Astronomy:Chandrasekhar limit

From HandWiki
Short description: Maximum mass of a stable white dwarf star

The Chandrasekhar limit (/ˌændrəˈʃkər/)[1] is the maximum mass of a stable white dwarf star. The currently accepted value of the Chandrasekhar limit is about 1.4 M (2.765×1030 kg).[2][3][4]

White dwarfs resist gravitational collapse primarily through electron degeneracy pressure, compared to main sequence stars, which resist collapse through thermal pressure. The Chandrasekhar limit is the mass above which electron degeneracy pressure in the star's core is insufficient to balance the star's own gravitational self-attraction. Consequently, a white dwarf with a mass greater than the limit is subject to further gravitational collapse, evolving into a different type of stellar remnant, such as a neutron star or black hole. White dwarfs with masses up to the Chandrasekhar limit remain stable.[5] The Tolman–Oppenheimer–Volkoff limit is theoretically a next level to reach in order for a neutron star to collapse into a denser form such as a black hole.

The limit was named after Subrahmanyan Chandrasekhar. Chandrasekhar improved upon the accuracy of the calculation in 1930 by calculating the limit for a polytrope model of a star in hydrostatic equilibrium, and comparing his limit to the earlier limit found by E. C. Stoner for a uniform density star. Importantly, the existence of a limit, based on the conceptual breakthrough of combining relativity with Fermi degeneracy, was indeed first established in separate papers published by Wilhelm Anderson and E. C. Stoner in 1929. The limit was initially ignored by the community of scientists because such a limit would logically require the existence of black holes, which were considered a scientific impossibility at the time. The fact that the roles of Stoner and Anderson are often overlooked in the astronomy community has been noted.[6][7]

The priority dispute has been discussed at length by Virginia Trimble who purports that: "Chandrasekhar famously, perhaps even notoriously did his critical calculation on board ship in 1930, and ... was not aware of either Stoner's or Anderson's work at the time. His work was therefore independent, but, more to the point, he adopted Eddington's polytropes for his models which could, therefore, be in hydrostatic equilibrium, which constant density stars cannot, and real ones must be."[8]. It is hard to document exactly what he was unaware of when he started working on the problem, but Chandrasekhar's mass limit paper [9] explicitly refers to Stoner's paper.[10], so there is no publication of a mass limit by Chandrasekhar that preceded Stoner's work.

Physics

Radius–mass relations for a model white dwarf. The green curve uses the general pressure law for an ideal Fermi gas, while the blue curve is for a non-relativistic ideal Fermi gas. The black line marks the ultrarelativistic limit.

Electron degeneracy pressure is a quantum-mechanical effect arising from the Pauli exclusion principle. Since electrons are fermions, no two electrons can be in the same state, so not all electrons can be in the minimum-energy level. Rather, electrons must occupy a band of energy levels. Compression of the electron gas increases the number of electrons in a given volume and raises the maximum energy level in the occupied band. Therefore, the energy of the electrons increases on compression, so pressure must be exerted on the electron gas to compress it, producing electron degeneracy pressure. With sufficient compression, electrons are forced into nuclei in the process of electron capture, relieving the pressure.

In the nonrelativistic case, electron degeneracy pressure gives rise to an equation of state of the form P = K1ρ5/3, where P is the pressure, ρ is the mass density, and K1 is a constant. Solving the hydrostatic equation leads to a model white dwarf that is a polytrope of index 3/2 – and therefore has radius inversely proportional to the cube root of its mass, and volume inversely proportional to its mass.[11]

As the mass of a model white dwarf increases, the typical energies to which degeneracy pressure forces the electrons are no longer negligible relative to their rest masses. The velocities of the electrons approach the speed of light, and special relativity must be taken into account. In the strongly relativistic limit, the equation of state takes the form P = K2ρ4/3. This yields a polytrope of index 3, which has a total mass, Mlimit, depending only on K2.[9]

For a fully relativistic treatment, the equation of state used interpolates between the equations P = K1ρ5/3 for small ρ and P = K2ρ4/3 for large ρ. When this is done, the model radius still decreases with mass, but becomes zero at Mlimit. This is the Chandrasekhar limit.[12] The curves of radius against mass for the non-relativistic and relativistic models are shown in the graph. They are colored blue and green, respectively. μe has been set equal to 2. Radius is measured in standard solar radii[13] or kilometers, and mass in standard solar masses.

Calculated values for the limit vary depending on the nuclear composition of the mass.[14] Chandrasekhar[15]:{{{1}}}[12]:{{{1}}}[16]:{{{1}}} gives the following expression, based on the equation of state for an ideal Fermi gas: [math]\displaystyle{ M_\text{limit} = \frac{\omega_3^0 \sqrt{3\pi}}{2} \left ( \frac{\hbar c}{G}\right )^\frac{3}{2} \frac{1}{(\mu_\text{e} m_\text{H})^2} }[/math] where:

As ħc/G is the Planck mass, the limit is of the order of [math]\displaystyle{ \frac{M_\text{Pl}^3}{m_\text{H}^2} }[/math] The limiting mass can be obtained formally from the Chandrasekhar's white dwarf equation by taking the limit of large central density.

A more accurate value of the limit than that given by this simple model requires adjusting for various factors, including electrostatic interactions between the electrons and nuclei and effects caused by nonzero temperature.[14] Lieb and Yau[17] have given a rigorous derivation of the limit from a relativistic many-particle Schrödinger equation.

History

In 1926, the British physicist Ralph H. Fowler observed that the relationship between the density, energy, and temperature of white dwarfs could be explained by viewing them as a gas of nonrelativistic, non-interacting electrons and nuclei that obey Fermi–Dirac statistics.[18] This Fermi gas model was then used by the British physicist Edmund Clifton Stoner in 1929 to calculate the relationship among the mass, radius, and density of white dwarfs, assuming they were homogeneous spheres.[19] Wilhelm Anderson applied a relativistic correction to this model, giving rise to a maximum possible mass of approximately 1.37×1030 kg.[20] In 1930, Stoner derived the internal energydensity equation of state for a Fermi gas, and was then able to treat the mass–radius relationship in a fully relativistic manner, giving a limiting mass of approximately 2.19×1030 kg (for μe = 2.5).[21] Stoner went on to derive the pressuredensity equation of state, which he published in 1932.[22] These equations of state were also previously published by the Soviet physicist Yakov Frenkel in 1928, together with some other remarks on the physics of degenerate matter.[23] Frenkel's work, however, was ignored by the astronomical and astrophysical community.[24]

A series of papers published between 1931 and 1935 had its beginning on a trip from India to England in 1930, where the Indian physicist Subrahmanyan Chandrasekhar worked on the calculation of the statistics of a degenerate Fermi gas.[25] In these papers, Chandrasekhar solved the hydrostatic equation together with the nonrelativistic Fermi gas equation of state,[11] and also treated the case of a relativistic Fermi gas, giving rise to the value of the limit shown above.[9][12][15][26] Chandrasekhar reviews this work in his Nobel Prize lecture.[16] This value was also computed in 1932 by the Soviet physicist Lev Landau,[27] who, however, did not apply it to white dwarfs and concluded that quantum laws might be invalid for stars heavier than 1.5 solar mass.

Chandrasekhar–Eddington dispute

Main page: Physics:Chandrasekhar–Eddington dispute

Chandrasekhar's work on the limit aroused controversy, owing to the opposition of the British astrophysicist Arthur Eddington. Eddington was aware that the existence of black holes was theoretically possible, and also realized that the existence of the limit made their formation possible. However, he was unwilling to accept that this could happen. After a talk by Chandrasekhar on the limit in 1935, he replied:

The star has to go on radiating and radiating and contracting and contracting until, I suppose, it gets down to a few km radius, when gravity becomes strong enough to hold in the radiation, and the star can at last find peace. ... I think there should be a law of Nature to prevent a star from behaving in this absurd way![28]

Eddington's proposed solution to the perceived problem was to modify relativistic mechanics so as to make the law P = K1ρ5/3 universally applicable, even for large ρ.[29] Although Niels Bohr, Fowler, Wolfgang Pauli, and other physicists agreed with Chandrasekhar's analysis, at the time, owing to Eddington's status, they were unwilling to publicly support Chandrasekhar.[30]:{{{1}}} Through the rest of his life, Eddington held to his position in his writings,[31][32][33][34][35] including his work on his fundamental theory.[36] The drama associated with this disagreement is one of the main themes of Empire of the Stars, Arthur I. Miller's biography of Chandrasekhar.[30] In Miller's view:

Chandra's discovery might well have transformed and accelerated developments in both physics and astrophysics in the 1930s. Instead, Eddington's heavy-handed intervention lent weighty support to the conservative community astrophysicists, who steadfastly refused even to consider the idea that stars might collapse to nothing. As a result, Chandra's work was almost forgotten.[30]:{{{1}}}

However, in 1983 in recognition for his work, Chandrasekhar shared a Nobel prize "for his theoretical studies of the physical processes of importance to the structure and evolution of the stars" with William Alfred Fowler.[37]

Applications

The core of a star is kept from collapsing by the heat generated by the fusion of nuclei of lighter elements into heavier ones. At various stages of stellar evolution, the nuclei required for this process are exhausted, and the core collapses, causing it to become denser and hotter. A critical situation arises when iron accumulates in the core, since iron nuclei are incapable of generating further energy through fusion. If the core becomes sufficiently dense, electron degeneracy pressure will play a significant part in stabilizing it against gravitational collapse.[38]

If a main-sequence star is not too massive (less than approximately 8 solar masses), it eventually sheds enough mass to form a white dwarf having mass below the Chandrasekhar limit, which consists of the former core of the star. For more-massive stars, electron degeneracy pressure does not keep the iron core from collapsing to very great density, leading to formation of a neutron star, black hole, or, speculatively, a quark star. (For very massive, low-metallicity stars, it is also possible that instabilities destroy the star completely.)[39][40][41][42] During the collapse, neutrons are formed by the capture of electrons by protons in the process of electron capture, leading to the emission of neutrinos.[38]:{{{1}}} The decrease in gravitational potential energy of the collapsing core releases a large amount of energy on the order of 1046 J (100 foes). Most of this energy is carried away by the emitted neutrinos[43] and the kinetic energy of the expanding shell of gas; only about 1% is emitted as optical light.[44] This process is believed responsible for supernovae of types Ib, Ic, and II.[38]

Type Ia supernovae derive their energy from runaway fusion of the nuclei in the interior of a white dwarf. This fate may befall carbonoxygen white dwarfs that accrete matter from a companion giant star, leading to a steadily increasing mass. As the white dwarf's mass approaches the Chandrasekhar limit, its central density increases, and, as a result of compressional heating, its temperature also increases. This eventually ignites nuclear fusion reactions, leading to an immediate carbon detonation, which disrupts the star and causes the supernova.[45]:{{{1}}}

A strong indication of the reliability of Chandrasekhar's formula is that the absolute magnitudes of supernovae of Type Ia are all approximately the same; at maximum luminosity, MV is approximately −19.3, with a standard deviation of no more than 0.3.[45]:{{{1}}} A 1-sigma interval therefore represents a factor of less than 2 in luminosity. This seems to indicate that all type Ia supernovae convert approximately the same amount of mass to energy.

Super-Chandrasekhar mass supernovas

In April 2003, the Supernova Legacy Survey observed a type Ia supernova, designated SNLS-03D3bb, in a galaxy approximately 4 billion light years away. According to a group of astronomers at the University of Toronto and elsewhere, the observations of this supernova are best explained by assuming that it arose from a white dwarf that had grown to twice the mass of the Sun before exploding. They believe that the star, dubbed the "Champagne Supernova"[46] may have been spinning so fast that a centrifugal tendency allowed it to exceed the limit. Alternatively, the supernova may have resulted from the merger of two white dwarfs, so that the limit was only violated momentarily. Nevertheless, they point out that this observation poses a challenge to the use of type Ia supernovae as standard candles.[47][48][49]

Since the observation of the Champagne Supernova in 2003, several more type Ia supernovae have been observed that are very bright, and thought to have originated from white dwarfs whose masses exceeded the Chandrasekhar limit. These include SN 2006gz, SN 2007if, and SN 2009dc.[50] The super-Chandrasekhar mass white dwarfs that gave rise to these supernovae are believed to have had masses up to 2.4–2.8 solar masses.[50] One way to potentially explain the problem of the Champagne Supernova was considering it the result of an aspherical explosion of a white dwarf. However, spectropolarimetric observations of SN 2009dc showed it had a polarization smaller than 0.3, making the large asphericity theory unlikely.[50]

Tolman–Oppenheimer–Volkoff limit

After a supernova explosion, a neutron star may be left behind (except Ia type supernova explosion, which never leaves any remnants behind). These objects are even more compact than white dwarfs and are also supported, in part, by degeneracy pressure. A neutron star, however, is so massive and compressed that electrons and protons have combined to form neutrons, and the star is thus supported by neutron degeneracy pressure (as well as short-range repulsive neutron-neutron interactions mediated by the strong force) instead of electron degeneracy pressure. The limiting value for neutron star mass, analogous to the Chandrasekhar limit, is known as the Tolman–Oppenheimer–Volkoff limit.[citation needed]

See also

References

  1. "Great Indians: Professor Subrahmanyan Chandrasekhar". https://www.youtube.com/watch?v=n-lJjR7pM7k&t=16s. 
  2. Hawking, S. W.; Israel, W., eds (1989). Three Hundred Years of Gravitation (1st pbk. corrected ed.). Cambridge: Cambridge University Press. ISBN 978-0-521-37976-2. 
  3. Bethe, Hans A.; Brown, Gerald (2003). "How A Supernova Explodes". in Bethe, Hans A.; Brown, Gerald; Lee, Chang-Hwan. Formation And Evolution of Black Holes in the Galaxy: Selected Papers with Commentary. River Edge, NJ: World Scientific. p. 55. ISBN 978-981-238-250-4. Bibcode2003febh.book.....B. https://archive.org/details/formationevoluti00beth_874. 
  4. Mazzali, P. A.; Röpke, F. K.; Benetti, S.; Hillebrandt, W. (2007). "A Common Explosion Mechanism for Type Ia Supernovae". Science 315 (5813): 825–828. doi:10.1126/science.1136259. PMID 17289993. Bibcode2007Sci...315..825M. 
  5. Sean Carroll, Ph.D., Caltech, 2007, The Teaching Company, Dark Matter, Dark Energy: The Dark Side of the Universe, Guidebook Part 2 page 44, Accessed Oct. 7, 2013, "...Chandrasekhar limit: The maximum mass of a white dwarf star, about 1.4 times the mass of the Sun. Above this mass, the gravitational pull becomes too great, and the star must collapse to a neutron star or black hole..."
  6. Eric G. Blackman, "Giants of physics found white-dwarf mass limits", Nature 440, 148 (2006)
  7. Michael Nauenberg, "Edmund C. Stoner and the Discovery of the Maximum Mass of White Dwarfs," Journal for the History of Astronomy, Vol. 39, p. 297-312, (2008)
  8. Virginia Trimble, "Chandrasekhar and the history of astronomy", Fluid Flows to Black Holes, pp. 49-53 (2011)
  9. 9.0 9.1 9.2 Chandrasekhar, S. (1931). "The Maximum Mass of Ideal White Dwarfs". Astrophysical Journal 74: 81–82. doi:10.1086/143324. Bibcode1931ApJ....74...81C. 
  10. Stoner, Edmund C. (1929). "The Limiting Density of White Dwarf Stars". Philosophical Magazine 7 (41): 63–70. doi:10.1080/14786440108564713. 
  11. 11.0 11.1 Chandrasekhar, S. (1931). "The Density of White Dwarf Stars". Philosophical Magazine 11 (70): 592–596. doi:10.1080/14786443109461710. 
  12. 12.0 12.1 12.2 Chandrasekhar, S. (1935). "The Highly Collapsed Configurations of a Stellar Mass (second paper)". Monthly Notices of the Royal Astronomical Society 95 (3): 207–225. doi:10.1093/mnras/95.3.207. Bibcode1935MNRAS..95..207C. 
  13. Standards for Astronomical Catalogues, Version 2.0 , section 3.2.2, web page, accessed 12-I-2007.
  14. 14.0 14.1 Timmes, F. X.; Woosley, S. E.; Weaver, Thomas A. (1996). "The Neutron Star and Black Hole Initial Mass Function". Astrophysical Journal 457: 834–843. doi:10.1086/176778. Bibcode1996ApJ...457..834T. 
  15. 15.0 15.1 Chandrasekhar, S. (1931). "The Highly Collapsed Configurations of a Stellar Mass". Monthly Notices of the Royal Astronomical Society 91 (5): 456–466. doi:10.1093/mnras/91.5.456. Bibcode1931MNRAS..91..456C. 
  16. 16.0 16.1 On Stars, Their Evolution and Their Stability , Nobel Prize lecture, Subrahmanyan Chandrasekhar, December 8, 1983.
  17. Lieb, Elliott H.; Yau, Horng-Tzer (1987). "A rigorous examination of the Chandrasekhar theory of stellar collapse". Astrophysical Journal 323: 140–144. doi:10.1086/165813. Bibcode1987ApJ...323..140L. https://dash.harvard.edu/bitstream/1/32706795/1/1987ApJ___323__140L.pdf. Retrieved 2019-09-04. 
  18. Fowler, R. H. (1926). "On Dense Matter". Monthly Notices of the Royal Astronomical Society 87 (2): 114–122. doi:10.1093/mnras/87.2.114. Bibcode1926MNRAS..87..114F. 
  19. Stoner, Edmund C. (1929). "The Limiting Density of White Dwarf Stars". Philosophical Magazine 7 (41): 63–70. doi:10.1080/14786440108564713. 
  20. Anderson, Wilhelm (1929). "Uber die Grenzdichte der Materie und der Energie". Zeitschrift für Physik 56 (11–12): 851–856. doi:10.1007/BF01340146. Bibcode1929ZPhy...56..851A. 
  21. Stoner, Edmund C. (1930). "The Equilibrium of Dense Stars". Philosophical Magazine 9: 944–963. 
  22. Stoner, E. C. (1932). "The minimum pressure of a degenerate electron gas". Monthly Notices of the Royal Astronomical Society 92 (7): 651–661. doi:10.1093/mnras/92.7.651. Bibcode1932MNRAS..92..651S. 
  23. Frenkel, J. (1928). "Anwendung der Pauli-Fermischen Elektronengastheorie auf das Problem der Kohäsionskräfte". Zeitschrift für Physik 50 (3–4): 234–248. doi:10.1007/BF01328867. Bibcode1928ZPhy...50..234F. .
  24. Yakovlev, D. G. (1994). "The article by Ya I Frenkel' on 'binding forces' and the theory of white dwarfs". Physics-Uspekhi 37 (6): 609–612. doi:10.1070/pu1994v037n06abeh000031. Bibcode1994PhyU...37..609Y. 
  25. Chandrasekhar's biographical memoir at the National Academy of Sciences , web page, accessed 12-01-2007.
  26. Chandrasekhar, S. (1934). "Stellar Configurations with degenerate Cores". The Observatory 57: 373–377. Bibcode1934Obs....57..373C. 
  27. On the Theory of Stars, in Collected Papers of L. D. Landau, ed. and with an introduction by D. ter Haar, New York: Gordon and Breach, 1965; originally published in Phys. Z. Sowjet. 1 (1932), 285.
  28. "Meeting of the Royal Astronomical Society, Friday, 1935 January 11". The Observatory 58: 33–41. 1935. Bibcode1935Obs....58...33.. 
  29. Eddington, A. S. (1935). "On "Relativistic Degeneracy"". Monthly Notices of the Royal Astronomical Society 95 (3): 194–206. doi:10.1093/mnras/95.3.194a. Bibcode1935MNRAS..95..194E. 
  30. 30.0 30.1 30.2 Empire of the Stars: Obsession, Friendship, and Betrayal in the Quest for Black Holes, Arthur I. Miller, Boston, New York: Houghton Mifflin, 2005, ISBN:0-618-34151-X; reviewed at The Guardian: The battle of black holes .
  31. "The International Astronomical Union meeting in Paris, 1935". The Observatory 58: 257–265 [259]. 1935. Bibcode1935Obs....58..257.. 
  32. Eddington, A. S. (1935). "Note on "Relativistic Degeneracy"". Monthly Notices of the Royal Astronomical Society 96: 20–21. doi:10.1093/mnras/96.1.20. Bibcode1935MNRAS..96...20E. 
  33. Eddington, Arthur (1935). "The Pressure of a Degenerate Electron Gas and Related Problems". Proceedings of the Royal Society of London. Series A, Mathematical and Physical Sciences 152 (876): 253–272. doi:10.1098/rspa.1935.0190. Bibcode1935RSPSA.152..253E. 
  34. Relativity Theory of Protons and Electrons, Sir Arthur Eddington, Cambridge: Cambridge University Press, 1936, chapter 13.
  35. Eddington, A. S. (1940). "The physics of white dwarf matter". Monthly Notices of the Royal Astronomical Society 100 (8): 582–594. doi:10.1093/mnras/100.8.582. Bibcode1940MNRAS.100..582E. 
  36. Fundamental Theory, Sir A. S. Eddington, Cambridge: Cambridge University Press, 1946, §43–45.
  37. "The Nobel Prize in Physics 1983" (in en-US). https://www.nobelprize.org/prizes/physics/1983/summary/. 
  38. 38.0 38.1 38.2 Woosley, S. E.; Heger, A.; Weaver, T. A. (2002). "The evolution and explosion of massive stars". Reviews of Modern Physics 74 (4): 1015–1071. doi:10.1103/revmodphys.74.1015. Bibcode2002RvMP...74.1015W. 
  39. Koester, D.; Reimers, D. (1996). "White dwarfs in open clusters. VIII. NGC 2516: a test for the mass-radius and initial-final mass relations". Astronomy and Astrophysics 313: 810–814. Bibcode1996A&A...313..810K. 
  40. Kurtis A. Williams, M. Bolte, and Detlev Koester 2004 An Empirical Initial-Final Mass Relation from Hot, Massive White Dwarfs in NGC 2168 (M35) , Astrophysical Journal 615, pp. L49–L52 arXiv astro-ph/0409447 .
  41. Heger, A.; Fryer, C. L.; Woosley, S. E.; Langer, N.; Hartmann, D. H. (2003). "How Massive Single Stars End Their Life". Astrophysical Journal 591 (1): 288–300. doi:10.1086/375341. Bibcode2003ApJ...591..288H. 
  42. Schaffner-Bielich, Jürgen (2005). "Strange quark matter in stars: a general overview]". Journal of Physics G: Nuclear and Particle Physics 31 (6): S651–S657. doi:10.1088/0954-3899/31/6/004. Bibcode2005JPhG...31S.651S. 
  43. Lattimer, James M.; Prakash, Madappa (2004). "The Physics of Neutron Stars". Science 304 (5670): 536–542. doi:10.1126/science.1090720. PMID 15105490. Bibcode2004Sci...304..536L. 
  44. Schneider, Stephen E.; and Arny, Thomas T.; Readings: Unit 66: End of a star's life , Astronomy 122: Birth and Death of Stars, University of Oregon
  45. 45.0 45.1 Hillebrandt, Wolfgang; Niemeyer, Jens C. (2000). "Type IA Supernova Explosion Models". Annual Review of Astronomy and Astrophysics 38: 191–230. doi:10.1146/annurev.astro.38.1.191. Bibcode2000ARA&A..38..191H. 
  46. Branch, David (21 September 2006). "Astronomy: Champagne supernova". Nature 443 (7109): 283–284. doi:10.1038/443283a. PMID 1698869. Bibcode2006Natur.443..283B. 
  47. "The weirdest type Ia supernova yet" (Press release). LBL. Archived from the original on 6 July 2017. Retrieved 13 January 2007.
  48. "Champagne supernova challenges ideas about how supernovae work". spacedaily.com (Press release). Archived from the original on 1 July 2017. Retrieved 13 January 2007.
  49. Howell, D. Andrew (2006). "The type Ia supernova SNLS-03D3bb from a super-Chandrasekhar-mass white dwarf star". Nature 443 (7109): 308–311. doi:10.1038/nature05103. PMID 16988705. Bibcode2006Natur.443..308H. 
  50. 50.0 50.1 50.2 Hachisu, Izumi et al. (2012). "A single degenerate progenitor model for type Ia supernovae highly exceeding the Chandrasekhar mass limit". The Astrophysical Journal 744 (1): 76–79. doi:10.1088/0004-637X/744/1/69. Article 69. Bibcode2012ApJ...744...69H. 

Further reading