Difference between revisions of "White dwarf" - New World Encyclopedia

From New World Encyclopedia
m
m (Robot: Remove claimed tag)
Line 1: Line 1:
{{Claimed}}
+
 
 
{{Otheruses}}
 
{{Otheruses}}
  

Revision as of 18:59, 2 April 2008

For other uses, see White dwarf (disambiguation).
Image of Sirius A and Sirius B taken by the Hubble Space Telescope. Sirius B, which is a white dwarf, can be seen as a faint dot to the lower left of the much brighter Sirius A.

A white dwarf, also called a degenerate dwarf, is the kind of star which a main-sequence star of low or medium mass will become in the last stage of its evolution. After its hydrogen-fusing lifetime, such a star will expand to a red giant which fuses helium to carbon and oxygen in its core by the triple-alpha process. If a red giant has insufficient mass to generate the core temperatures required to fuse carbon, an inert mass of carbon and oxygen will build up at its center. After shedding its outer layers to form a planetary nebula, it will leave behind this core, which forms the remnant white dwarf.[1] Usually, therefore, white dwarfs are composed of carbon and oxygen. It is also possible that core temperatures suffice to fuse carbon but not neon, in which case an oxygen-neon-magnesium white dwarf may be formed.[2] Also, some helium[3][4] white dwarfs appear to have been formed by mass loss in binary systems. White dwarfs are thought to be the final state of over 97% of all stars in our galaxy.[5], §1. They comprise roughly 6% of all known stars in the solar neighborhood.

The material in a white dwarf no longer undergoes fusion reactions, so the star has no source of energy, nor is it supported against gravitational collapse by the heat generated by fusion. It is supported only by electron degeneracy pressure and is therefore extremely dense, with a typical mass comparable to the Sun's and a volume comparable to the Earth's. The physics of degeneracy yields a maximum mass for a nonrotating white dwarf, the Chandrasekhar limit—approximately 1.4 solar masses—beyond which it cannot be supported by degeneracy pressure. A carbon-oxygen white dwarf that approaches this mass limit, typically by mass transfer from a companion star, may explode as a Type Ia supernova via a process known as carbon detonation.[1][6]

A white dwarf is very hot when it is formed, but since it has no source of energy, it will gradually radiate away its energy and cool down. This is the source of its faint luminosity, which initially has a high color temperature, but will dim and redden with time. Over a very long period of time, a white dwarf will cool to temperatures at which it is no longer visible and become a cold black dwarf.[1] However, since no white dwarf can be older than the age of the Universe (approximately 13.7 billion years)[7], even the oldest white dwarfs still radiate at temperatures of a few thousand kelvin, and no black dwarfs are thought to exist yet.[5][6]

Discovery

The first white dwarf discovered was in the triple star system of 40 Eridani, which contains the relatively bright main sequence star 40 Eridani A, orbited at a distance by the closer binary system of the white dwarf 40 Eridani B and the main sequence red dwarf 40 Eridani C. The pair 40 Eridani B/C was discovered by Friedrich Wilhelm Herschel on January 31, 1783;[8], p. 73 it was again observed by Friedrich Georg Wilhelm Struve in 1825 and by Otto Wilhelm von Struve in 1851.[9][10] In 1910, it was discovered by Henry Norris Russell, Edward Charles Pickering and Williamina Fleming that despite being a dim star, 40 Eridani B was of spectral type A, or white.[11] In 1939, Russell looked back on the discovery:[12], p. 1

I was visiting my friend and generous benefactor, Prof. Edward C. Pickering. With characteristic kindness, he had volunteered to have the spectra observed for all the stars—including comparison stars—which had been observed in the observations for stellar parallax which Hinks and I made at Cambridge, and I discussed. This piece of apparently routine work proved very fruitful—it led to the discovery that all the stars of very faint absolute magnitude were of spectral class M. In conversation on this subject (as I recall it), I asked Pickering about certain other faint stars, not on my list, mentioning in particular 40 Eridani B. Characteristically, he sent a note to the Observatory office and before long the answer came (I think from Mrs Fleming) that the spectrum of this star was A. I knew enough about it, even in these paleozoic days, to realize at once that there was an extreme inconsistency between what we would then have called ‘possible’ values of the surface brightness and density. I must have shown that I was not only puzzled but crestfallen, at this exception to what looked like a very pretty rule of stellar characteristics; but Pickering smiled upon me, and said: ‘It is just these exceptions that lead to an advance in our knowledge’, and so the white dwarfs entered the realm of study!

The spectral type of 40 Eridani B was officially described in 1914 by Walter Adams.[13]

The companion of Sirius, Sirius B, was next to be discovered. During the nineteenth century, positional measurements of some stars became precise enough to measure small changes in their location. Friedrich Bessel used just such precise measurements to determine that the stars Sirius (α Canis Majoris) and Procyon (α Canis Minoris) were changing their positions. In 1844 he predicted that both stars had unseen companions:[14]

If we were to regard Sirius and Procyon as double stars, the change of their motions would not surprise us; we should acknowledge them as necessary, and have only to investigate their amount by observation. But light is no real property of mass. The existence of numberless visible stars can prove nothing against the existence of numberless invisible ones.

Bessel roughly estimated the period of the companion of Sirius to be about half a century[14]; C. H. F. Peters computed an orbit for it in 1851.[15] It was not until January 31, 1862 that Alvan Graham Clark observed a previously unseen star close to Sirius, later identified as the predicted companion.[15] Walter Adams announced in 1915 that he had found the spectrum of Sirius B to be similar to that of Sirius.[16]

In 1917, Adriaan Van Maanen discovered Van Maanen's Star, an isolated white dwarf.[17] These three white dwarfs, the first discovered, are the so-called classical white dwarfs.[12], p. 2 Eventually, many faint white stars were found which had high proper motion, indicating that they could be suspected to be low-luminosity stars close to the Earth, and hence white dwarfs. Willem Luyten appears to have been the first to use the term white dwarf when he examined this class of stars in 1922;[11][18][19][20][21] the term was later popularized by Arthur Stanley Eddington.[22][11] Despite these suspicions, the first non-classical white dwarf was not definitely identified until the 1930s. By 1939, a total of 18 white dwarfs were known.[12], p. 3 Luyten and others continued to search for white dwarfs in the 1940s. By 1950, over a hundred were known[23], and by 1999, over 2,000 were known.[24] Since then the Sloan Digital Sky Survey has found over 9,000 white dwarfs, mostly new.[25]

Composition and structure

Although white dwarfs are known with estimated masses as low as 0.17[26] and as high as 1.33[27] solar masses, the mass distribution is strongly peaked at 0.6 solar masses, with the bulk of observed white dwarfs having masses from 0.5 to 0.7 solar masses.[27] The estimated radii of observed white dwarfs, however, are typically between 0.008 and 0.02 times the radius of the Sun[28]; this is comparable to the Earth's radius of approximately 0.009 solar radii. A white dwarf, then, packs mass comparable to the Sun's into a volume that is typically a million times smaller than the Sun's; the average density of matter in a white dwarf must therefore be, very roughly, 1,000,000 times greater than the average density of the Sun, or approximately 106 grams per cubic centimeter.[6] White dwarfs are composed of one of the densest forms of matter known, surpassed only by other compact stars such as neutron stars, black holes and, hypothetically, quark stars.[29]

White dwarfs were found to be extremely dense soon after their discovery. If a star is in a binary system, as is the case for Sirius B and 40 Eridani B, it is possible to estimate its mass from observations of the binary orbit. This was done for Sirius B by 1910,[30], yielding a mass estimate of 0.94 solar masses. (A more modern estimate is 1.00 solar masses.)[31] Since hotter bodies radiate more than colder ones, a star's surface brightness can be estimated from its effective surface temperature, and hence from its spectrum. If the star's distance is known, its overall luminosity can also be estimated. Comparison of the two figures yields the star's radius. Reasoning of this sort led to the realization, puzzling to astronomers at the time, that Sirius B and 40 Eridani B must be very dense. For example, when Ernst Öpik estimated the density of a number of visual binary stars in 1916, he found that 40 Eridani B had a density of over 25,000 times the Sun's, which was so high that he called it "impossible".[32] As Arthur Stanley Eddington put it later in 1927:[33], p. 50

We learn about the stars by receiving and interpreting the messages which their light brings to us. The message of the Companion of Sirius when it was decoded ran: ‘I am composed of material 3,000 times denser than anything you have ever come across; a ton of my material would be a little nugget that you could put in a matchbox.’ What reply can one make to such a message? The reply which most of us made in 1914 was—‘Shut up. Don't talk nonsense.’

As Eddington pointed out in 1924, densities of this order implied that, according to the theory of general relativity, the light from Sirius B should be gravitationally redshifted.[22] This was confirmed when Adams measured this redshift in 1925.[34]

Such densities are possible because white dwarf material is not composed of atoms bound by chemical bonds, but rather consists of a plasma of unbound nuclei and electrons. There is therefore no obstacle to placing nuclei closer to each other than electron orbitals would normally allow.[22] Eddington, however, wondered what would happen when this plasma cooled and the energy which kept the atoms ionized was no longer present.[35] This paradox was resolved by R. H. Fowler in 1926 by an application of the newly devised quantum mechanics. Since electrons were known to obey Fermi-Dirac statistics, also introduced in 1926, the Pauli exclusion principle meant that no two electrons could occupy the same state. At zero temperature, therefore, electrons had to occupy a band of energy levels at the bottom of the Fermi sea—a state called degenerate—meaning that a star could cool to zero temperature and still possess high energy. This also meant that compression of the electrons increased the number of electrons in a given volume and therefore raised the maximum energy level occupied by an electron, causing pressure.[35] This electron degeneracy pressure is what supports a white dwarf against gravitational collapse. It depends only on density and not on temperature. Degenerate matter is relatively compressible; this means that the density of a high-mass white dwarf is so much greater than that of a low-mass white dwarf that the radius of a white dwarf decreases as its mass increases.[6]

Another consequence of being supported by electron degeneracy pressure is the existence of a limiting mass which no white dwarf can exceed. These limiting masses were first published in 1929 by Wilhelm Anderson[36] and in 1930 by Edmund C. Stoner.[37] The modern value of the limit was first published in 1931 by Subrahmanyan Chandrasekhar in his paper "The Maximum Mass of Ideal White Dwarfs".[38] For a nonrotating white dwarf, it is equal to approximately 5.7/μe2 solar masses, where μe is the average molecular weight per electron of the star.[39], eq. (63) As the carbon-12 and oxygen-16 which predominantly compose a carbon-oxygen white dwarf both have atomic number equal to half their atomic weight, one should take μe equal to 2 for such a star, [40]leading to the commonly-quoted value of 1.4 solar masses. (Near the beginning of the 20th century, there was reason to believe that stars were composed chiefly of heavy elements,[37], p. 955 so, in his 1931 paper, Chandrasekhar set the average molecular weight per electron, μe, equal to 2.5, giving a limit of 0.91 solar masses.) Together with William Alfred Fowler, Chandrasekhar received the Nobel prize for this and other work in 1983.[41] The limiting mass is now called the Chandrasekhar limit.

If a white dwarf were to exceed the Chandrasekhar limit, and nuclear reactions did not take place, the pressure exerted by electrons would no longer be able to balance the force of gravity, and it would collapse into a denser object such as a neutron star or black hole.[42] However, carbon-oxygen white dwarfs accreting mass from a neighboring star undergo a runaway nuclear fusion reaction, which leads to a Type Ia supernova explosion in which the white dwarf is destroyed, just prior to reaching the limiting mass.[43]

White dwarfs have low luminosity and therefore occupy a strip at the bottom of the Hertzsprung-Russell diagram. They should not be confused with low-luminosity objects at the low-mass end of the main sequence, such as the hydrogen-fusing red dwarfs, whose cores are supported in part by thermal pressure[44], or the even lower-temperature brown dwarfs.[45]

Mass-radius relationship and mass limit

It is simple to derive a rough relationship between the mass and radii of white dwarfs using an energy minimization argument. The energy of the white dwarf can be approximated by taking it to be the sum of its gravitational potential energy and kinetic energy. The gravitational potential energy of a unit mass piece of white dwarf, Eg, will be on the order of -G M / R, where M is the mass of the white dwarf and R is its radius. The kinetic energy of the unit mass, Ek, will primarily come from the motion of electrons, so it will be approximately N p2/2m, where p is the average electron momentum, m is the electron mass, and N is the number of electrons per unit mass. Since the electrons are degenerate, we can estimate p to be on the order of the uncertainty in momentum, Δ p, given by the uncertainty principle, which says that Δ p Δ x is on the order of the reduced Planck constant, ħ. Δ x will be on the order of the average distance between electrons, which will be approximately n-1/3, i.e., the reciprocal of the cube root of the number density, n, of electrons per unit volume. Since there are N M electrons in the white dwarf and its volume is on the order of R3, n will be on the order of N M / R3.[40] Solving for the kinetic energy per unit mass, Ek, we find that

The white dwarf will be at equilibrium when its total energy, Eg + Ek, is minimized. At this point, the kinetic and gravitational potential energies should be comparable, so we may derive a rough mass-radius relationship by equating their magnitudes:

Solving this for the radius, R, gives[40]

Dropping N, which depends only on the composition of the white dwarf, and the universal constants leaves us with a relationship between mass and radius:

i.e., the radius of a white dwarf is inversely proportional to the cube root of its mass.

Since this analysis uses the non-relativistic formula p2/2m for the kinetic energy, it is non-relativistic. If we wish to analyze the situation where the electron velocity in a white dwarf is close to the speed of light, c, we should replace p2/2m by the extreme relativistic approximation p c for the kinetic energy. With this substitution, we find

If we equate this to the magnitude of Eg, we find that R drops out and the mass, M, is forced to be [40]

To interpret this result, observe that as we add mass to a white dwarf, its radius will decrease, so, by the uncertainty principle, the momentum, and hence the velocity, of its electrons will increase. As this velocity approaches c, the extreme relativistic analysis becomes more exact, meaning that the mass M of the white dwarf must approach Mlimit. Therefore, no white dwarf can be heavier than the limiting mass Mlimit.

File:ChandrasekharLimitGraph.png
Radius versus mass for a model white dwarf.

For a more accurate computation of the mass-radius relationship and limiting mass of a white dwarf, one must compute the equation of state which describes the relationship between density and pressure in the white dwarf material. If the density and pressure are both set equal to functions of the radius from the center of the star, the system of equations consisting of the hydrostatic equation together with the equation of state can then be solved to find the structure of the white dwarf at equilibrium. In the non-relativistic case, we will still find that the radius is inversely proportional to the cube root of the mass.[39], eq. (80) Relativistic corrections will alter the result so that the radius becomes zero at a finite value of the mass. This is the limiting value of the mass—called the Chandrasekhar limit—at which the white dwarf can no longer be supported by electron degeneracy pressure. The graph at the right shows the result of such a computation. It shows how radius varies with mass for non-relativistic (green curve) and relativistic (red curve) models of a white dwarf. Both models treat the white dwarf as a cold Fermi gas in hydrostatic equilibrium. The average molecular weight per electron, μe, has been set equal to 2. Radius is measured in standard solar radii and mass in standard solar masses.[46][39]

These computations all assume that the white dwarf is nonrotating. If the white dwarf is rotating, the equation of hydrostatic equilibrium must be modified to take into account the centrifugal pseudo-force arising from working in a rotating frame.[47] For a uniformly rotating white dwarf, the limiting mass increases only slightly. However, if the star is allowed to rotate nonuniformly, and viscosity is neglected, then, as was pointed out by Fred Hoyle in 1947[48], there is no limit to the mass for which it is possible for a model white dwarf to be in static equilibrium. Not all of these model stars, however, will be dynamically stable.[49]

Radiation and cooling

The visible radiation emitted by white dwarfs varies over a wide color range, from the blue-white color of an O-type main sequence star to the red of a M-type red dwarf.[50] White dwarf effective surface temperatures extend from over 150,000K[24] to under 4,000K.[51][52] In accordance with the Stefan-Boltzmann law, luminosity increases with increasing surface temperature; this surface temperature range corresponds to a luminosity from over 100 times the Sun's to under 1/10,000th that of the Sun's.[52] Hot white dwarfs, with surface temperatures in excess of 30,000K, have been observed to be sources of soft (i.e., lower-energy) X-rays. This enables the composition and structure of their atmospheres to be studied by soft X-ray and extreme ultraviolet observations.[53]

A comparison between the white dwarf IK Pegasi B (center), its A-class companion IK Pegasi A (left) and the Sun (right). This white dwarf has a surface temperature of 35,500 K.

Unless the white dwarf accretes matter from a companion star or other source, this radiation comes from its stored heat, which is not replenished. White dwarfs have an extremely small surface area to radiate this heat from, so they remain hot for a long period of time.[1] As a white dwarf cools, its surface temperature decreases, the radiation which it emits reddens, and its luminosity decreases. Since the white dwarf has no energy sink other than radiation, it follows that its cooling slows with time. A white dwarf may cool from a surface temperature of 20,000K to one of 5,000K in approximately the same amount of time it takes to cool from one of 5,000K to one of 4,000K.[54], eq. (2.3). Although white dwarf material is initially plasma—a fluid composed of nuclei and electrons—it was theoretically predicted in the 1960s that at a late stage of cooling, it should crystallize, starting at the center of the star.[55] The crystal structure is thought to be a body-centered cubic lattice.[56][5] In 1995 it was pointed out that asteroseismological observations of pulsating white dwarfs yielded a potential test of the crystallization theory,[57] and in 2004, Travis Metcalfe and a team of researchers at the Harvard-Smithsonian Center for Astrophysics estimated, on the basis of such observations, that approximately 90% of the mass of BPM 37093 had crystallized.[55][58][59][60] Other work gives a crystallized mass fraction of between 32% and 82%.[61]

Most observed white dwarfs have relatively high surface temperatures, between 8,000K and 40,000K.[62][25] A white dwarf, though, spends more of its lifetime at cooler temperatures than at hotter temperatures, so we should expect that there are more cool white dwarfs than hot white dwarfs. Once we adjust for the selection effect that hotter, more luminous white dwarfs are easier to observe, we do find that decreasing the temperature range examined results in finding more white dwarfs.[63] This trend stops when we reach extremely cool white dwarfs; few white dwarfs are observed with surface temperatures below 4,000K[64], and one of the coolest so far observed, WD 0346+246, has a surface temperature of approximately 3,900 K.[51] The reason for this is that, as the Universe's age is finite, there has not been time for white dwarfs to cool down below this temperature. The white dwarf luminosity function can therefore be used to find the time when stars started to form in a region; an estimate for the age of the Galactic disk found in this way is 8 billion years.[63]

A white dwarf will eventually cool and become a non-radiating black dwarf in approximate thermal equilibrium with its surroundings and with the cosmic background radiation. However, no black dwarfs are thought to exist yet.[6]

Atmosphere and spectra

Although most white dwarfs are thought to be composed of carbon and oxygen, spectroscopy typically shows that their emitted light comes from an atmosphere which is observed to be either hydrogen-dominated or helium-dominated. The dominant element is usually at least 1,000 times more abundant than all other elements. As explained by Schatzman in the 1940s, the high surface gravity is thought to cause this purity by gravitationally separating the atmosphere so that heavy elements are on the bottom and lighter ones on top.[65][66], §5–6 This atmosphere, the only part of the white dwarf visible to us, is thought to be the top of an envelope which is a residue of the star's envelope in the AGB phase and may also contain material accreted from the interstellar medium. The envelope is believed to consist of a helium-rich layer with mass no more than 1/100th of the star's total mass, which, if the atmosphere is hydrogen-dominated, is overlain by a hydrogen-rich layer with mass approximately 1/10,000th of the stars total mass.[52][67], §4–5.

Although thin, these outer layers determine the thermal evolution of the white dwarf. The degenerate electrons in the bulk of a white dwarf conduct heat well. Most of a white dwarf's mass is therefore almost isothermal, and it is also hot: a white dwarf with surface temperature between 8,000K and 16,000K will have a core temperature between approximately 5,000,000K and 20,000,000K. The white dwarf is kept from cooling very quickly only by its outer layers' opacity to radiation.[52]

Primary and secondary features
A H lines present; no He I or metal lines
B He I lines; no H or metal lines
C Continuous spectrum; no lines
O He II lines, accompanied by He I or H lines
Z Metal lines; no H or He I lines
Q Carbon lines present
X Unclear or unclassifiable spectrum
Secondary features only
P Magnetic white dwarf with detectable polarization
H Magnetic white dwarf without detectable polarization
E Emission lines present
V Variable
White dwarf spectral types[24]

The first attempt to classify white dwarf spectra appears to have been by G. P. Kuiper in 1941[50][68], and various classification schemes have been proposed and used since then.[69][70] The system currently in use was introduced by Edward M. Sion and his coauthors in 1983 and has been subsequently revised several times. It classifies a spectrum by a symbol which consists of an initial D, a letter describing the primary feature of the spectrum followed by an optional sequence of letters describing secondary features of the spectrum (as shown in the table to the right), and a temperature index number, computed by dividing 50,400K by the effective temperature. For example:

  • A white dwarf with only He I lines in its spectrum and an effective temperature of 15,000 K could be given the classification of DB3, or, if warranted by the precision of the temperature measurement, DB3.5.
  • A white dwarf with a polarized magnetic field, an effective temperature of 17,000 K, and a spectrum domainated by He I lines which also had hydrogen features could be given the classification of DBAP3.

The symbols ? and : may also be used if the correct classification is uncertain.[50][24]

White dwarfs whose primary spectral classification is DA have hydrogen-dominated atmospheres. They make up the majority (approximately three-quarters) of all observed white dwarfs.[52] The classifiable remainder (DB, DC, DO, DZ, and DQ) have helium-dominated atmospheres. Assuming that carbon and metals are not present, which spectral classification is seen depends on the effective temperature. Between approximately 100,000K to 45,000K, the spectrum will be classified DO, dominated by singly ionized helium. From 30,000K to 12,000K, the spectrum will be DB, showing neutral helium lines, and below about 12,000K, the spectrum will be featureless and classified DC.[67],§ 2.4[52] The reason for the absence of white dwarfs with helium-dominated atmospheres and effective temperatures between 30,000K and 45,000K, called the DB gap, is not clear. It is suspected to be due to competing atmospheric evolutionary processes, such as gravitational separation and convective mixing.[52]

Magnetic field

Magnetic fields in white dwarfs with a strength at the surface of ~1 million Gauss were predicted by P. M. S. Blackett in 1947 as a consequence of a physical law he had proposed which stated that an uncharged, rotating body should generate a magnetic field proportional to its angular momentum.[71] This putative law, sometimes called the Blackett effect, was never generally accepted, and by the 1950s even Blackett felt it had been refuted.[72], pp. 39–43 In the 1960s, it was proposed that white dwarfs might have magnetic fields because of conservation of total surface magnetic flux during the evolution of a non-degenerate star to a white dwarf. A surface magnetic field of ~100 Gauss in the progenitor star would thus become a surface magnetic field of ~100·1002=1 million Gauss once the star's radius had shrunk by a factor of 100.[66], §8;[73], p. 484 The first magnetic white dwarf to be observed was GJ 742, which was detected to have a magnetic field in 1970 by its emission of circularly polarized light.[74] It is thought to have a surface field of approximately 300 million Gauss.[66], §8 Since then magnetic fields have been discovered in well over 100 white dwarfs, ranging from 2,000 to 109 Gauss. Only a small number of white dwarfs have been examined for fields, and it has been estimated that at least 10% of white dwarfs have fields in excess of 1 million Gauss.[75][76]

Variability

DAV (GCVS: ZZA) DA spectral type, having only hydrogen absorption lines in its spectrum
DBV (GCVS: ZZB) DB spectral type, having only helium absorption lines in its spectrum
DOV (GCVS: ZZO) DO spectral type, showing He II and C IV absorption lines in its spectrum
Types of pulsating white dwarf[77]
Main article: Pulsating white dwarf
See also: Cataclysmic variables

Early calculations suggested that there might be white dwarfs whose luminosity varied with a period of around 10 seconds, but searches in the 1960s failed to observe this.[66], § 7.1.1;[78] The first variable white dwarf found was HL Tau 76; in 1965 and 1966, Arlo U. Landolt observed it to vary with a period of approximately 12.5 minutes.[79] The reason for this period being longer than predicted is that the variability of EGGR 265, like that of the other pulsating variable white dwarfs known, arises from non-radial gravity wave pulsations.[66], § 7. Known types of pulsating white dwarf include the DAV, or ZZ Ceti, stars, including HL Tau 76, with hydrogen-dominated atmospheres and the spectral type DA[66], pp. 891, 895; DBV, or V777 Her, stars, with helium-dominated atmospheres and the spectral type DB[52], p. 3525; and DOV, or GW Vir, stars, with the spectral type DO.[66], p. 894 DAV, DBV and DOV variables all exhibit small (1%–30%) variations in light output, arising from a superposition of vibrational modes with periods of hundreds to thousands of seconds. Observation of these variations gives asteroseismological evidence about the interiors of white dwarfs.[80]

Formation

White dwarfs are thought to represent the end point of stellar evolution for main-sequence stars with masses from about 0.07 to 10 solar masses.[81][5] The composition of the white dwarf produced will differ depending on the initial mass of the star.

Stars with very low mass

If the mass of a main-sequence star is lower than approximately half a solar mass, it will never become hot enough to fuse helium at its core. It is thought that, over a lifespan exceeding the age (~13.7 billion years[7]) of the Universe, such a star will eventually burn all its hydrogen and end its evolution as a helium white dwarf composed chiefly of helium-4 nuclei. Owing to the time this process takes, it is not thought to be the origin of observed helium white dwarfs. Rather, they are thought to be the product of mass loss in binary systems.[3][4][82][83][84][1]

Stars with low to medium mass

If the mass of a main-sequence star is between approximately 0.5 and 8 solar masses, its core will become sufficiently hot to fuse helium into carbon and oxygen via the triple-alpha process, but it will never become sufficiently hot to fuse carbon into neon. Near the end of the period in which it undergoes fusion reactions, such a star will have a carbon-oxygen core which does not undergo fusion reactions, surrounded by an inner helium-burning shell and an outer hydrogen-burning shell. On the Hertzsprung-Russell diagram, it will be found on the asymptotic giant branch. It will then expel most of its outer material, creating a planetary nebula, until only the carbon-oxygen core is left. This process is responsible for the carbon-oxygen white dwarfs which form the vast majority of observed white dwarfs.[82][85][86]

Stars with medium to high mass

If a star is sufficiently massive, its core will eventually become sufficiently hot to fuse carbon to neon, and then to fuse neon to iron. Such a star will not become a white dwarf as the mass of its central, non-fusing, core, supported by electron degeneracy pressure, will eventually exceed the largest possible mass supportable by degeneracy pressure. At this point the core of the star will collapse and it will explode in a core-collapse supernova which will leave behind a remnant neutron star, black hole, or possibly a more exotic form of compact star.[81][87] Some main-sequence stars, of perhaps 8 to 10 solar masses, although sufficiently massive to fuse carbon to neon and magnesium, may be insufficiently massive to fuse neon. Such a star may leave a remnant white dwarf composed chiefly of oxygen, neon, and magnesium, provided that its core does not collapse, and provided that fusion does not proceed so violently as to blow apart the star in a supernova.[88][89] Although some isolated white dwarfs have been identified which may be of this type, most evidence for the existence of such stars comes from the novae called ONeMg or neon novae. The spectra of these novae exhibit abundances of neon, magnesium, and other intermediate-mass elements which appear to be only explicable by the accretion of material onto an oxygen-neon-magnesium white dwarf.[2][90][91]

Fate

A white dwarf is stable once formed and will continue to cool almost indefinitely. Assuming that the Universe continues to expand, it is thought that in 1019 to 1020 years, the galaxies will evaporate as their stars escape into intergalactic space.[92], §IIIA. White dwarfs should generally survive this, although an occasional collision between white dwarfs may produce a new fusing star or a super-Chandrasekhar mass white dwarf which will explode in a type Ia supernova.[92], §IIIC, IV. The subsequent lifetime of white dwarfs is thought to be on the order of the lifetime of the proton, known to be at least 1032 years. Some simple grand unified theories predict a proton lifetime of no more than 1049 years. If these theories are not valid, the proton may decay by more complicated nuclear processes, or by quantum gravitational processes involving a virtual black hole; in these cases, the lifetime is estimated to be no more than 10200 years. If protons do decay, the mass of a white dwarf will decrease very slowly with time as its nuclei decay, until it loses so much mass as to become a nondegenerate lump of matter, and finally disappears completely.[92], §IV.

Stellar system

A white dwarf's stellar and planetary system is inherited from its progenitor star and may interact with the white dwarf in various ways. Infrared spectroscopic observations made by NASA's Spitzer Space Telescope of the central star of the Helix Nebula suggest the presence of a dust cloud, which may be caused by cometary collisions. It is possible that infalling material from this may cause X-ray emission from the central star.[93][94] Similarly, observations made in 2004 indicated the presence of a dust cloud around the young white dwarf star G29-38 (estimated to have formed from its AGB progenitor about 500 million years ago), which may have been created by tidal disruption of a comet passing close to the white dwarf.[95] If a white dwarf is in a binary system, it may accrete matter from its stellar companion. This leads to a variety of phenomena, including novae and Type Ia supernovae.

Type Ia supernovae

Multiwavelength X-ray image of SN 1572 or Tycho's Nova, the remnant of a Type Ia supernova.
Main article: Type Ia supernova

The mass of an isolated, nonrotating white dwarf cannot exceed the Chandrasekhar limit of ~1.4 solar masses. (This limit may increase if the white dwarf is rotating rapidly and nonuniformly.)[96] White dwarfs in binary systems, however, can increase in mass by accreting material from a companion star. If the accreted material were to push the mass of the white dwarf beyond the limit, degeneracy pressure would no longer support the star, and an electron-capture collapse would ensue.[42] During the accretion process, however, the central density and temperature of the star will increase. In a carbon-oxygen white dwarf, it is believed that the compressional heating of the core leads to ignition of carbon fusion as the mass approaches the limit.[43] Because the white dwarf is supported against gravity by quantum degeneracy pressure instead of by thermal pressure, adding heat to the star's interior increases its temperature but not its pressure, so the white dwarf does not expand and cool in response. Rather, the increased temperature accelerates the rate of the fusion reaction, in a runaway process that feeds on itself. The thermonuclear flame consumes much of the white dwarf in a few seconds, causing an explosion that obliterates the star.[6][43][97]

Cataclysmic variables

Main article: Cataclysmic variable star

When accretion of material does not push a white dwarf close to the Chandrasekhar limit, accreted hydrogen-rich material on the surface may still ignite in a thermonuclear explosion. Since the white dwarf's core remains intact, these surface explosions can be repeated as long as accretion continues. This weaker kind of repetitive cataclysmic phenomenon is called a (classical) nova. Astronomers have also observed dwarf novae, which have smaller, more frequent luminosity peaks than classical novae. These are thought to not be caused by fusion but rather by the release of gravitational potential energy during accretion. In general, binary systems with a white dwarf accreting matter from a stellar companion are called cataclysmic variables. As well as novae and dwarf novae, several other classes of these variables are known.[6][43][98][99] Both fusion- and accretion-powered cataclysmic variables have been observed to be X-ray sources.[99]

See also

  • Pulsating white dwarf
  • Stellar classification
  • Timeline of white dwarfs, neutron stars, and supernovae
  • Degenerate matter
  • Black dwarf
  • Supernova
  • Red dwarf
  • Brown dwarf

Notes

  1. 1.0 1.1 1.2 1.3 1.4 Richmond, Michael. Late stages of evolution for low-mass stars. Rochester Institute of Technology. Retrieved September 18, 2007.
  2. 2.0 2.1 Werner, K., N. J. Hammer, T. Nagel, T. Rauch, and S. Dreizler. 2005. 14th European Workshop on White Dwarfs; Proceedings of a meeting held at Kiel, July 19-23, 2004. D. Koester and S. Moehler, eds. San Francisco, CA: Astronomical Society of the Pacific. Retrieved September 18, 2007.
  3. 3.0 3.1 Liebert, J.; P. Bergeron; D. Eisenstein; H.C. Harris; S.J. Kleinman; A. Nitta; J. Krzesinski. 2004. A Helium White Dwarf of Extremely Low Mass. The Astrophysical Journal. 606:2:L147-L149. Retrieved September 18, 2007.
  4. 4.0 4.1 Cosmic weight loss: The lowest mass white dwarf. Harvard-Smithsonian Center for Astrophysics. Retrieved September 18, 2007.
  5. 5.0 5.1 5.2 5.3 Fontaine, G., P. Brassard, and P. Bergeron. 2001. The Potential of White Dwarf Cosmochronology. Publications of the Astronomical Society of the Pacific. 113:782:409–435. Retrieved September 18, 2007.
  6. 6.0 6.1 6.2 6.3 6.4 6.5 6.6 Johnson, Jennifer. Extreme Stars: White Dwarfs & Neutron Stars. Ohio State University. Retrieved September 18, 2007.
  7. 7.0 7.1 Spergel, D.N., R. Bean, O. Doré, M.R. Nolta, C.L. Bennett, J. Dunkley, G. Hinshaw, N. Jarosik, E. Komatsu, L. Page, H.V. Peiris, L. Verde, M. Halpern, R.S. Hill, A. Kogut, M. Limon, S.S. Meyer, N. Odegard, G.S. Tucker, J.L. Weiland, E. Wollack, and E.L. Wright. 2007. Wilkinson Microwave Anisotropy Probe (WMAP) Three Year Results: Implications for Cosmology. Retrieved September 18, 2007.
  8. Herschel, William. 1783. Catalogue of Double Stars. Philosophical Transactions of the Royal Society of London. 75:40–126. Retrieved September 18, 2007.
  9. van den Bos, W.H. 1926. The orbit and the masses of 40 Eridani BC. Bulletin of the Astronomical Institutes of the Netherlands. 3:98:128–132. Retrieved September 18, 2007.
  10. Heintz, W.D. 1974. Astrometric study of four visual binaries. Astronomical Journal. 79:7:819–825. Retrieved September 18, 2007.
  11. 11.0 11.1 11.2 Holberg, J.B. 2005. How Degenerate Stars Came to be Known as White Dwarfs. Bulletin of the American Astronomical Society. 37:1503. Retrieved September 18, 2007.
  12. 12.0 12.1 12.2 Schatzman, E. 1958. White Dwarfs. Amsterdam: North-Holland.
  13. Adams, Walter S. 1914. An A-Type Star of Very Low Luminosity. Publications of the Astronomical Society of the Pacific. 26:155:198. Retrieved September 18, 2007.
  14. 14.0 14.1 Bessel, F.W. communicated by J. F. W. Herschel. 1844. On the Variations of the Proper Motions of Procyon and Sirius. Monthly Notices of the Royal Astronomical Society. 6:136–141. Retrieved September 18, 2007.
  15. 15.0 15.1 Flammarion, Camille. 1877. The Companion of Sirius. The Astronomical Register. 15:176:186–189. Retrieved September 18, 2007.
  16. Adams, W.S. 1915. The Spectrum of the Companion of Sirius. Publications of the Astronomical Society of the Pacific. 27:161:236–237. Retrieved September 18, 2007.
  17. van Maanen, A. 1917. Two Faint Stars with Large Proper Motion. Publications of the Astronomical Society of the Pacific. 29:172:258–259. Retrieved September 18, 2007.
  18. Luyten, Willem, J. 1922. The Mean Parallax of Early-Type Stars of Determined Proper Motion and Apparent Magnitude. Publications of the Astronomical Society of the Pacific. 34:199:156–160. Retrieved September 18, 2007.
  19. Luyten, Willem, J. 1922. Note on Some Faint Early Type Stars with Large Proper Motions. Publications of the Astronomical Society of the Pacific. 34:197:54–55. Retrieved September 18, 2007.
  20. Luyten, Willem J. 1922. Additional Note on Faint Early-Type Stars with Large Proper-Motions Publications of the Astronomical Society of the Pacific. 34:198:132. Retrieved September 18, 2007.
  21. Luyten, Willem J. 1922. Third Note on Faint Early Type Stars with Large Proper Motion. Publications of the Astronomical Society of the Pacific. 34:202:356–357. Retrieved September 18, 2007.
  22. 22.0 22.1 22.2 Eddington, A.S. 1924. On the relation between the masses and luminosities of the stars. Monthly Notices of the Royal Astronomical Society. 84:308–332. Retrieved September 18, 2007.
  23. Luyten, W.J. 1950. The search for white dwarfs. Astronomical Journal. 55:1183:86–89. Retrieved September 18, 2007.
  24. 24.0 24.1 24.2 24.3 McCook, George P. and Edward M. Sion. 1999. A Catalog of Spectroscopically Identified White Dwarfs. Astrophysical Journal Supplement. 121:1:1–130. Retrieved September 18, 2007.
  25. 25.0 25.1 Eisenstein, Daniel J., James Liebert, Hugh C. Harris, S. J. Kleinman, Atsuko Nitta, Nicole Silvestri, Scott A. Anderson, J. C. Barentine, Howard J. Brewington, J. Brinkmann, Michael Harvanek, Jurek Krzesiński, Eric H. Neilsen, Jr., Dan Long, Donald P. Schneider, and Stephanie A. Snedden. 2006. A Catalog of Spectroscopically Confirmed White Dwarfs from the Sloan Digital Sky Survey Data Release 4. Astrophysical Journal Supplement. 167:1:40–58. Retrieved September 18, 2007.
  26. Kulic, Mukremin, Carlos Allende Prieto, Warren R. Brown, and D. Koester. 2007. The Lowest Mass White Dwarf. Astrophysical Journal. 660:2:1451–1461. Retrieved September 18, 2007.
  27. 27.0 27.1 Kepler, S.O., S.J. Kleinman, A. Nitta, D. Koester, B.G. Castanheira, O. Giovannini, A.F.M. Costa, and L. Althaus. 2007. White dwarf mass distribution in the SDSS. Monthly Notices of the Royal Astronomical Society. 375:4:1315-1324. Retrieved September 18, 2007.
  28. Shipman, H.L. 1979. Masses and radii of white-dwarf stars. III - Results for 110 hydrogen-rich and 28 helium-rich stars. Astrophysical Journal. 228:240–256. Retrieved September 18, 2007.
  29. Sandin, Fredrik. 2005. Exotic Phases of Matter in Compact Stars. Luleå University of Technology. Retrieved September 18, 2007.
  30. Boss, L. 1910. Preliminary General Catalogue. Washington, D.C.: Carnegie Institution.
  31. Liebert, J., P. A. Young, D. Arnett, J. B. Holberg, K. A. Williams. 2005. The Age and Progenitor Mass of Sirius B. The Astrophysical Journal. 630:1:L69-L72. Retrieved September 18, 2007.
  32. Öpik, E. 1916. The Densities of Visual Binary Stars. Astrophysical Journal. 44:292–302. Retrieved September 18, 2007.
  33. Eddington, A.S. 1927. Stars and Atoms. Oxford, UK: Clarendon Press.
  34. Adams, Walter S. 1925. The Relativity Displacement of the Spectral Lines in the Companion of Sirius. Proceedings of the National Academy of Sciences of the United States of America. 11:7:382–387. Retrieved September 18, 2007.
  35. 35.0 35.1 Fowler, R.H. 1926. On Dense Matter. Monthly Notices of the Royal Astronomical Society. 87:114–122. Retrieved September 18, 2007.
  36. Anderson, Wilhelm. 1929. Über die Grenzdichte der Materie und der Energie, Zeitschrift für Physik. 56:11–12:851–856.
  37. 37.0 37.1 Stoner, Edmund C. 1930. The Equilibrium of Dense Stars. Philosophical Magazine (7th series). 9:944–963.
  38. Chandrasekhar, S. 1931. The Maximum Mass of Ideal White Dwarfs. Astrophysical Journal. 74:81–82. Retrieved September 18, 2007.
  39. 39.0 39.1 39.2 Chandrasekhar, S. 1935. The Highly Collapsed Configurations of a Stellar Mass (second paper). Monthly Notices of the Royal Astronomical Society. 95:207—225. Retrieved September 18, 2007.
  40. 40.0 40.1 40.2 40.3 Estimating Stellar Parameters from Energy Equipartition. ScienceBits. Retrieved September 18, 2007.
  41. The Nobel Prize in Physics 1983. Nobel Foundation. Retrieved September 18, 2007.
  42. 42.0 42.1 Canal, R. and J. Gutierrez. The Possible White Dwarf-Neutron Star Connection. Retrieved September 18, 2007.
  43. 43.0 43.1 43.2 43.3 Hillebrandt, Wolfgang, and Jens C. Niemeyer. 2000. Type IA Supernova Explosion Models. Annual Review of Astronomy and Astrophysics. 38:191–230. Retrieved September 18, 2007.
  44. Chabrier, Gilles and Isabelle Baraffe. 2000. Theory of Low-Mass Stars and Substellar Objects. Annual Review of Astronomy and Astrophysics. 38:337–377. Retrieved September 18, 2007.
  45. Kaler, Jim. The Hertzsprung-Russell (HR) diagram. Retrieved September 18, 2007.
  46. Standards for Astronomical Catalogues, Version 2.0. Retrieved September 18, 2007.
  47. Tohline, Joel E. The Structure, Stability, and Dynamics of Self-Gravitating Systems. Retrieved September 18, 2007.
  48. Hoyle, F. 1947. Note on equilibrium configurations for rotating white dwarfs. Monthly Notices of the Royal Astronomical Society. 107:231–236. Retrieved September 18, 2007.
  49. Ostriker, Jeremiah P. and Peter Bodenheimer. 1968. Rapidly Rotating Stars. II. Massive White Dwarfs. Astrophysical Journal. 151:1089–1098. Retrieved September 18, 2007.
  50. 50.0 50.1 50.2 Sion, E.M., J.L. Greenstein, J.D. Landstreet, J. Liebert, H.L. Shipman, and G.A. Wegner. 1983. A proposed new white dwarf spectral classification system. Astrophysical Journal. 269:253–257. Retrieved September 18, 2007.
  51. 51.0 51.1 Hambly, N.C., S.J. Smartt, and S. Hodgkin. 1997. WD 0346+246: A Very Low Luminosity, Cool Degenerate in Taurus. Astrophysical Journal Letters. 489:L157–L160. Retrieved September 18, 2007.
  52. 52.0 52.1 52.2 52.3 52.4 52.5 52.6 52.7 Murdin, Paul. 2001. Encyclopedia of Astronomy and Astrophysics. Bristol and Philadelphia: Institute of Physics Publishing and London, New York and Tokyo: Nature Publishing Group. ISBN 0333750888. Cite error: Invalid <ref> tag; name "wden" defined multiple times with different content
  53. Heise, J. 1985. X-ray emission from isolated hot white dwarfs. Space Science Reviews. 40:79–90. Retrieved September 18, 2007.
  54. Kawaler, Steven D. White dwarf stars and the Hubble Deep Field. Retrieved September 18, 2007.
  55. 55.0 55.1 Metcalfe, T.S., M. H. Montgomery, and A. Kanaan. 2004. Testing White Dwarf Crystallization Theory with Asteroseismology of the Massive Pulsating DA Star BPM 37093. Astrophysical Journal. 605:2:L133–L136. Retrieved September 18, 2007.
  56. Barrat, J.L., J. P. Hansen, and R. Mochkovitch. 1988. Crystallization of carbon-oxygen mixtures in white dwarfs. Astronomy and Astrophysics. 199:1–2:L15–L18. Retrieved September 18, 2007.
  57. Winget, D.E. 1995. The Status of White Dwarf Asteroseismology and a Glimpse of the Road Ahead. Baltic Astronomy. 4:129–136. Retrieved September 18, 2007.
  58. Whitehouse, David. 2004. Diamond star thrills astronomers. BBC NEWS. Retrieved September 18, 2007.
  59. Press release. Harvard-Smithsonian Center for Astrophysics.
  60. Kanaan, A., A. Nitta, D. E. Winget, S. O. Kepler, M. H. Montgomery, T. S. Metcalfe, et al. Whole Earth Telescope observations of BPM 37093: a seismological test of crystallization theory in white dwarfs. Retrieved September 18, 2007.
  61. Brassard, P. and G. Fontaine. 2005. Asteroseismology of the Crystallized ZZ Ceti Star BPM 37093: A Different View. Astrophysical Journal. 622:1:572–576. Retrieved September 18, 2007.
  62. McCook, G.P. and E.M. Sion. III/235A: A Catalogue of Spectroscopically Identified White Dwarfs. Centre de Données astronomiques de Strasbourg. Retrieved September 18, 2007.
  63. 63.0 63.1 Leggett, S.K., Maria Teresa Ruiz, and P. Bergeron. 1998. The Cool White Dwarf Luminosity Function and the Age of the Galactic Disk. Astrophysical Journal. 497:294–302. Retrieved September 18, 2007.
  64. Gates, Evalyn, Geza Gyuk, Hugh C. Harris, Mark Subbarao, Scott Anderson, S. J. Kleinman, James Liebert, Howard Brewington, J. Brinkmann, Michael Harvanek, Jurek Krzesinski, Don Q. Lamb, Dan Long, Eric H. Neilsen, Jr., Peter R. Newman, Atsuko Nitta, and Stephanie A. Snedden. 2004. Discovery of New Ultracool White Dwarfs in the Sloan Digital Sky Survey. Astrophysical Journal. 612:2:L129–L132. Retrieved September 18, 2007.
  65. Schatzman, Evry. 1945. Théorie du débit d'énergie des naines blanches. Annales d'Astrophysique. 8:143–209. Retrieved September 18, 2007.
  66. 66.0 66.1 66.2 66.3 66.4 66.5 66.6 Koester, D. and G. Chanmugam. 1990. Physics of white dwarf stars. Reports on Progress in Physics. 53:837–915. Retrieved September 18, 2007.
  67. 67.0 67.1 Kawaler, Steven D., I. Novikov, and G. Srinivasan, edited by Georges Meynet and Daniel Schaerer. 1997. Stellar remnants. Berlin, DE: Springer. ISBN 3540615202.
  68. Kuiper, Gerard P. 1941. List of Known White Dwarfs. Publications of the Astronomical Society of the Pacific. 53:314:248–252. Retrieved September 18, 2007.
  69. Luyten, Willem J. 1952. The Spectra and Luminosities of White Dwarfs. Astrophysical Journal. 116:283–290. Retrieved September 18, 2007.
  70. Jesse Leonard Greenstein, Jesse Leaonard. 1960. Stellar atmospheres. vol. 6. Chicago, IL: University of Chicago Press. Retrieved September 18, 2007.
  71. Blackett, P.M.S. 1947. The magnetic field of massive rotating bodies. Nature. 159:658 ff. Retrieved September 18, 2007.
  72. Lovell, Bernard. 1975. Patrick Maynard Stuart Blackett, Baron Blackett, of Chelsea, 18 November 1897-13 July 1974. Biographical Memoirs of Fellows of the Royal Society. 21:1–115. Retrieved September 18, 2007.
  73. Ginzburg, V.L., V. V. Zheleznyakov, and V. V. Zaitsev. 1969. Coherent Mechanisms of Radio Emission and Magnetic Models of Pulsars. Astrophysics and Space Science. 4:464–504. Retrieved September 18, 2007.
  74. Kemp, James C., John B. Swedlund, J.D. Landstreet, and J.R.P. Angel. 1970. Discovery of Circularly Polarized Light from a White Dwarf. Astrophysical Journal. 161:L77–L79. Retrieved September 18, 2007.
  75. Jordan, S., R. Aznar Cuadrado, R. Napiwotzki, H. M. Schmid, and S. K. Solanki. 2007. The fraction of DA white dwarfs with kilo-Gauss magnetic fields. Astronomy and Astrophysics. 462:3:1097–1101. Retrieved September 18, 2007.
  76. Liebert, James, P. Bergeron, and J. B. Holberg. 2003. The True Incidence of Magnetism Among Field White Dwarfs. Astronomical Journal. 125:1:348–353. Retrieved September 18, 2007.
  77. Association Française des Observateurs d'Etoiles Variables. ZZ Ceti variables. Centre de Données astronomiques de Strasbourg. Retrieved September 18, 2007.
  78. Lawrence, George M., Jeremiah P. Ostriker, and James E. Hesser. 1967. Ultrashort-Period Stellar Oscillations. I. Results from White Dwarfs, Old Novae, Central Stars of Planetary Nebulae, 3C 273, and Scorpius XR-1. Astrophysical Journal. 148:3:L161–L163. Retrieved September 18, 2007.
  79. Landolt, Arlo U. 1968. A New Short-Period Blue Variable. Astrophysical Journal. 153:1:151–164. Retrieved September 18, 2007.
  80. Winget, D.E. 1998. Asteroseismology of white dwarf stars. Journal of Physics: Condensed Matter. 10:49:11247–11261. Retrieved September 18, 2007.
  81. 81.0 81.1 Heger, A., C. L. Fryer, S. E. Woosley, N. Langer, and D. H. Hartmann. 2003. How Massive Single Stars End Their Life. Astrophysical Journal. 591:1:288–300. Retrieved September 18, 2007.
  82. 82.0 82.1 Jeffery, Simon. Stars Beyond Maturity. Retrieved September 18, 2007.
  83. Sarna, M.J., E. Ergma, and J. Gerskevits. 2001. Helium core white dwarf evolution—including white dwarf companions to neutron stars. Astronomische Nachrichten. 322:5/6:405–410. Retrieved September 18, 2007.
  84. Benvenuto, O.G., M. A. De Vito. 2005. The formation of helium white dwarfs in close binary systems - II. Monthly Notices of the Royal Astronomical Society. 362:3:891-905. Retrieved September 18, 2007.
  85. Dhillon, Vik. the evolution of low-mass stars. University of Sheffield. Retrieved September 18, 2007.
  86. Dhillon, Vik. the evolution of high-mass stars. University of Sheffield. Retrieved September 18, 2007.
  87. Schaffner-Bielich, Jürgen. 2005. Strange quark matter in stars: a general overview. Journal of Physics G: Nuclear and Particle Physics. 31:6:S651–S657. Retrieved September 18, 2007.
  88. Nomoto, Ken'ichi. 1984. Evolution of 8-10 solar mass stars toward electron capture supernovae. I - Formation of electron-degenerate O + Ne + Mg cores. Astrophysical Journal. 277:791-805. Retrieved September 18, 2007.
  89. Woosley, S.E., A. Heger, and T. A. Weaver. 2002. The evolution and explosion of massive stars. Reviews of Modern Physics. 74:4:1015–1071. Retrieved September 18, 2007.
  90. Werner, K., T. Rauch, M. A. Barstow, and J. W. Kruk. 2004. Chandra and FUSE spectroscopy of the hot bare stellar core H 1504+65. Astronomy and Astrophysics. 421:1169–1183. Retrieved September 18, 2007.
  91. Livio, Mario and James W. Truran. 1994. On the interpretation and implications of nova abundances: an abundance of riches or an overabundance of enrichments. Astrophysical Journal. 425:2:797–801. Retrieved September 18, 2007.
  92. 92.0 92.1 92.2 Adams, Fred C. and Gregory Laughlin. 1997. A dying universe: the long-term fate and evolution of astrophysical objects. Reviews of Modern Physics. 69:2:337–372. Retrieved September 18, 2007.
  93. Comet clash kicks up dusty haze. BBC News. Retrieved September 18, 2007.
  94. Su, K. Y. L., Y.H. Chu, G.H. Rieke, P.J. Huggins, R. Gruendl, R. Napiwotzki, T. Rauch, W.B. Latter, K. Volk. 2007. A Debris Disk around the Central Star of the Helix Nebula?. Astrophysical Journal. 657:1:L41–L45. Retrieved September 18, 2007.
  95. Reach, William T., Marc J. Kuchner, Ted von Hippel, Adam Burrows, Fergal Mullally, Mukremin Kilic, and D. E. Winget. 2005. The Dust Cloud around the White Dwarf G29-38. Astrophysical Journal. 635:2:L161–L164. Retrieved September 18, 2007.
  96. Yoon, S.C.; N. Langer. 2004. Presupernova Evolution of Accreting White Dwarfs with Rotation. Astronomy and Astrophysics. 419:2:623–644. Retrieved September 18, 2007.
  97. Blinnikov, S.I., F.K. Röpke, E.I. Sorokina, M. Gieseler, M. Reinecke, C. Travaglio, W. Hillebrandt, and M. Stritzinger. 2006. Theoretical light curves for deflagration models of type Ia supernova. Astronomy and Astrophysics. 453:1:229–240. Retrieved September 18, 2007.
  98. Imagine the Universe! Cataclysmic Variables. NASA Goddard. Retrieved September 18, 2007.
  99. 99.0 99.1 Introduction to Cataclysmic Variables (CVs). NASA Goddard. Retrieved September 18, 2007.

References
ISBN links support NWE through referral fees

General

  • Kawaler, S.Sd., I. Novikov, G. Srinivasan, Georges Meynet, Daniel Schaerer. 1997. Stellar Remnants: Saas-Fee Advanced Course 25 Lecture Notes 1995 Swiss Society for Astrophysics and Astronomy (Saas-Fee Advanced Courses). Berlin, DE: Springer. ISBN 3540615202.

Physics

  • Shapiro, Stuart L. and Saul A. Teukolsky. 1983. Black holes, white dwarfs, and neutron stars: the physics of compact objects Hoboken, NJ: Wiley. ISBN 0471873179.

Variability

Magnetic field

Frequency

Observational

External links

Credits

New World Encyclopedia writers and editors rewrote and completed the Wikipedia article in accordance with New World Encyclopedia standards. This article abides by terms of the Creative Commons CC-by-sa 3.0 License (CC-by-sa), which may be used and disseminated with proper attribution. Credit is due under the terms of this license that can reference both the New World Encyclopedia contributors and the selfless volunteer contributors of the Wikimedia Foundation. To cite this article click here for a list of acceptable citing formats.The history of earlier contributions by wikipedians is accessible to researchers here:

The history of this article since it was imported to New World Encyclopedia:

Note: Some restrictions may apply to use of individual images which are separately licensed.