ART

A white dwarf, also called a degenerate dwarf, is a stellar core remnant composed mostly of electron-degenerate matter. A white dwarf is very dense: its mass is comparable to that of the Sun, while its volume is comparable to that of Earth. A white dwarf's faint luminosity comes from the emission of stored thermal energy; no fusion takes place in a white dwarf.[1] The nearest known white dwarf is Sirius B, at 8.6 light years, the smaller component of the Sirius binary star. There are currently thought to be eight white dwarfs among the hundred star systems nearest the Sun.[2] The unusual faintness of white dwarfs was first recognized in 1910.[3]:1 The name white dwarf was coined by Willem Luyten in 1922.

White dwarfs are thought to be the final evolutionary state of stars whose mass is not high enough to become a neutron star, which is about 10 solar masses. This includes over 97% of the other stars in the Milky Way.[4]:§1 After the hydrogen-fusing period of a main-sequence star of low or medium mass ends, such a star will expand to a red giant during which it 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 (around 1 billion K), an inert mass of carbon and oxygen will build up at its center. After such a star sheds its outer layers and forms a planetary nebula, it will leave behind a core, which is the remnant white dwarf.[5] Usually, white dwarfs are composed of carbon and oxygen. If the mass of the progenitor is between 8 and 10.5 solar masses (M☉), the core temperature will be sufficient to fuse carbon but not neon, in which case an oxygen–neon–magnesium white dwarf may form.[6] Stars of very low mass will not be able to fuse helium, hence, a helium white dwarf[7][8] may form by mass loss in binary systems.

The material in a white dwarf no longer undergoes fusion reactions, so the star has no source of energy. As a result, it cannot support itself by the heat generated by fusion against gravitational collapse, but is supported only by electron degeneracy pressure, causing it to be extremely dense. The physics of degeneracy yields a maximum mass for a non-rotating white dwarf, the Chandrasekhar limit—approximately 1.44 times M☉—beyond which it cannot be supported by electron 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][5] SN 1006 is thought to be a famous example.

A white dwarf is very hot when it forms, but because it has no source of energy, it will gradually cool as it radiates its energy. This means that its radiation, which initially has a high color temperature, will lessen and redden with time. Over a very long time, a white dwarf will cool and its material will begin to crystallize, starting with the core. The star's low temperature means it will no longer emit significant heat or light, and it will become a cold black dwarf.[5] Because the length of time it takes for a white dwarf to reach this state is calculated to be longer than the current age of the universe (approximately 13.8 billion years),[9] it is thought that no black dwarfs yet exist.[1][4] The oldest white dwarfs still radiate at temperatures of a few thousand kelvins.

Discovery
See also: List of white dwarfs

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 William Herschel on 31 January 1783.[10] In 1910, Henry Norris Russell, Edward Charles Pickering and Williamina Fleming discovered 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:[3]: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.[12]

The white dwarf 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 position measurements to determine that the stars Sirius (α Canis Majoris) and Procyon (α Canis Minoris) were changing their positions periodically. In 1844 he predicted that both stars had unseen companions:[13]

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

In 1917, Adriaan van Maanen discovered Van Maanen's Star, an isolated white dwarf.[16] These three white dwarfs, the first discovered, are the so-called classical white dwarfs.[3]: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][17][18][19][20] the term was later popularized by Arthur Stanley Eddington.[11][21] Despite these suspicions, the first non-classical white dwarf was not definitely identified until the 1930s. 18 white dwarfs had been discovered by 1939.[3]:3 Luyten and others continued to search for white dwarfs in the 1940s. By 1950, over a hundred were known,[22] and by 1999, over 2,000 were known.[23] Since then the Sloan Digital Sky Survey has found over 9,000 white dwarfs, mostly new.[24]

Composition and structure
Hertzsprung–Russell diagram
Spectral type
Brown dwarfs
White dwarfs
Red dwarfs
Subdwarfs
Main sequence
("dwarfs")
Subgiants
Giants
Bright giants
Supergiants
Hypergiants
absolute
magni-
tude
(MV)

Although white dwarfs are known with estimated masses as low as 0.17 M☉[25] and as high as 1.33 M☉,[26] the mass distribution is strongly peaked at 0.6 M☉, and the majority lie between 0.5 and 0.7 M☉.[26] The estimated radii of observed white dwarfs are typically 0.8–2% the radius of the Sun;[27] this is comparable to the Earth's radius of approximately 0.9% solar radius. 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 g/cm3, or 1 tonne per cubic centimetre.[1] A typical white dwarf has a density of between 104 and 107 g/cm3. White dwarfs are composed of one of the densest forms of matter known, surpassed only by other compact stars such as neutron stars, quark stars (hypothetically),[28] and black holes.

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 or 40 Eridani B, it is possible to estimate its mass from observations of the binary orbit. This was done for Sirius B by 1910,[29] yielding a mass estimate of 0.94 M☉, which compares well with a more modern estimate of 1.00 M☉.[30] Since hotter bodies radiate more energy than colder ones, a star's surface brightness can be estimated from its effective surface temperature, and that from its spectrum. If the star's distance is known, its absolute luminosity can also be estimated. From the absolute luminosity and distance, the star's surface area and its radius can be calculated. 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. 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".[31] As Arthur Stanley Eddington put it later in 1927:[32]: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.[21] This was confirmed when Adams measured this redshift in 1925.[33]

Material Density in kg/m3 Notes
Water (fresh) 1,000 At STP
Osmium 22,610 Near room temperature
The core of the Sun c. 150,000
White dwarf 1 × 109[1]
Atomic nuclei 2.3 × 1017[34] Does not depend strongly on size of nucleus
Neutron star core 8.4 × 1016 – 1 × 1018
Black hole 2 × 1030[35] Critical density of an Earth-mass black hole

Such densities are possible because white dwarf material is not composed of atoms joined by chemical bonds, but rather consists of a plasma of unbound nuclei and electrons. There is therefore no obstacle to placing nuclei closer than normally allowed by electron orbitals limited by normal matter.[21] Eddington wondered what would happen when this plasma cooled and the energy to keep the atoms ionized was no longer sufficient.[36] This paradox was resolved by R. H. Fowler in 1926 by an application of the newly devised quantum mechanics. Since electrons obey the Pauli exclusion principle, no two electrons can occupy the same state, and they must obey Fermi–Dirac statistics, also introduced in 1926 to determine the statistical distribution of particles which satisfy the Pauli exclusion principle.[37] At zero temperature, therefore, electrons can not all occupy the lowest-energy, or ground, state; some of them would have to occupy higher-energy states, forming a band of lowest-available energy states, the Fermi sea. This state of the electrons, called degenerate, meant that a white dwarf could cool to zero temperature and still possess high energy.[36][38]

Compression of a white dwarf will increase the number of electrons in a given volume. Applying the Pauli exclusion principle, this will increase the kinetic energy of the electrons, thereby increasing the pressure.[36][39] This electron degeneracy pressure supports a white dwarf against gravitational collapse. The pressure 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 much greater than that of a low-mass white dwarf and that the radius of a white dwarf decreases as its mass increases.[1]

The existence of a limiting mass that no white dwarf can exceed without collapsing to a neutron star is another consequence of being supported by electron degeneracy pressure. Such limiting masses were calculated for cases of an idealized, constant density star in 1929 by Wilhelm Anderson[40] and in 1930 by Edmund C. Stoner.[41] This value was corrected by considering hydrostatic equilibrium for the density profile, and the presently known value of the limit was first published in 1931 by Subrahmanyan Chandrasekhar in his paper "The Maximum Mass of Ideal White Dwarfs".[42] For a non-rotating white dwarf, it is equal to approximately 5.7M☉/μe2, where μe is the average molecular weight per electron of the star.[43]: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,[38] leading to the commonly quoted value of 1.4 M☉. (Near the beginning of the 20th century, there was reason to believe that stars were composed chiefly of heavy elements,[41]: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 M☉.) Together with William Alfred Fowler, Chandrasekhar received the Nobel prize for this and other work in 1983.[44] 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 called a neutron star.[45] 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 may be destroyed, before it reaches the limiting mass.[46]

New research indicates that many white dwarfs—at least in certain types of galaxies — may not approach that limit by way of accretion. It has been postulated that at least some of the white dwarfs that become supernovae attain the necessary mass by colliding with one another. It may be that in elliptical galaxies such collisions are the major source of supernovae. This hypothesis is based on the fact that the X-rays produced by those galaxies are 30 to 50 times less than what is expected to be produced by type Ia supernovas of that galaxy as matter accretes on the white dwarf from its encircling companion. It has been concluded that no more than 5 percent of the supernovae in such galaxies could be created by the process of accretion onto white dwarfs. The significance of this finding is that there could be two types of supernovae, which could mean that the Chandrasekhar limit might not always apply in determining when a white dwarf goes supernova, given that two colliding white dwarfs could have a range of masses. This in turn would confuse efforts to use exploding white dwarfs as standard candles in determining distances.[47]

White dwarfs have low luminosity and therefore occupy a strip at the bottom of the Hertzsprung–Russell diagram, a graph of stellar luminosity versus color or temperature. 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,[48] or the even lower-temperature brown dwarfs.[49]

Mass–radius relationship and mass limit

The relationship between the mass and radius of white dwarfs can be derived 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 G is the gravitational constant, M is the mass of the white dwarf, and R is its radius.
E g ≈ − G M R . {\displaystyle E_{g}\approx {\frac {-GM}{R}}.} {\displaystyle E_{g}\approx {\frac {-GM}{R}}.} \)

The kinetic energy of the unit mass, Ek, will primarily come from the motion of electrons, so it will be approximately Np2 ∕ 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, Δp, given by the uncertainty principle, which says that Δp Δx is on the order of the reduced Planck constant, ħ. Δx 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, where M is the star's mass and its volume is on the order of R3, n will be on the order of NM ∕ R3.[38]

Solving for the kinetic energy per unit mass, Ek, we find that
\( E_{k}\approx {\frac {N(\Delta p)^{2}}{2m}}\approx {\frac {N\hbar ^{2}n^{2/3}}{2m}}\approx {\frac {M^{2/3}N^{5/3}\hbar ^{2}}{2mR^{2}}}. \)

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:
\( |E_{g}|\approx {\frac {GM}{R}}=E_{k}\approx {\frac {M^{2/3}N^{5/3}\hbar ^{2}}{2mR^{2}}}. \)

Solving this for the radius, R, gives[38]
\( R\approx {\frac {N^{5/3}\hbar ^{2}}{2mGM^{1/3}}}.

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:
\( R\sim M^{-1/3}

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 pc for the kinetic energy. With this substitution, we find

\( E_{k\ {\rm {relativistic}}}\approx {\frac {M^{1/3}N^{4/3}\hbar c}{R}}. \)

If we equate this to the magnitude of Eg, we find that R drops out and the mass, M, is forced to be[38]
\( M_{\rm {limit}}\approx N^{2}\left({\frac {\hbar c}{G}}\right)^{3/2}. \)
Radius–mass relations for a model white dwarf. Mlimit is denoted as MCh

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 a limiting mass of Mlimit. Therefore, no white dwarf can be heavier than the limiting mass Mlimit, or 1.4 M☉.

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.[43]: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 on the right shows the result of such a computation. It shows how radius varies with mass for non-relativistic (blue curve) and relativistic (green 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.[43][50]

These computations all assume that the white dwarf is non-rotating. 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.[51] For a uniformly rotating white dwarf, the limiting mass increases only slightly. If the star is allowed to rotate nonuniformly, and viscosity is neglected, then, as was pointed out by Fred Hoyle in 1947,[52] 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 will be dynamically stable.[53]

Radiation and cooling

The degenerate matter that makes up the bulk of a white dwarf has a very low opacity, because any absorption of a photon requires that an electron must transition to a higher empty state, which may not be possible as the energy of the photon may not be a match for the possible quantum states available to that electron, hence radiative heat transfer within a white dwarf is low; it does, however, have a high thermal conductivity. As a result, the interior of the white dwarf maintains a uniform temperature, approximately 107 K. An outer shell of non-degenerate matter cools from approximately 107 K to 104 K. This matter radiates roughly as a black body. A white dwarf remains visible for a long time, as its tenuous outer atmosphere of normal matter begins to radiate at about 107 K, upon formation, while its greater interior mass is at 107  K but cannot radiate through its normal matter shell.[54]

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 an M-type red dwarf.[55] White dwarf effective surface temperatures extend from over 150,000 K[23] to barely under 4,000 K.[56][57] 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 ​110,000 that of the Sun's.[57] Hot white dwarfs, with surface temperatures in excess of 30,000 K, 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.[58]

White dwarfs also radiate neutrinos through the Urca process.[59]
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.

As was explained by Leon Mestel in 1952, unless the white dwarf accretes matter from a companion star or other source, its radiation comes from its stored heat, which is not replenished.[60][61]:§2.1 White dwarfs have an extremely small surface area to radiate this heat from, so they cool gradually, remaining hot for a long time.[5] 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. The rate of cooling has been estimated for a carbon white dwarf of 0.59 M☉ with a hydrogen atmosphere. After initially taking approximately 1.5 billion years to cool to a surface temperature of 7,140 K, cooling approximately 500 more kelvins to 6,590 K takes around 0.3 billion years, but the next two steps of around 500 kelvins (to 6,030 K and 5,550 K) take first 0.4 and then 1.1 billion years.[62]:Table 2

Most observed white dwarfs have relatively high surface temperatures, between 8,000 K and 40,000 K.[24][63] 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.[64] This trend stops when we reach extremely cool white dwarfs; few white dwarfs are observed with surface temperatures below 4,000 K,[65] and one of the coolest so far observed, WD 0346+246, has a surface temperature of approximately 3,900 K.[56] The reason for this is that the Universe's age is finite;[66][67] there has not been enough time for white dwarfs to cool 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 our Galactic disk found in this way is 8 billion years.[64] A white dwarf will eventually, in many trillions of years, cool and become a non-radiating black dwarf in approximate thermal equilibrium with its surroundings and with the cosmic background radiation. No black dwarfs are thought to exist yet.[1]
The white dwarf cooling sequence seen by ESA's Gaia mission

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 its center.[68] The crystal structure is thought to be a body-centered cubic lattice.[4][69] In 1995 it was suggested that asteroseismological observations of pulsating white dwarfs yielded a potential test of the crystallization theory,[70] and in 2004, observations were made that suggested approximately 90% of the mass of BPM 37093 had crystallized.[68][71][72] Other work gives a crystallized mass fraction of between 32% and 82%.[73] As a white dwarf core undergoes crystallization into a solid phase, latent heat is released which provides a source of thermal energy that delays its cooling.[74] This effect was first confirmed in 2019 after the identification of a pile up in the cooling sequence of more than 15,000 white dwarfs observed with the Gaia satellite.[75]

Low-mass helium white dwarfs (mass < 0.20 M☉), often referred to as "extremely low-mass white dwarfs, ELM WDs" are formed in binary systems. As a result of their hydrogen-rich envelopes, residual hydrogen burning via the CNO cycle may keep these white dwarfs hot on a long timescale. In addition, they remain in a bloated proto-white dwarf stage for up to 2 Gyr before they reach the cooling track.[76]
Atmosphere and spectra
Artist's impression of the WD J0914+1914 system.[77]

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 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 below and the lighter above.[78][79]:§§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 ​1⁄100 of the star's total mass, which, if the atmosphere is hydrogen-dominated, is overlain by a hydrogen-rich layer with mass approximately ​1⁄10,000 of the stars total mass.[57][80]:§§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 at almost the same temperature (isothermal), and it is also hot: a white dwarf with surface temperature between 8,000 K and 16,000 K will have a core temperature between approximately 5,000,000 K and 20,000,000 K. The white dwarf is kept from cooling very quickly only by its outer layers' opacity to radiation.[57]

White dwarf spectral types[23]
Primary and secondary features
A H lines present
B He I lines
C Continuous spectrum; no lines
O He II lines, accompanied by He I or H lines
Z Metal 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

The first attempt to classify white dwarf spectra appears to have been by G. P. Kuiper in 1941,[55][81] and various classification schemes have been proposed and used since then.[82][83] The system currently in use was introduced by Edward M. Sion, Jesse L. Greenstein and their 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 adjacent table), and a temperature index number, computed by dividing 50,400 K 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 dominated 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.[23][55]

White dwarfs whose primary spectral classification is DA have hydrogen-dominated atmospheres. They make up the majority, approximately 80%, of all observed white dwarfs.[57] The next class in number is of DBs, approximately 16%.[84] The hot, above 15,000 K, DQ class (roughly 0.1%) have carbon-dominated atmospheres.[85] Those classified as DB, DC, DO, DZ, and cool 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 100,000 K to 45,000 K, the spectrum will be classified DO, dominated by singly ionized helium. From 30,000 K to 12,000 K, the spectrum will be DB, showing neutral helium lines, and below about 12,000 K, the spectrum will be featureless and classified DC.[80]:§2.4[57]

Molecular hydrogen (H2) has been detected in spectra of the atmospheres of some white dwarfs.[86]
Metal-rich white dwarfs

Around 25–33% of white dwarfs have metal lines in their spectra, which is notable because any heavy elements in a white dwarf should sink into the star's interior in just a small fraction of the star's lifetime.[87] The prevailing explanation for metal-rich white dwarfs is that they have recently accreted rocky planetesimals.[87] The bulk composition of the accreted object can be measured from the strengths of the metal lines. For example, a 2015 study of the white dwarf Ton 345 concluded that its metal abundances were consistent with those of a differentiated, rocky planet whose mantle had been eroded by the host star's wind during its asymptotic giant branch phase.[88]
Magnetic field

Magnetic fields in white dwarfs with a strength at the surface of c. 1 million gauss (100 teslas) 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.[89] This putative law, sometimes called the Blackett effect, was never generally accepted, and by the 1950s even Blackett felt it had been refuted.[90]:39–43 In the 1960s, it was proposed that white dwarfs might have magnetic fields due to conservation of total surface magnetic flux that existed in its progenitor star phase.[91] A surface magnetic field of c. 100 gauss (0.01 T) in the progenitor star would thus become a surface magnetic field of c. 100·1002 = 1 million gauss (100 T) once the star's radius had shrunk by a factor of 100.[79]:§8[92]:484 The first magnetic white dwarf to be discovered was GJ 742 (also known as Grw +70 8247) which was identified by James Kemp, John Swedlund, John Landstreet and Roger Angel in 1970 to host a magnetic field by its emission of circularly polarized light.[93] It is thought to have a surface field of approximately 300 million gauss (30 kT).[79]:§8

Since 1970 magnetic fields have been discovered in well over 200 white dwarfs, ranging from 2 × 103 to 109 gauss (0.2 T to 100 kT).[94] The large number of presently known magnetic white dwarfs is due to the fact that most white dwarfs are identified by low-resolution spectroscopy, which is able to reveal the presence of a magnetic field of 1 megagauss or more. Thus the basic identification process also sometimes results in discovery of magnetic fields.[95] It has been estimated that at least 10% of white dwarfs have fields in excess of 1 million gauss (100 T).[96][97]

The highly magnetized white dwarf in the binary system AR Scorpii was identified in 2016 as the first pulsar in which the compact object is a white dwarf instead of a neutron star.[98]
Chemical bonds

The magnetic fields in a white dwarf may allow for the existence of a new type of chemical bond, perpendicular paramagnetic bonding, in addition to ionic and covalent bonds, resulting in what has been initially described as "magnetized matter" in research published in 2012.[99]
Variability
Main article: Pulsating white dwarf
See also: Cataclysmic variables

Types of pulsating white dwarf[100][101]:§§1.1, 1.2
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
GW Vir (GCVS: ZZO) Atmosphere mostly C, He and O; may be divided into DOV and PNNV stars

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.[79]:§7.1.1[102] The first variable white dwarf found was HL Tau 76; in 1965 and 1966, and was observed to vary with a period of approximately 12.5 minutes.[103] The reason for this period being longer than predicted is that the variability of HL Tau 76, like that of the other pulsating variable white dwarfs known, arises from non-radial gravity wave pulsations.[79]:§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;[79]:891, 895 DBV, or V777 Her, stars, with helium-dominated atmospheres and the spectral type DB;[57]:3525 and GW Vir stars, sometimes subdivided into DOV and PNNV stars, with atmospheres dominated by helium, carbon, and oxygen.[101][104] GW Vir stars are not, strictly speaking, white dwarfs, but are stars which are in a position on the Hertzsprung-Russell diagram between the asymptotic giant branch and the white dwarf region. They may be called pre-white dwarfs.[101][105] These 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.[106]
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 M☉.[4][107] The composition of the white dwarf produced will depend on the initial mass of the star. Current galactic models suggest the Milky Way galaxy currently contains about ten billion white dwarfs.[108]
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 in its core. It is thought that, over a lifespan that considerably exceeds the age of the Universe (c. 13.8 billion years),[9] such a star will eventually burn all its hydrogen, for a while becoming a blue dwarf, and end its evolution as a helium white dwarf composed chiefly of helium-4 nuclei.[109] Due to the very long time this process takes, it is not thought to be the origin of the observed helium white dwarfs. Rather, they are thought to be the product of mass loss in binary systems[5][7][8][110][111][112] or mass loss due to a large planetary companion.[113][114]
Stars with low to medium mass

If the mass of a main-sequence star is between 0.5 and 8 M☉ like our sun, 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.[110][115][116]
Stars with medium to high mass

If a star is massive enough, 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, because the mass of its central, non-fusing core, initially 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.[107][117] Some main-sequence stars, of perhaps 8 to 10 M☉, 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.[118][119] Although a few white dwarfs have been identified which may be of this type, most evidence for the existence of such 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.[6][120][121]
Type Iax supernova

Type Iax supernova, that involve helium accretion by a white dwarf, have been proposed to be a channel for transformation of this type of stellar remnant. In this scenario, the carbon detonation produced in a Type Ia supernova is too weak to destroy the white dwarf, expelling just a small part of its mass as ejecta, but produces an asymmetric explosion that kicks the star, often known as a zombie star, to high speeds of a hypervelocity star. The matter processed in the failed detonation is re-accreted by the white dwarf with the heaviest elements such as iron falling to its core where it accumulates.[122] These iron-core white dwarfs would be smaller than the carbon–oxygen kind of similar mass and would cool and crystallize faster than those.[123]
Fate
File:White Dwarf Ages.ogvPlay media
Artist's concept of white dwarf aging

A white dwarf is stable once formed and will continue to cool almost indefinitely, eventually to become a black dwarf. 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.[124]:§IIIA White dwarfs should generally survive galactic dispersion, 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.[124]:§§IIIC, IV The subsequent lifetime of white dwarfs is thought to be on the order of the hypothetical lifetime of the proton, known to be at least 1034–1035 years. Some grand unified theories predict a proton lifetime between 1030 and 1036 years. If these theories are not valid, the proton might still decay by complicated nuclear reactions or through quantum gravitational processes involving virtual black holes; 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 enough mass to become a nondegenerate lump of matter, and finally disappears completely.[124]:§IV

A white dwarf can also be cannibalized or evaporated by a companion star, causing the white dwarf to lose so much mass that it becomes a planetary mass object. The resultant object, orbiting the former companion, now host star, could be a helium planet or diamond planet.[125][126]
Debris disks and planets
Artist's impression of debris around a white dwarf[127]
Comet falling into white dwarf (artist's impression)[128]

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.[129][130] Similarly, observations made in 2004 indicated the presence of a dust cloud around the young (estimated to have formed from its AGB progenitor about 500 million years ago) white dwarf G29-38, which may have been created by tidal disruption of a comet passing close to the white dwarf.[131] Some estimations based on the metal content of the atmospheres of the white dwarfs consider that at least 15% of them may be orbited by planets and/or asteroids, or at least their debris.[132] Another suggested idea is that white dwarfs could be orbited by the stripped cores of rocky planets, that would have survived the red giant phase of their star but losing their outer layers and, given those planetary remnants would likely be made of metals, to attempt to detect them looking for the signatures of their interaction with the white dwarf's magnetic field.[133] Other suggested ideas of how white dwarfs are polluted with dust involve the scattering of asteroids by planets[134][135][136] or via planet-planet scattering.[137] Liberation of exomoons from their host planet could cause white dwarf pollution with dust. Either the liberation could cause asteroids to be scattered towards the white dwarf or the exomoon could be scattered into the Roche-Radius of the white dwarf.[138] The mechanism behind the pollution of white dwarfs in binaries was also explored as these systems are more likely to lack a major planet, but this idea cannot explain the presence of dust around single white dwarfs.[139] While old white dwarfs show evidence of dust accretion, white dwarfs older than ~1 billion years or >7000 K with dusty infrared excess were not detected[140] until the discovery of LSPM J0207+3331 in 2018, which has a cooling age of ~3 billion years. The white dwarf shows two dusty components that are being explained with two rings with different temperatures.[141]
Exoplanet orbits WD 1856+534
File:NASA-ExoplanetOrbitingWhiteDwarfStarWD1856+534.webmPlay media
(NASA; video; 2:10)

There is a planet in the white dwarf–pulsar binary system PSR B1620-26.

There are two circumbinary planets around the white dwarf–red dwarf binary NN Serpentis.

The metal-rich white dwarf WD 1145+017 is the first white dwarf observed with a disintegrating minor planet which transits the star.[142][143] The disintegration of the planetesimal generates a debris cloud which passes in front of the star every 4.5 hours, causing a 5-minute-long fade in the star's optical brightness.[143] The depth of the transit is highly variable.[143]

The white dwarf WD 0145+234 shows brightening in the mid-infrared, seen in NEOWISE data. The brightening is not seen before 2018. The brightening is interpreted as the tidal disruption of an exoasteroid. The first time such an event has been observed.[144]

WD 0806-661 has a Y-dwarf that orbits the white dwarf in a wide orbit with a projected distance of 2500 astronomical units. Considering the low mass and the wide orbit of this object, WD 0806-661 B can be interpreted as either a sub-brown dwarf or a directly imaged exoplanet.

WD J0914+1914 is the first single white dwarf star found to have a giant planet orbiting it. The giant planet is being evaporated by the strong ultraviolet radiation of the hot white dwarf. Part of the evaporated material is being accreted in a gaseous disk around the white dwarf. The weak hydrogen line as well as other lines in the spectrum of the white dwarf revealed the presence of the giant planet.[145]

In September 2020, astronomers reported the discovery, for the first time, of a very massive Jupiter-sized planet, named WD 1856 b, closely orbiting, every 36 hours, a white dwarf, named WD 1856+534.[146][147][148]
Habitability

It has been proposed that white dwarfs with surface temperatures of less than 10,000 Kelvins could harbor a habitable zone at a distance of c. 0.005 to 0.02 AU that would last upwards of 3 billion years. This is so close that any habitable planets would be tidally locked. The goal is to search for transits of hypothetical Earth-like planets that could have migrated inward and/or formed there. As a white dwarf has a size similar to that of a planet, these kinds of transits would produce strong eclipses.[149] Newer research casts some doubts on this idea, given that the close orbits of those hypothetical planets around their parent stars would subject them to strong tidal forces that could render them uninhabitable by triggering a greenhouse effect.[150] Another suggested constraint to this idea is the origin of those planets. Leaving aside formation from the accretion disk surrounding the white dwarf, there are two ways a planet could end in a close orbit around stars of this kind: by surviving being engulfed by the star during its red giant phase, and then spiralling inward, or inward migration after the white dwarf has formed. The former case is implausible for low-mass bodies, as they are unlikely to survive being absorbed by their stars. In the latter case, the planets would have to expel so much orbital energy as heat, through tidal interactions with the white dwarf, that they would likely end as uninhabitable embers.[151]
Binary stars and novae
The merger process of two co-orbiting white dwarfs produces gravitational waves

If a white dwarf is in a binary star system and is accreting matter from its companion, a variety of phenomena may occur, including novae and Type Ia supernovae. It may also be a super-soft x-ray source if it is able to take material from its companion fast enough to sustain fusion on its surface.[152] On the other hand, phenomena in binary systems as tidal interaction and star-disc interaction, moderated by magnetic fields or not, act on the rotation of accreting white dwarfs. In fact, the fastest-spinning, securely known white dwarfs, are members of binary systems (being the white dwarf in CTCV J2056-3014 the fastest one). [153] A close binary system of two white dwarfs can radiate energy in the form of gravitational waves, causing their mutual orbit to steadily shrink until the stars merge.[154][155]
Type Ia supernovae
Main article: Type Ia supernova

The mass of an isolated, nonrotating white dwarf cannot exceed the Chandrasekhar limit of ~1.4 M☉. This limit may increase if the white dwarf is rotating rapidly and nonuniformly.[156] White dwarfs in binary systems can accrete material from a companion star, increasing both their mass and their density. As their mass approaches the Chandrasekhar limit, this could theoretically lead to either the explosive ignition of fusion in the white dwarf or its collapse into a neutron star.[45]

Accretion provides the currently favored mechanism called the single-degenerate model for Type Ia supernovae. In this model, a carbon–oxygen white dwarf accretes mass and compresses its core by pulling mass from a companion star.[46]:14 It is believed that compressional heating of the core leads to ignition of carbon fusion as the mass approaches the Chandrasekhar limit.[46] 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 a Type Ia supernova explosion that obliterates the star.[1][46][157] In another possible mechanism for Type Ia supernovae, the double-degenerate model, two carbon–oxygen white dwarfs in a binary system merge, creating an object with mass greater than the Chandrasekhar limit in which carbon fusion is then ignited.[46]:14

Observations have failed to note signs of accretion leading up to Type Ia supernovae, and this is now thought to be because the star is first loaded up to above the Chandrasekhar limit while also being spun up to a very high rate by the same process. Once the accretion stops the star gradually slows until the spin is no longer enough to prevent the explosion.[158]

The historical bright SN 1006 is thought to have been a type Ia supernova from a white dwarf, possibly the merger of two white dwarfs.[159] Tycho's Supernova of 1572 was also a type Ia supernova, and its remnant has been detected.[160]
Post-common envelope binary
Main article: Post common envelope binary

A post-common envelope binary (PCEB) is a binary consisting out of a white dwarf and a closely tidally locked red dwarf (in other cases this might be a brown dwarf instead of a red dwarf). These binaries form when the red dwarf is engulfed in the red giant phase and as the red dwarf orbits inside the common envelope it is slowed down in the more dense environment. This slowed orbital speed is compensated with a decrease of the orbital distance between the red dwarf and the core of the red giant. The red dwarf spirals inwards towards the core and might merge with the core. If this does not happen and instead the common envelope is ejected, then the binary ends up in a close orbit, consisting of a white dwarf and a red dwarf. This type of binary is called post-common envelope binary. The evolution of the PCEB continues as the binary orbit closer and closer due to magnetic braking and by releasing gravitational waves. The binary might evolve at some point into a cataclysmic variable and therefore post-common envelope binaries are sometimes called pre-cataclysmic variables.
Cataclysmic variables
Main article: Cataclysmic variable star

Before accretion of material pushes a white dwarf close to the Chandrasekhar limit, accreted hydrogen-rich material on the surface may ignite in a less destructive type of thermonuclear explosion powered by hydrogen fusion. These surface explosions can be repeated as long as the white dwarf's core remains intact. 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 the classical novae. These are thought to be caused by the release of gravitational potential energy when part of the accretion disc collapses onto the star, rather than through a release of energy due to fusion. 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, including polars and intermediate polars, both of which feature highly magnetic white dwarfs.[1][46][161][162] Both fusion- and accretion-powered cataclysmic variables have been observed to be X-ray sources.[162]
Other non-pre-supernova binaries

Other non-pro-supernova binaries include binaries that consist of a main sequence star (or giant) and a white dwarf. The binary Sirius AB is probably the most famous example. White dwarfs can also exist as binaries or multiple star systems that only consist of white dwarfs. An example of a resolved triple white dwarf system is WD J1953-1019, discovered with Gaia DR2 data. One interesting field is the study of remnant planetary systems around white dwarfs. While stars are bright and often outshine the exoplanets and brown dwarfs that orbit them, the white dwarfs are faint. This allows astronomers to study these brown dwarfs or exoplanets in more detail. The sub-brown dwarf around the white dwarf WD 0806−661 is one such example.
Nearest

White Dwarfs within 25 Light Years[163]
Identifier WD Number Distance (ly) Type Absolute
magnitude
Mass
(M)
Luminosity
(L)
Age (Gyr) Objects in system
Sirius B 0642–166 8.66 DA 11.18 0.98 0.0295 0.10 2
Procyon B 0736+053 11.46 DQZ 13.20 0.63 0.00049 1.37 2
Van Maanen 2 0046+051 14.07 DZ 14.09 0.68 0.00017 3.30 1
LP 145-141 1142–645 15.12 DQ 12.77 0.61 0.00054 1.29 1
40 Eridani B 0413-077 16.39 DA 11.27 0.59 0.0141 0.12 3
Stein 2051 B 0426+588 17.99 DC 13.43 0.69 0.00030 2.02 2
G 240-72 1748+708 20.26 DQ 15.23 0.81 0.000085 5.69 1
Gliese 223.2 0552–041 21.01 DZ 15.29 0.82 0.000062 7.89 1
Gliese 3991 B[164] 1708+437 24.23 D?? >15 0.5 <0.000086 >6 2

See also

Black dwarf – Cold stellar remnant
Chandrasekhar's white dwarf equation
Degenerate matter – Collection of free, non-interacting particles with a pressure and other physical characteristics determined by quantum mechanical effects
List of white dwarfs – Wikipedia list article
Neutron star – Collapsed core of a massive star
PG 1159 star
Planetary nebula – Type of emission nebula
Robust associations of massive baryonic objects
Stellar classification – Classification of stars based on their spectral characteristics
Timeline of white dwarfs, neutron stars, and supernovae – Chronological list of developments in knowledge and records

References

Johnson, J. (2007). "Extreme Stars: White Dwarfs & Neutron Stars". Lecture notes, Astronomy 162. Ohio State University. Archived from the original on 31 March 2012. Retrieved 17 October 2011.
Henry, T. J. (1 January 2009). "The One Hundred Nearest Star Systems". Research Consortium on Nearby Stars. Archived from the original on 12 November 2007. Retrieved 21 July 2010.
White Dwarfs, E. Schatzman, Amsterdam: North-Holland, 1958.
Fontaine, G.; Brassard, P.; Bergeron, P. (2001). "The Potential of White Dwarf Cosmochronology". Publications of the Astronomical Society of the Pacific. 113 (782): 409–435. Bibcode:2001PASP..113..409F. doi:10.1086/319535.
Richmond, M. "Late stages of evolution for low-mass stars". Lecture notes, Physics 230. Rochester Institute of Technology. Archived from the original on 4 September 2017. Retrieved 3 May 2007.
Werner, K.; Hammer, N. J.; Nagel, T.; Rauch, T.; Dreizler, S. (2005). "On Possible Oxygen/Neon White Dwarfs: H1504+65 and the White Dwarf Donors in Ultracompact X-ray Binaries". 14th European Workshop on White Dwarfs. 334: 165. arXiv:astro-ph/0410690. Bibcode:2005ASPC..334..165W.
Liebert, J.; Bergeron, P.; Eisenstein, D.; Harris, H. C.; Kleinman, S. J.; Nitta, A.; Krzesinski, J. (2004). "A Helium White Dwarf of Extremely Low Mass". The Astrophysical Journal. 606 (2): L147. arXiv:astro-ph/0404291. Bibcode:2004ApJ...606L.147L. doi:10.1086/421462. S2CID 118894713.
"Cosmic weight loss: The lowest mass white dwarf" (Press release). Harvard-Smithsonian Center for Astrophysics. 17 April 2007. Archived from the original on 22 April 2007. Retrieved 20 April 2007.
Spergel, D. N.; Bean, R.; Doré, O.; Nolta, M. R.; Bennett, C. L.; Dunkley, J.; Hinshaw, G.; Jarosik, N.; et al. (2007). "Wilkinson Microwave Anisotropy Probe (WMAP) Three Year Results: Implications for Cosmology". The Astrophysical Journal Supplement Series. 170 (2): 377–408. arXiv:astro-ph/0603449. Bibcode:2007ApJS..170..377S. doi:10.1086/513700. S2CID 1386346.
Herschel, W. (1785). "Catalogue of Double Stars. By William Herschel, Esq. F. R. S". Philosophical Transactions of the Royal Society of London. 75: 40–126. Bibcode:1785RSPT...75...40H. doi:10.1098/rstl.1785.0006. JSTOR 106749. S2CID 186209747.
Holberg, J. B. (2005). How Degenerate Stars Came to be Known as White Dwarfs. American Astronomical Society Meeting 207. 207. p. 1503. Bibcode:2005AAS...20720501H.
Adams, W. S. (1914). "An A-Type Star of Very Low Luminosity". Publications of the Astronomical Society of the Pacific. 26 (155): 198. Bibcode:1914PASP...26..198A. doi:10.1086/122337.
Bessel, F. W. (1844). "On the variations of the proper motions of Procyon and Sirius". Monthly Notices of the Royal Astronomical Society. 6 (11): 136–141. Bibcode:1844MNRAS...6R.136B. doi:10.1093/mnras/6.11.136a.
Flammarion, Camille (1877). "The Companion of Sirius". Astronomical Register. 15: 186. Bibcode:1877AReg...15..186F.
Adams, W. S. (1915). "The Spectrum of the Companion of Sirius". Publications of the Astronomical Society of the Pacific. 27 (161): 236. Bibcode:1915PASP...27..236A. doi:10.1086/122440.
Van Maanen, A. (1917). "Two Faint Stars with Large Proper Motion". Publications of the Astronomical Society of the Pacific. 29 (172): 258. Bibcode:1917PASP...29..258V. doi:10.1086/122654.
Luyten, W. 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. Bibcode:1922PASP...34..156L. doi:10.1086/123176.
Luyten, W. J. (1922). "Note on Some Faint Early Type Stars with Large Proper Motions". Publications of the Astronomical Society of the Pacific. 34 (197): 54. Bibcode:1922PASP...34...54L. doi:10.1086/123146.
Luyten, W. J. (1922). "Additional Note on Faint Early-Type Stars with Large Proper-Motions". Publications of the Astronomical Society of the Pacific. 34 (198): 132. Bibcode:1922PASP...34..132L. doi:10.1086/123168.
Aitken, R. G. (1922). "Comet c 1922 (Baade)". Publications of the Astronomical Society of the Pacific. 34 (202): 353. Bibcode:1922PASP...34..353A. doi:10.1086/123244.
Eddington, A. S. (1924). "On the relation between the masses and luminosities of the stars". Monthly Notices of the Royal Astronomical Society. 84 (5): 308–333. Bibcode:1924MNRAS..84..308E. doi:10.1093/mnras/84.5.308.
Luyten, W. J. (1950). "The search for white dwarfs". The Astrophysical Journal. 55: 86. Bibcode:1950AJ.....55...86L. doi:10.1086/106358.
McCook, George P.; Sion, Edward M. (1999). "A Catalog of Spectroscopically Identified White Dwarfs". The Astrophysical Journal Supplement Series. 121 (1): 1–130. Bibcode:1999ApJS..121....1M. doi:10.1086/313186.
Eisenstein, Daniel J.; Liebert, James; Harris, Hugh C.; Kleinman, S. J.; Nitta, Atsuko; Silvestri, Nicole; Anderson, Scott A.; Barentine, J. C.; Brewington, Howard J.; Brinkmann, J.; Harvanek, Michael; Krzesiński, Jurek; Neilsen, Jr., Eric H.; Long, Dan; Schneider, Donald P.; Snedden, Stephanie A. (2006). "A Catalog of Spectroscopically Confirmed White Dwarfs from the Sloan Digital Sky Survey Data Release 4". The Astrophysical Journal Supplement Series. 167 (1): 40–58. arXiv:astro-ph/0606700. Bibcode:2006ApJS..167...40E. doi:10.1086/507110. S2CID 13829139.
Kilic, M.; Allende Prieto, C.; Brown, Warren R.; Koester, D. (2007). "The Lowest Mass White Dwarf". The Astrophysical Journal. 660 (2): 1451–1461. arXiv:astro-ph/0611498. Bibcode:2007ApJ...660.1451K. doi:10.1086/514327. S2CID 18587748.
Kepler, S. O.; Kleinman, S. J.; Nitta, A.; Koester, D.; Castanheira, B. G.; Giovannini, O.; Costa, A. F. M.; Althaus, L. (2007). "White dwarf mass distribution in the SDSS". Monthly Notices of the Royal Astronomical Society. 375 (4): 1315–1324. arXiv:astro-ph/0612277. Bibcode:2007MNRAS.375.1315K. doi:10.1111/j.1365-2966.2006.11388.x. S2CID 10892288.
Shipman, H. L. (1979). "Masses and radii of white-dwarf stars. III – Results for 110 hydrogen-rich and 28 helium-rich stars". The Astrophysical Journal. 228: 240. Bibcode:1979ApJ...228..240S. doi:10.1086/156841.
Sandin, F. (2005). "Exotic Phases of Matter in Compact Stars" (PDF). Licentiate thesis. Luleå University of Technology. Archived (PDF) from the original on 15 August 2011. Retrieved 20 August 2011.
Boss, L. (1910). Preliminary General Catalogue of 6188 stars for the epoch 1900. Carnegie Institution of Washington. Bibcode:1910pgcs.book.....B. LCCN 10009645.
Liebert, J.; Young, P. A.; Arnett, D.; Holberg, J. B.; Williams, K. A. (2005). "The Age and Progenitor Mass of Sirius B". The Astrophysical Journal. 630 (1): L69. arXiv:astro-ph/0507523. Bibcode:2005ApJ...630L..69L. doi:10.1086/462419. S2CID 8792889.
Öpik, E. (1916). "The Densities of Visual Binary Stars". The Astrophysical Journal. 44: 292. Bibcode:1916ApJ....44..292O. doi:10.1086/142296.
Eddington, A. S. (1927). Stars and Atoms. Clarendon Press. LCCN 27015694.
Adams, W. S. (1925). "The Relativity Displacement of the Spectral Lines in the Companion of Sirius". Proceedings of the National Academy of Sciences. 11 (7): 382–387. Bibcode:1925PNAS...11..382A. doi:10.1073/pnas.11.7.382. PMC 1086032. PMID 16587023.
Nave, C. R. "Nuclear Size and Density". HyperPhysics. Georgia State University. Archived from the original on 6 July 2009. Retrieved 26 June 2009.
Adams, Steve (1997). Relativity: an introduction to space-time physics. Relativity : An Introduction to Space-Time Physics. London ; Bristol: CRC Press. p. 240. Bibcode:1997rist.book.....A. ISBN 978-0-7484-0621-0.
Fowler, R. H. (1926). "On dense matter". Monthly Notices of the Royal Astronomical Society. 87 (2): 114–122. Bibcode:1926MNRAS..87..114F. doi:10.1093/mnras/87.2.114.
Hoddeson, L. H.; Baym, G. (1980). "The Development of the Quantum Mechanical Electron Theory of Metals: 1900–28". Proceedings of the Royal Society of London. 371 (1744): 8–23. Bibcode:1980RSPSA.371....8H. doi:10.1098/rspa.1980.0051. JSTOR 2990270. S2CID 120476662.
"Estimating Stellar Parameters from Energy Equipartition". ScienceBits. Archived from the original on 30 June 2012. Retrieved 9 May 2007.
Bean, R. "Lecture 12 – Degeneracy pressure" (PDF). Lecture notes, Astronomy 211. Cornell University. Archived from the original (PDF) on 25 September 2007. Retrieved 21 September 2007.
Anderson, W. (1929). "Über die Grenzdichte der Materie und der Energie". Zeitschrift für Physik. 56 (11–12): 851–856. Bibcode:1929ZPhy...56..851A. doi:10.1007/BF01340146. S2CID 122576829.
Stoner, C. (1930). "The Equilibrium of Dense Stars". Philosophical Magazine. 9: 944.
Chandrasekhar, S. (1931). "The Maximum Mass of Ideal White Dwarfs". The Astrophysical Journal. 74: 81. Bibcode:1931ApJ....74...81C. doi:10.1086/143324.
Chandrasekhar, S. (1935). "The highly collapsed configurations of a stellar mass (Second paper)". Monthly Notices of the Royal Astronomical Society. 95 (3): 207–225. Bibcode:1935MNRAS..95..207C. doi:10.1093/mnras/95.3.207.
"The Nobel Prize in Physics 1983". The Nobel Foundation. Archived from the original on 6 May 2007. Retrieved 4 May 2007.
Canal, R.; Gutierrez, J. (1997). "The Possible White Dwarf-Neutron Star Connection". White Dwarfs. Astrophysics and Space Science Library. 214. pp. 49–55. arXiv:astro-ph/9701225. Bibcode:1997ASSL..214...49C. doi:10.1007/978-94-011-5542-7_7. ISBN 978-94-010-6334-0. S2CID 9288287.
Hillebrandt, W.; Niemeyer, J. C. (2000). "Type IA supernova explosion models". Annual Review of Astronomy and Astrophysics. 38: 191–230. arXiv:astro-ph/0006305. Bibcode:2000ARA&A..38..191H. doi:10.1146/annurev.astro.38.1.191. S2CID 10210550.
Overbye, D. (22 February 2010). "From the Clash of White Dwarfs, the Birth of a Supernova". New York Times. Archived from the original on 25 February 2010. Retrieved 22 February 2010.
Chabrier, G.; Baraffe, I. (2000). "Theory of low-Mass stars and substellar objects". Annual Review of Astronomy and Astrophysics. 38: 337–377.arXiv:astro-ph/0006383. Bibcode:2000ARA&A..38..337C. doi:10.1146/annurev.astro.38.1.337. S2CID 59325115.
Kaler, J. "The Hertzsprung-Russell (HR) diagram". Archived from the original on 31 August 2009. Retrieved 5 May 2007.
"Basic symbols". Standards for Astronomical Catalogues, Version 2.0. VizieR. Archived from the original on 8 May 2017. Retrieved 12 January 2007.
Tohline, J. E. "The Structure, Stability, and Dynamics of Self-Gravitating Systems". Archived from the original on 27 June 2010. Retrieved 30 May 2007.
Hoyle, F. (1947). "Stars, Distribution and Motions of, Note on equilibrium configurations for rotating white dwarfs". Monthly Notices of the Royal Astronomical Society. 107 (2): 231–236. Bibcode:1947MNRAS.107..231H. doi:10.1093/mnras/107.2.231.
Ostriker, J. P.; Bodenheimer, P. (1968). "Rapidly Rotating Stars. II. Massive White Dwarfs". The Astrophysical Journal. 151: 1089. Bibcode:1968ApJ...151.1089O. doi:10.1086/149507.
Kutner, M. L. (2003). Astronomy: A physical perspective. Cambridge University Press. p. 189. ISBN 978-0-521-52927-3. Retrieved 28 February 2016.
Sion, E. M.; Greenstein, J. L.; Landstreet, J. D.; Liebert, J.; Shipman, H. L.; Wegner, G. A. (1983). "A proposed new white dwarf spectral classification system". The Astrophysical Journal. 269: 253. Bibcode:1983ApJ...269..253S. doi:10.1086/161036.
Hambly, N. C.; Smartt, S. J.; Hodgkin, S. T. (1997). "WD 0346+246: A Very Low Luminosity, Cool Degenerate in Taurus". The Astrophysical Journal. 489 (2): L157. Bibcode:1997ApJ...489L.157H. doi:10.1086/316797.
Fontaine, G.; Wesemael, F. (2001). "White dwarfs". In Murdin, P. (ed.). Encyclopedia of Astronomy and Astrophysics. IOP Publishing/Nature Publishing Group. ISBN 978-0-333-75088-9.
Heise, J. (1985). "X-ray emission from isolated hot white dwarfs". Space Science Reviews. 40 (1–2): 79–90. Bibcode:1985SSRv...40...79H. doi:10.1007/BF00212870. S2CID 120431159.
Lesaffre, P.; Podsiadlowski, Ph.; Tout, C. A. (2005). "A two-stream formalism for the convective Urca process". Monthly Notices of the Royal Astronomical Society. 356 (1): 131–144.arXiv:astro-ph/0411016. Bibcode:2005MNRAS.356..131L. doi:10.1111/j.1365-2966.2004.08428.x. S2CID 15797437.
Mestel, L. (1952). "On the theory of white dwarf stars. I. The energy sources of white dwarfs". Monthly Notices of the Royal Astronomical Society. 112 (6): 583–597. Bibcode:1952MNRAS.112..583M. doi:10.1093/mnras/112.6.583.
Kawaler, S. D. (1998). "White Dwarf Stars and the Hubble Deep Field". The Hubble Deep Field : Proceedings of the Space Telescope Science Institute Symposium. The Hubble Deep Field. p. 252.arXiv:astro-ph/9802217. Bibcode:1998hdf..symp..252K. ISBN 978-0-521-63097-9.
Bergeron, P.; Ruiz, M. T.; Leggett, S. K. (1997). "The Chemical Evolution of Cool White Dwarfs and the Age of the Local Galactic Disk". The Astrophysical Journal Supplement Series. 108 (1): 339–387. Bibcode:1997ApJS..108..339B. doi:10.1086/312955.
McCook, G. P.; Sion, E. M. (1999). "A Catalogue of Spectroscopically Identified White Dwarfs". The Astrophysical Journal Supplement Series. 121 (1): 1–130. Bibcode:1999ApJS..121....1M. doi:10.1086/313186.
Leggett, S. K.; Ruiz, M. T.; Bergeron, P. (1998). "The Cool White Dwarf Luminosity Function and the Age of the Galactic Disk". The Astrophysical Journal. 497 (1): 294–302. Bibcode:1998ApJ...497..294L. doi:10.1086/305463.
Gates, E.; Gyuk, G.; Harris, H. C.; Subbarao, M.; Anderson, S.; Kleinman, S. J.; Liebert, J.; Brewington, H.; et al. (2004). "Discovery of New Ultracool White Dwarfs in the Sloan Digital Sky Survey". The Astrophysical Journal. 612 (2): L129.arXiv:astro-ph/0405566. Bibcode:2004ApJ...612L.129G. doi:10.1086/424568. S2CID 7570539.
Winget, D. E.; Hansen, C. J.; Liebert, J.; Van Horn, H. M.; Fontaine, G.; Nather, R. E.; Kepler, S. O.; Lamb, D. Q. (1987). "An independent method for determining the age of the universe". The Astrophysical Journal. 315: L77. Bibcode:1987ApJ...315L..77W. doi:10.1086/184864. hdl:10183/108730.
Trefil, J. S. (2004). The Moment of Creation: Big Bang Physics from Before the First Millisecond to the Present Universe. Dover Publications. ISBN 978-0-486-43813-9.
Metcalfe, T. S.; Montgomery, M. H.; Kanaan, A. (2004). "Testing White Dwarf Crystallization Theory with Asteroseismology of the Massive Pulsating DA Star BPM 37093". The Astrophysical Journal. 605 (2): L133.arXiv:astro-ph/0402046. Bibcode:2004ApJ...605L.133M. doi:10.1086/420884. S2CID 119378552.
Barrat, J. L.; Hansen, J. P.; Mochkovitch, R. (1988). "Crystallization of carbon-oxygen mixtures in white dwarfs". Astronomy and Astrophysics. 199 (1–2): L15. Bibcode:1988A&A...199L..15B.
Winget, D. E. (1995). "The Status of White Dwarf Asteroseismology and a Glimpse of the Road Ahead". Baltic Astronomy. 4 (2): 129. Bibcode:1995BaltA...4..129W. doi:10.1515/astro-1995-0209.
Whitehouse, David (16 February 2004). "Diamond star thrills astronomers". BBC News. Archived from the original on 5 February 2007. Retrieved 6 January 2007.
Kanaan, A.; Nitta, A.; Winget, D. E.; Kepler, S. O.; Montgomery, M. H.; Metcalfe, T. S.; Oliveira, H.; Fraga, L.; et al. (2005). "Whole Earth Telescope observations of BPM 37093: A seismological test of crystallization theory in white dwarfs". Astronomy and Astrophysics. 432 (1): 219–224.arXiv:astro-ph/0411199. Bibcode:2005A&A...432..219K. doi:10.1051/0004-6361:20041125. S2CID 7297628.
Brassard, P.; Fontaine, G. (2005). "Asteroseismology of the Crystallized ZZ Ceti Star BPM 37093: A Different View". The Astrophysical Journal. 622 (1): 572–576. Bibcode:2005ApJ...622..572B. doi:10.1086/428116.
Hansen, B.M.S.; Liebert, J. (2003). "Cool White Dwarfs". Annual Review of Astronomy and Astrophysics. 41: 465. Bibcode:2003ARA&A..41..465H. doi:10.1146/annurev.astro.41.081401.155117.
Tremblay, P.-E.; Fontaine, G.; Fusillo, N. P. G.; Dunlap, B. H.; Gänsicke, B. T.; Hollands, M. H.; Hermes, J. J.; Marsh, T. R.; Cukanovaite, E.; Cunningham, T. (2019). "Core crystallization and pile-up in the cooling sequence of evolving white dwarfs" (PDF). Nature. 565 (7738): 202–205.arXiv:1908.00370. Bibcode:2019Natur.565..202T. doi:10.1038/s41586-018-0791-x. PMID 30626942. S2CID 58004893. Archived (PDF) from the original on 23 July 2019. Retrieved 23 July 2019.
Istrate; et al. (2014). "The timescale of low-mass proto-helium white dwarf evolution". Astronomy and Astrophysics. 571: L3.arXiv:1410.5471. Bibcode:2014A&A...571L...3I. doi:10.1051/0004-6361/201424681. S2CID 55152203.
"First Giant Planet around White Dwarf Found – ESO observations indicate the Neptune-like exoplanet is evaporating". www.eso.org. Archived from the original on 4 December 2019. Retrieved 4 December 2019.
Schatzman, E. (1945). "Théorie du débit d'énergie des naines blanches". Annales d'Astrophysique. 8: 143. Bibcode:1945AnAp....8..143S.
Koester, D.; Chanmugam, G. (1990). "Physics of white dwarf stars". Reports on Progress in Physics. 53 (7): 837–915. Bibcode:1990RPPh...53..837K. doi:10.1088/0034-4885/53/7/001. S2CID 122582479.
Kawaler, S. D. (1997). "White Dwarf Stars". In Kawaler, S. D.; Novikov, I.; Srinivasan, G. (eds.). Stellar remnants. 1997. ISBN 978-3-540-61520-0.
Kuiper, G. P. (1941). "List of Known White Dwarfs". Publications of the Astronomical Society of the Pacific. 53 (314): 248. Bibcode:1941PASP...53..248K. doi:10.1086/125335.
Luyten, W. J. (1952). "The Spectra and Luminosities of White Dwarfs". The Astrophysical Journal. 116: 283. Bibcode:1952ApJ...116..283L. doi:10.1086/145612.
Greenstein, J. L. (1960). Stellar atmospheres. University of Chicago Press. Bibcode:1960stat.book.....G. LCCN 61-9138.
Kepler, S. O.; Kleinman, S. J.; Nitta, A.; Koester, D.; Castanheira, B. G.; Giovannini, O.; Costa, A. F. M.; Althaus, L. (2007). "White dwarf mass distribution in the SDSS". Monthly Notices of the Royal Astronomical Society. 375 (4): 1315–1324.arXiv:astro-ph/0612277. Bibcode:2007MNRAS.375.1315K. doi:10.1111/j.1365-2966.2006.11388.x. S2CID 10892288.
Dufour, P.; Liebert, J.; Fontaine, G.; Behara, N. (2007). "White dwarf stars with carbon atmospheres". Nature. 450 (7169): 522–4.arXiv:0711.3227. Bibcode:2007Natur.450..522D. doi:10.1038/nature06318. PMID 18033290. S2CID 4398697.
"Discovery of Molecular Hydrogen in White Dwarf Atmospheres". IOPscience. Archived from the original on 24 June 2014. Retrieved 24 October 2013.
Jura, M.; Young, E.D. (1 January 2014). "Extrasolar Cosmochemistry". Annual Review of Earth and Planetary Sciences. 42 (1): 45–67. Bibcode:2014AREPS..42...45J. doi:10.1146/annurev-earth-060313-054740.
Wilson, D.J.; Gänsicke, B.T.; Koester, D.; Toloza, O.; Pala, A. F.; Breedt, E.; Parsons, S.G. (11 August 2015). "The composition of a disrupted extrasolar planetesimal at SDSS J0845+2257 (Ton 345)". Monthly Notices of the Royal Astronomical Society. 451 (3): 3237–3248.arXiv:1505.07466. Bibcode:2015MNRAS.451.3237W. doi:10.1093/mnras/stv1201. S2CID 54049842.
Blackett, P. M. S. (1947). "The Magnetic Field of Massive Rotating Bodies". Nature. 159 (4046): 658–66. Bibcode:1947Natur.159..658B. doi:10.1038/159658a0. PMID 20239729. S2CID 4133416.
Lovell, B. (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. doi:10.1098/rsbm.1975.0001. JSTOR 769678. S2CID 74674634.
Landstreet, John D. (1967). "Synchrotron radiation of neutrinos and its astrophysical significance". Physical Review. 153 (5): 1372–1377. Bibcode:1967PhRv..153.1372L. doi:10.1103/PhysRev.153.1372.
Ginzburg, V. L.; Zheleznyakov, V. V.; Zaitsev, V. V. (1969). "Coherent mechanisms of radio emission and magnetic models of pulsars". Astrophysics and Space Science. 4 (4): 464–504. Bibcode:1969Ap&SS...4..464G. doi:10.1007/BF00651351. S2CID 119003761.
Kemp, J. C.; Swedlund, J. B.; Landstreet, J. D.; Angel, J. R. P. (1970). "Discovery of Circularly Polarized Light from a White Dwarf". The Astrophysical Journal. 161: L77. Bibcode:1970ApJ...161L..77K. doi:10.1086/180574.
Ferrario, Lilia; de Martino, Domtilla; Gaensicke, Boris (2015). "Magnetic white dwarfs". Space Science Reviews. 191 (1–4): 111–169.arXiv:1504.08072. Bibcode:2015SSRv..191..111F. doi:10.1007/s11214-015-0152-0. S2CID 119057870.
Kepler, S.O.; Pelisoli, I.; Jordan, S.; Kleinman, S.J.; Koester, D.; Kuelebi, B.; Pecanha, V.; Castanhiera, B.G.; Nitta, A.; Costa, J.E.S.; Winget, D.E.; Kanaan, A.; Fraga, L. (2013). "Magnetic white dwarf stars in the Sloan Digital Sky Survey". Monthly Notices of the Royal Astronomical Society. 429 (4): 2934–2944.arXiv:1211.5709. Bibcode:2013MNRAS.429.2934K. doi:10.1093/mnras/sts522. S2CID 53316287.
Landstreet, J.D.; Bagnulo, S.; Valyavin, G.G.; Fossati, L.; Jordan, S.; Monin, D.; Wade, G.A. (2012). "On the incidence of weak magnetic fields in DA white dwarfs". Astronomy and Astrophysics. 545 (A30): 9pp.arXiv:1208.3650. Bibcode:2012A&A...545A..30L. doi:10.1051/0004-6361/201219829. S2CID 55153825.
Liebert, James; Bergeron, P.; Holberg, J. B. (2003). "The True Incidence of Magnetism Among Field White Dwarfs". The Astronomical Journal. 125 (1): 348–353.arXiv:astro-ph/0210319. Bibcode:2003AJ....125..348L. doi:10.1086/345573. S2CID 9005227.
Buckley, D. A. H.; Meintjes, P. J.; Potter, S. B.; Marsh, T. R.; Gänsicke, B. T. (23 January 2017). "Polarimetric evidence of a white dwarf pulsar in the binary system AR Scorpii". Nature Astronomy. 1 (2): 0029.arXiv:1612.03185. Bibcode:2017NatAs...1E..29B. doi:10.1038/s41550-016-0029. S2CID 15683792.
"Stars draw atoms closer together". Nature News & Comment. Archived from the original on 20 July 2012. Retrieved 21 July 2012.
"ZZ Ceti variables". Centre deDonnées astronomiques de Strasbourg. Association Française des Observateurs d'Etoiles Variables. Archived from the original on 5 February 2007. Retrieved 6 June 2007.
Quirion, P.‐O.; Fontaine, G.; Brassard, P. (2007). "Mapping the Instability Domains of GW Vir Stars in the Effective Temperature–Surface Gravity Diagram". The Astrophysical Journal Supplement Series. 171 (1): 219–248. Bibcode:2007ApJS..171..219Q. doi:10.1086/513870.
Lawrence, G. M.; Ostriker, J. P.; Hesser, J. E. (1967). "Ultrashort-Period Stellar Oscillations. I. Results from White Dwarfs, Old Novae, Central Stars of Planetary Nebulae, 3c 273, and Scorpius XR-1". The Astrophysical Journal. 148: L161. Bibcode:1967ApJ...148L.161L. doi:10.1086/180037.
Landolt, A. U. (1968). "A New Short-Period Blue Variable". The Astrophysical Journal. 153: 151. Bibcode:1968ApJ...153..151L. doi:10.1086/149645.
Nagel, T.; Werner, K. (2004). "Detection of non-radial g-mode pulsations in the newly discovered PG 1159 star HE 1429-1209". Astronomy and Astrophysics. 426 (2): L45.arXiv:astro-ph/0409243. Bibcode:2004A&A...426L..45N. doi:10.1051/0004-6361:200400079. S2CID 9481357.
O'Brien, M. S. (2000). "The Extent and Cause of the Pre–White Dwarf Instability Strip". The Astrophysical Journal. 532 (2): 1078–1088.arXiv:astro-ph/9910495. Bibcode:2000ApJ...532.1078O. doi:10.1086/308613. S2CID 115958740.
Winget, D. E. (1998). "Asteroseismology of white dwarf stars". Journal of Physics: Condensed Matter. 10 (49): 11247–11261. Bibcode:1998JPCM...1011247W. doi:10.1088/0953-8984/10/49/014.
Heger, A.; Fryer, C. L.; Woosley, S. E.; Langer, N.; Hartmann, D. H. (2003). "How Massive Single Stars End Their Life". The Astrophysical Journal. 591 (1): 288–300.arXiv:astro-ph/0212469. Bibcode:2003ApJ...591..288H. doi:10.1086/375341. S2CID 59065632.
Napiwotzki, Ralf (2009). "The galactic population of white dwarfs". Journal of Physics. Conference Series. 172 (1): 012004.arXiv:0903.2159. Bibcode:2009JPhCS.172a2004N. doi:10.1088/1742-6596/172/1/012004. S2CID 17521113.
Laughlin, G.; Bodenheimer, P.; Adams, Fred C. (1997). "The End of the Main Sequence". The Astrophysical Journal. 482 (1): 420–432. Bibcode:1997ApJ...482..420L. doi:10.1086/304125.
Jeffery, Simon. "Stars Beyond Maturity". Archived from the original on 4 April 2015. Retrieved 3 May 2007.
Sarna, M. J.; Ergma, E.; Gerškevitš, J. (2001). "Helium core white dwarf evolution – including white dwarf companions to neutron stars". Astronomische Nachrichten. 322 (5–6): 405–410. Bibcode:2001AN....322..405S. doi:10.1002/1521-3994(200112)322:5/6<405::AID-ASNA405>3.0.CO;2-6.
Benvenuto, O. G.; De Vito, M. A. (2005). "The formation of helium white dwarfs in close binary systems – II". Monthly Notices of the Royal Astronomical Society. 362 (3): 891–905. Bibcode:2005MNRAS.362..891B. doi:10.1111/j.1365-2966.2005.09315.x.
Nelemans, G.; Tauris, T. M. (1998). "Formation of undermassive single white dwarfs and the influence of planets on late stellar evolution". Astronomy and Astrophysics. 335: L85.arXiv:astro-ph/9806011. Bibcode:1998A&A...335L..85N.
"Planet diet helps white dwarfs stay young and trim". New Scientist. No. 2639. 18 January 2008. Archived from the original on 20 April 2010. Retrieved 18 September 2017.
Dhillon, Vik. "The evolution of low-mass stars". lecture notes, Physics 213. University of Sheffield. Archived from the original on 7 November 2012. Retrieved 3 May 2007.
Dhillon, Vik. "The evolution of high-mass stars". lecture notes, Physics 213. University of Sheffield. Archived from the original on 7 November 2012. Retrieved 3 May 2007.
Schaffner-Bielich, Jürgen (2005). "Strange quark matter in stars: A general overview". Journal of Physics G: Nuclear and Particle Physics. 31 (6): S651–S657.arXiv:astro-ph/0412215. Bibcode:2005JPhG...31S.651S. doi:10.1088/0954-3899/31/6/004. S2CID 118886040.
Nomoto, K. (1984). "Evolution of 8–10 solar mass stars toward electron capture supernovae. I – Formation of electron-degenerate O + NE + MG cores". The Astrophysical Journal. 277: 791. Bibcode:1984ApJ...277..791N. doi:10.1086/161749.
Woosley, S. E.; Heger, A.; Weaver, T. A. (2002). "The evolution and explosion of massive stars". Reviews of Modern Physics. 74 (4): 1015–1071. Bibcode:2002RvMP...74.1015W. doi:10.1103/RevModPhys.74.1015.
Werner, K.; Rauch, T.; Barstow, M. A.; Kruk, J. W. (2004). "Chandra and FUSE spectroscopy of the hot bare stellar core H?1504+65". Astronomy and Astrophysics. 421 (3): 1169–1183.arXiv:astro-ph/0404325. Bibcode:2004A&A...421.1169W. doi:10.1051/0004-6361:20047154. S2CID 2983893.
Livio, Mario; Truran, James W. (1994). "On the interpretation and implications of nova abundances: An abundance of riches or an overabundance of enrichments". The Astrophysical Journal. 425: 797. Bibcode:1994ApJ...425..797L. doi:10.1086/174024.
Jordan, George C. IV.; Perets, Hagai B.; Fisher, Robert T.; van Rossum, Daniel R. (2012). "Failed-detonation Supernovae: Subluminous Low-velocity Ia Supernovae and their Kicked Remnant White Dwarfs with Iron-rich Cores". The Astrophysical Journal Letters. 761 (2): L23.arXiv:1208.5069. Bibcode:2012ApJ...761L..23J. doi:10.1088/2041-8205/761/2/L23. S2CID 119203015.
Panei, J. A.; Althaus, L. G.; Benvenuto, O. G. (2000). "The evolution of iron-core white dwarfs". Monthly Notices of the Royal Astronomical Society. 312 (3): 531–539.arXiv:astro-ph/9911371. Bibcode:2000MNRAS.312..531P. doi:10.1046/j.1365-8711.2000.03236.x. S2CID 17854858.
Adams, Fred C.; Laughlin, Gregory (1997). "A dying universe: The long-term fate and evolution of astrophysical objects". Reviews of Modern Physics. 69 (2): 337–372.arXiv:astro-ph/9701131. Bibcode:1997RvMP...69..337A. doi:10.1103/RevModPhys.69.337. S2CID 12173790.
Seager, S.; Kuchner, M.; Hier-Majumder, C.; Militzer, B. (19 July 2007). "Mass-Radius Relationships for Solid Exoplanets". The Astrophysical Journal (published November 2007). 669 (2): 1279–1297.arXiv:0707.2895. Bibcode:2007ApJ...669.1279S. doi:10.1086/521346. S2CID 8369390.
Lemonick, Michael (26 August 2011). "Scientists Discover a Diamond as Big as a Planet". Time Magazine. Archived from the original on 24 August 2013. Retrieved 18 June 2015.
"Hubble finds dead stars "polluted" with planetary debris". ESA/Hubble Press Release. Archived from the original on 9 June 2013. Retrieved 10 May 2013.
"Comet falling into white dwarf (artist's impression)". www.spacetelescope.org. Archived from the original on 15 February 2017. Retrieved 14 February 2017.
"Comet clash kicks up dusty haze". BBC News. 13 February 2007. Archived from the original on 16 February 2007. Retrieved 20 September 2007.
Su, K. Y. L.; Chu, Y.-H.; Rieke, G. H.; Huggins, P. J.; Gruendl, R.; Napiwotzki, R.; Rauch, T.; Latter, W. B.; Volk, K. (2007). "A Debris Disk around the Central Star of the Helix Nebula?". The Astrophysical Journal. 657 (1): L41.arXiv:astro-ph/0702296. Bibcode:2007ApJ...657L..41S. doi:10.1086/513018. S2CID 15244406.
Reach, William T.; Kuchner, Marc J.; Von Hippel, Ted; Burrows, Adam; Mullally, Fergal; Kilic, Mukremin; Winget, D. E. (2005). "The Dust Cloud around the White Dwarf G29-38". The Astrophysical Journal. 635 (2): L161.arXiv:astro-ph/0511358. Bibcode:2005ApJ...635L.161R. doi:10.1086/499561. S2CID 119462589.
Sion, Edward M.; Holberg, J.B.; Oswalt, Terry D.; McCook, George P.; Wasatonic, Richard (2009). "The White Dwarfs Within 20 Parsecs of the Sun: Kinematics and Statistics". The Astronomical Journal. 138 (6): 1681–1689.arXiv:0910.1288. Bibcode:2009AJ....138.1681S. doi:10.1088/0004-6256/138/6/1681. S2CID 119284418.
Li, Jianke; Ferrario, Lilia; Wickramasinghe, Dayal (1998). "Planets around White Dwarfs". Astrophysical Journal Letters. 503 (1): L151. Bibcode:1998ApJ...503L.151L. doi:10.1086/311546. p. L51.
Debes, John H.; Walsh, Kevin J.; Stark, Christopher (24 February 2012). "The Link Between Planetary Systems, Dusty White Dwarfs, and Metal-Polluted White Dwarfs". The Astrophysical Journal. 747 (2): 148.arXiv:1201.0756. Bibcode:2012ApJ...747..148D. doi:10.1088/0004-637X/747/2/148. ISSN 0004-637X. S2CID 118688656.
Veras, Dimitri; Gänsicke, Boris T. (21 February 2015). "Detectable close-in planets around white dwarfs through late unpacking". Monthly Notices of the Royal Astronomical Society. 447 (2): 1049–1058.arXiv:1411.6012. Bibcode:2015MNRAS.447.1049V. doi:10.1093/mnras/stu2475. ISSN 0035-8711. S2CID 119279872.
Frewen, S. F. N.; Hansen, B. M. S. (11 April 2014). "Eccentric planets and stellar evolution as a cause of polluted white dwarfs". Monthly Notices of the Royal Astronomical Society. 439 (3): 2442–2458.arXiv:1401.5470. Bibcode:2014MNRAS.439.2442F. doi:10.1093/mnras/stu097. ISSN 0035-8711. S2CID 119257046.
Bonsor, Amy; Gänsicke, Boris T.; Veras, Dimitri; Villaver, Eva; Mustill, Alexander J. (21 May 2018). "Unstable low-mass planetary systems as drivers of white dwarf pollution". Monthly Notices of the Royal Astronomical Society. 476 (3): 3939–3955.arXiv:1711.02940. Bibcode:2018MNRAS.476.3939M. doi:10.1093/mnras/sty446. ISSN 0035-8711. S2CID 4809366.
Gänsicke, Boris T.; Holman, Matthew J.; Veras, Dimitri; Payne, Matthew J. (21 March 2016). "Liberating exomoons in white dwarf planetary systems". Monthly Notices of the Royal Astronomical Society. 457 (1): 217–231.arXiv:1603.09344. Bibcode:2016MNRAS.457..217P. doi:10.1093/mnras/stv2966. ISSN 0035-8711. S2CID 56091285.
Rebassa-Mansergas, Alberto; Xu (许偲艺), Siyi; Veras, Dimitri (21 January 2018). "The critical binary star separation for a planetary system origin of white dwarf pollution". Monthly Notices of the Royal Astronomical Society. 473 (3): 2871–2880.arXiv:1708.05391. Bibcode:2018MNRAS.473.2871V. doi:10.1093/mnras/stx2141. ISSN 0035-8711. S2CID 55764122.
Becklin, E. E.; Zuckerman, B.; Farihi, J. (10 February 2008). "Spitzer IRAC Observations of White Dwarfs. I. Warm Dust at Metal-Rich Degenerates". The Astrophysical Journal. 674 (1): 431–446.arXiv:0710.0907. Bibcode:2008ApJ...674..431F. doi:10.1086/521715. ISSN 0004-637X. S2CID 17813180.
Debes, John H.; Thévenot, Melina; Kuchner, Marc J.; Burgasser, Adam J.; Schneider, Adam C.; Meisner, Aaron M.; Gagné, Jonathan; Faherty, Jacqueline K.; Rees, Jon M. (19 February 2019). "A 3 Gyr White Dwarf with Warm Dust Discovered via the Backyard Worlds: Planet 9 Citizen Science Project". The Astrophysical Journal. 872 (2): L25.arXiv:1902.07073. Bibcode:2019ApJ...872L..25D. doi:10.3847/2041-8213/ab0426. ISSN 2041-8213. S2CID 119359995.
Lemonick, Michael D. (21 October 2015). "Zombie Star Caught Feasting on Asteroids". National Geographic News. Archived from the original on 24 October 2015. Retrieved 22 October 2015.
Vanderburg, Andrew; Johnson, John Asher; Rappaport, Saul; Bieryla, Allyson; Irwin, Jonathan; Lewis, John Arban; Kipping, David; Brown, Warren R.; Dufour, Patrick (22 October 2015). "A disintegrating minor planet transiting a white dwarf". Nature. 526 (7574): 546–549.arXiv:1510.06387. Bibcode:2015Natur.526..546V. doi:10.1038/nature15527. PMID 26490620. S2CID 4451207.
Wang, Ting-Gui; Jiang, Ning; Ge, Jian; Cutri, Roc M.; Jiang, Peng; Sheng, Zhengfeng; Zhou, Hongyan; Bauer, James; Mainzer, Amy; Wright, Edward L. (9 October 2019). "An On-going Mid-infrared Outburst in the White Dwarf 0145+234: Catching in Action of Tidal Disruption of an Exoasteroid?".arXiv:1910.04314 [astro-ph.SR].
Gänsicke, Boris T.; Schreiber, Matthias R.; Toloza, Odette; Gentile Fusillo, Nicola P.; Koester, Detlev; Manser, Christopher J. "Accretion of a giant planet onto a white dwarf" (PDF). ESO. Archived (PDF) from the original on 4 December 2019. Retrieved 11 December 2019.
Vanderburg, Andrew; et al. (16 September 2020). "A giant planet candidate transiting a white dwarf". Nature. 585 (7825): 363–367.arXiv:2009.07282. Bibcode:2020Natur.585..363V. doi:10.1038/s41586-020-2713-y. PMID 32939071. S2CID 221738865. Retrieved 17 September 2020.
Chou, felicia; Andreoli, Claire; Cofield, Calia (16 September 2020). "NASA Missions Spy First Possible Planet Hugging a Stellar Cinder". NASA. Retrieved 17 September 2020.
Gary, Bruce L. (17 September 2020). "WD 1856+534 Transit Light Curve Photometry". BruceGary.net. Retrieved 17 September 2020.
Agol, Eric (2011). "Transit Surveys for Earths in the Habitable Zones of White Dwarfs". The Astrophysical Journal Letters. 635 (2): L31.arXiv:1103.2791. Bibcode:2011ApJ...731L..31A. doi:10.1088/2041-8205/731/2/L31. S2CID 118739494.
Barnes, Rory; Heller, René (2011). "Habitable Planets Around White and Brown Dwarfs: The Perils of a Cooling Primary". Astrobiology. 13 (3): 279–291.arXiv:1211.6467. Bibcode:2013AsBio..13..279B. doi:10.1089/ast.2012.0867. PMC 3612282. PMID 23537137.
Nordhaus, J.; Spiegel, D.S. (2013). "On the orbits of low-mass companions to white dwarfs and the fates of the known exoplanets". Monthly Notices of the Royal Astronomical Society. 432 (1): 500–505.arXiv:1211.1013. Bibcode:2013MNRAS.432..500N. doi:10.1093/mnras/stt569. S2CID 119227364.
Di Stefano, R.; Nelson, L. A.; Lee, W.; Wood, T. H.; Rappaport, S. (1997). P. Ruiz-Lapuente; R. Canal; J. Isern (eds.). Luminous Supersoft X-ray Sources as Type Ia Progenitors. Nato Advanced Science Institutes (Asi) Series C. NATO ASI series: Mathematical and physical sciences. 486. Springer. pp. 148–149. Bibcode:1997ASIC..486..147D. doi:10.1007/978-94-011-5710-0_10. ISBN 978-0-7923-4359-2.
{{cite journal |bibcode=2020ApJ...898L..40L |arxiv= 2007.13932 |doi= 10.3847/2041-8213/aba618 |title=CTCV J2056-3014: An X-Ray-faint Intermediate Polar Harboring an Extremely Fast-spinning White Dwarf |date=2020 |last1=Lopes de Oliveira |first1=R. |last2=Bruch |first2=A. |last3=Rodrigues |first3=C. V. |last4=de Oliveira |first4=A. S. |last5=Mukai |first5=K. |journal=The Astrophysical Journal Letters |volume=898 |issue=2 |pages=L40 }}
Aguilar, David A.; Pulliam, Christine (16 November 2010). "Astronomers Discover Merging Star Systems that Might Explode". Harvard-Smithsonian Center for Astrophysics. Archived from the original on 9 April 2011. Retrieved 16 February 2011.
Aguilar, David A.; Pulliam, Christine (13 July 2011). "Evolved Stars Locked in Fatalistic Dance". Harvard-Smithsonian Center for Astrophysics. Archived from the original on 15 July 2011. Retrieved 17 July 2011.
Yoon, S.-C.; Langer, N. (2004). "Presupernova evolution of accreting white dwarfs with rotation". Astronomy and Astrophysics. 419 (2): 623–644.arXiv:astro-ph/0402287. Bibcode:2004A&A...419..623Y. doi:10.1051/0004-6361:20035822. S2CID 2963085.
Blinnikov, S. I.; Röpke, F. K.; Sorokina, E. I.; Gieseler, M.; Reinecke, M.; Travaglio, C.; Hillebrandt, W.; Stritzinger, M. (2006). "Theoretical light curves for deflagration models of type Ia supernova". Astronomy and Astrophysics. 453 (1): 229–240.arXiv:astro-ph/0603036. Bibcode:2006A&A...453..229B. doi:10.1051/0004-6361:20054594. S2CID 15493284.
O'Neill, Ian (6 September 2011). "Don't slow down white dwarf, you might explode". Discovery Communications, LLC. Archived from the original on 24 January 2012.
González Hernández, J.I.; Ruiz-Lapuente, P.; Tabernero, H. M.; Montes, D.; Canal, R.; Méndez, J.; Bedin, L. R. (2012). "No surviving evolved companions of the progenitor of SN 1006". Nature. 489 (7417): 533–536.arXiv:1210.1948. Bibcode:2012Natur.489..533G. doi:10.1038/nature11447. PMID 23018963. S2CID 4431391.
Krause, Oliver; et al. (2008). "Tycho Brahe's 1572 supernova as a standard type Ia as revealed by its light-echo spectrum". Nature. 456 (7222): 617–619.arXiv:0810.5106. Bibcode:2008Natur.456..617K. doi:10.1038/nature07608. PMID 19052622. S2CID 4409995.
"Cataclysmic Variables". fact sheet. Imagine the Universe!. NASA Goddard. Archived from the original on 9 July 2007. Retrieved 4 May 2007.
"Introduction to Cataclysmic Variables (CVs)". fact sheet. NASA Goddard. Archived from the original on 8 June 2012. Retrieved 4 May 2007.
Giammichele, N.; Bergeron, P.; Dufour, P. (April 2012), "Know Your Neighborhood: A Detailed Model Atmosphere Analysis of Nearby White Dwarfs", The Astrophysical Journal Supplement, 199 (2): 35,arXiv:1202.5581, Bibcode:2012ApJS..199...29G, doi:10.1088/0067-0049/199/2/29, S2CID 118304737, 29.

Delfosse, X.; Forveille, T.; Beuzit, J.-L.; Udry, S.; Mayor, M.; Perrier, C. (1 December 1998). "New neighbours. I. 13 new companions to nearby M dwarfs". Astronomy & Astrophysics. 334: 897.arXiv:astro-ph/9812008. Bibcode:1999A&A...344..897D.

External links and further reading
General

Kawaler, S. D. (1997). "White Dwarf Stars". In Kawaler, S. D.; Novikov, I.; Srinivasan, G. (eds.). Stellar remnants. 1997. ISBN 978-3-540-61520-0.
Kepler, S. O.; Pelisoli, I.; Koester, D.; Ourique, G.; Kleinman, S. J.; Romero, A. D.; Nitta, A.; Eisenstein, D.J.; CostaJ. E. S.; Külebi, B.; Jordan, S.; Dufour, P.; Paolo Giommi, P.; Rebassa-Mansergas, A. «New white dwarf stars in the Sloan Digital Sky Survey Data Release 10». Monthly Notices of the Royal Astronomical Society, 446, 4, 01-02-2015, pàg. 4078–4087. DOI: 10.1093/mnras/stu2388. ISSN: 1365–2966.
Rebassa-Mansergas, A.; Gänsicke, B. T.; Rodríguez-Gil, P.; Schreiber, M. R.; Koester, D. (28 November 2007). "Post-common-envelope binaries from SDSS – I. 101 white dwarf main-sequence binaries with multiple Sloan Digital Sky Survey spectroscopy: Post-common-envelope binaries from SDSS". Monthly Notices of the Royal Astronomical Society. 382 (4): 1377–1393. doi:10.1111/j.1365-2966.2007.12288.x.

Physics

Black holes, white dwarfs, and neutron stars: the physics of compact objects, Stuart L. Shapiro and Saul A. Teukolsky, New York: Wiley, 1983. ISBN 0-471-87317-9.
Koester, D; Chanmugam, G (1990). "Physics of white dwarf stars". Reports on Progress in Physics. 53 (7): 837–915. Bibcode:1990RPPh...53..837K. doi:10.1088/0034-4885/53/7/001. S2CID 122582479.
Gentile, Dave (1995). White dwarf stars and the Chandrasekhar limit (Master's thesis). DePaul University.
"Estimating Stellar Parameters from Energy Equipartition". sciencebits.com. — Discusses how to find mass-radius relations and mass limits for white dwarfs using simple energy arguments.

Variability

Winget, D.E. (1998). "Asteroseismology of white dwarf stars". Journal of Physics: Condensed Matter. 10 (49): 11247–11261. Bibcode:1998JPCM...1011247W. doi:10.1088/0953-8984/10/49/014.

Magnetic field

Wickramasinghe, D. T.; Ferrario, Lilia (2000). "Magnetism in Isolated and Binary White Dwarfs". Publications of the Astronomical Society of the Pacific. 112 (773): 873–924. Bibcode:2000PASP..112..873W. doi:10.1086/316593.

Frequency

Gibson, B. K.; Flynn, C (2001). "White Dwarfs and Dark Matter". Science. 292 (5525): 2211a.arXiv:astro-ph/0104255. doi:10.1126/science.292.5525.2211a. PMID 11423620. S2CID 14080941.

Observational

Provencal, J. L.; Shipman, H. L.; Hog, Erik; Thejll, P. (1998). "Testing the White Dwarf Mass‐Radius Relation with Hipparcos". The Astrophysical Journal. 494 (2): 759–767. Bibcode:1998ApJ...494..759P. doi:10.1086/305238.
Gates, Evalyn; Gyuk, Geza; Harris, Hugh C.; Subbarao, Mark; Anderson, Scott; Kleinman, S. J.; Liebert, James; Brewington, Howard; et al. (2004). "Discovery of New Ultracool White Dwarfs in the Sloan Digital Sky Survey". The Astrophysical Journal. 612 (2): L129.arXiv:astro-ph/0405566. Bibcode:2004ApJ...612L.129G. doi:10.1086/424568. S2CID 7570539.
McCook, G.P.; Sion, E.M. (eds.). "White Dwarf Catalogue WD". Villanova University.
Dufour, P.; Liebert, J.; Fontaine, G.; Behara, N. (2007). "White dwarf stars with carbon atmospheres". Nature. 450 (7169): 522–4.arXiv:0711.3227. Bibcode:2007Natur.450..522D. doi:10.1038/nature06318. PMID 18033290. S2CID 4398697.

Images

Astronomy Picture of the Day
NGC 2440: Cocoon of a New White Dwarf. Astronomy Picture of the Day (photograph). NASA. 21 February 2010.
Dust and the Helix Nebula. Astronomy Picture of the Day (photograph). NASA. 31 December 2009.
The Helix Nebula from La Silla Observatory. Astronomy Picture of the Day (photograph). NASA. 3 March 2009.
IC 4406: A Seemingly Square Nebula. Astronomy Picture of the Day (photograph). NASA. 27 July 2008.
A Nearby Supernova in Spiral Galaxy M100. Astronomy Picture of the Day (photograph). NASA. 7 March 2006.
White Dwarf Star Spiral. Astronomy Picture of the Day (photograph). NASA. 1 June 2005.

vte

White dwarf
Formation

Chandrasekhar limit PG 1159 star Stellar evolution Hertzsprung–Russell diagram Mira variable

Fate

Black dwarf Type Ia supernova
Candidates Neutron star
Pulsar Magnetar Related links Stellar black hole
Related links Compact star
Quark star Exotic star Extreme helium star Subdwarf B star Helium planet

In binary
systems

Nova
Remnant List Dwarf nova Symbiotic nova Cataclysmic variable star
AM CVn star Polar Intermediate polar X-ray binary
Super soft X-ray source Binary pulsar
X-ray pulsar List Helium flash Carbon detonation

Properties

Pulsating Urca process Electron-degenerate matter Quasi-periodic oscillations

Related

Planetary nebula
List RAMBOs White dwarf luminosity function Timeline of white dwarfs, neutron stars, and supernovae

vte

Neutron star
Types

Radio-quiet Pulsar

Single pulsars

Magnetar
Soft gamma repeater Anomalous X-ray Rotating radio transient

Binary pulsars

Binary X-ray pulsar
X-ray binary X-ray burster List Millisecond Be/X-ray Spin-up

Properties

Blitzar
Fast radio burst Bondi accretion Chandrasekhar limit Gamma-ray burst Glitch Neutronium Neutron-star oscillation Optical Pulsar kick Quasi-periodic oscillation Relativistic Rp-process Starquake Timing noise Tolman–Oppenheimer–Volkoff limit Urca process

Related

Gamma-ray burst progenitors Asteroseismology Compact star
Quark star Exotic star Supernova
Supernova remnant Related links Hypernova Kilonova Neutron star merger Quark-nova White dwarf
Related links Stellar black hole
Related links Radio star Pulsar planet Pulsar wind nebula Thorne–Żytkow object

Discovery

LGM-1 Centaurus X-3 Timeline of white dwarfs, neutron stars, and supernovae

Satellite
investigation

Rossi X-ray Timing Explorer Fermi Gamma-ray Space Telescope Compton Gamma Ray Observatory Chandra X-ray Observatory

Other

X-ray pulsar-based navigation Tempo software program Astropulse The Magnificent Seven

vte

Stars
Formation

Accretion Molecular cloud Bok globule Young stellar object
Protostar Pre-main-sequence Herbig Ae/Be T Tauri FU Orionis Herbig–Haro object Hayashi track Henyey track

Evolution

Main sequence Red-giant branch Horizontal branch
Red clump Asymptotic giant branch
super-AGB Blue loop Protoplanetary nebula Planetary nebula PG1159 Dredge-up OH/IR Instability strip Luminous blue variable Blue straggler Stellar population Supernova Superluminous supernova / Hypernova

Spectral classification

Early Late Main sequence
O B A F G K M Brown dwarf WR OB Subdwarf
O B Subgiant Giant
Blue Red Yellow Bright giant Supergiant
Blue Red Yellow Hypergiant
Yellow Carbon
S CN CH White dwarf Chemically peculiar
Am Ap/Bp HgMn Helium-weak Barium Extreme helium Lambda Boötis Lead Technetium Be
Shell B[e]

Remnants

White dwarf
Helium planet Black dwarf Neutron
Radio-quiet Pulsar
Binary X-ray Magnetar Stellar black hole X-ray binary
Burster

Hypothetical

Blue dwarf Green Black dwarf Exotic
Boson Electroweak Strange Preon Planck Dark Dark-energy Quark Q Black Gravastar Frozen Quasi-star Thorne–Żytkow object Iron Blitzar

Stellar nucleosynthesis

Deuterium burning Lithium burning Proton–proton chain CNO cycle Helium flash Triple-alpha process Alpha process Carbon burning Neon burning Oxygen burning Silicon burning S-process R-process Fusor Nova
Symbiotic Remnant Luminous red nova

Structure

Core Convection zone
Microturbulence Oscillations Radiation zone Atmosphere
Photosphere Starspot Chromosphere Stellar corona Stellar wind
Bubble Bipolar outflow Accretion disk Asteroseismology
Helioseismology Eddington luminosity Kelvin–Helmholtz mechanism

Properties

Designation Dynamics Effective temperature Luminosity Kinematics Magnetic field Absolute magnitude Mass Metallicity Rotation Starlight Variable Photometric system Color index Hertzsprung–Russell diagram Color–color diagram

Star systems

Binary
Contact Common envelope Eclipsing Symbiotic Multiple Cluster
Open Globular Super Planetary system

Earth-centric
observations

Sun
Solar System Sunlight Pole star Circumpolar Constellation Asterism Magnitude
Apparent Extinction Photographic Radial velocity Proper motion Parallax Photometric-standard

Lists

Proper names
Arabic Chinese Extremes Most massive Highest temperature Lowest temperature Largest volume Smallest volume Brightest
Historical Most luminous Nearest
Nearest bright With exoplanets Brown dwarfs White dwarfs Milky Way novae Supernovae
Candidates Remnants Planetary nebulae Timeline of stellar astronomy

Related articles

Substellar object
Brown dwarf Sub-brown dwarf Planet Galactic year Galaxy Guest Gravity Intergalactic Planet-hosting stars Tidal disruption event

Brilliant Proof of a 90 Year Old White Dwarf Theory (V)

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