Supernova

This article is about the astronomical event. For other uses, see Supernova (disambiguation).
SN 1994D (bright spot on the lower left), a type Ia supernova in the NGC 4526 galaxy

A supernova is an astronomical event that occurs during the last stellar evolutionary stages of a massive star's life, whose dramatic and catastrophic destruction is marked by one final titanic explosion. For a short time, this causes the sudden appearance of a 'new' bright star, before slowly fading from sight over several weeks or months.

Only three Milky Way naked-eye supernova events have been observed during the last thousand years, though many have been telescopically seen in other galaxies. The most recent directly observed supernova in the Milky Way was Kepler's Supernova in 1604, but remnants of two more recent supernovae have been found retrospectively.[1] Statistical observations of supernovae in other galaxies suggest they should occur on average about three times every century in the Milky Way, and that any galactic supernova would almost certainly be observable in modern astronomical equipment.[2]

Supernovae are more energetic than novae. In Latin, nova means "new", referring astronomically to what appears to be a temporary new bright star. Adding the prefix "super-" distinguishes supernovae from ordinary novae, which are far less luminous. The word supernova was coined by Walter Baade and Fritz Zwicky in 1931.[3] It is pronounced /ˌspərnvə/ with the plural supernovae /ˌspərnv/ or supernovas (abbreviated SN, plural SNe after "supernovae").

During maximum brightness, the total equivalent radiant energies produced by supernovae may briefly outshine an entire output of a typical galaxy and emit energies equal to that created over the lifetime of any solar-like star.[4] Such extreme catastrophes may also expel much, if not all, of its stellar material away from the star,[5] at velocities up to 30,000 km/s or 10% of the speed of light. This drives an expanding and fast-moving shock wave[6] into the surrounding interstellar medium, and in turn, sweeping up an expanding shell of gas and dust, which is observed as a supernova remnant. Supernovae create, fuse and eject the bulk of the chemical elements produced by nucleosynthesis.[7] Supernovae play a significant role in enriching the interstellar medium with the heavier atomic mass chemical elements.[8] Furthermore, the expanding shock waves from supernova explosions can trigger the formation of new stars.[9][10] Supernova remnants are expected to accelerate a large fraction of galactic primary cosmic rays, but direct evidence for cosmic ray production was found only in a few of them so far.[11] They are also potentially strong galactic sources of gravitational waves.[12]

Theoretical studies of many supernovae indicate that most are triggered by one of two basic mechanisms: the sudden re-ignition of nuclear fusion in a degenerate star or the sudden gravitational collapse of a massive star's core. In the first instance, a degenerate white dwarf may accumulate sufficient material from a binary companion, either through accretion or via a merger, to raise its core temperature enough to trigger runaway nuclear fusion, completely disrupting the star. In the second case, the core of a massive star may undergo sudden gravitational collapse, releasing gravitational potential energy as a supernova. While some observed supernovae are more complex than these two simplified theories, the astrophysical collapse mechanics have been established and accepted by most astronomers for some time.

Due to the wide range of astrophysical consequences of these events, astronomers now deem supernovae research, across the fields of stellar and galactic evolution, as an especially important area for investigation.

Etymology

In Latin, Nova means "new", referring astronomically to what appears to be a temporary new bright star. Adding the prefix "super-" distinguishes supernovae from ordinary novae, which are far less luminous. The word supernova was coined by Walter Baade and Fritz Zwicky in 1931.[3] It is pronounced /ˌspərnvə/ with the plural supernovae /ˌspərnv/ or supernovas (abbreviated SN, plural SNe after "supernovae").

Observation history

The Crab Nebula is a pulsar wind nebula associated with the 1054 supernova
The highlighted passages refer to the Chinese observation of SN 1054

The earliest recorded supernova, SN 185, was viewed by Chinese astronomers in 185 AD, with the brightest recorded supernova being SN 1006, which occurred in 1006 AD and was described in detail by Chinese and Islamic astronomers.[13] The widely observed supernova SN 1054 produced the Crab Nebula. Supernovae SN 1572 and SN 1604, the latest to be observed with the naked eye in the Milky Way galaxy, had notable effects on the development of astronomy in Europe because they were used to argue against the Aristotelian idea that the universe beyond the Moon and planets was static and unchanging.[14] Johannes Kepler began observing SN 1604 at its peak on October 17, 1604, and continued to make estimates of its brightness until it faded from naked eye view a year later.[15] It was the second supernova to be observed in a generation (after SN 1572 seen by Tycho Brahe in Cassiopeia).[16]

Before the development of the telescope, there have only been five supernovae seen in the last millennium. In the perspective of how long a star's lifetime is, its death is very brief. In fact, a star's death may only last a few months. Due to this, a typical human will only experience this rarity, on average, once in their lifetime. This is a microscopic fraction in comparison to the 100 billion stars that compose a galaxy.[17] However, since the use of modern equipment, particularly in this millennium, professional and amateur astronomers have been finding several hundreds of supernovae each year.

The field of supernova discovery has extended to other galaxies, starting with SN 1885A in the Andromeda galaxy. American astronomers Rudolph Minkowski and Fritz Zwicky developed the modern supernova classification scheme beginning in 1941.[18] In the 1960s, astronomers found that the maximum intensities of supernova explosions could be used as standard candles, hence indicators of astronomical distances.[19] Some of the most distant supernovae recently observed appeared dimmer than expected. This supports the view that the expansion of the universe is accelerating.[20] Techniques were developed for reconstructing supernova explosions that have no written records of being observed. The date of the Cassiopeia A supernova event was determined from light echoes off nebulae,[21] while the age of supernova remnant RX J0852.0-4622 was estimated from temperature measurements[22] and the gamma ray emissions from the decay of titanium-44.[23]

The brightest observed supernova, ASASSN-15lh, was detected in June 2015.[24] With a brightness of 570 billion Suns, ASASSN-15lh's peak luminosity was twice that of the previous record holder.[25]

Discovery

Early work on what was originally believed to be simply a new category of novae was performed during the 1930s by two astronomers named Walter Baade and Fritz Zwicky at Mount Wilson Observatory.[26] The name super-novae was first used during 1931 lectures held at Caltech by Baade and Zwicky, then used publicly in 1933 at a meeting of the American Physical Society.[3] By 1938, the hyphen had been lost and the modern name was in use.[27] Because supernovae are relatively rare events within a galaxy, occurring about three times a century in the Milky Way,[1] obtaining a good sample of supernovae to study requires regular monitoring of many galaxies.

Supernovae in other galaxies cannot be predicted with any meaningful accuracy. Normally, when they are discovered, they are already in progress.[28] Most scientific interest in supernovae—as standard candles for measuring distance, for example—require an observation of their peak luminosity. It is therefore important to discover them well before they reach their maximum. Amateur astronomers, who greatly outnumber professional astronomers, have played an important role in finding supernovae, typically by looking at some of the closer galaxies through an optical telescope and comparing them to earlier photographs.[29]

Toward the end of the 20th century astronomers increasingly turned to computer-controlled telescopes and CCDs for hunting supernovae. While such systems are popular with amateurs, there are also professional installations such as the Katzman Automatic Imaging Telescope.[30] Recently the Supernova Early Warning System (SNEWS) project has begun using a network of neutrino detectors to give early warning of a supernova in the Milky Way galaxy.[31][32] Neutrinos are particles that are produced in great quantities by a supernova explosion,[33] and they are not significantly absorbed by the interstellar gas and dust of the galactic disk.

"A star set to explode", the SBW1 nebula surrounds a massive blue supergiant in the Carina Nebula.

Supernova searches fall into two classes: those focused on relatively nearby events and those looking for explosions farther away. Because of the expansion of the universe, the distance to a remote object with a known emission spectrum can be estimated by measuring its Doppler shift (or redshift); on average, more distant objects recede with greater velocity than those nearby, and so have a higher redshift. Thus the search is split between high redshift and low redshift, with the boundary falling around a redshift range of z = 0.1–0.3[34]—where z is a dimensionless measure of the spectrum's frequency shift.

High redshift searches for supernovae usually involve the observation of supernova light curves. These are useful for standard or calibrated candles to generate Hubble diagrams and make cosmological predictions. Supernova spectroscopy, used to study the physics and environments of supernovae, is more practical at low than at high redshift.[35][36] Low redshift observations also anchor the low-distance end of the Hubble curve, which is a plot of distance versus redshift for visible galaxies.[37][38] (See also Hubble's law).

Naming convention

Multiwavelength X-ray, infrared, and optical compilation image of Kepler's supernova remnant, SN 1604.

Supernova discoveries are reported to the International Astronomical Union's Central Bureau for Astronomical Telegrams, which sends out a circular with the name it assigns to that supernova. The name is the marker SN followed by the year of discovery, suffixed with a one or two-letter designation. The first 26 supernovae of the year are designated with a capital letter from A to Z. Afterward pairs of lower-case letters are used: aa, ab, and so on. Hence, for example, SN 2003C designates the third supernova reported in the year 2003.[39] The last supernova of 2005 was SN 2005nc, indicating that it was the 367th[nb 1] supernova found in 2005. Since 2000, professional and amateur astronomers have been finding several hundreds of supernovae each year (572 in 2007, 261 in 2008, 390 in 2009; 231 in 2013).[40][41]

Historical supernovae are known simply by the year they occurred: SN 185, SN 1006, SN 1054, SN 1572 (called Tycho's Nova) and SN 1604 (Kepler's Star). Since 1885 the additional letter notation has been used, even if there was only one supernova discovered that year (e.g. SN 1885A, SN 1907A, etc.) — this last happened with SN 1947A. SN, for SuperNova, is a standard prefix. Until 1987, two-letter designations were rarely needed; since 1988, however, they have been needed every year.

Classification

Artist's impression of supernova 1993J.[42]

As part of the attempt to understand supernovae, astronomers have classified them according to their light curves and the absorption lines of different chemical elements that appear in their spectra. The first element for division is the presence or absence of a line caused by hydrogen. If a supernova's spectrum contains lines of hydrogen (known as the Balmer series in the visual portion of the spectrum) it is classified Type II; otherwise it is Type I. In each of these two types there are subdivisions according to the presence of lines from other elements or the shape of the light curve (a graph of the supernova's apparent magnitude as a function of time).[43][44]

Supernova taxonomy[43][44]
Type I
No hydrogen
Type Ia
Presents a singly ionized silicon (Si II) line at 615.0 nm (nanometers), near peak light
Thermal runaway
Type Ib/c
Weak or no silicon absorption feature
Type Ib
Shows a non-ionized helium (He I) line at 587.6 nm
Core collapse
Type Ic
Weak or no helium
Type II
Shows hydrogen
Type II-P/L/N
Type II spectrum throughout
Type II-P/L
No narrow lines
Type II-P
Reaches a "plateau" in its light curve
Type II-L
Displays a "linear" decrease in its light curve (linear in magnitude versus time).[45]
Type IIn
Some narrow lines
Type IIb
Spectrum changes to become like Type Ib

Type I

Type I supernovae are subdivided on the basis of their spectra, with type Ia showing a strong ionised silicon absorption line. Type I supernovae without this strong line are classified as Type Ib and Ic, with Type Ib showing strong neutral helium lines and Type Ic lacking them. The light curves are all similar, although Type Ia are generally brighter at peak luminosity, but the light curve is not important for classification of Type I supernovae.

A small number of Type Ia supernovae exhibit unusual features such as non-standard luminosity or broadened light curves, and these are typically classified by referring to the earliest example showing similar features. For example, the sub-luminous SN 2008ha is often referred to as SN 2002cx-like or class Ia-2002cx.

Type II

Light curves are used to classify type II-P and type II-L supernovae

The supernovae of Type II can also be sub-divided based on their spectra. While most Type II supernovae show very broad emission lines which indicate expansion velocities of many thousands of kilometres per second, some, such as SN 2005gl, have relatively narrow features in their spectra. These are called Type IIn, where the 'n' stands for 'narrow'.

A few supernovae, such as SN 1987K and SN 1993J, appear to change types: they show lines of hydrogen at early times, but, over a period of weeks to months, become dominated by lines of helium. The term "Type IIb" is used to describe the combination of features normally associated with Types II and Ib.[44]

Type II supernovae with normal spectra dominated by broad hydrogen lines that remain for the life of the decline are classified on the basis of their light curves. The most common type shows a distinctive "plateau" in the light curve shortly after peak brightness where the visual luminosity stays relatively constant for several months before the decline resumes. These are called type II-P referring to the plateau. Less common are type II-L supernovae that lack a distinct plateau. The "L" signifies "linear" although the light curve is not actually a straight line.

Supernovae that do not fit into the normal classifications are designated peculiar, or 'pec'.[44]

Types III, IV, and V

Fritz Zwicky defined additional supernovae types, although based on a very few examples that didn't cleanly fit the parameters for a Type I or Type II supernova. SN 1961i in NGC 4303 was the prototype and only member of the Type III supernova class, noted for its broad light curve maximum and broad hydrogen Balmer lines that were slow to develop in the spectrum. SN 1961f in NGC 3003 was the prototype and only member of the Type IV class, with a light curve similar to a Type II-P supernova, with hydrogen absorption lines but weak hydrogen emission lines. The Type V class was coined for SN 1961V in NGC 1058, an unusual faint supernova or supernova imposter with a slow rise to brightness, a maximum lasting many months, and an unusual emission spectrum. The similarity of SN 1961V to the Eta Carinae Great Outburst was noted.[46] Supernovae in M101 (1909) and M83 (1923 and 1957) were also suggested as possible type IV or type V supernovae.[47]

These types would now all be treated as peculiar Type II supernovae, of which many more examples have been discovered, although it is still debated whether SN 1961V was a true supernova following an LBV outburst or an imposter.[45]

Current models

Sequence shows the rapid brightening and slower fading of a supernova explosion in the galaxy NGC 1365[48]

The type codes, described above given to supernovae, are taxonomic in nature: the type number describes the light observed from the supernova, not necessarily its cause. For example, type Ia supernovae are produced by runaway fusion ignited on degenerate white dwarf progenitors while the spectrally similar type Ib/c are produced from massive Wolf-Rayet progenitors by core collapse. The following summarizes what is currently believed to be the most plausible explanations for supernovae.

Thermal runaway

Main article: Type Ia supernova
Formation of a type Ia supernova

A white dwarf star may accumulate sufficient material from a stellar companion to raise its core temperature enough to ignite carbon fusion, at which point it undergoes runaway nuclear fusion, completely disrupting it. There are three avenues by which this detonation is theorized to happen: stable accretion of material from a companion, the collision of two white dwarfs, or accretion that causes ignition in a shell that then ignites. The dominant mechanism by which Type Ia supernovae are produced remains unclear.[49] Despite this uncertainty in how Type Ia supernovae are produced, Type Ia supernovae have very uniform properties, and are useful standard candles over intergalactic distances. Some calibrations are required to compensate for the gradual change in properties or different frequencies of abnormal luminosity supernovae at high red shift, and for small variations in brightness identified by light curve shape or spectrum.[50][51]

Normal Type Ia

There are several means by which a supernova of this type can form, but they share a common underlying mechanism. If a carbon-oxygen[nb 2] white dwarf accreted enough matter to reach the Chandrasekhar limit of about 1.44 solar masses (M)[52] (for a non-rotating star), it would no longer be able to support the bulk of its mass through electron degeneracy pressure[53][54] and would begin to collapse. However, the current view is that this limit is not normally attained; increasing temperature and density inside the core ignite carbon fusion as the star approaches the limit (to within about 1%[55]), before collapse is initiated.[52]

Within a few seconds, a substantial fraction of the matter in the white dwarf undergoes nuclear fusion, releasing enough energy (1–2×1044 J)[56] to unbind the star in a supernova explosion.[57] An outwardly expanding shock wave is generated, with matter reaching velocities on the order of 5,000–20,000 km/s, or roughly 3% of the speed of light. There is also a significant increase in luminosity, reaching an absolute magnitude of −19.3 (or 5 billion times brighter than the Sun), with little variation.[58]

The model for the formation of this category of supernova is a closed binary star system. The larger of the two stars is the first to evolve off the main sequence, and it expands to form a red giant. The two stars now share a common envelope, causing their mutual orbit to shrink. The giant star then sheds most of its envelope, losing mass until it can no longer continue nuclear fusion. At this point it becomes a white dwarf star, composed primarily of carbon and oxygen.[59] Eventually the secondary star also evolves off the main sequence to form a red giant. Matter from the giant is accreted by the white dwarf, causing the latter to increase in mass. Despite widespread acceptance of the basic model, the exact details of initiation and of the heavy elements produced in the explosion are still unclear.

Type Ia supernovae follow a characteristic light curve—the graph of luminosity as a function of time—after the explosion. This luminosity is generated by the radioactive decay of nickel-56 through cobalt-56 to iron-56.[58] The peak luminosity of the light curve is extremely consistent across normal Type Ia supernovae, having a maximum absolute magnitude of about −19.3. This allows them to be used as a secondary[60] standard candle to measure the distance to their host galaxies.[61]

Non-standard Type Ia

Another model for the formation of a Type Ia explosion involves the merger of two white dwarf stars, with the combined mass momentarily exceeding the Chandrasekhar limit.[62] There is much variation in this type of explosion,[63] and in many cases there may be no supernova at all, but it is expected that they will have a broader and less luminous light curve than the more normal Type Ia explosions.

Abnormally bright Type Ia supernovae are expected when the white dwarf already has a mass higher than the Chandrasekhar limit,[64] possibly enhanced further by asymmetry,[65] but the ejected material will have less than normal kinetic energy.

There is no formal sub-classification for the non-standard Type Ia supernovae. It has been proposed that a group of sub-luminous supernovae that occur when helium accretes onto a white dwarf should be classified as type Iax.[66][67] This type of supernova may not always completely destroy the white dwarf progenitor and could leave behind a zombie star.[68]

One specific type of non-standard Type Ia supernova develops hydrogen, and other, emission lines and gives the appearance of mixture between a normal Type Ia and a Type IIn supernova. Examples are SN 2002ic and SN 2005gj. These supernova have been dubbed Type Ia/IIn, Type Ian, Type IIa and Type IIan.[69]

Core collapse

Supernova types by initial mass-metallicity
The onion-like layers of a massive, evolved star just prior to core collapse (Not to scale)

Very massive stars can undergo core collapse when nuclear fusion suddenly becomes unable to sustain the core against its own gravity; this is the cause of all types of supernova except type Ia. The collapse may cause violent expulsion of the outer layers of the star resulting in a supernova, or the release of gravitational potential energy may be insufficient and the star may collapse into a black hole or neutron star with little radiated energy.

Core collapse can be caused by several different mechanisms: electron capture; exceeding the Chandrasekhar limit; pair-instability; or photodisintegration.[5][70] When a massive star develops an iron core larger than the Chandrasekhar mass it will no longer be able to support itself by electron degeneracy pressure and will collapse further to a neutron star or black hole. Electron capture by magnesium in a degenerate O/Ne/Mg core causes gravitational collapse followed by explosive oxygen fusion, with very similar results. Electron-positron pair production in a large post-helium burning core removes thermodynamic support and causes initial collapse followed by runaway fusion, resulting in a pair-instability supernova. A sufficiently large and hot stellar core may generate gamma-rays energetic enough to initiate photodisintegration directly, which will cause a complete collapse of the core.

The table below lists the known reasons for core collapse in massive stars, the types of star that they occur in, their associated supernova type, and the remnant produced. The metallicity is the proportion of elements other than hydrogen or helium, as compared to the Sun. The initial mass is the mass of the star prior to the supernova event, given in multiples of the Sun's mass, although the mass at the time of the supernova may be much lower.

Type IIn supernovae are not listed in the table. They can potentially be produced by various types of core collapse in different progenitor stars, possibly even by type Ia white dwarf ignitions, although it seems that most will be from iron core collapse in luminous supergiants or hypergiants (including LBVs). The narrow spectral lines for which they are named occur because the supernova is expanding into a small dense cloud of circumstellar material.[71] It appears that a significant proportion of supposed type IIn supernovae are actually supernova imposters, massive eruptions of LBV-like stars similar to the Great Eruption of Eta Carinae. In these events, material previously ejected from the star creates the narrow absorption lines and causes a shock wave through interaction with the newly ejected material.[72]

Core collapse scenarios by mass and metallicity[5]
Cause of collapse Progenitor star approximate initial mass Supernova Type Remnant
Electron capture in a degenerate O+Ne+Mg core 8–10 Faint II-P Neutron star
Iron core collapse 10–25 Faint II-P Neutron star
25–40 with low or solar metallicity Normal II-P Black hole after fallback of material onto an initial neutron star
25–40 with very high metallicity II-L or II-b Neutron star
40–90 with low metallicity None Black hole
≥40 with near-solar metallicity Faint Ib/c, or hypernova with gamma-ray burst (GRB) Black hole after fallback of material onto an initial neutron star
≥40 with very high metallicity Ib/c Neutron star
≥90 with low metallicity None, possible GRB Black hole
Pair instability 140–250 with low metallicity II-P, sometimes a hypernova, possible GRB No remnant
Photodisintegration ≥250 with low metallicity None (or luminous supernova?), possible GRB Massive black hole
Remnants of single massive stars
Within a massive, evolved star (a) the onion-layered shells of elements undergo fusion, forming an iron core (b) that reaches Chandrasekhar-mass and starts to collapse. The inner part of the core is compressed into neutrons (c), causing infalling material to bounce (d) and form an outward-propagating shock front (red). The shock starts to stall (e), but it is re-invigorated by a process that may include neutrino interaction. The surrounding material is blasted away (f), leaving only a degenerate remnant.

When a stellar core is no longer supported against gravity it collapses in on itself with velocities reaching 70,000 km/s (0.23c),[73] resulting in a rapid increase in temperature and density. What follows next depends on the mass and structure of the collapsing core, with low mass degenerate cores forming neutron stars, higher mass degenerate cores mostly collapsing completely to black holes, and non-degenerate cores undergoing runaway fusion.

The initial collapse of degenerate cores is accelerated by beta decay, photodisintegration and electron capture, which causes a burst of electron neutrinos. As the density increases, neutrino emission is cut off as they become trapped in the core. The inner core eventually reaches typically 30 km diameter[74] and a density comparable to that of an atomic nucleus, and neutron degeneracy pressure tries to halt the collapse. If the core mass is more than about 15 M then neutron degeneracy is insufficient to stop the collapse and a black hole forms directly with no supernova explosion.

In lower mass cores the collapse is stopped and the newly formed neutron core has an initial temperature of about 100 billion kelvin, 6000 times the temperature of the sun's core.[75] At this temperature, neutrino-antineutrino pairs of all flavors are efficiently formed by thermal emission. These thermal neutrinos are several times more abundant than the electron-capture neutrinos.[76] About 1046 joules, approximately 10% of the star's rest mass, is converted into a ten-second burst of neutrinos which is the main output of the event.[74][77] The suddenly halted core collapse rebounds and produces a shock wave that stalls within milliseconds[78] in the outer core as energy is lost through the dissociation of heavy elements. A process that is not clearly understood is necessary to allow the outer layers of the core to reabsorb around 1044 joules[77] (1 foe) from the neutrino pulse, producing the visible explosion, although there are also other theories on how to power the explosion.[74]

Some material from the outer envelope falls back onto the neutron star, and for cores beyond about 8 M there is sufficient fallback to form a black hole. This fallback will reduce the kinetic energy of the explosion and the mass of expelled radioactive material, but in some situations it may also generate relativistic jets that result in a gamma-ray burst or an exceptionally luminous supernova.

Collapse of massive non-degenerate cores will ignite further fusion. When the core collapse is initiated by pair instability, oxygen fusion begins and the collapse may be halted. For core masses of 40–60 M, the collapse halts and the star remains intact, but core collapse will occur again when a larger core has formed. For cores of around 60–130 M, the fusion of oxygen and heavier elements is so energetic that the entire star is disrupted, causing a supernova. At the upper end of the mass range, the supernova is unusually luminous and extremely long-lived due to many solar masses of ejected 56Ni. For even larger core masses, the core temperature becomes high enough to allow photodisintegration and the core collapses completely into a black hole.[79]

Type II

Main article: Type II supernova
The atypical subluminous type II SN 1997D

Stars with initial masses less than about eight times the sun never develop a core large enough to collapse and they eventually lose their atmospheres to become white dwarfs. Stars with at least 9 M (possibly as much as 12 M[80]) evolve in a complex fashion, progressively burning heavier elements at hotter temperatures in their cores.[74][81] The star becomes layered like an onion, with the burning of more easily fused elements occurring in larger shells.[5][82] Although popularly described as an onion with an iron core, the least massive supernova progenitors only have oxygen-neon(-magnesium) cores. These super AGB stars may form the majority of core collapse supernovae, although less luminous and so less commonly observed than those from more massive progenitors.[80]

If core collapse occurs during a supergiant phase when the star still has a hydrogen envelope, the result is a Type II supernova. The rate of mass loss for luminous stars depends on the metallicity and luminosity. Extremely luminous stars at near solar metallicity will lose all their hydrogen before they reach core collapse and so will not form a type II supernova. At low metallicity, all stars will reach core collapse with a hydrogen envelope but sufficiently massive stars collapse directly to a black hole without producing a visible supernova.

Stars with an initial mass up to about 90 times the sun, or a little less at high metallicity, are expected to result in a type II-P supernova which is the most commonly observed type. At moderate to high metallicity, stars near the upper end of that mass range will have lost most of their hydrogen when core collapse occurs and the result will be a Type II-L supernova. At very low metallicity, stars of around 140–250 M will reach core collapse by pair instability while they still have a hydrogen atmosphere and an oxygen core and the result will be a supernova with Type II characteristics but a very large mass of ejected 56Ni and high luminosity.

Type Ib and Ic

SN 2008D, a Type Ib[83] supernova, shown in X-ray (left) and visible light (right) at the far upper end of the galaxy[84]

These supernovae, like those of Type II, are massive stars that undergo core collapse. However the stars which become Types Ib and Ic supernovae have lost most of their outer (hydrogen) envelopes due to strong stellar winds or else from interaction with a companion.[85] These stars are known as Wolf-Rayet stars, and they occur at moderate to high metallicity where continuum driven winds cause sufficiently high mass loss rates. Observations of type Ib/c supernova do not match the observed or expected occurrence of Wolf Rayet stars and alternate explanations for this type of core collapse supernova involve stars stripped of their hydrogen by binary interactions. Binary models provide a better match for the observed supernovae, with the proviso that no suitable binary helium stars have ever been observed.[86] Since a supernova explosion can occur whenever the mass of the star at the time of core collapse is low enough not to cause complete fallback to a black hole, any massive star may result in a supernova if it loses enough mass before core collapse occurs.

Type Ib supernovae are the more common and result from Wolf-Rayet stars of type WC which still have helium in their atmospheres. For a narrow range of masses, stars evolve further before reaching core collapse to become WO stars with very little helium remaining and these are the progenitors of type Ic supernovae.

A few percent of the Type Ic supernovae are associated with gamma-ray bursts (GRB), though it is also believed that any hydrogen-stripped Type Ib or Ic supernova could produce a GRB, depending on the geometry of the explosion.[87] The mechanism for producing this type of GRB is the jets produced by the magnetic field of the rapidly spinning magnetar formed at the collapsing core of the star. The jets would also transfer energy into the expanding outer shell of the explosion to produce a super-luminous supernova.[88][89]

Ultra-stripped supernovae occur when the exploding star has been stripped (almost) all the way to the metal core, via mass transfer in a close binary.[90] As a result, very little material is ejected from the exploding star (~0.1 MSun). In the most extreme cases, ultra-stripped supernovae can occur in naked metal cores, barely above the Chandrasekhar mass limit. SN 2005ek[91] might be an observational example of an ultra-stripped supernova, giving rise to a relatively dim and fast decaying light curve. The nature of ultra-stripped supernovae can be both iron core-collapse and electron capture supernovae, depending on the mass of the collapsing core.

Failed

The core collapse of some massive stars may not result in a visible supernova. The main model for this is a sufficiently massive core that the explosion is insufficient to reverse the infall of the outer layers onto a black hole. These events are difficult to detect, but large surveys have detected possible candidates.[92][93]

Light curves

Comparative supernova type light curves

A historic puzzle concerned the source of energy that can maintain the optical supernova glow for months. Although the energy that disrupts each type of supernovae is delivered promptly, the light curves are mostly dominated by subsequent radioactive heating of the rapidly expanding ejecta. Some have considered rotational energy from the central pulsar. The ejecta gases would dim quickly without some energy input to keep it hot. The intensely radioactive nature of the ejecta gases, which is now known to be correct for most supernovae, was first calculated on sound nucleosynthesis grounds in the late 1960s.[94] It was not until SN 1987A that direct observation of gamma-ray lines unambiguously identified the major radioactive nuclei.[95]

It is now known by direct observation that much of the light curve (the graph of luminosity as a function of time) after the explosion of a Type II Supernova such as SN 1987A is provided its energy by those predicted radioactive decays. Although the luminous emission consists of optical photons, it is the radioactive power absorbed by the ejected gases that keeps the remnant hot enough to radiate light. The radioactive decay of 56Ni through its daughters 56Co to 56Fe produces gamma-ray photons , primarily of 847keV and 1238keV, that are absorbed and dominate the heating and thus the luminosity of the ejecta at intermediate times (several weeks) to late times (several months).[96] Energy for the peak of the light curve of SN1987A was provided by the decay of 56Ni to 56Co (half life 6 days) while energy for the later light curve in particular fit very closely with the 77.3 day half-life of 56Co decaying to 56Fe. Later measurements by space gamma-ray telescopes of the small fraction of the 56Co and 57Co gamma rays that escaped the SN 1987A remnant without absorption confirmed earlier predictions that those two radioactive nuclei were the power sources.[95]

The visual light curves of the different supernova types all depend at late times on radioactive heating, but they vary in shape and amplitude on the underlying mechanisms of the explosion, the way that visible radiation is produced, the epoch of its observation, and the transparency of the ejected material. The light curves can be significantly different at other wavelengths. For example, at ultraviolet wavelengths there is an early extremely luminous peak lasting only a few hours corresponding to the breakout of the shock launched by the initial explosion, but that breakout is hardly detectable optically.

The light curves for type Ia are mostly very uniform, with a consistent maximum absolute magnitude and a relatively steep decline in luminosity. Their optical energy output is driven by radioactive decay of nickel-56 (half life 6 days), which then decays to radioactive cobalt-56 (half life 77 days). These radioisotopes from material ejected in the explosion excite surrounding material to incandescence. Studies of cosmology today rely on 56Ni radioactivity providing the energy for the optical brightness of supernovae of Type Ia, which are the "standard candles" of cosmology but whose diagnostic 847keV and 1238keV gamma rays were first detected only in 2014.[97] The initial phases of the light curve decline steeply as the effective size of the photosphere decreases and trapped electromagnetic radiation is depleted. The light curve continues to decline in the B band while it may show a small shoulder in the visual at about 40 days, but this is only a hint of a secondary maximum that occurs in the infra-red as certain ionised heavy elements recombine to produce infra-red radiation and the ejecta become transparent to it. The visual light curve continues to decline at a rate slightly greater than the decay rate of the radioactive cobalt (which has the longer half life and controls the later curve), because the ejected material becomes more diffuse and less able to convert the high energy radiation into visual radiation. After several months, the light curve changes its decline rate again as positron emission becomes dominant from the remaining cobalt-56, although this portion of the light curve has been little-studied.

Type Ib and Ic light curves are basically similar to type Ia although with a lower average peak luminosity. The visual light output is again due to radioactive decay being converted into visual radiation, but there is a much lower mass of nickel-56 produced in these types of explosion. The peak luminosity varies considerably and there are even occasional type Ib/c supernovae orders of magnitude more and less luminous than the norm. The most luminous type Ic supernovae are referred to as hypernovae and tend to have broadened light curves in addition to the increased peak luminosity. The source of the extra energy is thought to be relativistic jets driven by the formation of a rotating black hole, which also produce gamma-ray bursts.

The light curves for type II supernovae are characterised by a much slower decline than type I, on the order of 0.05 magnitudes per day,[98] excluding the plateau phase. The visual light output is dominated by kinetic energy rather than radioactive decay for several months, due primarily to the existence of hydrogen in the ejecta from the atmosphere of the supergiant progenitor star. In the initial explosion this hydrogen becomes heated and ionised. The majority of type II supernovae show a prolonged plateau in their light curves as this hydrogen recombines, emitting visible light and becoming more transparent. This is then followed by a declining light curve driven by radioactive decay although slower than in type I supernovae, due to the efficiency of conversion into light by all the hydrogen.[45]

In type II-L the plateau is absent because the progenitor had relatively little hydrogen left in its atmosphere, sufficient to appear in the spectrum but insufficient to produce a noticeable plateau in the light output. In type IIb supernovae the hydrogen atmosphere of the progenitor is so depleted (thought to be due to tidal stripping by a companion star) that the light curve is closer to a type I supernova and the hydrogen even disappears from the spectrum after several weeks.[45]

Type IIn supernovae are characterised by additional narrow spectral lines produced in a dense shell of circumstellar material. Their light curves are generally very broad and extended, occasionally also extremely luminous and referred to as a hypernova. These light curves are produced by the highly efficient conversion of kinetic energy of the ejecta into electromagnetic radiation by interaction with the dense shell of material. This only occurs when the material is sufficiently dense and compact, indicating that it has been produced by the progenitor star itself only shortly before the supernova occurs.

Large numbers of supernovae have been catalogued and classified to provide distance candles and test models. Average characteristics vary somewhat with distance and type of host galaxy, but can broadly be specified for each supernova type.

Physical properties of supernovae by type[99][100]
Typea Average peak absolute magnitudeb Approximate energy (foe)c Days to peak luminosity Days from peak to 10% luminosity
Ia −19 1 approx. 19 around 60
Ib/c (faint) around −15 0.1 15–25 unknown
Ib around −17 1 15–25 40–100
Ic around −16 1 15–25 40–100
Ic (bright) to −22 above 5 roughly 25 roughly 100
II-b around −17 1 around 20 around 100
II-L around −17 1 around 13 around 150
II-P (faint) around −14 0.1 roughly 15 unknown
II-P around −16 1 around 15 Plateau then around 50
IInd around −17 1 12–30 or more 50–150
IIn (bright) to −22 above 5 above 50 above 100

Notes:

  • a. ^ Faint types may be a distinct sub-class. Bright types may be a continuum from slightly over-luminous to hypernovae.
  • b. ^ These magnitudes are measured in the R band. Measurements in V or B bands are common and will be around half a magnitude brighter for supernovae.
  • c. ^ Order of magnitude kinetic energy. Total electromagnetic radiated energy is usually lower, (theoretical) neutrino energy much higher.
  • d. ^ Probably a heterogeneous group, any of the other types embedded in nebulosity.

Asymmetry

The pulsar in the Crab nebula is travelling at 375 km/s relative to the nebula.[101]

A long-standing puzzle surrounding Type II supernovae is why the compact object remaining after the explosion is given a large velocity away from the epicentre;[102] pulsars, and thus neutron stars, are observed to have high velocities, and black holes presumably do as well, although they are far harder to observe in isolation. The initial impetus can be substantial, propelling an object of more than a solar mass at a velocity of 500 km/s or greater. This indicates an asymmetry in the explosion, but the mechanism by which momentum is transferred to the compact object remains a puzzle. Proposed explanations for this kick include convection in the collapsing star and jet production during neutron star formation.

One possible explanation for the asymmetry in the explosion is large-scale convection above the core. The convection can create variations in the local abundances of elements, resulting in uneven nuclear burning during the collapse, bounce and resulting explosion.[103]

Another possible explanation is that accretion of gas onto the central neutron star can create a disk that drives highly directional jets, propelling matter at a high velocity out of the star, and driving transverse shocks that completely disrupt the star. These jets might play a crucial role in the resulting supernova explosion.[104][105] (A similar model is now favored for explaining long gamma-ray bursts.)

Initial asymmetries have also been confirmed in Type Ia supernova explosions through observation. This result may mean that the initial luminosity of this type of supernova depends on the viewing angle. However, the explosion becomes more symmetrical with the passage of time. Early asymmetries are detectable by measuring the polarization of the emitted light.[106]

Energy output

The radioactive decays of nickel-56 and cobalt-56 that produce a supernova visible light curve

Although we are used to thinking of supernovae primarily as luminous visible events, the electromagnetic radiation they release is almost a minor side-effect of the explosion. Particularly in the case of core collapse supernovae, the emitted electromagnetic radiation is a tiny fraction of the total energy released during the event.

There is a fundamental difference between the balance of energy production in the different types of supernova. In type Ia white dwarf detonations, most of the energy is directed into heavy element synthesis and the kinetic energy of the ejecta. In core collapse supernovae, the vast majority of the energy is directed into neutrino emission, and while some of this apparently powers the following main explosion 99%+ of the neutrinos escape the star in the first few minutes following the start of the collapse.

Type Ia supernovae derive their energy from a runaway nuclear fusion of a carbon-oxygen white dwarf. The details of the energetics are still not fully understood, but the end result is the ejection of the entire mass of the original star at high kinetic energy. Around half a solar mass of that mass is 56Ni generated from silicon burning. 56Ni is radioactive and decays into 56Co by beta plus decay (with a half life of six days) and gamma rays. 56Co itself decays by the beta plus (an anti-electron) path with a half life of 77 days into stable 56Fe. These two processes are responsible for the electromagnetic radiation from type Ia supernovae. In combination with the changing transparency of the ejected material, they produce the rapidly declining light curve.[107]

Core collapse supernovae are on average visually fainter than type Ia supernovae, but the total energy released is far higher. In the case of supernovae, the energy of gravitational potential energy is converted into kinetic energy that compresses and collapses the core, initially producing electron neutrinos from disintegrating nucleons, followed by all flavours of thermal neutrinos from the super-heated neutron star core. Around 1% of these neutrinos are thought to deposit sufficient energy into the outer layers of the star to drive the resulting explosion, but again the details cannot be reproduced exactly in current models. Kinetic energies and nickel yields are somewhat lower than type Ia supernovae, hence the lower visual luminosity of supernovae, but energy from the de-ionisation of the many solar masses of remaining hydrogen can contribute to a much slower decline in luminosity and produce the plateau phase seen in the majority of core collapse supernovae.

Energetics of supernovae
Supernova Approximate total energy
1044 joules (foe)c
Ejected Ni
(solar masses)
Neutrino energy
(foe)
Kinetic energy
(foe)
Electromagnetic radiation
(foe)
Type Ia[107][108][109] 1.5 0.4 – 0.8 0.1 1.3 – 1.4 ~0.01
Core collapse[110][111] 100 (0.01) – 1 100 1 0.001 – 0.01
Hypernova 100 ~1 1–100 1–100 ~0.1
Pair instability[79] 5–100 0.5 – 50 low? 1–100 0.01 – 0.1

In some core collapse supernovae, fallback onto a black hole drives relativistic jets which may produce a brief energetic and directional burst of gamma rays and also transfers substantial further energy into the ejected material. This is one scenario for producing high luminosity supernovae and is thought to be the cause of type Ic hypernovae and long duration gamma-ray bursts. If the relativistic jets are too brief and fail to penetrate the stellar envelope then a low luminosity gamma-ray burst may be produced and the supernova may be sub-luminous.

When a supernova occurs inside a small dense cloud of circumstellar material, it will produce a shock wave that can efficiently convert a high fraction of the kinetic energy into electromagnetic radiation. Even though the initial explosion energy was entirely normal the resulting supernova will have high luminosity and extended duration since it does not rely on exponential radioactive decay. This type of event may cause type IIn hypernovae.

Although pair-instability supernovae are core collapse supernovae with spectra and light curves similar to type II-P, the nature of that explosion following core collapse is more like that of a giant type Ia with runaway fusion of carbon, oxygen, and silicon. The total energy released by the highest mass events is comparable to other core collapse supernovae but neutrino production is thought to be very low, hence the kinetic and electromagnetic energy released is very high. The cores of these stars are much larger than any white dwarf and the amount of radioactive nickel and other heavy elements ejected from their cores can be orders of magnitude higher, with consequently high visual luminosity.

Progenitor

Shown in this sped-up artist's impression, is a collection of distant galaxies, the occasional supernova can be seen. Each of these exploding stars briefly rivals the brightness of its host galaxy.

The supernova classification type is closely tied to the type of star at the time of the explosion. The occurrence of each type of supernova depends dramatically on the metallicity and hence the age of the host galaxy.

Type Ia supernovae are produced from white dwarf stars in binary systems and occur in all galaxy types. Core collapse supernovae are only found in galaxies undergoing current or very recent star formation, since they result from short-lived massive stars. They are most commonly found in type Sc spirals, but also in the arms of other spiral galaxies and in irregular galaxies, especially starburst galaxies.

Type Ib/c and II-L, and possibly most type IIn, supernovae are only thought to be produced from stars having near-solar metallicity levels that result in high mass loss from massive stars, hence they are less common in older more distant galaxies. The table shows the expected progenitor for the main types of core collapse supernova, and the approximate proportions of each in the local neighbourhood.

Fraction of core collapse supernovae types by progenitor[86]
Type Progenitor star Fraction
Ib WC Wolf-Rayet 9%
Ic WO Wolf-Rayet 17%
II-P Supergiant 55%
II-L Supergiant with a depleted hydrogen shell 3.0%
IIn Supergiant in a dense cloud of expelled material (such as LBV) 2.4%
IIb Supergiant with highly depleted hydrogen (stripped by companion?) 12%
IIpec Blue supergiant? 1.0%

There are a number of difficulties reconciling modelled and observed stellar evolution leading up to core collapse supernovae. Red supergiants are the expected progenitors for the vast majority of core collapse supernovae, and these have been observed but only at relatively low masses and luminosities, below about 18 M and 100,000 L respectively. Most progenitors of type II supernovae are not detected and must be considerably fainter, and presumably less massive. It is now proposed that higher mass red supergiants do not explode as supernovae, but instead evolve back towards hotter temperatures. Several progenitors of type IIb supernovae have been confirmed, and these were K and G supergiants, plus one A supergiant.[112] Yellow hypergiants or LBVs are proposed progenitors for type IIb supernovae, and almost all type IIb supernovae near enough to observe have shown such progenitors.[113][114]

Until just a few decades ago, hot supergiants were not considered likely to explode, but observations have shown otherwise. Blue supergiants form an unexpectedly high proportion of confirmed supernova progenitors, partly due to their high luminosity and easy detection, while not a single Wolf-Rayet progenitor has yet been clearly identified.[112][115] Models have had difficulty showing how blue supergiants lose enough mass to reach supernova without progressing to a different evolutionary stage. One study has shown a possible route for low-luminosity post-red supergiant luminous blue variables to collapse, most likely as a type IIn supernova.[116]

The expected progenitors of type Ib supernovae, luminous WC stars, are not observed at all. Instead WC stars are found at lower luminosities, apparently post-red supergiant stars. WO stars are extremely rare and visually relatively faint, so it is difficult to say whether such progenitors are missing or just yet to be observed. Very luminous progenitors, despite numerous supernovae being observed near enough that such progenitors would have been clearly imaged.[117] Several examples of hot luminous progenitors of type IIn supernovae have been detected: SN 2005gy and SN 2010jl were both apparently massive luminous stars, but are very distant; and SN 2009ip had a highly luminous progenitor likely to have been an LBV, but is a peculiar supernova whose exact nature is disupted.[112]

Interstellar impact

Source of heavy elements

Supernovae are the major source of elements heavier than oxygen.[118] These elements are produced by nuclear fusion for nuclei up to 34S, by silicon photodisintegration rearrangement and quasiequilibrium (see Supernova nucleosynthesis) during silicon burning for nuclei between 36Ar and 56Ni, and by rapid captures of neutrons during the supernova explosion for elements heavier than iron. Nucleosynthesis during silicon burning yields nuclei roughly 1000-100,000 times more abundant than the r-process isotopes heavier than iron.[119] Supernovae are the most likely, although not undisputed, candidate sites for the r-process, which is the rapid capture of neutrons that occurs at high temperature and high density of neutrons. Those reactions produce highly unstable nuclei that are rich in neutrons and that rapidly beta decay into more stable forms. The r-process produces about half of all the heavier isotopes of the elements beyond iron, including plutonium and uranium.[120] The only other major competing process for producing elements heavier than iron is the s-process in large, old, red-giant AGB stars, which produces these elements slowly over longer epochs and which cannot produce elements heavier than lead.[121]

Role in stellar evolution

Main article: Supernova remnant

The remnant of a supernova explosion consists of a compact object and a rapidly expanding shock wave of material. This cloud of material sweeps up the surrounding interstellar medium during a free expansion phase, which can last for up to two centuries. The wave then gradually undergoes a period of adiabatic expansion, and will slowly cool and mix with the surrounding interstellar medium over a period of about 10,000 years.[122]

Supernova remnant N 63A lies within a clumpy region of gas and dust in the Large Magellanic Cloud.

The Big Bang produced hydrogen, helium, and traces of lithium, while all heavier elements are synthesized in stars and supernovae. Supernovae tend to enrich the surrounding interstellar medium with metals—elements other than hydrogen and helium.

These injected elements ultimately enrich the molecular clouds that are the sites of star formation.[123] Thus, each stellar generation has a slightly different composition, going from an almost pure mixture of hydrogen and helium to a more metal-rich composition. Supernovae are the dominant mechanism for distributing these heavier elements, which are formed in a star during its period of nuclear fusion. The different abundances of elements in the material that forms a star have important influences on the star's life, and may decisively influence the possibility of having planets orbiting it.

The kinetic energy of an expanding supernova remnant can trigger star formation due to compression of nearby, dense molecular clouds in space.[124] The increase in turbulent pressure can also prevent star formation if the cloud is unable to lose the excess energy.[9]

Evidence from daughter products of short-lived radioactive isotopes shows that a nearby supernova helped determine the composition of the Solar System 4.5 billion years ago, and may even have triggered the formation of this system.[125] Supernova production of heavy elements over astronomic periods of time ultimately made the chemistry of life on Earth possible.

Effect on Earth

Main article: Near-Earth supernova

A near-Earth supernova is a supernova close enough to the Earth to have noticeable effects on its biosphere. Depending upon the type and energy of the supernova, it could be as far as 3000 light-years away. Gamma rays from a supernova would induce a chemical reaction in the upper atmosphere converting molecular nitrogen into nitrogen oxides, depleting the ozone layer enough to expose the surface to harmful solar radiation. This has been proposed as the cause of the Ordovician–Silurian extinction, which resulted in the death of nearly 60% of the oceanic life on Earth.[126] In 1996 it was theorized that traces of past supernovae might be detectable on Earth in the form of metal isotope signatures in rock strata. Iron-60 enrichment was later reported in deep-sea rock of the Pacific Ocean.[127][128][129] In 2009, elevated levels of nitrate ions were found in Antarctic ice, which coincided with the 1006 and 1054 supernovae. Gamma rays from these supernovae could have boosted levels of nitrogen oxides, which became trapped in the ice.[130]

Type Ia supernovae are thought to be potentially the most dangerous if they occur close enough to the Earth. Because these supernovae arise from dim, common white dwarf stars in binary systems, it is likely that a supernova that can affect the Earth will occur unpredictably and in a star system that is not well studied. The closest known candidate is IK Pegasi (see below).[131] Recent estimates predict that a Type II supernova would have to be closer than eight parsecs (26 light-years) to destroy half of the Earth's ozone layer, and there are no such candidates closer than about 500 light years.[132]

Milky Way candidates

The next supernova in the Milky Way will likely be detectable even if it occurs on the far side of the galaxy. It is likely to be produced by the collapse of an unremarkable red supergiant and it is very probable that it will already have been catalogued in infrared surveys such as 2MASS. There is a smaller chance that the next core collapse supernova will be produced by a different type of massive star such as a yellow hypergiant, luminous blue variable, or Wolf-Rayet. The chances of the next supernova being a type Ia produced by a white dwarf are calculated to be about a third of those for a core collapse supernova. Again it should be observable wherever it occurs, but it is less likely that the progenitor will ever have been observed prior to the explosion. It isn't even known exactly what a type Ia progenitor system looks like, and difficult to detect them beyond a few parsecs. The total supernova rate in our galaxy is estimated to be about 4.6 per century, or one every 22 years, although we haven't actually observed one for several centuries.[2]

The nebula around Wolf–Rayet star WR124, which is located at a distance of about 21,000 light years.[133]

Statistically, the next supernova is likely to be produced from an otherwise unremarkable red supergiant, but it is difficult to identify which of those supergiants are in the final stages of heavy element fusion in their cores and which have millions of years left. The most massive red supergiants are expected to shed their atmospheres and evolve to Wolf-Rayet stars before their cores collapse. All Wolf-Rayet stars are expected to end their lives from the Wolf-Rayet phase within a million years or so, but again it is difficult to identify those that are closest to core collapse. One class that is expected to have no more than a few thousand years before exploding are the WO Wolf-Rayet stars, which are known to have exhausted their core helium.[134] Only eight of them are known, and only four of those are in the Milky Way.[135]

A number of close or well known stars have been identified as possible core collapse supernova candidates: the red supergiants Antares and Betelgeuze;[136] the yellow hypergiant Rho Cassiopeiae;[137] the luminous blue variable Eta Carinae that has already produced a supernova imposter explosion;[138] and the brightest component, a Wolf–Rayet star in the Regor or Gamma Velorum system,[139] Others have gained notoriety as possible, although not very likely, progenitors for a gamma-ray burst; for example WR 104.[140]

Identification of candidates for a type Ia supernova explosion is much more speculative. Any binary with an accreting white dwarf might produce a supernova although the exact mechanism and timescale is still debated. These systems are faint and difficult to identify, but the novae and recurrent novae are such systems that conveniently advertise themselves. One examples is U Scorpii,[141] The nearest known type Ia supernova candidate is IK Pegasi (HR 8210), located at a distance of 150 light-years,[142] but observations suggest it will be several million years before the white dwarf can accrete the critical mass required to become a Type Ia supernova.[143]

See also

References

  1. 1 2 Reynolds, Stephen P.; et al. (2008). "The Youngest Galactic Supernova Remnant: G1.9+0.3". The Astrophysical Journal Letters. 680 (1): L41–L44. arXiv:0803.1487Freely accessible. Bibcode:2008ApJ...680L..41R. doi:10.1086/589570.
  2. 1 2 Adams, Scott M.; Kochanek, Christopher S.; Beacom, John F.; Vagins, Mark R.; Stanek, Krzysztof Z. (2013). "Observing the Next Galactic Supernova". The Astrophysical Journal. 778 (2): 164. arXiv:1306.0559Freely accessible. Bibcode:2013ApJ...778..164A. doi:10.1088/0004-637X/778/2/164.
  3. 1 2 3 Osterbrock, Donald E. (2001). "Who Really Coined the Word Supernova? Who First Predicted Neutron Stars?". Bulletin of the American Astronomical Society. 33: 1330. Bibcode:2001AAS...199.1501O.
  4. Giacobbe, Frederick W. (2005). "How a Type II Supernova Explodes". Electronic Journal of Theoretical Physics. 2 (6): 30–38. Bibcode:2005EJTP....2f..30G.
  5. 1 2 3 4 Heger, Alexander; Fryer, Chris L.; Woosley, Stanford E.; Langer, Norbert; Hartmann, Dieter H. (2003). "How Massive Single Stars End Their Life". Astrophysical Journal. 591: 288. arXiv:astro-ph/0212469Freely accessible. Bibcode:2003ApJ...591..288H. doi:10.1086/375341.
  6. Schawinski, Kevin; et al. (2008). "Supernova Shock Breakout from a Red Supergiant". Science. 321 (5886): 223–226. arXiv:0803.3596Freely accessible. Bibcode:2008Sci...321..223S. doi:10.1126/science.1160456. PMID 18556514.
  7. Timmes, F. X.; Woosley, S. E.; Weaver, Thomas A. (1995). "Galactic chemical evolution: Hydrogen through zinc". Astrophysical Journal Supplement Series. 98: 617. arXiv:astro-ph/9411003Freely accessible. Bibcode:1995ApJS...98..617T. doi:10.1086/192172.
  8. Whittet, Doug C. B. (2003). Dust in the Galactic Environment. CRC Press. pp. 45–46. ISBN 0-7503-0624-6.
  9. 1 2 Krebs, J.; Hillebrandt, Wolfgang (1983). "The interaction of supernova shockfronts and nearby interstellar clouds". Astronomy and Astrophysics. 128: 411. Bibcode:1983A&A...128..411K.
  10. Boss, A. P.; Ipatov, S. I.; Keiser, S. A.; Myhill, E. A.; Vanhala, H. A. T. (2008). "Simultaneous Triggered Collapse of the Presolar Dense Cloud Core and Injection of Short-Lived Radioisotopes by a Supernova Shock Wave". The Astrophysical Journal Letters. 686 (2): L119–L122. arXiv:0809.3045Freely accessible. Bibcode:2008ApJ...686L.119B. doi:10.1086/593057.
  11. Ackermann, M.; et al. (2013). "Detection of the Characteristic Pion-Decay Signature in Supernova Remnants". Science. 339 (6121): 807–11. arXiv:1302.3307Freely accessible. Bibcode:2013Sci...339..807A. doi:10.1126/science.1231160. PMID 23413352.
  12. Ott, Christian D.; et al. (2012). "Core-Collapse Supernovae, Neutrinos, and Gravitational Waves". Nuclear Physics B: Proceedings Supplement. 235: 381. arXiv:1212.4250Freely accessible. Bibcode:2013NuPhS.235..381O. doi:10.1016/j.nuclphysbps.2013.04.036.
  13. Winkler, P. Frank; Gupta, Gaurav; Long, Knox S. (2003). "The SN 1006 Remnant: Optical Proper Motions, Deep Imaging, Distance, and Brightness at Maximum". Astrophysical Journal. 585 (1): 324. arXiv:astro-ph/0208415Freely accessible. Bibcode:2003ApJ...585..324W. doi:10.1086/345985.
  14. Clark, David H.; Stephenson, Francis Richard (1982). "The Historical Supernovae". Supernovae: A survey of current research; Proceedings of the Advanced Study Institute, Cambridge, England, June 29 – July 10, 1981. Dordrecht: D. Reidel. pp. 355–370. Bibcode:1982sscr.conf..355C.
  15. Baade, Walter (1943). "No. 675. Nova Ophiuchi of 1604 as a supernova". Contributions from the Mount Wilson Observatory / Carnegie Institution of Washington. 675: 1–9. Bibcode:1943CMWCI.675....1B.
  16. Motz, Lloyd; Weaver, Jefferson Hane (2001). The Story of Astronomy. Basic Books. p. 76. ISBN 0-7382-0586-9.
  17. Murdin, Paul & Lesley (1978). Supernovae. New York, NY: Press Syndicate of the University of Cambridge. pp. 1–3. ISBN 052130038X.
  18. da Silva, Luiz Augusto L. (1993). "The Classification of Supernovae". Astrophysics and Space Science. 202 (2): 215–236. Bibcode:1993Ap&SS.202..215D. doi:10.1007/BF00626878.
  19. Kowal, Charles T. (1968). "Absolute magnitudes of supernovae". Astronomical Journal. 73: 1021–1024. Bibcode:1968AJ.....73.1021K. doi:10.1086/110763.
  20. Leibundgut, B.; et al. (2003). "A cosmological surprise: The universe accelerates". The Astrophysical Journal. 32 (4): 121. Bibcode:2001ENews..32..121L. doi:10.1086/378560.
  21. Fabian, Andrew C. (2008). "A Blast from the Past". Science. 320 (5880): 1167–1168. doi:10.1126/science.1158538. PMID 18511676.
  22. Aschenbach, Bernd (1998). "Discovery of a young nearby supernova remnant". Nature. 396 (6707): 141–142. Bibcode:1998Natur.396..141A. doi:10.1038/24103.
  23. Iyudin, A. F.; et al. (1998). "Emission from 44Ti associated with a previously unknown Galactic supernova". Nature. 396 (6707): 142–144. Bibcode:1998Natur.396..142I. doi:10.1038/24106.
  24. Subo Dong, B. J.; et al. (2016). "ASASSN-15lh: A highly super-luminous supernova". Science. 351 (6270): 257–260. arXiv:1507.03010Freely accessible. Bibcode:2016Sci...351..257D. doi:10.1126/science.aac9613. PMID 26816375.
  25. Cowen, Ron (14 January 2016). "Brightest-ever supernova still baffles astronomers". nature.com. Retrieved 15 January 2016.
  26. Baade, Walter; Zwicky, Fritz (1934). "On Super-novae". Proceedings of the National Academy of Sciences. 20 (5): 254–259. Bibcode:1934PNAS...20..254B. doi:10.1073/pnas.20.5.254. PMC 1076395Freely accessible. PMID 16587881.
  27. Murdin, Paul; Murdin, Lesley (1985). Supernovae (2nd ed.). Cambridge University Press. p. 42. ISBN 0-521-30038-X.
  28. Colgate, Stirling A.; McKee, Chester (1969). "Early Supernova Luminosity". The Astrophysical Journal. 157: 623. Bibcode:1969ApJ...157..623C. doi:10.1086/150102.
  29. Zuckerman, Ben; Malkan, Matthew A. (1996). The Origin and Evolution of the Universe. Jones & Bartlett Learning. p. 68. ISBN 0-7637-0030-4.
  30. Filippenko, Alexei V.; Li, Wei-Dong; Treffers, Richard R.; Modjaz, Maryam (2001). "The Lick Observatory Supernova Search with the Katzman Automatic Imaging Telescope". In Paczynski, B.; Chen, W.-P.; Lemme, C. Small Telescope Astronomy on Global Scale. ASP Conference Series. 246. San Francisco: Astronomical Society of the Pacific. p. 121. Bibcode:2001ASPC..246..121F. ISBN 978-1-58381-084-2.
  31. Antonioli, P.; et al. (2004). "SNEWS: The SuperNova Early Warning System". New Journal of Physics. 6: 114. arXiv:astro-ph/0406214Freely accessible. Bibcode:2004NJPh....6..114A. doi:10.1088/1367-2630/6/1/114.
  32. Scholberg, Kate (2000). "SNEWS: The supernova early warning system". AIP Conference Proceedings. 523: 355. arXiv:astro-ph/9911359Freely accessible. Bibcode:2000AIPC..523..355S. doi:10.1063/1.1291879.
  33. Beacom, John F. (1999). "Supernova neutrinos and the neutrino masses". Revista Mexicana de Fisica. 45 (2): 36. arXiv:hep-ph/9901300Freely accessible. Bibcode:1999RMxF...45...36B.
  34. Frieman, J. A.; et al. (2008). "The Sloan Digital Sky Survey-Ii Supernova Survey: Technical Summary". The Astronomical Journal. 135: 338. arXiv:0708.2749Freely accessible. Bibcode:2008AJ....135..338F. doi:10.1088/0004-6256/135/1/338.
  35. Perlmutter, Saul A.; et al. (1997). "Scheduled discovery of 7+ high-redshift SNe: First cosmology results and bounds on q0". In Ruiz-Lapuente, P.; Canal, R.; Isern, J. Thermonuclear Supernovae, Proceedings of the NATO Advanced Study Institute. NATO Advanced Science Institutes Series C. 486. Dordrecth: Kluwer Academic Publishers. p. 749. arXiv:astro-ph/9602122Freely accessible. Bibcode:1997ASIC..486..749P.
  36. Linder, Eric V.; Huterer, Dragan (2003). "Importance of supernovae at z > 1.5 to probe dark energy". Physical Review D. 67 (8): 081303. arXiv:astro-ph/0208138Freely accessible. Bibcode:2002astro.ph..8138L. doi:10.1103/PhysRevD.67.081303.
  37. Perlmutter, Saul A.; et al. (1997). "Measurements of the Cosmological Parameters Ω and Λ from the First Seven Supernovae at z  0.35". The Astrophysical Journal. 483 (2): 565. arXiv:astro-ph/9608192Freely accessible. Bibcode:1997ApJ...483..565P. doi:10.1086/304265.
  38. Copin, Y.; et al. (2006). "The Nearby Supernova Factory". New Astronomy Reviews. 50 (4–5): 436. Bibcode:2006NewAR..50..436C. doi:10.1016/j.newar.2006.02.035.
  39. Kirshner, Robert P. (1980). "Type I supernovae: An observer's view". AIP Conference Proceedings. AIP Conference Proceedings. 63: 33. Bibcode:1980AIPC...63...33K. doi:10.1063/1.32212.
  40. "List of Supernovae". IAU Central Bureau for Astronomical Telegrams. Retrieved 2010-10-25.
  41. "The Padova-Asiago supernova catalogue". Osservatorio Astronomico di Padova. Retrieved 2014-01-10.
  42. "Artist's impression of supernova 1993J". SpaceTelescope.org. Retrieved 2014-09-12.
  43. 1 2 Cappellaro, Enrico; Turatto, Massimo (2001). "Supernova Types and Rates". Influence of Binaries on Stellar Population Studies. 264. Dordrecht: Kluwer Academic Publishers. p. 199. arXiv:astro-ph/0012455Freely accessible. Bibcode:2001ASSL..264..199C. doi:10.1007/978-94-015-9723-4_16. ISBN 978-0-7923-7104-5.
  44. 1 2 3 4 Turatto, Massimo (2003). "Classification of Supernovae". Supernovae and Gamma-Ray Bursters. Lecture Notes in Physics. 598. p. 21. doi:10.1007/3-540-45863-8_3. ISBN 978-3-540-44053-6.
  45. 1 2 3 4 Doggett, Jesse B.; Branch, David (1985). "A comparative study of supernova light curves". The Astronomical Journal. 90: 2303. Bibcode:1985AJ.....90.2303D. doi:10.1086/113934.
  46. Zwicky, Fritz (1964). "NGC 1058 and its Supernova 1961". The Astrophysical Journal. 139: 514. Bibcode:1964ApJ...139..514Z. doi:10.1086/147779.
  47. Zwicky, Fritz (1962). "New Observations of Importance to Cosmology". In McVittie, G. C. Problems of Extra-Galactic Research, Proceedings from IAU Symposium. 15. New York: Macmillan Press. p. 347. Bibcode:1962IAUS...15..347Z.
  48. "The Rise and Fall of a Supernova". ESO Picture of the Week. Retrieved 2013-06-14.
  49. Piro, Anthony L.; Thompson, Todd A.; Kochanek, Christopher S. (2014). "Reconciling 56Ni production in Type Ia supernovae with double degenerate scenarios". Monthly Notices of the Royal Astronomical Society. 438 (4): 3456. arXiv:1308.0334Freely accessible. Bibcode:2014MNRAS.438.3456P. doi:10.1093/mnras/stt2451.
  50. Chen, Wen-Cong; Li, Xiang-Dong (2009). "On the Progenitors of Super-Chandrasekhar Mass Type Ia Supernovae". The Astrophysical Journal. 702: 686. arXiv:0907.0057Freely accessible. Bibcode:2009ApJ...702..686C. doi:10.1088/0004-637X/702/1/686.
  51. Howell, D. Andrew; Sullivan, Mark; Conley, Alexander J.; Carlberg, Raymond G. (2007). "Predicted and Observed Evolution in the Mean Properties of Type Ia Supernovae with Redshift". Astrophysical Journal Letters. 667 (1): L37–L40. arXiv:astro-ph/0701912Freely accessible. Bibcode:2007ApJ...667L..37H. doi:10.1086/522030.
  52. 1 2 Mazzali, Paolo A.; Röpke, Friedrich K.; Benetti, Stefano; Hillebrandt, Wolfgang (2007). "A Common Explosion Mechanism for Type Ia Supernovae". Science. 315 (5813): 825–828. arXiv:astro-ph/0702351Freely accessible. Bibcode:2007Sci...315..825M. doi:10.1126/science.1136259. PMID 17289993.
  53. Lieb, Elliott H.; Yau, Horng-Tzer (1987). "A rigorous examination of the Chandrasekhar theory of stellar collapse". The Astrophysical Journal. 323 (1): 140–144. Bibcode:1987ApJ...323..140L. doi:10.1086/165813.
  54. Canal, Ramon; Gutiérrez, Jordi L. (1997). "The possible white dwarf-neutron star connection". In Isern, J.; Hernanz, M.; Gracia-Berro, E. Proceedings of the 10th European Workshop on White Dwarfs. 214. Dordrecht: Kluwer Academic Publishers. p. 49. Bibcode:1997astro.ph..1225C. ISBN 978-0-7923-4585-5.
  55. Wheeler, J. Craig (2000). Cosmic Catastrophes: Supernovae, Gamma-Ray Bursts, and Adventures in Hyperspace. Cambridge University Press. p. 96. ISBN 978-0-521-65195-0.
  56. Khokhlov, Alexei M.; Mueller, Ewald; Höflich, Peter A. (1993). "Light curves of Type IA supernova models with different explosion mechanisms". Astronomy and Astrophysics. 270 (1–2): 223–248. Bibcode:1993A&A...270..223K.
  57. Röpke, Friedrich K.; Hillebrandt, Wolfgang (2004). "The case against the progenitor's carbon-to-oxygen ratio as a source of peak luminosity variations in Type Ia supernovae". Astronomy and Astrophysics Letters. 420 (1): L1–L4. arXiv:astro-ph/0403509Freely accessible. Bibcode:2004A&A...420L...1R. doi:10.1051/0004-6361:20040135.
  58. 1 2 Hillebrandt, Wolfgang; Niemeyer, Jens C. (2000). "Type IA Supernova Explosion Models". Annual Review of Astronomy and Astrophysics. 38 (1): 191–230. arXiv:astro-ph/0006305Freely accessible. Bibcode:2000ARA&A..38..191H. doi:10.1146/annurev.astro.38.1.191.
  59. Paczyński, Bohdan (1976). "Common Envelope Binaries". In Eggleton, P.; Mitton, S.; Whelan, J. Structure and Evolution of Close Binary Systems. IAU Symposium No. 73. Dordrecht: D. Reidel. pp. 75–80. Bibcode:1976IAUS...73...75P.
  60. Macri, Lucas M.; Stanek, Krzysztof Z.; Bersier, David; Greenhill, Lincoln J.; Reid, Mark J. (2006). "A New Cepheid Distance to the Maser-Host Galaxy NGC 4258 and Its Implications for the Hubble Constant". The Astrophysical Journal. 652 (2): 1133–1149. arXiv:astro-ph/0608211Freely accessible. Bibcode:2006ApJ...652.1133M. doi:10.1086/508530.
  61. Colgate, Stirling A. (1979). "Supernovae as a standard candle for cosmology". The Astrophysical Journal. 232 (1): 404–408. Bibcode:1979ApJ...232..404C. doi:10.1086/157300.
  62. Ruiz-Lapuente, P.; et al. (2000). "Type IA supernova progenitors". Memorie della Societa Astronomica Italiana. 71: 435. Bibcode:2000MmSAI..71..435R.
  63. Dan, Marius; Rosswog, Stephan; Guillochon, James; Ramirez-Ruiz, Enrico (2012). "How the merger of two white dwarfs depends on their mass ratio: Orbital stability and detonations at contact". Monthly Notices of the Royal Astronomical Society. 422 (3): 2417. arXiv:1201.2406Freely accessible. Bibcode:2012MNRAS.422.2417D. doi:10.1111/j.1365-2966.2012.20794.x.
  64. Howell, D. Andrew; et al. (2006). "The type Ia supernova SNLS-03D3bb from a super-Chandrasekhar-mass white dwarf star". Nature. 443 (7109): 308–311. arXiv:astro-ph/0609616Freely accessible. Bibcode:2006Natur.443..308H. doi:10.1038/nature05103. PMID 16988705.
  65. Tanaka, M.; et al. (2010). "Spectropolarimetry of Extremely Luminous Type Ia Supernova 2009dc: Nearly Spherical Explosion of Super-Chandrasekhar Mass White Dwarf". The Astrophysical Journal. 714 (2): 1209. arXiv:0908.2057Freely accessible. Bibcode:2010ApJ...714.1209T. doi:10.1088/0004-637X/714/2/1209.
  66. Wang, B.; Liu, D.; Jia, S.; Han, Z. (2014). "Helium double-detonation explosions for the progenitors of type Ia supernovae". Proceedings of the International Astronomical Union. 9 (S298): 442. arXiv:1301.1047Freely accessible. Bibcode:2014IAUS..298..442W. doi:10.1017/S1743921313007072.
  67. Foley, R. J.; et al. (2013). "Type Iax Supernovae: A New Class of Stellar Explosion". The Astrophysical Journal. 767: 57. arXiv:1212.2209Freely accessible. Bibcode:2013ApJ...767...57F. doi:10.1088/0004-637X/767/1/57.
  68. McCully, Curtis; et al. (2014). "A luminous, blue progenitor system for the type Iax supernova 2012Z". Nature. 512 (7512): 54–56. arXiv:1408.1089Freely accessible. Bibcode:2014Natur.512...54M. doi:10.1038/nature13615. PMID 25100479.
  69. Silverman, J. M.; et al. (2013). "Type Ia Supernovae strongle interaction with their circumstellar medium". The Astrophysical Journal Supplement Series. 207 (1): 3. arXiv:1304.0763Freely accessible. Bibcode:2013ApJS..207....3S. doi:10.1088/0067-0049/207/1/3.
  70. Nomoto, Ken′ichi; Tanaka, Masaomi; Tominaga, Nozomu; Maeda, Keiichi (2010). "Hypernovae, gamma-ray bursts, and first stars". New Astronomy Reviews. 54 (3–6): 191. Bibcode:2010NewAR..54..191N. doi:10.1016/j.newar.2010.09.022.
  71. Moriya, Takashi J. (2012). "Progenitors of Recombining Supernova Remnants". The Astrophysical Journal. 750 (1): L13. arXiv:1203.5799Freely accessible. Bibcode:2012ApJ...750L..13M. doi:10.1088/2041-8205/750/1/L13.
  72. Smith, N.; et al. (2009). "Sn 2008S: A Cool Super-Eddington Wind in a Supernova Impostor". The Astrophysical Journal. 697: L49. arXiv:0811.3929Freely accessible. Bibcode:2009ApJ...697L..49S. doi:10.1088/0004-637X/697/1/L49.
  73. Fryer, Chris L.; New, Kimberly C. B. (2003). "Gravitational Waves from Gravitational Collapse". Living Reviews in Relativity. 6. doi:10.12942/lrr-2003-2.
  74. 1 2 3 4 Woosley, Stanford E.; Janka, Hans-Thomas (2005). "The Physics of Core-Collapse Supernovae". Nature Physics. 1 (3): 147–154. arXiv:astro-ph/0601261Freely accessible. Bibcode:2005NatPh...1..147W. doi:10.1038/nphys172.
  75. Janka, Hans-Thomas; Langanke, Karlheinz; Marek, Andreas; Martínez-Pinedo, Gabriel; Müller, Bernhard (2007). "Theory of core-collapse supernovae". Physics Reports. 442: 38. arXiv:astro-ph/0612072Freely accessible. Bibcode:2007PhR...442...38J. doi:10.1016/j.physrep.2007.02.002.
  76. Gribbin, J. R.; Gribbin, M. (2000). Stardust: Supernovae and Life – The Cosmic Connection. Yale University Press. p. 173. ISBN 978-0-300-09097-0.
  77. 1 2 Barwick, S. W.; et al. (2004). "APS Neutrino Study: Report of the Neutrino Astrophysics and Cosmology Working Group". arXiv:astro-ph/0412544Freely accessible.
  78. Myra, Eric S.; Burrows, Adam (1990). "Neutrinos from type II supernovae- The first 100 milliseconds". Astrophysical Journal. 364: 222–231. Bibcode:1990ApJ...364..222M. doi:10.1086/169405.
  79. 1 2 Kasen, D.; Woosley, Stanford E.; Heger, Alexander (2011). "Pair Instability Supernovae: Light Curves, Spectra, and Shock Breakout". The Astrophysical Journal. 734 (2): 102. arXiv:1101.3336Freely accessible. Bibcode:2011ApJ...734..102K. doi:10.1088/0004-637X/734/2/102.
  80. 1 2 Poelarends, Arend J. T.; Herwig, Falk; Langer, Norbert; Heger, Alexander (2008). "The Supernova Channel of Super‐AGB Stars". The Astrophysical Journal. 675: 614. arXiv:0705.4643Freely accessible. Bibcode:2008ApJ...675..614P. doi:10.1086/520872.
  81. Gilmore, Gerry (2004). "ASTRONOMY: The Short Spectacular Life of a Superstar". Science. 304 (5679): 1915–1916. doi:10.1126/science.1100370. PMID 15218132.
  82. Faure, Gunter; Mensing, Teresa M. (2007). "Life and Death of Stars". Introduction to Planetary Science. pp. 35–48. doi:10.1007/978-1-4020-5544-7_4. ISBN 978-1-4020-5233-0.
  83. Malesani, D.; et al. (2009). "Early Spectroscopic Identification of SN 2008D". The Astrophysical Journal Letters. 692 (2): L84. arXiv:0805.1188Freely accessible. Bibcode:2009ApJ...692L..84M. doi:10.1088/0004-637X/692/2/L84.
  84. Svirski, Gilad; Nakar, Ehud (2014). "Sn 2008D: A Wolf-Rayet Explosion Through a Thick Wind". The Astrophysical Journal. 788: L14. arXiv:1403.3400Freely accessible. Bibcode:2014ApJ...788L..14S. doi:10.1088/2041-8205/788/1/L14.
  85. Pols, Onno (1997). "Close Binary Progenitors of Type Ib/Ic and IIb/II-L Supernovae". In Leung, K.-C. Proceedings of The Third Pacific Rim Conference on Recent Development on Binary Star Research. ASP Conference Series. 130. pp. 153–158. Bibcode:1997rdbs.conf..153P.
  86. 1 2 Eldridge, John J.; Fraser, Morgan; Smartt, Stephen J.; Maund, Justyn R.; Crockett, R. Mark (2013). "The death of massive stars – II. Observational constraints on the progenitors of Type Ibc supernovae". Monthly Notices of the Royal Astronomical Society. 436: 774. arXiv:1301.1975Freely accessible. Bibcode:2013MNRAS.436..774E. doi:10.1093/mnras/stt1612.
  87. Ryder, Stuart D.; et al. (2004). "Modulations in the radio light curve of the Type IIb supernova 2001ig: evidence for a Wolf-Rayet binary progenitor?". Monthly Notices of the Royal Astronomical Society. 349 (3): 1093–1100. arXiv:astro-ph/0401135Freely accessible. Bibcode:2004MNRAS.349.1093R. doi:10.1111/j.1365-2966.2004.07589.x.
  88. Inserra, C.; et al. (2013). "Super-luminous Type Ic Supernovae: Catching a Magnetar by the Tail". The Astrophysical Journal. 770 (2): 28. arXiv:1304.3320Freely accessible. Bibcode:2013ApJ...770..128I. doi:10.1088/0004-637X/770/2/128.
  89. Nicholl, M.; et al. (2013). "Slowly fading super-luminous supernovae that are not pair-instability explosions". Nature. 502 (7471): 346–349. arXiv:1310.4446Freely accessible. Bibcode:2013Natur.502..346N. doi:10.1038/nature12569. PMID 24132291.
  90. Tauris, Thomas M.; Langer, Norbert; Moriya, Takashi J.; Podsiadlowski, Philipp; Yoon, Sung-Chul; Blinnikov, Sergey I. (2013). "Ultra-stripped Type Ic supernovae from close binary evolution". Astrophysical Journal Letters. 778 (2): L23. arXiv:1310.6356Freely accessible. Bibcode:2013ApJ...778L..23T. doi:10.1088/2041-8205/778/2/L23.
  91. Drout, M. R.; Soderberg, A. M.; Mazzali, P. A.; Parrent, J. T.; Margutti, R.; Milisavljevic, D.; Sanders, N. E.; Chornock, R.; Foley, R. J.; Kirshner, Robert P.; Filippenko, Alexei V.; Li, W.; Brown, P. J.; Cenko, S. B.; Chakraborti, S.; Challis, P.; Friedman, A.; Ganeshalingam, M.; Hicken, M.; Jensen, C.; Modjaz, Maryam; Perets, H. B.; Silverman, J. M.; Wong, D. S. (2013). "The Fast and Furious Decay of the Peculiar Type Ic Supernova 2005ek". Astrophysical Journal. 774 (58): 44. arXiv:1306.2337Freely accessible. Bibcode:2013ApJ...774...58D. doi:10.1088/0004-637X/774/1/58.
  92. Reynolds, Thomas M.; Fraser, Morgan; Gilmore, Gerard (2015). "Gone without a bang: an archivalHSTsurvey for disappearing massive stars". Monthly Notices of the Royal Astronomical Society. 453 (3): 2886–2901. arXiv:1507.05823Freely accessible. Bibcode:2015MNRAS.453.2885R. doi:10.1093/mnras/stv1809. ISSN 0035-8711.
  93. Gerke, J. R.; Kochanek, Christopher S.; Stanek, Krzysztof Z. (2015). "The search for failed supernovae with the Large Binocular Telescope: first candidates". Monthly Notices of the Royal Astronomical Society. 450 (3): 3289–3305. arXiv:1411.1761Freely accessible. Bibcode:2015MNRAS.450.3289G. doi:10.1093/mnras/stv776. ISSN 0035-8711.
  94. Bodansky, David; Clayton, Donald D.; Fowler, William A. (1968). "Nucleosynthesis During Silicon Burning". Physical Review Letters. 20 (4): 161. Bibcode:1968PhRvL..20..161B. doi:10.1103/PhysRevLett.20.161.
  95. 1 2 Matz, S. M.; Share, G. H.; Leising, M. D.; Chupp, E. L.; Vestrand, W. T.; Purcell, W.R.; Strickman, M.S.; Reppin, C. (1988). "Gamma-ray line emission from SN1987A". Nature. 331 (6155): 416. Bibcode:1988Natur.331..416M. doi:10.1038/331416a0.
  96. Kasen, Daniel; Woosley, S. E. (2009). "Type Ii Supernovae: Model Light Curves and Standard Candle Relationships". The Astrophysical Journal. 703 (2): 2205. arXiv:0910.1590Freely accessible. Bibcode:2009ApJ...703.2205K. doi:10.1088/0004-637X/703/2/2205.
  97. Churazov, E.; Sunyaev, R.; Isern, J.; Knödlseder, J.; Jean, P.; Lebrun, F.; Chugai, N.; Grebenev, S.; Bravo, E.; Sazonov, S.; Renaud, M. (2014). "Cobalt-56 γ-ray emission lines from the type Ia supernova 2014J". Nature. 512 (7515): 406. arXiv:1405.3332Freely accessible. Bibcode:2014Natur.512..406C. doi:10.1038/nature13672. PMID 25164750.
  98. Barbon, Roberto; Ciatti, Franco; Rosino, Leonida (1979). "Photometric properties of type II supernovae". Astronomy and Astrophysics. 72: 287. Bibcode:1979A&A....72..287B.
  99. Li, W.; Leaman, J.; Chornock, R.; Filippenko, A. V.; Poznanski, D.; Ganeshalingam, M.; Wang, X.; Modjaz, M.; Jha, S.; Foley, R. J.; Smith, N. (2011). "Nearby supernova rates from the Lick Observatory Supernova Search – II. The observed luminosity functions and fractions of supernovae in a complete sample". Monthly Notices of the Royal Astronomical Society. 412 (3): 1441. arXiv:1006.4612Freely accessible. Bibcode:2011MNRAS.412.1441L. doi:10.1111/j.1365-2966.2011.18160.x.
  100. Richardson, D.; Branch, D.; Casebeer, D.; Millard, J.; Thomas, R. C.; Baron, E. (2002). "A Comparative Study of the Absolute Magnitude Distributions of Supernovae". The Astronomical Journal. 123 (2): 745. arXiv:astro-ph/0112051Freely accessible. Bibcode:2002AJ....123..745R. doi:10.1086/338318.
  101. Frail, Dale A.; Giacani, Elsa B.; Goss, W. Miller; Dubner, Gloria M. (1996). "The Pulsar Wind Nebula Around PSR B1853+01 in the Supernova Remnant W44". Astrophysical Journal Letters. 464 (2): L165–L168. arXiv:astro-ph/9604121Freely accessible. Bibcode:1996ApJ...464L.165F. doi:10.1086/310103.
  102. Höflich, Peter A.; Kumar, Pawan; Wheeler, J. Craig (2004). "Neutron star kicks and supernova asymmetry". Cosmic explosions in three dimensions: Asymmetries in supernovae and gamma-ray bursts. Cambridge University Press. p. 276. Bibcode:2004cetd.conf..276L.
  103. Fryer, Chris L. (2004). "Neutron Star Kicks from Asymmetric Collapse". Astrophysical Journal. 601 (2): L175–L178. arXiv:astro-ph/0312265Freely accessible. Bibcode:2004ApJ...601L.175F. doi:10.1086/382044.
  104. Gilkis, Avishai; Soker, Noam (2014). "Implications of turbulence for jets in core-collapse supernova explosions". arXiv:1412.4984Freely accessible.
  105. Khokhlov, Alexei M.; et al. (1999). "Jet-induced Explosions of Core Collapse Supernovae". The Astrophysical Journal. 524 (2): L107. arXiv:astro-ph/9904419Freely accessible. Bibcode:1999ApJ...524L.107K. doi:10.1086/312305.
  106. Wang, L.; et al. (2003). "Spectropolarimetry of SN 2001el in NGC 1448: Asphericity of a Normal Type Ia Supernova". The Astrophysical Journal. 591 (2): 1110. arXiv:astro-ph/0303397Freely accessible. Bibcode:2003ApJ...591.1110W. doi:10.1086/375444.
  107. 1 2 Mazzali, P. A.; Nomoto, K. I.; Cappellaro, E.; Nakamura, T.; Umeda, H.; Iwamoto, K. (2001). "Can Differences in the Nickel Abundance in Chandrasekhar‐Mass Models Explain the Relation between the Brightness and Decline Rate of Normal Type Ia Supernovae?". The Astrophysical Journal. 547 (2): 988. arXiv:astro-ph/0009490Freely accessible. Bibcode:2001ApJ...547..988M. doi:10.1086/318428.
  108. Iwamoto, K. (2006). "Neutrino Emission from Type Ia Supernovae". AIP Conference Proceedings. 847. p. 406. doi:10.1063/1.2234440.
  109. Hayden, B. T.; Garnavich, P. M.; Kessler, R.; Frieman, J. A.; Jha, S. W.; Bassett, B.; Cinabro, D.; Dilday, B.; Kasen, D.; Marriner, J.; Nichol, R. C.; Riess, A. G.; Sako, M.; Schneider, D. P.; Smith, M.; Sollerman, J. (2010). "The Rise and Fall of Type Ia Supernova Light Curves in the SDSS-II Supernova Survey". The Astrophysical Journal. 712: 350. arXiv:1001.3428Freely accessible. Bibcode:2010ApJ...712..350H. doi:10.1088/0004-637X/712/1/350.
  110. Janka, Hans-Thomas (2012). "Explosion Mechanisms of Core-Collapse Supernovae". Annual Review of Nuclear and Particle Science. 62: 407. arXiv:1206.2503Freely accessible. Bibcode:2012ARNPS..62..407J. doi:10.1146/annurev-nucl-102711-094901.
  111. Smartt (2009). "Progenitors of core-collapse supernovae". Annual Review of Astronomy and Astrophysics. 47: 63–106. arXiv:0908.0700v2Freely accessible. Bibcode:2009ARA&A..47...63S. doi:10.1146/annurev-astro-082708-101737.
  112. 1 2 3 Smartt, Stephen J. (2009). "Progenitors of Core-Collapse Supernovae". Annual Review of Astronomy & Astrophysics. 47: 63. arXiv:0908.0700Freely accessible. Bibcode:2009ARA&A..47...63S. doi:10.1146/annurev-astro-082708-101737.
  113. Walmswell, J. J.; Eldridge, J. J. (2012). "Circumstellar dust as a solution to the red supergiant supernova progenitor problem". Monthly Notices of the Royal Astronomical Society. 419 (3): 2054. arXiv:1109.4637Freely accessible. Bibcode:2012MNRAS.419.2054W. doi:10.1111/j.1365-2966.2011.19860.x.
  114. Georgy, C. (2012). "Yellow supergiants as supernova progenitors: An indication of strong mass loss for red supergiants?". Astronomy & Astrophysics. 538: L8–L2. arXiv:1111.7003Freely accessible. Bibcode:2012A&A...538L...8G. doi:10.1051/0004-6361/201118372.
  115. Yoon, S. -C.; Gräfener, G.; Vink, J. S.; Kozyreva, A.; Izzard, R. G. (2012). "On the nature and detectability of Type Ib/c supernova progenitors". Astronomy & Astrophysics. 544: L11. arXiv:1207.3683Freely accessible. Bibcode:2012A&A...544L..11Y. doi:10.1051/0004-6361/201219790.
  116. Groh, J. H.; Meynet, G.; Ekström, S. (2013). "Massive star evolution: Luminous blue variables as unexpected supernova progenitors". Astronomy & Astrophysics. 550: L7. arXiv:1301.1519Freely accessible. Bibcode:2013A&A...550L...7G. doi:10.1051/0004-6361/201220741.
  117. Yoon, S.-C.; Gräfener, G.; Vink, J. S.; Kozyreva, A.; Izzard, R. G. (2012). "On the nature and detectability of Type Ib/c supernova progenitors". Astronomy & Astrophysics. 544: L11. arXiv:1207.3683Freely accessible. Bibcode:2012A&A...544L..11Y. doi:10.1051/0004-6361/201219790.
  118. François, P.; et al. (2004). "The evolution of the Milky Way from its earliest phases: Constraints on stellar nucleosynthesis". Astronomy and Astrophysics. 421 (2): 613–621. arXiv:astro-ph/0401499Freely accessible. Bibcode:2004A&A...421..613F. doi:10.1051/0004-6361:20034140.
  119. Woosley, Stanford E.; Arnett, W. D.; Clayton, D. D. (1973). "The Explosive Burning of Oxygen and Silicon". Astrophysical Journal Supplement. 26: 231–312. Bibcode:1973ApJS...26..231W. doi:10.1086/190282.
  120. Qian, Y.-Z.; Vogel, P.; Wasserburg, G. J. (1998). "Diverse Supernova Sources for the r-Process". Astrophysical Journal. 494 (1): 285–296. arXiv:astro-ph/9706120Freely accessible. Bibcode:1998ApJ...494..285Q. doi:10.1086/305198.
  121. Gonzalez, Guillermo; Brownlee, Donald; Ward, Peter (2001). "The Galactic Habitable Zone: Galactic Chemical Evolution". Icarus. 152: 185. arXiv:astro-ph/0103165Freely accessible. Bibcode:2001Icar..152..185G. doi:10.1006/icar.2001.6617.
  122. Cox, Donald P. (1972). "Cooling and Evolution of a Supernova Remnant". Astrophysical Journal. 178: 159. Bibcode:1972ApJ...178..159C. doi:10.1086/151775.
  123. Sandstrom, Karin M.; Bolatto, Alberto D.; Stanimirović, Snežana; Van Loon, Jacco Th.; Smith, J. D. T. (2009). "Measuring Dust Production in the Small Magellanic Cloud Core-Collapse Supernova Remnant 1E 0102.2–7219". The Astrophysical Journal. 696 (2): 2138. arXiv:0810.2803Freely accessible. Bibcode:2009ApJ...696.2138S. doi:10.1088/0004-637X/696/2/2138.
  124. Preibisch, T.; Zinnecker, H. (2001). "Triggered Star Formation in the Scorpius-Centaurus OB Association (Sco OB2)". From Darkness to Light: Origin and Evolution of Young Stellar Clusters. 243: 791. arXiv:astro-ph/0008013Freely accessible. Bibcode:2001ASPC..243..791P.
  125. Cameron, A.G.W.; Truran, J.W. (1977). "The supernova trigger for formation of the solar system". Icarus. 30 (3): 447. Bibcode:1977Icar...30..447C. doi:10.1016/0019-1035(77)90101-4.
  126. Melott, A.; et al. (2004). "Did a gamma-ray burst initiate the late Ordovician mass extinction?". International Journal of Astrobiology. 3 (2): 55–61. arXiv:astro-ph/0309415Freely accessible. Bibcode:2004IJAsB...3...55M. doi:10.1017/S1473550404001910.
  127. Fields, Brian D.; Hochmuth, Kathrin A.; Ellis, John (2005). "Deep‐Ocean Crusts as Telescopes: Using Live Radioisotopes to Probe Supernova Nucleosynthesis". The Astrophysical Journal. 621 (2): 902. arXiv:astro-ph/0410525Freely accessible. Bibcode:2005ApJ...621..902F. doi:10.1086/427797.
  128. Knie, K.; et al. (2004). "60Fe Anomaly in a Deep-Sea Manganese Crust and Implications for a Nearby Supernova Source". Physical Review Letters. 93 (17): 171103–171106. Bibcode:2004PhRvL..93q1103K. doi:10.1103/PhysRevLett.93.171103.
  129. Fields, B. D.; Ellis, J. (1999). "On Deep-Ocean Fe-60 as a Fossil of a Near-Earth Supernova". New Astronomy. 4 (6): 419–430. arXiv:astro-ph/9811457Freely accessible. Bibcode:1999NewA....4..419F. doi:10.1016/S1384-1076(99)00034-2.
  130. "In Brief". Scientific American. 300 (5): 28. 2009. doi:10.1038/scientificamerican0509-28a.
  131. Gorelick, M. (2007). "The Supernova Menace". Sky & Telescope. 113: 26. Bibcode:2007S&T...113c..26G.
  132. Gehrels, Neil; et al. (2003). "Ozone Depletion from Nearby Supernovae". Astrophysical Journal. 585 (2): 1169–1176. arXiv:astro-ph/0211361Freely accessible. Bibcode:2003ApJ...585.1169G. doi:10.1086/346127.
  133. Van Der Sluys, M. V.; Lamers, H. J. G. L. M. (2003). "The dynamics of the nebula M1-67 around the run-away Wolf-Rayet star WR 124". Astronomy and Astrophysics. 398: 181. arXiv:astro-ph/0211326Freely accessible. Bibcode:2003A&A...398..181V. doi:10.1051/0004-6361:20021634.
  134. Tramper, F.; Straal, S. M.; Sanyal, D.; Sana, H.; De Koter, A.; Gräfener, G.; Langer, N.; Vink, J. S.; De Mink, S. E.; Kaper, L. (2015). "Massive stars on the verge of exploding: The properties of oxygen sequence Wolf-Rayet stars". Astronomy & Astrophysics. 581: A110. arXiv:1507.00839Freely accessible. Bibcode:2015A&A...581A.110T. doi:10.1051/0004-6361/201425390.
  135. Tramper, F.; Gräfener, G.; Hartoog, O. E.; Sana, H.; De Koter, A.; Vink, J. S.; Ellerbroek, L. E.; Langer, N.; Garcia, M.; Kaper, L.; De Mink, S. E. (2013). "On the nature of WO stars: A quantitative analysis of the WO3 star DR1 in IC 1613". Astronomy & Astrophysics. 559: A72. arXiv:1310.2849Freely accessible. Bibcode:2013A&A...559A..72T. doi:10.1051/0004-6361/201322155.
  136. Inglis, Michael (2015). "Star Death: Supernovae, Neutron Stars & Black Holes". Astrophysics is Easy!. The Patrick Moore Practical Astronomy Series. p. 203. doi:10.1007/978-3-319-11644-0_12. ISBN 978-3-319-11643-3.
  137. Lobel, A.; et al. (2004). "Spectroscopy of the Millennium Outburst and Recent Variability of the Yellow Hypergiant Rho Cassiopeiae". Stars as suns : activity. 219: 903. arXiv:astro-ph/0312074Freely accessible. Bibcode:2004IAUS..219..903L.
  138. Van Boekel, R.; et al. (2003). "Direct measurement of the size and shape of the present-day stellar wind of eta Carinae". Astronomy and Astrophysics. 410 (3): L37. arXiv:astro-ph/0310399Freely accessible. Bibcode:2003A&A...410L..37V. doi:10.1051/0004-6361:20031500.
  139. Thielemann, F.-K.; Hirschi, R.; Liebendörfer, M.; Diehl, R. (2011). "Massive Stars and Their Supernovae". Astronomy with Radioactivities. Lecture Notes in Physics. 812. p. 153. doi:10.1007/978-3-642-12698-7_4. ISBN 978-3-642-12697-0.
  140. Tuthill, Peter G.; et al. (2008). "The Prototype Colliding‐Wind Pinwheel WR 104". The Astrophysical Journal. 675: 698–710. arXiv:0712.2111Freely accessible. Bibcode:2008ApJ...675..698T. doi:10.1086/527286.
  141. Thoroughgood, T. D.; et al. (2002). "The recurrent nova U Scorpii — A type Ia supernova progenitor". The Physics of Cataclysmic Variables and Related Objects. 261. San Francisco, CA: Astronomical Society of the Pacific. Bibcode:2002ASPC..261...77T.
  142. Landsman, W.; Simon, T.; Bergeron, P. (1999). "The hot white-dwarf companions of HR 1608, HR 8210, and HD 15638". Publications of the Astronomical Society of the Pacific. 105 (690): 841–847. Bibcode:1993PASP..105..841L. doi:10.1086/133242.
  143. Vennes, S.; Kawka, A. (2008). "On the empirical evidence for the existence of ultramassive white dwarfs". Monthly Notices of the Royal Astronomical Society. 389 (3): 1367. arXiv:0806.4742Freely accessible. Bibcode:2008MNRAS.389.1367V. doi:10.1111/j.1365-2966.2008.13652.x.

Notes

    1. The value is obtained by converting the suffix "nc" from bijective base-26, with a = 1, b = 2, c =3,  ... z = 26. Thus nc = n × 26 + c = 14 × 26 + 3 = 367.
    2. For a core primarily composed of oxygen, neon and magnesium, the collapsing white dwarf will typically form a neutron star. In this case, only a fraction of the star's mass will be ejected during the collapse.

    Further reading

    External links

    Wikimedia Commons has media related to Supernovae.

    This article is issued from Wikipedia - version of the 12/5/2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.