦ)\mathcal{K}\Phi=\nabla_{\mu}\left(K^{\mu\nu}\nabla_{\nu}\Phi\right) and wave operator Φ=μ(gμννΦ)\Box\Phi=\nabla_{\mu}\left(g^{\mu\nu}\nabla_{\nu}\Phi\right) one has

[𝒦,]Φ=23(μ[R,K]μ)ννΦ.\left[\mathcal{K},\Box\right]\Phi=\frac{2}{3}\left(\nabla_{\mu}\left[R,K\right]^{\mu}{}_{\nu}\right)\nabla^{\nu}\Phi\ . (3.26)

This operator commutator will certainly vanish when the tensor commutator [R,K]μ:=νRμKαανKμRααν\left[R,K\right]^{\mu}{}_{\nu}:=R^{\mu}{}_{\alpha}K^{\alpha}{}_{\nu}-K^{\mu}{}_{\alpha}R^{\alpha}{}_{\nu} vanishes, and this tensor commutator certainly vanishes for the black-bounce–Kerr–Newman spacetime considered herein. Hence the wave equation (not just the Hamilton–Jacobi equation) separates on the black-bounce–Kerr–Newman spacetime.

In the Kerr–Newman spacetime we started from, the Killing tensor is part of a “Killing tower” which ultimately descends from the existence of a closed conformal Killing–Yano tensor; called a principal tensor for short [58]. Such a principal tensor is a rank-2, antisymmetric tensor hμνh_{\mu\nu} satisfying (in four spacetime dimensions) the equation:

μhνα=13[gμνβhβαgμαβhβν].\nabla_{\mu}h_{\nu\alpha}=\frac{1}{3}\left[g_{\mu\nu}\nabla^{\beta}h_{\beta\alpha}-g_{\mu\alpha}\nabla^{\beta}h_{\beta\nu}\right]. (3.27)

In the language of forms, 𝐡\mathbf{h} is a non-degenerate two-form satisfying

𝐘𝐡=𝐘𝐗,𝐗=13𝐡\mathbf{\nabla_{Y}h}=\mathbf{Y}\wedge\mathbf{X},\qquad\mathbf{X}=\frac{1}{3}\divergence{\mathbf{h}} (3.28)

with 𝐘\mathbf{Y} any vector. (The equation above implies, incidentally, that 𝐡\mathbf{h} is closed: d𝐛=0\mathbf{\differential b}=0 so that locally 𝐡=d𝐛\mathbf{h}=\mathbf{\differential b}.) The Hodge dual of a principal tensor is a Killing–Yano tensor, i.e.

𝐟=𝐡is such thatμfνα+νfμα=0.\displaystyle\mathbf{f}=\mathbf{\ast h}\qquad\text{is such that}\qquad\nabla_{\mu}f_{\nu\alpha}+\nabla_{\nu}f_{\mu\alpha}=0. (3.29)

A Killing–Yano tensor, in turn, squares to a tensor

kμν:=fμαfναk_{\mu\nu}:=f_{\mu\alpha}f_{\nu}^{\ \alpha} (3.30)

that is a Killing tensor; k(μν;λ)=0k_{(\mu\nu;\lambda)}=0.

We may thus wonder whether the Killing tensor (3.24) derives from a principal tensor, as in the Kerr–Newman case. Naively, one may want to apply the usual trick rr2+2r\to\sqrt{r^{2}+\ell^{2}} to the Kerr–Newman principal tensor, or to the potential 𝐛\mathbf{b} (the two options are not equivalent). By adopting the first strategy, one finds a “would-be” Killing–Yano tensor that does indeed square to (3.24) but fails to satisfy eq. (3.29). The second approach also fails.

In fact, one can prove that no principal tensor can exist in this spacetime. The system (3.27) is overdetermined and has a solution only if a certain integrability condition is satisfied: This condition implies that the corresponding spacetime be of Petrov type D. However, three of the current authors proved, in ref. [7], that the black-bounce–Kerr spacetime is not algebraically special, hence neither can the black-bounce–Kerr–Newman spacetime be algebraically special. More prosaically, the non-existence of the Killing tower can be seen as a side effect of the fact that our black-bounce–Kerr–Newman geometry does not fall into Carter’s “off shell” 2-free-function distortion of Kerr [58].

For reference, here is the would-be Killing–Yano tensor:

fμν\displaystyle f_{\mu\nu} =(0acosθ00acosθ00a2cosθsin2θ00000a2cosθsin2θ00)\displaystyle=\begin{pmatrix}0&-a\cos\theta&0&0\\ a\cos\theta&0&0&-a^{2}\cos\theta\sin^{2}\theta\\ 0&0&0&0\\ 0&a^{2}\cos\theta\sin^{2}\theta&0&0\end{pmatrix}
+r2+2sinθ(00a00000a00(r2+2+a2)00(r2+2+a2)0).\displaystyle+\sqrt{r^{2}+\ell^{2}}\sin\theta\begin{pmatrix}0&0&a&0\\ 0&0&0&0\\ -a&0&0&(r^{2}+\ell^{2}+a^{2})\\ 0&0&-(r^{2}+\ell^{2}+a^{2})&0\end{pmatrix}. (3.31)

This would-be Killing–Yano tensor is taken from [58, eq. (3.22), p. 47], with coordinates changed to Boyer–Lindquist form, and with the substitution rr2+2r\to\sqrt{r^{2}+\ell^{2}} in the tensor components. It is easy to check that f2=Kf^{2}=K, but

(μfν)α=(r2+2r)×[tensor that is finite as 0].\nabla_{(\mu}f_{\nu)\alpha}=\left(\sqrt{r^{2}+\ell^{2}}-r\right)\times\left[\text{tensor that is finite as }\ell\to 0\right]. (3.32)

(This manifestly vanishes when 0\ell\to 0, as it should to recover the Killing–Yano tensor of the Kerr–Newman spacetime.) Its divergence is in fact particularly simple:

μfμν=[(r2+2r)2acosθρ2](1,0,0,0).\nabla_{\mu}f^{\mu\nu}=\left[\left(\sqrt{r^{2}+\ell^{2}}-r\right)\frac{2a\cos\theta}{\rho^{2}}\right]\left(1,0,0,0\right). (3.33)

(This again manifestly vanishes when 0\ell\to 0 as it should.)

Note that if one instead takes

bμdxμ=12(r2+2a2cos2θ)dt12[r22+(r2+2+a2)cos2θ]adϕ,b_{\mu}\differential x^{\mu}=-\frac{1}{2}(r^{2}+\ell^{2}-a^{2}\cos^{2}\theta)\differential t-\frac{1}{2}\left[-r^{2}-\ell^{2}+(r^{2}+\ell^{2}+a^{2})\cos^{2}\theta\right]a\differential\phi, (3.34)

as in ref. [58, eq. (3.21), p. 47], converted to Boyer–Lindquist coordinates, and subjected to the substitution rr2+2r\to\sqrt{r^{2}+\ell^{2}}, one finds

fμνμbννbμ.f_{\mu\nu}\neq\nabla_{\mu}b_{\nu}-\nabla_{\nu}b_{\mu}. (3.35)

(This is not surprising since derivatives are involved.)

Having now completed a purely geometrical treatment of the properties of our spacetime we can move on to discuss the implications of choosing a definite geometrodynamics, namely, as before, that of general relativity.

3.5 Stress-energy tensor

We may again exploit the orthonormal tetrad (3.7) to characterise the distribution of stress-energy in our spacetime. Assuming standard general relativity holds, the Einstein tensor is proportional to the stress-energy tensor: we may thus interpret the one component of Gμ^ν^G_{\hat{\mu}\hat{\nu}} that corresponds to the timelike direction as an energy density ε\varepsilon, and all the other non-zero components as principal pressures pip_{i}.

In particular, outside any horizon (technically, whenever Δ>0\Delta>0), we have

ε\displaystyle\varepsilon =2(2a2z+2mρ26mz2+2z3ρ2z)8πρ6zQ2(32ρ2)8πρ6,\displaystyle=-\frac{\ell^{2}\left(2a^{2}z+2m\rho^{2}-6mz^{2}+2z^{3}-\rho^{2}z\right)}{8\pi\rho^{6}z}-\frac{Q^{2}\left(3\ell^{2}-\rho^{2}\right)}{8\pi\rho^{6}}, (3.36a)
p1\displaystyle p_{1} =2[ρ2(2mz)2mz2]8πρ6z+Q2(2ρ2)8πρ6,\displaystyle=\frac{\ell^{2}\left[\rho^{2}(2m-z)-2mz^{2}\right]}{8\pi\rho^{6}z}+\frac{Q^{2}\left(\ell^{2}-\rho^{2}\right)}{8\pi\rho^{6}}, (3.36b)
p2\displaystyle p_{2} =2[mρ4+ρ2z2(z2m)+2mz4]8πρ6z3+Q2(ρ22)8πρ6,\displaystyle=\frac{\ell^{2}\left[-m\rho^{4}+\rho^{2}z^{2}(z-2m)+2mz^{4}\right]}{8\pi\rho^{6}z^{3}}+\frac{Q^{2}\left(\rho^{2}-\ell^{2}\right)}{8\pi\rho^{6}}, (3.36c)
p3\displaystyle p_{3} =2[2a2z3+m(ρ42ρ2z2+2z4)ρ2z3+2z5]8πρ6z3+Q2(ρ22)8πρ6.\displaystyle=\frac{\ell^{2}\left[2a^{2}z^{3}+m\left(-\rho^{4}-2\rho^{2}z^{2}+2z^{4}\right)-\rho^{2}z^{3}+2z^{5}\right]}{8\pi\rho^{6}z^{3}}+\frac{Q^{2}\left(\rho^{2}-\ell^{2}\right)}{8\pi\rho^{6}}. (3.36d)

The expressions above prove, incidentally, that our black-bounce–Kerr–Newman spacetime is Hawking–Ellis type I [59, 46, 60, 61].

Note that

ε+p1=2Δ8πρ6.\varepsilon+p_{1}=-\frac{\ell^{2}\Delta}{{8\pi}\rho^{6}}. (3.37)

This is the same result one gets in the black-bounce–Kerr spacetime, modulo the redefinition of Δ\Delta. Thus, in particular, the NEC is violated. Note that on the horizon Δ=0\Delta=0 so (ε+p1)|H=0\left.(\varepsilon+p_{1})\right|_{H}=0. This on-horizon simplification is a useful consistency check [46, 47, 48, 42].

We now wish to characterise the spacetime by means of some variant of curved-spacetime Maxwell-like electromagnetism, that is, by assuming that some variant of Maxwell’s equations hold. By doing so, as in the non-rotating case, we find that the matter content is made up of two different components: one electrically neutral and one charged, with the charged component further subdividing into Maxwell-like and charged dust contributions. If we isolate the QQ-dependent contribution to the total stress-energy, we find

[TQ]μ^=ν^18πQ2(ρ22)ρ6[diag(1,1,1,1)+22ρ22diag(1,0,0,0)].[T_{Q}]^{\hat{\mu}}{}_{\hat{\nu}}=\frac{1}{8\pi}\frac{Q^{2}\left(\rho^{2}-\ell^{2}\right)}{\rho^{6}}\left[\operatorname{diag}\left(-1,-1,1,1\right)+\frac{2\ell^{2}}{\rho^{2}{-\ell^{2}}}\operatorname{diag}\left(1,0,0,0\right)\right]. (3.38)

This is structurally the same as what we saw happening for the black-bounce–Reissner–Nordström spacetime, cfr. eq. (2.5), with the substitutions

Q2r2(r2+2)3Q2(ρ22)ρ6and22r222ρ22.\frac{Q^{2}r^{2}}{(r^{2}+\ell^{2})^{3}}\longleftrightarrow\frac{Q^{2}\left(\rho^{2}-\ell^{2}\right)}{\rho^{6}}\qquad\hbox{and}\qquad\frac{2\ell^{2}}{r^{2}}\longleftrightarrow\frac{2\ell^{2}}{\rho^{2}{-\ell^{2}}}. (3.39)

The first term in eq. (3.38) is structurally of the form of the Maxwell stress-energy tensor, and the second term is structurally of the form of charged dust. At first sight this seems to suggest that a similar treatment as the one presented for the black-bounce–Reissner–Nordström spacetime should lead to a consistent picture. However, we shall soon see below, that this case is quite trickier than the previous one.

3.5.1 Electromagnetic potential and field-strength tensor

The first step in carrying on the same interpretation for the stress-energy tensor as that applied in the black-bounce–Reissner–Nordström spacetime, is to introduce the electromagnetic potential. Of course, also in this case there is no obvious way to derive it, since we are not a priori specifying the equations of motion for the electromagnetic sector. Therefore, we shall choose to modify the Kerr–Newman potential in a minimal way as we did before for the Reissner–Nordström case, i.e. by performing the usual substitution rr2+2r\to\sqrt{r^{2}+\ell^{2}}. Thus, our proposal in Boyer–Lindquist coordinates reads

Aμ=Qr2+2ρ2(1,0,0,asin2θ).A_{\mu}=-\frac{Q\sqrt{r^{2}+\ell^{2}}}{\rho^{2}}\left(1,0,0,-a\sin^{2}\theta\right). (3.40)

In the orthonormal basis one has

Aμ^=eμ^Aνν=Qr2+2ρ2|Δ|(1,0,0,0).A_{\hat{\mu}}=e_{\hat{\mu}}{}^{\nu}\;A_{\nu}=-\frac{Q\sqrt{r^{2}+\ell^{2}}}{\sqrt{\rho^{2}|\Delta|}}\left(1,0,0,0\right). (3.41)

This is a minimal modification in the sense that, when we put a0a\to 0, the corresponding electrostatic potential is that of the black-bounce–Reissner–Nordström spacetime (2.38), and when 0\ell\to 0 we regain the usual result for standard Kerr–Newman. The potential (3.40) is also compatible with the Newman–Janis procedure as outlined in ref. [12], and as applied to the black-bounce–Reissner–Nordström geometry.

We can now compute the electromagnetic field-strength tensor FμνF_{\mu\nu}. In the orthonormal basis, its only non-zero components are

Ft^r^=Fr^t^\displaystyle F_{\hat{t}\hat{r}}=-F_{\hat{r}\hat{t}} =Qρ4r2r2+2(r2+2a2cos2θ),\displaystyle=-\frac{Q}{\rho^{4}}\sqrt{\frac{r^{2}}{r^{2}+\ell^{2}}}(r^{2}+\ell^{2}-a^{2}\cos^{2}\theta)\ , (3.42)
Fθ^ϕ^=Fϕ^θ^\displaystyle F_{\hat{\theta}\hat{\phi}}=-F_{\hat{\phi}\hat{\theta}} =2aQcosθr2+2ρ4.\displaystyle=\frac{2aQ\cos\theta\sqrt{r^{2}+\ell^{2}}}{\rho^{4}}\ . (3.43)

The homogeneous Maxwell equation is trivially satisfied F[μν,σ]=0F_{[\mu\nu,\sigma]}=0. For the inhomogeneous Maxwell equation we find

μ^Fμ^ν^\displaystyle\nabla^{\hat{\mu}}F_{\hat{\mu}\hat{\nu}} =Jν^=Q2ρ7z(Δ(ρ4+2ρ2z24z4)z2|Δ|,0,0,2asinθ(ρ22z2)).\displaystyle=J_{\hat{\nu}}=\frac{Q\ell^{2}}{\rho^{7}z}\left(-\frac{\Delta\left(\rho^{4}+2\rho^{2}z^{2}-4z^{4}\right)}{z^{2}\sqrt{|\Delta|}},0,0,{2a\sin\theta\left(\rho^{2}-2z^{2}\right)}{}\right). (3.44)

We interpret the right-hand side of eq. (3.44) as an effective electromagnetic source. Note that in terms of the (orthonormal) components of the electric and magnetic fields we have Er^=Ft^r^E_{\hat{r}}=F_{\hat{t}\hat{r}} and Br^=Fθ^ϕ^B_{\hat{r}}=F_{\hat{\theta}\hat{\phi}}. It is then easy to check that this implies that the Maxwell stress-energy tensor (2.32) is diagonal in this orthonormal basis and that

[TMaxwell]μ^=ν^Er^2+Br^28πdiag(1,1,1,1)\displaystyle\left[T_{\text{Maxwell}}\right]^{\hat{\mu}}{}_{{\hat{\nu}}}=\frac{E_{\hat{r}}^{2}+B_{\hat{r}}^{2}}{8\pi}\operatorname{diag}\left(-1,-1,1,1\right) (3.45)

independent of the specific values of Er^E_{\hat{r}} and Br^B_{\hat{r}}. It is also useful to check that

Er^2+Br^2=Q2ρ4r2r2+2+4Q22a2cos2θρ8.E_{\hat{r}}^{2}+B_{\hat{r}}^{2}=\frac{Q^{2}}{\rho^{4}}\frac{r^{2}}{r^{2}+\ell^{2}}+\frac{4Q^{2}\ell^{2}a^{2}\cos^{2}\theta}{\rho^{8}}. (3.46)

3.5.2 Interpreting the black-bounce–Kerr–Newman stress-energy

All of the above treatment is a relatively straightforward generalisation of the Reissner–Nordström case and also provides the correct limits for 0\ell\to 0 and/or a0a\to 0. However, when one attempts to interpret eq. (3.38) as the sum of the Maxwell stress-energy tensor (3.45) and a charged dust, an inconsistency appears in the form of extra terms. Assuming some generalisation of the energy density of the charged dust and working out the needed electromagnetic potential also does not lead to satisfactory results.

In what follows, we shall present two alternative interpretations of the stress-energy tensor, one based on a generalisation of the Maxwell dynamics to a non-linear one, the other consisting of a generalisation of the charged dust fluid to one with anisotropic pressure.

Nonlinear electrodynamics

An alternative to identifying a Maxwell stress-energy tensor in eq. (3.38) consists in generalising the decomposition of the charged part of the stress energy tensor adopted in the Reissner–Nordström case to

[TQ]μ^=ν^𝒜[TMaxwell]μ^+ν^ΞVμ^Vν^.\left[T_{Q}\right]^{\hat{\mu}}{}_{\hat{\nu}}={\mathcal{A}}\left[T_{\text{Maxwell}}\right]^{\hat{\mu}}{}_{\hat{\nu}}+\Xi\;V^{\hat{\mu}}V_{\hat{\nu}}. (3.47)

The multiplicative factor 𝒜{\mathcal{A}} will soon be seen to be position-dependent, and to depend on the spin parameter aa and regularisation parameter \ell, but to be independent of the total charge QQ. This sort of behaviour is strongly reminiscent of nonlinear electrodynamics (NLED) where quite generically one finds [TNLED]μ^ν^[TMaxwell]μ^ν^\left[T_{\text{NLED}}\right]^{\hat{\mu}}{}_{\hat{\nu}}\propto\left[T_{\text{Maxwell}}\right]^{\hat{\mu}}{}_{\hat{\nu}}. (For various proposals regarding the use of NLED in regular black hole contexts, see [62, 63, 64, 65, 66, 67, 68, 69].) The contribution ΞVμ^Vν^\Xi\;V^{\hat{\mu}}V_{\hat{\nu}} is again that appropriate to charged dust. The 4-velocity Vμ^V^{\hat{\mu}} is now the (non-geodesic) unit vector parallel to the timelike leg of the tetrad.

If we now compare eq. (3.45) with [TQ]μ^ν^\left[T_{Q}\right]^{\hat{\mu}}{}_{{\hat{\nu}}} as defined in (3.47) we identify

𝒜=Q2(ρ22)/ρ6Er^2+Br^2=ρ2(ρ22)(r2+2)ρ4r2+4a22(r2+2)cos2θ.{\mathcal{A}}=\frac{Q^{2}(\rho^{2}-\ell^{2})/\rho^{6}}{E_{\hat{r}}^{2}+B_{\hat{r}}^{2}}=\frac{\rho^{2}(\rho^{2}-\ell^{2})(r^{2}+\ell^{2})}{\rho^{4}r^{2}+4a^{2}\ell^{2}(r^{2}+\ell^{2})\cos^{2}\theta}\ . (3.48)

We note that at small \ell

𝒜=1a2cos2θ(3r2a2cos2θ)r2(r2+a2cos2θ)22+𝒪(4){\mathcal{A}}=1-\frac{a^{2}\cos^{2}\theta\left(3r^{2}-a^{2}\cos^{2}\theta\right)}{r^{2}(r^{2}+a^{2}\cos^{2}\theta)^{2}}\;\ell^{2}+{\mathcal{O}}(\ell^{4}) (3.49)

So in the limit as 0\ell\rightarrow 0, we see that 𝒜1\mathcal{A}\rightarrow 1, restoring standard Maxwell electromagnetism as would be expected for ordinary Kerr–Newman.

Also, we observe the large distance limit

𝒜=132a2cos2θr4+𝒪(r6).{\mathcal{A}}=1-\frac{3\ell^{2}a^{2}\cos^{2}\theta}{r^{4}}+{\mathcal{O}}(r^{-6}). (3.50)

That is, at sufficiently large distances, [TQ]μ^ν^\left[T_{Q}\right]^{\hat{\mu}}{}_{\hat{\nu}} can safely be approximated as a Maxwell-like contribution plus a charged dust, while at small rr we have

𝒜=2+a2cosθ242+𝒪(r2).{\mathcal{A}}=\frac{\ell^{2}+a^{2}\cos\theta^{2}}{4\ell^{2}}+{\mathcal{O}}(r^{2}). (3.51)

This indicates a simple rescaling of the Maxwell stress-energy, (similar to what happens in NLED), deep in the core of the black bounce.

Indeed, it is possible to further characterise the departure from Maxwell-like behaviour by decomposing 𝒜=1a22\mathcal{A}=1-a^{2}\ell^{2}\mathcal{F}, where

=cos2θ[4(r2+2)ρ2]ρ4r2+4a22cos2θ(r2+2).\mathcal{F}=\frac{\cos^{2}\theta\left[4(r^{2}+\ell^{2})-{\rho^{2}}\right]}{\rho^{4}r^{2}+4a^{2}\ell^{2}\cos^{2}\theta(r^{2}+\ell^{2})}\ . (3.52)

The motivation for doing so is to make utterly transparent the correct limiting behaviour for 𝒜\mathcal{A} both for a0a\rightarrow 0, and for 0\ell\rightarrow 0.

Anisotropic fluid

As an alternative to the NLED interpretation, we can instead generalise the pressureless dust fluid we had introduced in the Reissner–Nordström case and impose

[TQ]μ^ν^[TMaxwell]μ^ν^=diag(εf,pf,pf,pf),\left[T_{Q}\right]_{\hat{\mu}\hat{\nu}}-\left[T_{\text{Maxwell}}\right]_{\hat{\mu}\hat{\nu}}=\operatorname{diag}\left(\varepsilon_{f},-p_{f},p_{f},p_{f}\right), (3.53)

which can be satisfied if

εf=Q22z2ρ8(4z47z2ρ2+ρ4),pf=Q22z2ρ8(4z45z2ρ2+ρ4).\varepsilon_{f}=\frac{Q^{2}\ell^{2}}{z^{2}\rho^{8}}\left(4z^{4}-7z^{2}\rho^{2}+\rho^{4}\right),\quad p_{f}=\frac{Q^{2}\ell^{2}}{z^{2}\rho^{8}}\left(4z^{4}-5z^{2}\rho^{2}+\rho^{4}\right). (3.54)

This implies that the right-hand side of eq. (3.53) can be interpreted, formally, as the stress-energy of an anisotropic fluid. Specifically, it can be written as

εfVμ^Vν^+pf3(gμ^ν^+Vμ^Vν^)+πμ^ν^\varepsilon_{f}V_{\hat{\mu}}V_{\hat{\nu}}+\frac{p_{f}}{3}\left(g_{\hat{\mu}\hat{\nu}}+V_{\hat{\mu}}V_{\hat{\nu}}\right)+\pi_{\hat{\mu}\hat{\nu}} (3.55)

with Vμ^=(1,0,0,0)V^{\hat{\mu}}=\left(1,0,0,0\right)i.e. (et^)μ^(e_{\hat{t}})^{\hat{\mu}} — the velocity of the fluid and

πμ^ν^=2pf3diag(0,2,1,1)\pi_{\hat{\mu}\hat{\nu}}=\frac{2p_{f}}{3}\operatorname{diag}\left(0,-2,1,1\right) (3.56)

the (traceless) anisotropic shear [59]. Note that

pf(4z45z2ρ2+ρ4)=(4z2ρ2)(z2ρ2)a2cos2θ.p_{f}\propto\left(4z^{4}-5z^{2}\rho^{2}+\rho^{4}\right)=\left(4z^{2}-\rho^{2}\right)\left(z^{2}-\rho^{2}\right)\propto a^{2}\cos^{2}\theta. (3.57)

So in the limit a0a\to 0 this anisotropic fluid reduces to the usual charged dust.

4 Discussion and conclusions

We have seen that adding an electromagnetic charge to the black-bounce–Schwarzschild and black-bounce–Kerr spacetimes leads to the black-bounce–Reissner–Nordström and black-bounce–Kerr–Newman spacetimes, which are minimalist, one-parameter, deformations of the entire Kerr–Newman family which have the desirable properties that they simultaneously (i) pass all weak-field observational tests, (ii) are globally regular (no curvature singularities), and (iii) neatly interpolate between regular black holes and charged traversable wormholes.

While adding an electromagnetic charge to the black-bounce–Schwarzschild and black-bounce–Kerr spacetimes in this manner is maybe not of direct astrophysical importance (since in any plausible astrophysical situation |Q|/m1|Q|/m\ll 1), it is of considerable theoretical importance, as it gives us an entirely new class of relatively clean everywhere regular black holes to work with. Indeed, we have seen that such geometries present interesting theoretical features such as the existence of a Killing tensor without the presence of the full Killing tower (principal tensor, Killing–Yano tensor) or the fact that the charge-dependent component of the stress-energy for the black-bounce–Kerr–Newman spacetime has a rather non-trivial physical interpretation. In particular, we found that it can either be interpreted as charged dust together with a non-linear modification to standard Maxwell electromagnetism, or as standard Maxwell electromagnetism together with an anisotropic fluid.

While there is no simple way to remove this ambiguity, we can however notice that it appears at least problematic to justify from a physical point of view the introduction of a NLED for the black-bounce–Kerr–Newman spacetime, given that the latter is not required for consistency with general relativity of the black-bounce–Reissner–Nordström or the black-bounce–Kerr spacetime discussed in ref. [7]. This seems to suggest that the anisotropic fluid interpretation might be more natural. Overall, we hope that this work will help the understanding of the rich structure of these families of black hole mimickers and stimulate further investigations about their possible realisation in physical settings.

Acknowledgments

EF, SL, and JM acknowledge funding from the Italian Ministry of Education and Scientific Research (MIUR) under the grant PRIN MIUR 2017-MB8AEZ.

AS acknowledges financial support via a PhD Doctoral Scholarship provided by Victoria University of Wellington. AS is also indirectly supported by the Marsden fund, administered by the Royal Society of New Zealand.

MV was directly supported by the Marsden Fund, via a grant administered by the Royal Society of New Zealand.

Appendix A Other curvature invariants for the black-bounce–Reissner–Nordström geometry

Here we display the explicit forms for the various non-zero curvature invariants for the black-bounce–Reissner–Nordström geometry. Notably, all non-zero curvature invariants (and orthonormal tensor components) are globally finite, as was immediately guaranteed via examination of the Kretschmann scalar in section 2.1.

For the Ricci scalar

R=22(r2+2)3[Q2+r2+23mr2+2],R=-\frac{2\ell^{2}}{(r^{2}+\ell^{2})^{3}}\left[Q^{2}+r^{2}+\ell^{2}-3m\sqrt{r^{2}+\ell^{2}}\right]\ , (A.1)

and as r0r\rightarrow 0

limr0R=2(Q2+23m)4.\lim_{r\rightarrow 0}R=-\frac{2(Q^{2}+\ell^{2}-3m\ell)}{\ell^{4}}\ . (A.2)

The quadratic Ricci invariant RμνRμνR^{\mu\nu}R_{\mu\nu} is

RμνRμν\displaystyle R^{\mu\nu}R_{\mu\nu} =4(r42r2+4)(r2+2)6Q4+42[m(r242)+(r2+2)3/2](r2+2)11/2Q2\displaystyle=\frac{4\left(r^{4}-\ell^{2}r^{2}+\ell^{4}\right)}{(r^{2}+\ell^{2})^{6}}Q^{4}+\frac{4\ell^{2}\left[m(r^{2}-4\ell^{2})+(r^{2}+\ell^{2})^{3/2}\right]}{(r^{2}+\ell^{2})^{11/2}}Q^{2}
+24[2r46m(r2+2)3/2+(9m2+42)(r2+2)24](r2+2)6,\displaystyle+\frac{2\ell^{4}\left[2r^{4}-6m(r^{2}+\ell^{2})^{3/2}+(9m^{2}+4\ell^{2})(r^{2}+\ell^{2})-2\ell^{4}\right]}{(r^{2}+\ell^{2})^{6}}\ , (A.3)

and as r0r\rightarrow 0 we see

limr0RμνRμν=48Q4+4(4m)7Q2+4212m+18m26.\lim_{r\rightarrow 0}R^{\mu\nu}R_{\mu\nu}=\frac{4}{\ell^{8}}\,Q^{4}+\frac{4(\ell-4m)}{\ell^{7}}\,Q^{2}+\frac{4\ell^{2}-12m\ell+18m^{2}}{\ell^{6}}\ . (A.4)

For the Weyl contraction CμνρσCμνρσC^{\mu\nu\rho\sigma}C_{\mu\nu\rho\sigma} we obtain

CμνρσCμνρσ\displaystyle C^{\mu\nu\rho\sigma}C_{\mu\nu\rho\sigma}
=16(3r22)23(r2+2)6Q4+16(23r2)[22r2+23m(22r2)]3(r2+2)11/2Q2+48m2r6(r2+2)6\displaystyle\quad=\frac{16(3r^{2}-\ell^{2})^{2}}{3(r^{2}+\ell^{2})^{6}}Q^{4}+\frac{16(\ell^{2}-3r^{2})\left[2\ell^{2}\sqrt{r^{2}+\ell^{2}}-3m(\ell^{2}-2r^{2})\right]}{3(r^{2}+\ell^{2})^{11/2}}Q^{2}+\frac{48m^{2}r^{6}}{(r^{2}+\ell^{2})^{6}}
+44[44+(9m2+8r2)2+4r427m2r212mr2+2(1+r22)(22r2)]3(r2+2)6,\displaystyle\quad+\frac{4\ell^{4}\left[4\ell^{4}+(9m^{2}+8r^{2})\ell^{2}+4r^{4}-27m^{2}r^{2}-12m\sqrt{r^{2}+\ell^{2}}(1+\frac{r^{2}}{\ell^{2}})(\ell^{2}-2r^{2})\right]}{3(r^{2}+\ell^{2})^{6}}\ , (A.5)

and as r0r\rightarrow 0

limr0CμνρσCμνρσ=4(2Q2+223m)238.\lim_{r\rightarrow 0}C^{\mu\nu\rho\sigma}C_{\mu\nu\rho\sigma}=\frac{4\left(2Q^{2}+2\ell^{2}-3\ell m\right)^{2}}{3\ell^{8}}\ . (A.6)

References