ART

A binary black hole (BBH) is a system consisting of two black holes in close orbit around each other. Like black holes themselves, binary black holes are often divided into stellar binary black holes, formed either as remnants of high-mass binary star systems or by dynamic processes and mutual capture, and binary supermassive black holes believed to be a result of galactic mergers.

For many years, proving the existence of binary black holes was made difficult because of the nature of black holes themselves, and the limited means of detection available. However, in the event that a pair of black holes were to merge, an immense amount of energy should be given off as gravitational waves, with distinctive waveforms that can be calculated using general relativity.[2][3][4] Therefore, during the late 20th and early 21st century, binary black holes became of great interest scientifically as a potential source of such waves, and a means by which gravitational waves could be proven to exist. Binary black hole mergers would be one of the strongest known sources of gravitational waves in the Universe, and thus offer a good chance of directly detecting such waves. As the orbiting black holes give off these waves, the orbit decays, and the orbital period decreases. This stage is called binary black hole inspiral. The black holes will merge once they are close enough. Once merged, the single hole settles down to a stable form, via a stage called ringdown, where any distortion in the shape is dissipated as more gravitational waves.[5] In the final fraction of a second the black holes can reach extremely high velocity, and the gravitational wave amplitude reaches its peak.

The existence of stellar-mass binary black holes (and gravitational waves themselves) was finally confirmed when LIGO detected GW150914 (detected September 2015, announced February 2016), a distinctive gravitational wave signature of two merging stellar-mass black holes of around 30 solar masses each, occurring about 1.3 billion light years away. In its final 20 ms of spiraling inward and merging, GW150914 released around 3 solar masses as gravitational energy, peaking at a rate of 3.6×1049 watts — more than the combined power of all light radiated by all the stars in the observable universe put together.[6][7][8] Supermassive binary black hole candidates have been found, but not yet categorically proven.[9]

Occurrence

Supermassive black-hole binaries are believed to form during galaxy mergers. Some likely candidates for binary black holes are galaxies with double cores still far apart. An example double nucleus is NGC 6240.[10] Much closer black-hole binaries are likely in single core galaxies with double emission lines. Examples include SDSS J104807.74+005543.5[11] and EGSD2 J142033.66 525917.5.[10] Other galactic nuclei have periodic emissions suggesting large objects orbiting a central black hole, for example in OJ287.[12]

The quasar PG 1302-102 appears to have a binary black hole with an orbital period of 1900 days.[13]

Stellar mass binary black holes have been demonstrated to exist, by the first detection of a black hole merger event GW150914 by LIGO.[14]
Final parsec problem

When two galaxies collide, the supermassive black holes at their centers are very unlikely to hit head-on, and would in fact most likely shoot past each other on hyperbolic trajectories if some mechanism did not bring them together. The most important mechanism is dynamical friction, which transfers kinetic energy from the black holes to nearby matter. As a black hole passes a star, the gravitational slingshot accelerates the star while decelerating the black hole.

This slows the black holes enough that they form a bound, binary, system, and further dynamical friction steals orbital energy from the pair until they are orbiting within a few parsecs of each other. However, this process also ejects matter from the orbital path, and as the orbits shrink, the volume of space the black holes pass through reduces, until there is so little matter remaining that it could not cause merger within the age of the universe.

Gravitational waves can cause significant loss of orbital energy, but not until the separation shrinks to a much smaller value, roughly 0.01–0.001 parsec.

Nonetheless, supermassive black holes appear to have merged, and what appears to be a pair in this intermediate range has been observed, in PKS 1302-102.[15][16] The question of how this happens is the "final parsec problem".[17]

A number of solutions to the final parsec problem have been proposed. Most involve mechanisms to bring additional matter, either stars or gas, close enough to the binary pair to extract energy from the binary and cause it to shrink. If enough stars pass close by to the orbiting pair, their gravitational ejection can bring the two black holes together in an astronomically plausible time.[18]

One mechanism that is known to work, although infrequently, is a third supermassive black hole from a second galactic collision.[19] With three black holes in close proximity, the orbits are chaotic and allow three additional energy loss mechanisms:

The black holes orbit through a substantially larger volume of the galaxy, interacting with (and losing energy to) a much greater amount of matter,
The orbits can become highly eccentric, allowing energy loss by gravitational radiation at the point of closest approach, and
Two of the black holes can transfer energy to the third, possibly ejecting it.[20]

Lifecycle
Inspiral

The first stage of the life of a binary black hole is the inspiral, a gradually shrinking orbit. The first stages of the inspiral take a very long time, as the gravitation waves emitted are very weak when the black holes are distant from each other. In addition to the orbit shrinking due to the emission of gravitational waves, extra angular momentum may be lost due to interactions with other matter present, such as other stars.

As the black holes’ orbit shrinks, the speed increases, and gravitational wave emission increases. When the black holes are close the gravitational waves cause the orbit to shrink rapidly.

The last stable orbit or innermost stable circular orbit (ISCO) is the innermost complete orbit before the transition from inspiral to merger.
Merger

This is followed by a plunging orbit in which the two black holes meet, followed by the merger. Gravitational wave emission peaks at this time.
Ringdown

Immediately following the merger, the now single black hole will “ring”. This ringing is damped in the next stage, called the ringdown, by the emission of gravitational waves. The ringdown phase starts when the black hole approach each other within the photon sphere. In this region most of the emitted gravitational waves go towards the event horizon, and the amplitude escaping reduces. Remotely detected gravitational waves have a fast reducing oscillation, as echos of the merger event result from tighter and tighter spirals around the resulting black hole.

Observation

The first observation of stellar mass binary black holes merging was performed by the LIGO detector.[14][21][22] As observed from Earth, a pair of black holes with estimated masses around 36 and 29 times that of the Sun spun into each other and merged to form a 62 solar mass black hole (approximate) on 14 September 2015, at 09:50 UTC.[23] Three solar masses were converted to gravitational radiation in the final fraction of a second, with a peak power 3.6×1056 ergs/second (200 solar masses per second),[14] which is 50 times the total output power of all the stars in the observable universe.[24] The merger took place at 1.3 billion light years from Earth,[21] and therefore 1.3 billion years ago. The observed signal is consistent with the predictions of numerical relativity.[2][3][4]
Dynamics modelling

Some simplified algebraic models can be used for the case where the black holes are far apart, during the inspiral stage, and also to solve for the final ringdown.

Post-Newtonian approximations can be used for the inspiral. These approximate the general relativity field equations adding extra terms to equations in Newtonian gravity. Orders used in these calculations may be termed 2PN (second order post Newtonian) 2.5PN or 3PN (third order post Newtonian). Effective-one-body (EOB) solves the dynamics of the binary black hole system by transforming the equations to those of a single object. This is especially useful where mass ratios are large, such as a stellar mass black hole merging with a galactic core black hole, but can also be used for equal mass systems.

For the ringdown, black hole perturbation theory can be used. The final Kerr black hole is distorted, and the spectrum of frequencies it produces can be calculated.

To solve for the entire evolution, including merger, requires solving the full equations of general relativity. This can be done in numerical relativity simulations. Numerical relativity models space-time and simulates its change over time. In these calculations it is important to have enough fine detail close into the black holes, and yet have enough volume to determine the gravitation radiation that propagates to infinity. In order to make this have few enough points to be tractable to calculation in a reasonable time, special coordinate systems can be used such as Boyer-Lindquist coordinates or fish-eye coordinates.

Numerical relativity techniques steadily improved from the initial attempts in the 1960s and 1970s.[25][26] Long-term simulations of orbiting black holes, however, were not possible until three groups independently developed groundbreaking new methods to model the inspiral, merger, and ringdown of binary black holes [2][3][4] in 2005.

In the full calculations of an entire merger, several of the above methods can be used together. It is then important to fit the different pieces of the model that were worked out using different algorithms. The Lazarus Project linked the parts on a spacelike hypersurface at the time of the merger.[27]

Results from the calculations can include the binding energy. In a stable orbit the binding energy is a local minimum relative to parameter perturbation. At the innermost stable circular orbit the local minimum becomes an inflection point.

The gravitational waveform produced is important for observation prediction and confirmation. When inspiralling reaches the strong zone of the gravitational field, the waves scatter within the zone producing what is called the post Newtonian tail (PN tail).[27]

In the ringdown phase of a Kerr black hole, frame-dragging produces a gravitation wave with the horizon frequency. In contrast the Schwarzschild black-hole ringdown looks like the scattered wave from the late inspiral, but with no direct wave.[27]

The radiation reaction force can be calculated by Padé resummation of gravitational wave flux. A technique to establish the radiation is the Cauchy characteristic extraction technique CCE which gives a close estimate of the flux at infinity, without having to calculate at larger and larger finite distances.

The final mass of the resultant black hole depends on the definition of mass in general relativity. The Bondi mass MB is calculated from the Bondi-Sach mass loss formula. \( {\displaystyle {\frac {dM_{B}}{dU}}=-f(U)} \). With f(U) the gravitational wave flux at retarded time U. f is a surface integral of the News function at null infinity varied by solid angle. The Arnowitt-Deser-Misner (ADM) energy or ADM mass is the mass as measured at infinite distance and includes all the gravitational radiation emitted. \( {\displaystyle M_{ADM}=M_{B}(U)+\int _{-\infty }^{U}F(V)dV} \).

Angular momentum is also lost in the gravitational radiation. This is primarily in the z axis of the initial orbit. It is calculated by integrating the product of the multipolar metric waveform with the news function complement over retarded time.[28]

Shape

One of the problems to solve is the shape or topology of the event horizon during a black-hole merger.

In numerical models, test geodesics are inserted to see if they encounter an event horizon. As two black holes approach each other, a ‘duckbill’ shape protrudes from each of the two event horizons towards the other one. This protrusion extends longer and narrower until it meets the protrusion from the other black hole. At this point in time the event horizon has a very narrow X-shape at the meeting point. The protrusions are drawn out into a thin thread.[29] The meeting point expands to a roughly cylindrical connection called a bridge.[29]

Simulations as of 2011 had not produced any event horizons with toroidal topology (ring-shaped). Some researchers suggested that it would be possible if, for example, several black holes in the same nearly-circular orbit coalesce.[29]

Black-hole merger recoil

An unexpected result can occur with binary black holes that merge, in that the gravitational waves carry momentum and the merging black-hole pair accelerates seemingly violating Newton's third law. The center of gravity can add over 1000 km/s of kick velocity.[30] The greatest kick velocities (approaching 5000 km/s) occur for equal-mass and equal-spin-magnitude black-hole binaries, when the spins directions are optimally oriented to be counter-aligned, parallel to the orbital plane or nearly aligned with the orbital angular momentum.[31] This is enough to escape large galaxies. With more likely orientations a smaller effect takes place, perhaps only a few hundred kilometers per second. This sort of speed will eject merging binary black holes from globular clusters, thus preventing the formation of massive black holes in globular cluster cores. In turn this reduces the chances of subsequent mergers, and thus the chance of detecting gravitational waves. For non spinning black holes a maximum recoil velocity of 175 km/s occurs for masses in the ratio of five to one. When spins are aligned in the orbital plane a recoil of 5000 km/s is possible with two identical black holes.[32] Parameters that may be of interest include the point at which the black holes merge, the mass ratio which produces maximum kick, and how much mass/energy is radiated via gravitational waves. In a head-on collision this fraction is calculated at 0.002 or 0.2%.[33] One of the best candidates of the recoiled supermassive black holes is CXO J101527.2+625911.[34]

Halo drive for space travel

It has been hypothesized that binary black holes could transfer energy and momentum to a spacecraft using a "halo drive", exploiting the holographic reflection created by a set of null geodesics looping behind and then around one of the black holes before returning to the spacecraft. The reflection passing through these null geodesics would form one end of a laser cavity, with a mirror on the spacecraft forming the other end of the laser cavity. Even a planet-sized spacecraft would thereby accelerate to speeds exceeding the approaching black hole's relative speed. If true, a network of these binary black holes might permit travel across the galaxy.[35]

References

Credits: SXS (Simulating eXtreme Spacetimes) project
Pretorius, Frans (2005). "Evolution of Binary Black-Hole Spacetimes". Physical Review Letters. 95 (12): 121101.arXiv:gr-qc/0507014. Bibcode:2005PhRvL..95l1101P. doi:10.1103/PhysRevLett.95.121101. ISSN 0031-9007. PMID 16197061. S2CID 24225193.
Campanelli, M.; Lousto, C. O.; Marronetti, P.; Zlochower, Y. (2006). "Accurate Evolutions of Orbiting Black-Hole Binaries without Excision". Physical Review Letters. 96 (11): 111101.arXiv:gr-qc/0511048. Bibcode:2006PhRvL..96k1101C. doi:10.1103/PhysRevLett.96.111101. ISSN 0031-9007. PMID 16605808. S2CID 5954627.
Baker, John G.; Centrella, Joan; Choi, Dae-Il; Koppitz, Michael; van Meter, James (2006). "Gravitational-Wave Extraction from an Inspiraling Configuration of Merging Black Holes". Physical Review Letters. 96 (11): 111102.arXiv:gr-qc/0511103. Bibcode:2006PhRvL..96k1102B. doi:10.1103/PhysRevLett.96.111102. ISSN 0031-9007. PMID 16605809. S2CID 23409406.
Abadie, J.; LIGO Scientific Collaboration; The Virgo Collaboration; Abernathy, M.; Accadia, T.; Acernese, F.; Adams, C.; Adhikari, R.; Ajith, P.; Allen, B.; Allen, G. S.; Amador Ceron, E.; Amin, R. S.; Anderson, S. B.; Anderson, W. G.; Antonucci, F.; Arain, M. A.; Araya, M. C.; Aronsson, M.; Aso, Y.; Aston, S. M.; Astone, P.; Atkinson, D.; Aufmuth, P.; Aulbert, C.; Babak, S.; Baker, P.; Ballardin, G.; Ballinger, T.; et al. (2011). "Search for gravitational waves from binary black hole inspiral, merger and ringdown". Physical Review D. 83 (12): 122005.arXiv:1102.3781. Bibcode:2011PhRvD..83l2005A. doi:10.1103/PhysRevD.83.122005. S2CID 174250.
"Observation Of Gravitational Waves From A Binary Black Hole Merger" (PDF). LIGO. 11 February 2016. Archived from the original (PDF) on 16 February 2016. Retrieved 11 February 2016.
Harwood, W. (11 February 2016). "Einstein was right: Scientists detect gravitational waves in breakthrough". CBS News. Archived from the original on 12 February 2016. Retrieved 12 February 2016.
Drake, Nadia (11 February 2016). "Found! Gravitational Waves, or a Wrinkle in Spacetime". National Geographic News. Archived from the original on 12 February 2016. Retrieved 12 February 2016.
Liu, Fukun; Komossa, Stefanie; Schartel, Norbert (22 April 2014). "Unique Pair of Hidden Black Holes Discovered yy XMM-Newton". A milli-parsec supermassive black hole binary candidate in the galaxy SDSS J120136.02+300305.5. Retrieved 23 December 2014.
Gerke, Brian F.; Newman, Jeffrey A.; Lotz, Jennifer; Yan, Renbin; Barmby, P.; Coil, Alison L.; Conselice, Christopher J.; Ivison, R. J.; Lin, Lihwai; Koo, David C.; Nandra, Kirpal; Salim, Samir; Small, Todd; Weiner, Benjamin J.; Cooper, Michael C.; Davis, Marc; Faber, S. M.; Guhathakurta, Puragra; et al. (6 April 2007). "The DEEP2 Galaxy Redshift Survey: AEGIS Observations of a Dual AGN AT z p 0.7". The Astrophysical Journal Letters. 660 (1): L23–L26.arXiv:astro-ph/0608380. Bibcode:2007ApJ...660L..23G. doi:10.1086/517968. S2CID 14320681.
Hongyan Zhou; Tinggui Wang; Xueguang Zhang; Xiaobo Dong; Cheng Li (26 February 2004). "Obscured Binary Quasar Cores in SDSS J104807.74+005543.5?". The Astrophysical Journal Letters. The American Astronomical Society. 604 (1): L33–L36.arXiv:astro-ph/0411167. Bibcode:2004ApJ...604L..33Z. doi:10.1086/383310. S2CID 14297940.
Valtonen, M. V.; Mikkola, S.; Merritt, D.; Gopakumar, A.; Lehto, H. J.; Hyvönen, T.; Rampadarath, H.; Saunders, R.; Basta, M.; Hudec, R. (February 2010). "Measuring the Spin of the Primary Black Hole in OJ287". The Astrophysical Journal. 709 (2): 725–732.arXiv:0912.1209. Bibcode:2010ApJ...709..725V. doi:10.1088/0004-637X/709/2/725. S2CID 119276181.
Graham, Matthew J.; Djorgovski, S. G.; Stern, Daniel; Glikman, Eilat; Drake, Andrew J.; Mahabal, Ashish A.; Donalek, Ciro; Larson, Steve; Christensen, Eric (7 January 2015). "A possible close supermassive black-hole binary in a quasar with optical periodicity". Nature. 518 (7537): 74–6.arXiv:1501.01375. Bibcode:2015Natur.518...74G. doi:10.1038/nature14143. ISSN 0028-0836. PMID 25561176. S2CID 4459433.
B. P. Abbott; et al. (LIGO Scientific Collaboration and Virgo Collaboration) (2016). "Observation of Gravitational Waves from a Binary Black Hole Merger". Physical Review Letters. 116 (6): 061102.arXiv:1602.03837. Bibcode:2016PhRvL.116f1102A. doi:10.1103/PhysRevLett.116.061102. PMID 26918975. S2CID 124959784.
D'Orazio, Daniel J.; Haiman, Zoltán; Schiminovich, David (17 September 2015). "Relativistic boost as the cause of periodicity in a massive black-hole binary candidate". Nature. 525 (7569): 351–353.arXiv:1509.04301. Bibcode:2015Natur.525..351D. doi:10.1038/nature15262. PMID 26381982. S2CID 205245606.
Overbye, Dennis (16 September 2015). "More Evidence for Coming Black Hole Collision". The New York Times.
Milosavljević, Miloš; Merritt, David (October 2003). "The Final Parsec Problem" (PDF). AIP Conference Proceedings. American Institute of Physics. 686 (1): 201–210.arXiv:astro-ph/0212270. Bibcode:2003AIPC..686..201M. doi:10.1063/1.1629432. S2CID 12124842.
Merritt, David (2013). Dynamics and Evolution of Galactic Nuclei. Princeton: Princeton University Press. ISBN 978-0-691-12101-7.
Ryu, Taeho; Perna, Rosalba; Haiman, Zoltán; Ostriker, Jeremiah P.; Stone, Nicholas C. (2018). "Interactions between multiple supermassive black holes in galactic nuclei: a solution to the final parsec problem". Monthly Notices of the Royal Astronomical Society. 473 (3): 3410–3433.arXiv:1709.06501. Bibcode:2018MNRAS.473.3410R. doi:10.1093/mnras/stx2524. S2CID 119083047.
Iwasawa, Masaki; Funato, Yoko; Makino, Junichiro (2006). "Evolution of Massive Blackhole Triples I – Equal-mass binary-single systems". Astrophys. J. 651 (2): 1059–1067.arXiv:astro-ph/0511391. Bibcode:2006ApJ...651.1059I. doi:10.1086/507473. S2CID 14816623. "We found that in most cases two of the three BHs merge through gravitational wave (GW) radiation in the timescale much shorter than the Hubble time, before ejecting one BH through a slingshot."
Castelvecchi, Davide; Witze, Witze (February 11, 2016). "Einstein's gravitational waves found at last". Nature News. doi:10.1038/nature.2016.19361. S2CID 182916902. Retrieved 2016-02-11.
"Gravitational waves detected 100 years after Einstein's prediction | NSF - National Science Foundation". www.nsf.gov. Retrieved 2016-02-11.
Abbott, Benjamin P.; et al. (LIGO Scientific Collaboration and Virgo Collaboration) (11 February 2016). "Properties of the binary black hole merger GW150914". Physical Review Letters. 116 (24): 241102.arXiv:1602.03840. Bibcode:2016PhRvL.116x1102A. doi:10.1103/PhysRevLett.116.241102. PMID 27367378. S2CID 217406416.
Kramer, Sarah (11 February 2016). "This collision was 50 times more powerful than all the stars in the universe combined". Tech Insider. Retrieved 12 February 2016.
Hahn, Susan G; Lindquist, Richard W (1964). "The two-body problem in geometrodynamics". Annals of Physics. 29 (2): 304–331. Bibcode:1964AnPhy..29..304H. doi:10.1016/0003-4916(64)90223-4. ISSN 0003-4916.
Smarr, Larry; Čadež, Andrej; DeWitt, Bryce; Eppley, Kenneth (1976). "Collision of two black holes: Theoretical framework". Physical Review D. 14 (10): 2443–2452. Bibcode:1976PhRvD..14.2443S. doi:10.1103/PhysRevD.14.2443. ISSN 0556-2821.
Nichols, David A.; Yanbei Chen (2012). "Hybrid method for understanding black-hole mergers: Inspiralling case". Physical Review D. 85 (4): 044035.arXiv:1109.0081. Bibcode:2012PhRvD..85d4035N. doi:10.1103/PhysRevD.85.044035. S2CID 30890236.
Thibault
Cohen, Michael I.; Jeffrey D. Kaplan; Mark A. Scheel (2012). "On Toroidal Horizons in Binary Black Hole Inspirals". Physical Review D. 85 (2): 024031.arXiv:1110.1668. Bibcode:2012PhRvD..85b4031C. doi:10.1103/PhysRevD.85.024031. S2CID 37654897.
Pietilä, Harri; Heinämäki, Pekka; Mikkola, Seppo; Valtonen, Mauri J. (10 January 1996). Anisotropic Gravitational Radiation In The Merger Of Black Holes. Relativistic Astrophysics Conference. CiteSeerX 10.1.1.51.2616.
Campanelli, Manuela; Lousto, Carlos; Zlochower, Yosef; Merritt, David (7 June 2007). "Maximum Gravitational Recoil". Physical Review Letters. 98 (23): 231102.arXiv:gr-qc/0702133. Bibcode:2007PhRvL..98w1102C. doi:10.1103/PhysRevLett.98.231102. PMID 17677894. S2CID 29246347.
Lousto, Carlos; Zlochower, Yosef (2011). "Hangup Kicks: Still Larger Recoils by Partial Spin-Orbit Alignment of Black-Hole Binaries". Physical Review Letters. 107 (23): 231102.arXiv:1108.2009. Bibcode:2011PhRvL.107w1102L. doi:10.1103/PhysRevLett.107.231102. PMID 22182078. S2CID 15546595.
Pietilä, Harri; Heinämäki, Pekka; Mikkola, Seppo; Valtonen, Mauri J. (1995). "Anisotropic gravitational radiation in the problems of three and four black holes". Celestial Mechanics and Dynamical Astronomy. 62 (4): 377–394. Bibcode:1995CeMDA..62..377P. CiteSeerX 10.1.1.51.2616. doi:10.1007/BF00692287. S2CID 122956625.
Kim, D.-C.; et al. (2017). "A Potential Recoiling Supermassive Black Hole CXO J101527.2+625911". Astrophysical Journal. 840 (2): 71–77.arXiv:1704.05549. Bibcode:2017ApJ...840...71K. doi:10.3847/1538-4357/aa6030. S2CID 119401892.

Kipping, David (2019). "The Halo Drive: Fuel-Free Relativistic Propulsion of Large Masses via Recycled Boomerang Photons".arXiv:1903.03423 [gr-qc].

External links
Wikimedia Commons has media related to Binary black holes.

Binary Black Holes Orbit and Collide
Merritt, David; Milosavljević, Miloš (2005). "Massive Black Hole Binary Evolution". Living Reviews in Relativity. 8: 8.arXiv:astro-ph/0410364. Bibcode:2005LRR.....8....8M. doi:10.12942/lrr-2005-8. S2CID 119367453. Archived from the original on 2012-03-30.

vte

Black holes
Types

Schwarzschild Rotating Charged Virtual Kugelblitz Primordial Planck particle


Black hole - Messier 87 crop max res.jpg
Size

Micro
Extremal Electron Stellar
Microquasar Intermediate-mass Supermassive
Active galactic nucleus Quasar Blazar

Formation

Stellar evolution Gravitational collapse Neutron star
Related links Tolman–Oppenheimer–Volkoff limit White dwarf
Related links Supernova
Related links Hypernova Gamma-ray burst Binary black hole

Properties

Gravitational singularity
Ring singularity Theorems Event horizon Photon sphere Innermost stable circular orbit Ergosphere
Penrose process Blandford–Znajek process Accretion disk Hawking radiation Gravitational lens Bondi accretion M–sigma relation Quasi-periodic oscillation Thermodynamics
Immirzi parameter Schwarzschild radius Spaghettification

Issues

Black hole complementarity Information paradox Cosmic censorship ER=EPR Final parsec problem Firewall (physics) Holographic principle No-hair theorem

Metrics

Schwarzschild (Derivation) Kerr Reissner–Nordström Kerr–Newman Hayward

Alternatives

Nonsingular black hole models Black star Dark star Dark-energy star Gravastar Magnetospheric eternally collapsing object Planck star Q star Fuzzball

Analogs

Optical black hole Sonic black hole

Lists

Black holes Most massive Nearest Quasars Microquasars

Related

Black Hole Initiative Black hole starship Compact star Exotic star
Quark star Preon star Gamma-ray burst progenitors Gravity well Hypercompact stellar system Membrane paradigm Naked singularity Quasi-star Rossi X-ray Timing Explorer Timeline of black hole physics White hole Wormhole

Physics Encyclopedia

World

Index

Hellenica World - Scientific Library

Retrieved from "http://en.wikipedia.org/"
All text is available under the terms of the GNU Free Documentation License