Reissner–Nordström metric

Spherically symmetric metric with electric charge
General relativity
Spacetime curvature schematic
G μ ν + Λ g μ ν = κ T μ ν {\displaystyle G_{\mu \nu }+\Lambda g_{\mu \nu }={\kappa }T_{\mu \nu }}
Fundamental concepts
  • Equivalence principle
  • Special relativity
  • World line
  • Pseudo-Riemannian manifold
Phenomena
  • Kepler problem
  • Gravitational lensing
  • Gravitational redshift
  • Gravitational time dilation
  • Gravitational waves
  • Frame-dragging
  • Geodetic effect
  • Event horizon
  • Singularity
  • Black hole
Spacetime
  • icon Physics portal
  •  Category
  • v
  • t
  • e

In physics and astronomy, the Reissner–Nordström metric is a static solution to the Einstein–Maxwell field equations, which corresponds to the gravitational field of a charged, non-rotating, spherically symmetric body of mass M. The analogous solution for a charged, rotating body is given by the Kerr–Newman metric.

The metric was discovered between 1916 and 1921 by Hans Reissner,[1] Hermann Weyl,[2] Gunnar Nordström[3] and George Barker Jeffery[4] independently.[5]

The metric

In spherical coordinates ( t , r , θ , φ ) {\displaystyle (t,r,\theta ,\varphi )} , the Reissner–Nordström metric (i.e. the line element) is d s 2 = c 2 d τ 2 = ( 1 r s r + r Q 2 r 2 ) c 2 d t 2 ( 1 r s r + r Q 2 r 2 ) 1 d r 2 r 2 d θ 2 r 2 sin 2 θ d φ 2 , {\displaystyle ds^{2}=c^{2}\,d\tau ^{2}=\left(1-{\frac {r_{\text{s}}}{r}}+{\frac {r_{\rm {Q}}^{2}}{r^{2}}}\right)c^{2}\,dt^{2}-\left(1-{\frac {r_{\text{s}}}{r}}+{\frac {r_{Q}^{2}}{r^{2}}}\right)^{-1}\,dr^{2}-r^{2}\,d\theta ^{2}-r^{2}\sin ^{2}\theta \,d\varphi ^{2},}

  • Where c {\displaystyle c} is the speed of light.
  • τ {\displaystyle \tau } is the proper time.
  • t {\displaystyle t} is the time coordinate (measured by a stationary clock at infinity).
  • r {\displaystyle r} is the radial coordinate.
  • ( θ , φ ) {\displaystyle (\theta ,\varphi )} are the spherical angles.
  • r s {\displaystyle r_{\text{s}}} is the Schwarzschild radius of the body given by

r s = 2 G M c 2 , {\displaystyle r_{\text{s}}={\frac {2GM}{c^{2}}},} .

  • r Q {\displaystyle r_{Q}} is a characteristic length scale given by

r Q 2 = Q 2 G 4 π ε 0 c 4 . {\displaystyle r_{Q}^{2}={\frac {Q^{2}G}{4\pi \varepsilon _{0}c^{4}}}.}

The total mass of the central body and its irreducible mass are related by[6][7] M i r r = c 2 G r + 2 2     M = Q 2 16 π ε 0 G M i r r + M i r r . {\displaystyle M_{\rm {irr}}={\frac {c^{2}}{G}}{\sqrt {\frac {r_{+}^{2}}{2}}}\ \to \ M={\frac {Q^{2}}{16\pi \varepsilon _{0}GM_{\rm {irr}}}}+M_{\rm {irr}}.}

The difference between M {\displaystyle M} and M i r r {\displaystyle M_{\rm {irr}}} is due to the equivalence of mass and energy, which makes the electric field energy also contribute to the total mass.

In the limit that the charge Q {\displaystyle Q} (or equivalently, the length scale r Q {\displaystyle r_{Q}} ) goes to zero, one recovers the Schwarzschild metric. The classical Newtonian theory of gravity may then be recovered in the limit as the ratio r s / r {\displaystyle r_{\text{s}}/r} goes to zero. In the limit that both r Q / r {\displaystyle r_{Q}/r} and r s / r {\displaystyle r_{\text{s}}/r} go to zero, the metric becomes the Minkowski metric for special relativity.

In practice, the ratio r s / r {\displaystyle r_{\text{s}}/r} is often extremely small. For example, the Schwarzschild radius of the Earth is roughly 9 mm (3/8 inch), whereas a satellite in a geosynchronous orbit has an orbital radius r {\displaystyle r} that is roughly four billion times larger, at 42,164 km (26,200 miles). Even at the surface of the Earth, the corrections to Newtonian gravity are only one part in a billion. The ratio only becomes large close to black holes and other ultra-dense objects such as neutron stars.

Charged black holes

Although charged black holes with rQ ≪ rs are similar to the Schwarzschild black hole, they have two horizons: the event horizon and an internal Cauchy horizon.[8] As with the Schwarzschild metric, the event horizons for the spacetime are located where the metric component g r r {\displaystyle g_{rr}} diverges; that is, where 1 r s r + r Q 2 r 2 = 1 g r r = 0. {\displaystyle 1-{\frac {r_{\rm {s}}}{r}}+{\frac {r_{\rm {Q}}^{2}}{r^{2}}}=-{\frac {1}{g_{rr}}}=0.}

This equation has two solutions: r ± = 1 2 ( r s ± r s 2 4 r Q 2 ) . {\displaystyle r_{\pm }={\frac {1}{2}}\left(r_{\rm {s}}\pm {\sqrt {r_{\rm {s}}^{2}-4r_{\rm {Q}}^{2}}}\right).}

These concentric event horizons become degenerate for 2rQ = rs, which corresponds to an extremal black hole. Black holes with 2rQ > rs cannot exist in nature because if the charge is greater than the mass there can be no physical event horizon (the term under the square root becomes negative).[9] Objects with a charge greater than their mass can exist in nature, but they can not collapse down to a black hole, and if they could, they would display a naked singularity.[10] Theories with supersymmetry usually guarantee that such "superextremal" black holes cannot exist.

The electromagnetic potential is A α = ( Q / r , 0 , 0 , 0 ) . {\displaystyle A_{\alpha }=(Q/r,0,0,0).}

If magnetic monopoles are included in the theory, then a generalization to include magnetic charge P is obtained by replacing Q2 by Q2 + P2 in the metric and including the term P cos θ  in the electromagnetic potential.[clarification needed]

Gravitational time dilation

The gravitational time dilation in the vicinity of the central body is given by γ = | g t t | = r 2 Q 2 + ( r 2 M ) r , {\displaystyle \gamma ={\sqrt {|g^{tt}|}}={\sqrt {\frac {r^{2}}{Q^{2}+(r-2M)r}}},} which relates to the local radial escape velocity of a neutral particle v e s c = γ 2 1 γ . {\displaystyle v_{\rm {esc}}={\frac {\sqrt {\gamma ^{2}-1}}{\gamma }}.}

Christoffel symbols

The Christoffel symbols Γ j k i = s = 0 3   g i s 2 ( g j s x k + g s k x j g j k x s ) {\displaystyle \Gamma _{jk}^{i}=\sum _{s=0}^{3}\ {\frac {g^{is}}{2}}\left({\frac {\partial g_{js}}{\partial x^{k}}}+{\frac {\partial g_{sk}}{\partial x^{j}}}-{\frac {\partial g_{jk}}{\partial x^{s}}}\right)} with the indices { 0 ,   1 ,   2 ,   3 } { t ,   r ,   θ ,   φ } {\displaystyle \{0,\ 1,\ 2,\ 3\}\to \{t,\ r,\ \theta ,\ \varphi \}} give the nonvanishing expressions Γ t r t = M r Q 2 r ( Q 2 + r 2 2 M r ) Γ t t r = ( M r Q 2 ) ( r 2 2 M r + Q 2 ) r 5 Γ r r r = Q 2 M r r ( Q 2 2 M r + r 2 ) Γ θ θ r = r 2 2 M r + Q 2 r Γ φ φ r = sin 2 θ ( r 2 2 M r + Q 2 ) r Γ θ r θ = 1 r Γ φ φ θ = sin θ cos θ Γ φ r φ = 1 r Γ φ θ φ = cot θ {\displaystyle {\begin{aligned}\Gamma _{tr}^{t}&={\frac {Mr-Q^{2}}{r(Q^{2}+r^{2}-2Mr)}}\\[6pt]\Gamma _{tt}^{r}&={\frac {(Mr-Q^{2})\left(r^{2}-2Mr+Q^{2}\right)}{r^{5}}}\\[6pt]\Gamma _{rr}^{r}&={\frac {Q^{2}-Mr}{r(Q^{2}-2Mr+r^{2})}}\\[6pt]\Gamma _{\theta \theta }^{r}&=-{\frac {r^{2}-2Mr+Q^{2}}{r}}\\[6pt]\Gamma _{\varphi \varphi }^{r}&=-{\frac {\sin ^{2}\theta \left(r^{2}-2Mr+Q^{2}\right)}{r}}\\[6pt]\Gamma _{\theta r}^{\theta }&={\frac {1}{r}}\\[6pt]\Gamma _{\varphi \varphi }^{\theta }&=-\sin \theta \cos \theta \\[6pt]\Gamma _{\varphi r}^{\varphi }&={\frac {1}{r}}\\[6pt]\Gamma _{\varphi \theta }^{\varphi }&=\cot \theta \end{aligned}}}

Given the Christoffel symbols, one can compute the geodesics of a test-particle.[11][12]

Tetrad form

Instead of working in the holonomic basis, one can perform efficient calculations with a tetrad.[13] Let e I = e μ I {\displaystyle {\bf {e}}_{I}=e_{\mu I}} be a set of one-forms with internal Minkowski index I { 0 , 1 , 2 , 3 } {\displaystyle I\in \{0,1,2,3\}} , such that η I J e μ I e ν J = g μ ν {\displaystyle \eta ^{IJ}e_{\mu I}e_{\nu J}=g_{\mu \nu }} . The Reissner metric can be described by the tetrad

e 0 = G 1 / 2 d t {\displaystyle {\bf {e}}_{0}=G^{1/2}\,dt} ,
e 1 = G 1 / 2 d r {\displaystyle {\bf {e}}_{1}=G^{-1/2}\,dr} ,
e 2 = r d θ {\displaystyle {\bf {e}}_{2}=r\,d\theta }
e 3 = r sin θ d φ {\displaystyle {\bf {e}}_{3}=r\sin \theta \,d\varphi }

where G ( r ) = 1 r s r 1 + r Q 2 r 2 {\displaystyle G(r)=1-r_{s}r^{-1}+r_{Q}^{2}r^{-2}} . The parallel transport of the tetrad is captured by the connection one-forms ω I J = ω J I = ω μ I J = e I ν μ e J ν {\displaystyle {\boldsymbol {\omega }}_{IJ}=-{\boldsymbol {\omega }}_{JI}=\omega _{\mu IJ}=e_{I}^{\nu }\nabla _{\mu }e_{J\nu }} . These have only 24 independent components compared to the 40 components of Γ μ ν λ {\displaystyle \Gamma _{\mu \nu }^{\lambda }} . The connections can be solved for by inspection from Cartan's equation d e I = e J ω I J {\displaystyle d{\bf {e}}_{I}={\bf {e}}^{J}\wedge {\boldsymbol {\omega }}_{IJ}} , where the left hand side is the exterior derivative of the tetrad, and the right hand side is a wedge product.

ω 10 = 1 2 r G d t {\displaystyle {\boldsymbol {\omega }}_{10}={\frac {1}{2}}\partial _{r}G\,dt}
ω 20 = ω 30 = 0 {\displaystyle {\boldsymbol {\omega }}_{20}={\boldsymbol {\omega }}_{30}=0}
ω 21 = G 1 / 2 d θ {\displaystyle {\boldsymbol {\omega }}_{21}=-G^{1/2}\,d\theta }
ω 31 = sin θ G 1 / 2 d φ {\displaystyle {\boldsymbol {\omega }}_{31}=-\sin \theta G^{1/2}d\varphi }
ω 32 = cos θ d φ {\displaystyle {\boldsymbol {\omega }}_{32}=-\cos \theta \,d\varphi }

The Riemann tensor R I J = R μ ν I J {\displaystyle {\bf {R}}_{IJ}=R_{\mu \nu IJ}} can be constructed as a collection of two-forms by the second Cartan equation R I J = d ω I J + ω I K ω K J , {\displaystyle {\bf {R}}_{IJ}=d{\boldsymbol {\omega }}_{IJ}+{\boldsymbol {\omega }}_{IK}\wedge {\boldsymbol {\omega }}^{K}{}_{J},} which again makes use of the exterior derivative and wedge product. This approach is significantly faster than the traditional computation with Γ μ ν λ {\displaystyle \Gamma _{\mu \nu }^{\lambda }} ; note that there are only four nonzero ω I J {\displaystyle {\boldsymbol {\omega }}_{IJ}} compared with nine nonzero components of Γ μ ν λ {\displaystyle \Gamma _{\mu \nu }^{\lambda }} .

Equations of motion

[14]

Because of the spherical symmetry of the metric, the coordinate system can always be aligned in a way that the motion of a test-particle is confined to a plane, so for brevity and without restriction of generality we use θ instead of φ. In dimensionless natural units of G = M = c = K = 1 the motion of an electrically charged particle with the charge q is given by x ¨ i = j = 0 3   k = 0 3   Γ j k i   x ˙ j   x ˙ k + q   F i k   x ˙ k {\displaystyle {\ddot {x}}^{i}=-\sum _{j=0}^{3}\ \sum _{k=0}^{3}\ \Gamma _{jk}^{i}\ {{\dot {x}}^{j}}\ {{\dot {x}}^{k}}+q\ {F^{ik}}\ {{\dot {x}}_{k}}} which yields t ¨ =   2 ( Q 2 M r ) r ( r 2 2 M r + Q 2 ) r ˙ t ˙ + q Q ( r 2 2 m r + Q 2 )   r ˙ {\displaystyle {\ddot {t}}={\frac {\ 2(Q^{2}-Mr)}{r(r^{2}-2Mr+Q^{2})}}{\dot {r}}{\dot {t}}+{\frac {qQ}{(r^{2}-2mr+Q^{2})}}\ {\dot {r}}} r ¨ = ( r 2 2 M r + Q 2 ) ( Q 2 M r )   t ˙ 2 r 5 + ( M r Q 2 ) r ˙ 2 r ( r 2 2 M r + Q 2 ) + ( r 2 2 M r + Q 2 )   θ ˙ 2 r + q Q ( r 2 2 m r + Q 2 ) r 4   t ˙ {\displaystyle {\ddot {r}}={\frac {(r^{2}-2Mr+Q^{2})(Q^{2}-Mr)\ {\dot {t}}^{2}}{r^{5}}}+{\frac {(Mr-Q^{2}){\dot {r}}^{2}}{r(r^{2}-2Mr+Q^{2})}}+{\frac {(r^{2}-2Mr+Q^{2})\ {\dot {\theta }}^{2}}{r}}+{\frac {qQ(r^{2}-2mr+Q^{2})}{r^{4}}}\ {\dot {t}}} θ ¨ = 2   θ ˙   r ˙ r . {\displaystyle {\ddot {\theta }}=-{\frac {2\ {\dot {\theta }}\ {\dot {r}}}{r}}.}

All total derivatives are with respect to proper time a ˙ = d a d τ {\displaystyle {\dot {a}}={\frac {da}{d\tau }}} .

Constants of the motion are provided by solutions S ( t , t ˙ , r , r ˙ , θ , θ ˙ , φ , φ ˙ ) {\displaystyle S(t,{\dot {t}},r,{\dot {r}},\theta ,{\dot {\theta }},\varphi ,{\dot {\varphi }})} to the partial differential equation[15] 0 = t ˙ S t + r ˙ S r + θ ˙ S θ + t ¨ S t ˙ + r ¨ S r ˙ + θ ¨ S θ ˙ {\displaystyle 0={\dot {t}}{\dfrac {\partial S}{\partial t}}+{\dot {r}}{\frac {\partial S}{\partial r}}+{\dot {\theta }}{\frac {\partial S}{\partial \theta }}+{\ddot {t}}{\frac {\partial S}{\partial {\dot {t}}}}+{\ddot {r}}{\frac {\partial S}{\partial {\dot {r}}}}+{\ddot {\theta }}{\frac {\partial S}{\partial {\dot {\theta }}}}} after substitution of the second derivatives given above. The metric itself is a solution when written as a differential equation S 1 = 1 = ( 1 r s r + r Q 2 r 2 ) c 2 t ˙ 2 ( 1 r s r + r Q 2 r 2 ) 1 r ˙ 2 r 2 θ ˙ 2 . {\displaystyle S_{1}=1=\left(1-{\frac {r_{s}}{r}}+{\frac {r_{\rm {Q}}^{2}}{r^{2}}}\right)c^{2}\,{\dot {t}}^{2}-\left(1-{\frac {r_{s}}{r}}+{\frac {r_{Q}^{2}}{r^{2}}}\right)^{-1}\,{\dot {r}}^{2}-r^{2}\,{\dot {\theta }}^{2}.}

The separable equation S r 2 r θ ˙ S θ ˙ = 0 {\displaystyle {\frac {\partial S}{\partial r}}-{\frac {2}{r}}{\dot {\theta }}{\frac {\partial S}{\partial {\dot {\theta }}}}=0} immediately yields the constant relativistic specific angular momentum S 2 = L = r 2 θ ˙ ; {\displaystyle S_{2}=L=r^{2}{\dot {\theta }};} a third constant obtained from S r 2 ( M r Q 2 ) r ( r 2 2 M r + Q 2 ) t ˙ S t ˙ = 0 {\displaystyle {\frac {\partial S}{\partial r}}-{\frac {2(Mr-Q^{2})}{r(r^{2}-2Mr+Q^{2})}}{\dot {t}}{\frac {\partial S}{\partial {\dot {t}}}}=0} is the specific energy (energy per unit rest mass)[16] S 3 = E = t ˙ ( r 2 2 M r + Q 2 ) r 2 + q Q r . {\displaystyle S_{3}=E={\frac {{\dot {t}}(r^{2}-2Mr+Q^{2})}{r^{2}}}+{\frac {qQ}{r}}.}

Substituting S 2 {\displaystyle S_{2}} and S 3 {\displaystyle S_{3}} into S 1 {\displaystyle S_{1}} yields the radial equation c d τ = r 2 d r r 4 ( E 1 ) + 2 M r 3 ( Q 2 + L 2 ) r 2 + 2 M L 2 r Q 2 L 2 . {\displaystyle c\int d\,\tau =\int {\frac {r^{2}\,dr}{\sqrt {r^{4}(E-1)+2Mr^{3}-(Q^{2}+L^{2})r^{2}+2ML^{2}r-Q^{2}L^{2}}}}.}

Multiplying under the integral sign by S 2 {\displaystyle S_{2}} yields the orbital equation c L r 2 d θ = L d r r 4 ( E 1 ) + 2 M r 3 ( Q 2 + L 2 ) r 2 + 2 M L 2 r Q 2 L 2 . {\displaystyle c\int Lr^{2}\,d\theta =\int {\frac {L\,dr}{\sqrt {r^{4}(E-1)+2Mr^{3}-(Q^{2}+L^{2})r^{2}+2ML^{2}r-Q^{2}L^{2}}}}.}

The total time dilation between the test-particle and an observer at infinity is γ = q   Q   r 3 + E   r 4 r 2   ( r 2 2 r + Q 2 ) . {\displaystyle \gamma ={\frac {q\ Q\ r^{3}+E\ r^{4}}{r^{2}\ (r^{2}-2r+Q^{2})}}.}

The first derivatives x ˙ i {\displaystyle {\dot {x}}^{i}} and the contravariant components of the local 3-velocity v i {\displaystyle v^{i}} are related by x ˙ i = v i ( 1 v 2 )   | g i i | , {\displaystyle {\dot {x}}^{i}={\frac {v^{i}}{\sqrt {(1-v^{2})\ |g_{ii}|}}},} which gives the initial conditions r ˙ = v r 2 2 M + Q 2 r ( 1 v 2 ) {\displaystyle {\dot {r}}={\frac {v_{\parallel }{\sqrt {r^{2}-2M+Q^{2}}}}{r{\sqrt {(1-v^{2})}}}}} θ ˙ = v r ( 1 v 2 ) . {\displaystyle {\dot {\theta }}={\frac {v_{\perp }}{r{\sqrt {(1-v^{2})}}}}.}

The specific orbital energy E = Q 2 2 r M + r 2 r 1 v 2 + q Q r {\displaystyle E={\frac {\sqrt {Q^{2}-2rM+r^{2}}}{r{\sqrt {1-v^{2}}}}}+{\frac {qQ}{r}}} and the specific relative angular momentum L = v   r 1 v 2 {\displaystyle L={\frac {v_{\perp }\ r}{\sqrt {1-v^{2}}}}} of the test-particle are conserved quantities of motion. v {\displaystyle v_{\parallel }} and v {\displaystyle v_{\perp }} are the radial and transverse components of the local velocity-vector. The local velocity is therefore v = v 2 + v 2 = ( E 2 1 ) r 2 Q 2 r 2 + 2 r M E 2 r 2 . {\displaystyle v={\sqrt {v_{\perp }^{2}+v_{\parallel }^{2}}}={\sqrt {\frac {(E^{2}-1)r^{2}-Q^{2}-r^{2}+2rM}{E^{2}r^{2}}}}.}

Alternative formulation of metric

The metric can be expressed in Kerr–Schild form like this: g μ ν = η μ ν + f k μ k ν f = G r 2 [ 2 M r Q 2 ] k = ( k x , k y , k z ) = ( x r , y r , z r ) k 0 = 1. {\displaystyle {\begin{aligned}g_{\mu \nu }&=\eta _{\mu \nu }+fk_{\mu }k_{\nu }\\[5pt]f&={\frac {G}{r^{2}}}\left[2Mr-Q^{2}\right]\\[5pt]\mathbf {k} &=(k_{x},k_{y},k_{z})=\left({\frac {x}{r}},{\frac {y}{r}},{\frac {z}{r}}\right)\\[5pt]k_{0}&=1.\end{aligned}}}

Notice that k is a unit vector. Here M is the constant mass of the object, Q is the constant charge of the object, and η is the Minkowski tensor.

See also

Notes

  1. ^ Reissner, H. (1916). "Über die Eigengravitation des elektrischen Feldes nach der Einsteinschen Theorie". Annalen der Physik. 355 (9): 106–120. Bibcode:1916AnP...355..106R. doi:10.1002/andp.19163550905. ISSN 0003-3804.
  2. ^ Weyl, Hermann (1917). "Zur Gravitationstheorie". Annalen der Physik. 359 (18): 117–145. Bibcode:1917AnP...359..117W. doi:10.1002/andp.19173591804. ISSN 0003-3804.
  3. ^ Nordström, G. (1918). "On the Energy of the Gravitational Field in Einstein's Theory". Koninklijke Nederlandsche Akademie van Wetenschappen Proceedings. 20 (2): 1238–1245. Bibcode:1918KNAB...20.1238N.
  4. ^ Jeffery, G. B. (1921). "The field of an electron on Einstein's theory of gravitation". Proceedings of the Royal Society of London. Series A, Containing Papers of a Mathematical and Physical Character. 99 (697): 123–134. Bibcode:1921RSPSA..99..123J. doi:10.1098/rspa.1921.0028. ISSN 0950-1207.
  5. ^ Siegel, Ethan (2021-10-13). "Surprise: the Big Bang isn't the beginning of the universe anymore". Big Think. Retrieved 2024-09-03.
  6. ^ Thibault Damour: Black Holes: Energetics and Thermodynamics, S. 11 ff.
  7. ^ Qadir, Asghar (December 1983). "Reissner-Nordstrom repulsion". Physics Letters A. 99 (9): 419–420. Bibcode:1983PhLA...99..419Q. doi:10.1016/0375-9601(83)90946-5.
  8. ^ Chandrasekhar, Subrahmanyan (2009). The mathematical theory of black holes. Oxford classic texts in the physical sciences (Reprinted ed.). Oxford: Clarendon Press. p. 205. ISBN 978-0-19-850370-5. And finally, the fact that the Reissner–Nordström solution has two horizons, an external event horizon and an internal 'Cauchy horizon,' provides a convenient bridge to the study of the Kerr solution in the subsequent chapters.
  9. ^ Andrew Hamilton: The Reissner Nordström Geometry (Casa Colorado)
  10. ^ Carter, Brandon (25 October 1968). "Global Structure of the Kerr Family of Gravitational Fields". Physical Review. 174 (5): 1559–1571. doi:10.1103/PhysRev.174.1559. ISSN 0031-899X.
  11. ^ Leonard Susskind: The Theoretical Minimum: Geodesics and Gravity, (General Relativity Lecture 4, timestamp: 34m18s)
  12. ^ Hackmann, Eva; Xu, Hongxiao (2013). "Charged particle motion in Kerr-Newmann space-times". Physical Review D. 87 (12): 124030. arXiv:1304.2142. doi:10.1103/PhysRevD.87.124030. ISSN 1550-7998.
  13. ^ Wald, Robert M. (2009). General relativity (Repr. ed.). Chicago: Univ. of Chicago Press. ISBN 978-0-226-87033-5.
  14. ^ Nordebo, Jonatan. "The Reissner-Nordström metric" (PDF). diva-portal. Retrieved 8 April 2021.
  15. ^ Smith, B. R. (December 2009). "First-order partial differential equations in classical dynamics". American Journal of Physics. 77 (12): 1147–1153. Bibcode:2009AmJPh..77.1147S. doi:10.1119/1.3223358. ISSN 0002-9505.
  16. ^ Misner, Charles W.; Thorne, Kip S.; Wheeler, John Archibald; Kaiser, David; et al. (2017). Gravitation. Princeton, N.J: Princeton University Press. pp. 656–658. ISBN 978-0-691-17779-3. OCLC 1006427790.

References

  • Adler, R.; Bazin, M.; Schiffer, M. (1965). Introduction to General Relativity. New York: McGraw-Hill Book Company. pp. 395–401. ISBN 978-0-07-000420-7.
  • Wald, Robert M. (1984). General Relativity. Chicago: The University of Chicago Press. pp. 158, 312–324. ISBN 978-0-226-87032-8.
  • v
  • t
  • e
Types
Size
Formation
Properties
Issues
Metrics
Alternatives
Analogs
Lists
Related
Notable
  • Category
  • Commons
  • v
  • t
  • e
Special
relativity
Background
Fundamental
concepts
Formulation
Phenomena
Spacetime
General
relativity
Background
Fundamental
concepts
Formulation
Phenomena
Advanced
theories
Solutions
Scientists
Category