Betelgeuse

Not to be confused with Beetlejuice.
For other uses, see Betelgeuse (disambiguation).
Betelgeuse

The pink arrow at the star on left labeled α indicates Betelgeuse in Orion.
Observation data
Epoch J2000.0      Equinox J2000.0
Constellation Orion
Pronunciation /ˈbtəlz/, /ˈbɛtəlz/[1] or
/ˈbtəls/[2]
Right ascension 05h 55m 10.30536s[3]
Declination +07° 24 25.4304[3]
Characteristics
Evolutionary stage Red supergiant
Spectral type M1–M2 Ia–ab[4]
Apparent magnitude (V) 0.50[5] (0.0 - 1.3[6])
Apparent magnitude (J) −3.00[7]
Apparent magnitude (K) −4.05[7]
U−B color index +2.06[5]
B−V color index +1.85[5]
Variable type SRc[6]
Astrometry
Radial velocity (Rv)+21.91[8] km/s
Proper motion (μ) RA: 24.95 ± 0.08[9] mas/yr
Dec.: 9.56 ± 0.15[9] mas/yr
Parallax (π)5.07 ± 1.10[9] mas
Distance643 ± 146[9] ly
(197 ± 45[9] pc)
Absolute magnitude (MV)−5.85[9]
Details
Mass11.6+5.0
−3.9
[10] M
Radius887±203[11] R
Luminosity90000150000[12] L
Surface gravity (log g)−0.5[13] cgs
Temperature3590[10] K
Metallicity [Fe/H]+0.05[14] dex
Rotational velocity (v sin i)5[15] km/s
Age8.0 – 8.5[11] Myr
Other designations
Betelgeuse, α Ori, 58 Ori, HR 2061, BD+7°1055, HD 39801, FK5 224, HIP 27989, SAO 113271, GC 7451, CCDM J05552+0724AP, AAVSO 0549+07
Database references
SIMBADdata

Coordinates: 05h 55m 10.3053s, +07° 24′ 25.426″

Betelgeuse, also designated Alpha Orionis (α Orionis, abbreviated Alpha Ori, α Ori), is the ninth-brightest star in the night sky and second-brightest in the constellation of Orion. Distinctly reddish, it is a semiregular variable star whose apparent magnitude varies between 0.0 and 1.3, the widest range of any first-magnitude star. Betelgeuse is one of three stars that make up the Winter Triangle asterism, and it marks the center of the Winter Hexagon. It would be the brightest star in the sky if the human eye could view all wavelengths of radiation.

The star is classified as a red supergiant of spectral type M1-2 and is one of the largest and most luminous stars visible to the naked eye. If Betelgeuse were at the center of the Solar System, its surface would extend past the asteroid belt, wholly engulfing Mercury, Venus, Earth and Mars. Calculations of its mass range from slightly under ten to a little over twenty times that of the Sun. It is calculated to be 640 light-years away, yielding an absolute magnitude of about −6. Less than 10 million years old, Betelgeuse has evolved rapidly because of its high mass. Having been ejected from its birthplace in the Orion OB1 Association—which includes the stars in Orion's Belt—this crimson runaway has been observed moving through the interstellar medium at a supersonic speed of 30 km/s, creating a bow shock over 4 light-years wide. Currently in a late stage of stellar evolution, the supergiant is expected to explode as a supernova within the next million years.

In 1920, Betelgeuse became the second star (after the Sun) to have the angular size of its photosphere measured. Recent studies have reported an angular diameter (apparent size) ranging from 0.042 to 0.056 arcseconds, with the differences ascribed to the non-sphericity, limb darkening, pulsations, and varying appearance at different wavelengths. It is also surrounded by a complex, asymmetric envelope roughly 250 times the size of the star, caused by mass loss from the star itself. The angular diameter of Betelgeuse is only exceeded by R Doradus (and the Sun).

Nomenclature

α Orionis (Latinised to Alpha Orionis) is the star's Bayer designation. The traditional name Betelgeuse is derived from the Arabic إبط الجوزاء Ibt al-Jauzā', meaning "the axilla of Orion", or يد الجوزاء Yad al-Jauzā', meaning "the hand of Orion" (see below). In 2016, the International Astronomical Union organized a Working Group on Star Names (WGSN)[16] to catalog and standardize proper names for stars. The WGSN's first bulletin of July 2016[17] included a table of the first two batches of names approved by the WGSN; which included Betelgeuse for this star. It is now so entered in the IAU Catalog of Star Names.[18]

Observational history

Betelgeuse and its red coloration have been noted since antiquity; the classical astronomer Ptolemy described its color as ὑπόκιρρος (hypókirrhos), a term that was later described by a translator of Ulugh Beg's Zij-i Sultani as rubedo, Latin for "ruddiness".[19][20] In the nineteenth century, before modern systems of stellar classification, Angelo Secchi included Betelgeuse as one of the prototypes for his Class III (orange to red) stars.[21] By contrast, three centuries before Ptolemy, Chinese astronomers observed Betelgeuse as having a yellow coloration, suggesting that the star may have spent time as a yellow supergiant around the beginning of the common era,[22] a possibility given current research into the complex circumstellar environment of these stars.[23]

Nascent discoveries

The variation in Betelgeuse's brightness was first described in 1836 by Sir John Herschel, when he published his observations in Outlines of Astronomy. From 1836 to 1840, he noticed significant changes in magnitude when Betelgeuse outshone Rigel in October 1837 and again in November 1839.[24] A 10-year quiescent period followed; then in 1849, Herschel noted another short cycle of variability, which peaked in 1852. Later observers recorded unusually high maxima with an interval of years, but only small variations from 1957 to 1967. The records of the American Association of Variable Star Observers (AAVSO) show a maximum brightness of 0.2 in 1933 and 1942, and a minimum of 1.2, observed in 1927 and 1941.[25][26] This variability in brightness may explain why Johann Bayer, with the publication of his Uranometria in 1603, designated the star alpha as it may have rivaled the usually brighter Rigel (beta).[27] From Arctic latitudes, Betelgeuse's red colour and higher location in the sky than Rigel meant the Inuit regarded it as brighter, and one local name was Ulluriajjuaq "large star".[28]

In 1920, Albert Michelson and Francis Pease mounted a 6-meter interferometer on the front of the 2.5-meter telescope at Mount Wilson Observatory. Helped by John Anderson, the trio measured the angular diameter of Betelgeuse at 0.047", a figure which resulted in a diameter of 3.84 × 108 km (2.58 AU) based on the parallax value of 0.018".[29] However, limb darkening and measurement errors resulted in uncertainty about the accuracy of these measurements.

The 1950s and 1960s saw two developments that would impact stellar convection theory in red supergiants: the Stratoscope projects and the 1958 publication of Structure and Evolution of the Stars, principally the work of Martin Schwarzschild and his colleague at Princeton University, Richard Härm.[30][31] This book disseminated ideas on how to apply computer technologies to create stellar models, while the Stratoscope projects, by taking balloon-borne telescopes above the Earth's turbulence, produced some of the finest images of solar granules and sunspots ever seen, thus confirming the existence of convection in the solar atmosphere.[30]

Imaging breakthroughs

1988/9 UV HST images of Betelgeuse showing asymmetrical pulsations with corresponding spectral line profiles

Astronomers in the 1970s saw some major advances in astronomical imaging technology beginning with Antoine Labeyrie's invention of speckle interferometry, a process that significantly reduced the blurring effect caused by astronomical seeing. It increased the optical resolution of ground-based telescopes, allowing for more precise measurements of Betelgeuse's photosphere.[32][33] With improvements in infrared telescopy atop Mount Wilson, Mount Locke and Mauna Kea in Hawaii, astrophysicists began peering into the complex circumstellar shells surrounding the supergiant,[34][35][36] causing them to suspect the presence of huge gas bubbles resulting from convection.[37] But it was not until the late 1980s and early 1990s, when Betelgeuse became a regular target for aperture masking interferometry, that breakthroughs occurred in visible-light and infrared imaging. Pioneered by John E. Baldwin and colleagues of the Cavendish Astrophysics Group, the new technique employed a small mask with several holes in the telescope pupil plane, converting the aperture into an ad-hoc interferometric array.[38] The technique contributed some of the most accurate measurements of Betelgeuse while revealing bright spots on the star's photosphere.[39][40][41] These were the first optical and infrared images of a stellar disk other than the Sun, taken first from ground-based interferometers and later from higher-resolution observations of the COAST telescope. The "bright patches" or "hotspots" observed with these instruments appeared to corroborate a theory put forth by Schwarzschild decades earlier of massive convection cells dominating the stellar surface.[42][43]

In 1995, the Hubble Space Telescope's Faint Object Camera captured an ultraviolet image with a resolution superior to that obtained by ground-based interferometers—the first conventional-telescope image (or "direct-image" in NASA terminology) of the disk of another star.[44] Because ultraviolet light is absorbed by the Earth's atmosphere, observations at these wavelengths are best performed by space telescopes.[45] Like earlier pictures, this image contained a bright patch indicating a region in the southwestern quadrant 2000 K hotter than the stellar surface.[46] Subsequent ultraviolet spectra taken with the Goddard High Resolution Spectrograph suggested that the hot spot was one of Betelgeuse's poles of rotation. This would give the rotational axis an inclination of about 20° to the direction of Earth, and a position angle from celestial North of about 55°.[47]

Recent studies

In a study published in December 2000, the star's diameter was measured with the Infrared Spatial Interferometer (ISI) at mid-infrared wavelengths producing a limb-darkened estimate of 55.2 ± 0.5 milliarcseconds (mas)—a figure entirely consistent with Michelson's findings eighty years earlier.[29][48] At the time of its publication, the estimated parallax from the Hipparcos mission was 7.63 ± 1.64 mas, yielding an estimated radius for Betelgeuse of 3.6 AU. However, numerous interferometric studies in the near-infrared made at the Paranal Observatory in Chile argue for much tighter diameters. On 9 June 2009, it was announced that the star had shrunk by 15% since 1993 at an increasing rate without a significant diminution in magnitude.[49][50] Subsequent observations suggest that the apparent contraction may be due to shell activity in the star's extended atmosphere.[51]

In addition to the star's diameter, questions have arisen about the complex dynamics of Betelgeuse's extended atmosphere. The mass that makes up galaxies is recycled as stars are formed and destroyed, and red supergiants are major contributors, yet the process by which mass is lost remain a mystery.[52] With advances in interferometric methodologies, astronomers may be close to resolving this conundrum. In July 2009, images released by the European Southern Observatory, taken by the ground-based Very Large Telescope Interferometer (VLTI), showed a vast plume of gas extending 30 AU from the star into the surrounding atmosphere.[15] This mass ejection was equal to the distance between the Sun and Neptune and is one of multiple events occurring in Betelgeuse's surrounding atmosphere. Astronomers have identified at least six shells surrounding Betelgeuse. Solving the mystery of mass loss in the late stages of a star's evolution may reveal those factors that precipitate the explosive deaths of these stellar giants.[49]

Visibility

Image showing Betelgeuse and the dense nebulae of the Orion Molecular Cloud Complex (Rogelio Bernal Andreo)

In the night sky, Betelgeuse is easy to spot with the naked eye owing to its distinctive orange-red color. In the Northern Hemisphere, beginning in January of each year, it can be seen rising in the east just after sunset. By mid-September to mid-March (best in mid-December), it is visible to virtually every inhabited region of the globe, except for a few research stations in Antarctica at latitudes south of 82°. In May (moderate northern latitudes) or June (southern latitudes), the red supergiant can be seen briefly on the western horizon after sunset, reappearing again a few months later on the eastern horizon before sunrise. In the intermediate period (June–July) it is invisible to the naked eye (visible only with a telescope in daylight), unless around midday (when the Sun is below horizon) on Antarctic regions between 70° and 80° south latitude.

Betelgeuse is a variable star whose brightness ranges between 0.0 and 1.3. There are periods when it will surpass Procyon to become the seventh brightest star, and occasionally even brighter. At its faintest Betelgeuse can fall behind Deneb and Mimosa, themselves both slightly variable, to be the 20th brightest star.

Betelgeuse has a color index (B–V) of 1.85—a figure which points to its advanced "redness". The photosphere has an extended atmosphere, which displays strong lines of emission rather than absorption, a phenomenon that occurs when a star is surrounded by a thick gaseous envelope (rather than ionized). This extended gaseous atmosphere has been observed moving away from and towards Betelgeuse, depending on radial velocity fluctuations in the photosphere. Betelgeuse is the brightest near-infrared source in the sky with a J band magnitude of −2.99.[53] As a result, only about 13% of the star's radiant energy is emitted in the form of visible light. If human eyes were sensitive to radiation at all wavelengths, Betelgeuse would appear as the brightest star in the sky.[26]

Star system

Various catalogues list up to nine faint visual companions to Betelgeuse. They are at distances of about one to four arc-minutes and all are fainter than 10th magnitude.[54][55] Betelgeuse is generally considered to be a single isolated star and a runaway star, not currently associated with any cluster or star-forming region, although its birthplace is unclear.[56]

Two spectroscopic companions have been proposed to the red supergiant star. Analysis of polarization data from 1968 through 1983 indicated a close companion with a periodic orbit of about 2.1 years. Using speckle interferometry, the team concluded that the closer of the two companions was located at 0.06″±0.01″ (~9 AU) from the main star with a position angle (PA) of 273 degrees, an orbit that would potentially place it within the star's chromosphere. The more distant companion was estimated at 0.51″±0.01″ (~77 AU) with a PA of 278 degrees.[57][58] Further studies have found no evidence for these companions or have actively refuted their existence,[59] but the possibility of a close companion contributing to the overall flux has never been fully ruled out.[60] High resolution interferometry of Betelgeuse and its vicinity, far beyond the technology of the 1980s and 90s, have not detected any companions.[15][61]

Distance measurements

NRAO's Very Large Array used to derive Betelgeuse's 2008 distance estimate

Parallax is the apparent change of the position of an object, measured in seconds of arc, caused by the change of position of the observer of that object. As the Earth orbits the Sun, every star is seen to shift by a fraction of an arc second, which measure, combined with the baseline provided by the Earth's orbit gives the distance to that star. Since the first successful parallax measurement by Friedrich Bessel in 1838, astronomers have been puzzled by Betelgeuse's apparent distance. Knowledge of the star's distance improves the accuracy of other stellar parameters, such as luminosity that, when combined with an angular diameter, can be used to calculate the physical radius and effective temperature; luminosity and isotopic abundances can also be used to estimate the stellar age and mass.[9] In 1920, when the first interferometric studies were performed on the star's diameter, the assumed parallax was 0.0180 arcseconds. This equated to a distance of 56 parsecs (pc) or roughly 180 light-years (ly), producing not only an inaccurate radius for the star but every other stellar characteristic. Since then, there has been ongoing work to measure the distance of Betelgeuse, with proposed distances as high as 400 pc or about 1300 ly.[9]

Before the publication of the Hipparcos Catalogue (1997), there were two conflicting parallax measurements for Betelgeuse. The first, in 1991, gave a parallax of π = 9.8 ± 4.7 mas, yielding a distance of roughly 102 pc or 330 ly.[62] The second was the Hipparcos Input Catalogue (1993) with a trigonometric parallax of π = 5 ± 4 mas, a distance of 200 pc or 650 ly.[63] Given this uncertainty, researchers were adopting a wide range of distance estimates, leading to significant variances in the calculation of the star's attributes.[9]

The results from the Hipparcos mission were released in 1997. The measured parallax of Betelgeuse was π = 7.63 ± 1.64 mas, which equated to a distance of 131 pc or roughly 430 ly, and had a smaller reported error than previous measurements.[64] However, later evaluation of the Hipparcos parallax measurements for variable stars like Betelgeuse found that the uncertainty of these measurements had been underestimated.[65] In 2007, an improved figure of π = 6.55±0.83 was calculated, hence a much tighter error factor yielding a distance of roughly 152±20 pc or 520±73 ly.[3]

In 2008, using the Very Large Array (VLA), produced a radio solution of π = 5.07±1.10 mas, equalling a distance of 197±45 pc or 643±146 ly.[9] As the researcher, Harper, points out: "The revised Hipparcos parallax leads to a larger distance (152±20 pc) than the original; however, the astrometric solution still requires a significant cosmic noise of 2.4 mas. Given these results it is clear that the Hipparcos data still contain systematic errors of unknown origin." Although the radio data also have systematic errors, the Harper solution combines the datasets in the hope of mitigating such errors.[9] The European Space Agency's current Gaia mission may not improve over the measurements of Betelgeuse by the earlier Hipparcos mission as Betelgeuse is brighter than the approximately V=6 saturation limit of the mission's instruments.[66]

Variability

AAVSO V-band light curve of Betelgeuse (Alpha Orionis) from Dec 1988 to Aug 2002

Betelgeuse is classified as a semiregular variable star, indicating that some periodicity is noticeable in the brightness changes, but amplitudes may vary, cycles may have different lengths, and there may be standstills or periods of irregularity. It is placed in subgroup SRc; these are pulsating red supergiants with amplitudes around one magnitude and periods from tens to hundreds of days.[6]

Betelgeuse typically shows only small brightness changes near to magnitude +0.5, although at its extremes it can become as bright as magnitude 0.0 or as faint as magnitude +1.3. Betelgeuse is listed in the General Catalogue of Variable Stars with a possible period of 2,335 days.[6] More detailed analyses have shown a main period near 400 days and a longer secondary period around 2,100 days.[61][67]

Radial pulsations of red supergiants are well-modelled and show that periods of a few hundred days are typically due to fundamental and first overtone pulsation.[68] Lines in the spectrum of Betelgeuse show doppler shifts indicating radial velocity changes corresponding, very roughly, to the brightness changes. This demonstrates the nature of the pulsations in size, although corresponding temperature and spectral variations are not clearly seen.[69] Variations in the diameter of Betelgeuse have also been measured directly.[51]

The source of the long secondary periods is unknown, but they certainly aren't due to radial pulsations.[67] Interferometric observations of Betelgeuse have shown hotspots that are thought to be created by massive convection cells, a significant fraction of the diameter of the star and each emitting 5-10% of the total light of the star.[60][61] One theory to explain long secondary periods is that they are caused by the evolution of such cells combined with the rotation of the star.[67] Other theories include close binary interactions, chromospheric magnetic activity influencing mass loss, or non-radial pulsations such as g-modes.[70]

In addition to the discrete dominant periods, small-amplitude stochastic variations are seen. It is proposed that this is due to granulation, similar to the same effect on the sun but on a much larger scale.[67]

Diameter

On 13 December 1920, Betelgeuse became the first star outside the Solar System to have the angular size of its photosphere measured.[29] Although interferometry was still in its infancy, the experiment proved a success. The researchers, using a uniform disk model, determined that Betelgeuse had a diameter of 0.047 arcseconds, although the stellar disk was likely 17% larger due to the limb darkening, resulting in an estimate for its angular diameter of about 0.055".[29][50] Since then, other studies have produced angular diameters that range from 0.042 to 0.069 arcseconds.[33][48][71] Combining these data with historical distance estimates of 180 to 815 ly yields a projected radius of the stellar disk of anywhere from 1.2 to 8.9 AU.[note 1] Using the Solar System for comparison, the orbit of Mars is about 1.5 AU, Ceres in the asteroid belt 2.7 AU, Jupiter 5.5 AU—so, assuming Betelgeuse occupying the place of the Sun, its photosphere might extend beyond the Jovian orbit, not quite reaching Saturn at 9.5 AU.

Radio image from 1998 (pre-Harper) showing the size of Betelgeuse's photosphere (circle) and the effect of convective forces on the star's atmosphere

The precise diameter has been hard to define for several reasons:

  1. Betelgeuse is a pulsating star, so its diameter changes with time;
  2. The star has no definable "edge" as limb darkening causes the optical emissions to vary in color and decrease the farther one extends out from the center;
  3. Betelgeuse is surrounded by a circumstellar envelope composed of matter ejected from the star—matter which absorbs and emits light—making it difficult to define the photosphere of the star;[49]
  4. Measurements can be taken at varying wavelengths within the electromagnetic spectrum and the difference in reported diameters can be as much as 30–35%, yet comparing one finding with another is difficult as the star's apparent size differs depending on the wavelength used.[49] Studies have shown that the measured angular diameter is considerably larger at ultraviolet wavelengths, decreases through the visible to a minimum in the near-infrared, and increase again in the mid-infrared spectrum;[44][72][73]
  5. Atmospheric twinkling limits the resolution obtainable from ground-based telescopes since turbulence degrades angular resolution.[39]

To overcome these challenges, researchers have employed various solutions. Astronomical interferometry, first conceived by Hippolyte Fizeau in 1868, was the seminal concept that has enabled major improvements in modern telescopy and led to the creation of the Michelson interferometer in the 1880s, and the first successful measurement of Betelgeuse.[74] Just as human depth perception increases when two eyes instead of one perceive an object, Fizeau proposed the observation of stars through two apertures instead of one to obtain interferences that would furnish information on the star's spatial intensity distribution. The science evolved quickly and multiple-aperture interferometers are now used to capture speckled images, which are synthesized using Fourier analysis to produce a portrait of high resolution.[75] It was this methodology that identified the hotspots on Betelgeuse in the 1990s.[76] Other technological breakthroughs include adaptive optics,[77] space observatories like Hipparcos, Hubble and Spitzer,[44][78] and the Astronomical Multi-BEam Recombiner (AMBER), which combines the beams of three telescopes simultaneously, allowing researchers to achieve milliarcsecond spatial resolution.[79][80]

Which part of the electromagnetic spectrum—the visible, near-infrared (NIR) or mid-infrared (MIR)—produces the most accurate angular measurement is still debated.[note 1] In 1996, Betelgeuse was shown to have a uniform disk of 56.6 ± 1.0 mas. In 2000, the SSL team produced another measure of 54.7 ± 0.3 mas, ignoring any possible contribution from hotspots, which are less noticeable in the mid-infrared.[48] Also included was a theoretical allowance for limb darkening, yielding a diameter of 55.2 ± 0.5 mas. The earlier estimate equates to a radius of roughly 5.6 AU or 1200 R, assuming the 2008 Harper distance of 197.0 ± 45 pc,[12] a figure roughly the size of the Jovian orbit of 5.5 AU, published in 2009 in Astronomy Magazine and a year later in NASA's Astronomy Picture of the Day.[81][82]

A team of astronomers working in the near-infrared announced in 2004, that the more accurate photospheric measurement was 43.33 ± 0.04 mas.[72] The study also put forth an explanation as to why varying wavelengths from the visible to mid-infrared produce different diameters: the star is seen through a thick, warm extended atmosphere. At short wavelengths (the visible spectrum) the atmosphere scatters light, thus slightly increasing the star's diameter. At near-infrared wavelengths (K and L bands), the scattering is negligible, so the classical photosphere can be directly seen; in the mid-infrared the scattering increases once more, causing the thermal emission of the warm atmosphere to increase the apparent diameter.[72]

Infrared image of Betelgeuse, Meissa and Bellatrix with surrounding nebulae

Studies with the IOTA and VLTI published in 2009 brought strong support to Perrin's analysis and yielded diameters ranging from 42.57 to 44.28 mas with comparatively insignificant margins of error.[60][83] In 2011, a third estimate in the near-infrared corroborating the 2009 numbers, this time showing a limb-darkened disk diameter of 42.49 ± 0.06 mas.[84] Consequently, if one combines the smaller Hipparcos distance from van Leeuwen of 152 ± 20 pc with Perrin's angular measurement of 43.33 mas, a near-infrared photospheric estimate would equate to about 3.4 AU or 730 R.[85] A 2014 paper derives an angular diameter of 42.28 mas (equivalent to a 41.01 mas uniform disc) using H and K band observations made with the VLTI AMBER instrument.[86]

Central to this discussion, it was announced in 2009, that the radius of Betelgeuse had shrunk from 1993 to 2009 by 15%, with the 2008 angular measurement equal to 47.0 mas, not too far from Perrin's estimate.[50][87] Unlike most earlier papers, this study encompassed a 15-year period at one specific wavelength. Earlier studies have typically lasted one to two years by comparison and have explored multiple wavelengths, often yielding vastly different results. The diminution in Betelgeuse's apparent size equates to a range of values between 56.0 ± 0.1 mas seen in 1993 to 47.0 ± 0.1 mas seen in 2008—a contraction of almost 0.9 AU in 15 years. What is not fully known is whether this observation is evidence of a rhythmic expansion and contraction of the star's photosphere as astronomers have theorized, and if so, what the periodic cycle might be, although Townes suggested that if a cycle does exist, it is probably a few decades long.[50] Other possible explanations are photospheric protrusions due to convection or a star that is not spherical but asymmetric causing the appearance of expansion and contraction as the star rotates on its axis.[88]

The debate about differences between measurements in the mid-infrared, which suggest a possible expansion and contraction of the star, and the near-infrared, which advocates a relatively constant photospheric diameter, remains to be resolved. In a paper published in 2012, the Berkeley team reported that their measurements were "dominated by the behavior of cool, optically thick material above the stellar photosphere," indicating that the apparent expansion and contraction may be due to activity in the star's outer shells and not the photosphere itself.[51] This conclusion, if further corroborated, would suggest an average angular diameter for Betelgeuse closer to Perrin's estimate at 43.33 arcseconds, hence a stellar radius of about 3.4 AU (730 R) assuming the shorter Hipparcos distance of 498 ± 73 ly in lieu of Harper's estimate at 643 ± 146 ly. The Gaia spacecraft may clarify assumptions presently used in calculating the size of Betelgeuse's stellar disk.

Once considered as having the largest angular diameter of any star in the sky after the Sun, Betelgeuse lost that distinction in 1997 when a group of astronomers measured R Doradus with a diameter of 57.0 ± 0.5 mas, although R Doradus, being much closer to Earth at about 200 ly, has a linear diameter roughly one-third that of Betelgeuse.[89]

The generally reported radii of large cool stars are Rosseland radii, defined as the radius of the photosphere at a specific optical depth of two thirds. This corresponds to the radius calculated from the effective temperature and bolometric luminosity. The Rosseland radius differs from directly measured radii, but there are widely used conversion factors depending on the wavelength used for the angular measurements.[90] For example, a measured angular diameter of 55.6 mas corresponds to a Rosseland mean diameter of 56.2 mas. The Rosseland radius derived from angular measurements of the star's photosphere rather than an extended envelope is 887 R.[11]

Properties

Relative sizes of the planets in the Solar System and several stars, including Betelgeuse
1. Mercury < Mars < Venus < Earth
2. Earth < Neptune < Uranus < Saturn < Jupiter
3. Jupiter < Proxima Centauri < Sun < Sirius
4. Sirius < Pollux < Arcturus < Aldebaran
5. Aldebaran < Rigel < Antares < Betelgeuse
6. Betelgeuse < VY CMa < NML Cyg < UY Sct.

Betelgeuse is a very large, luminous but cool star classified as an M1-2 Ia-ab red supergiant. The letter "M" in this designation means that it is a red star belonging to the M spectral class and therefore has a relatively low photospheric temperature; the "Ia-ab" suffix luminosity class indicates that it is an intermediate-luminosity supergiant, with properties partway between a normal supergiant and a luminous supergiant. Since 1943, the spectrum of Betelgeuse has served as one of the stable anchor points by which other stars are classified.[91]

Uncertainty in the star's surface temperature, diameter, and distance make it difficult to achieve a precise measurement of Betelgeuse's luminosity, but research from 2012 quotes a luminosity of around 126000 L, assuming a distance of 200 pc.[92] Studies since 2001 report effective temperatures ranging from 3250 to 3690 K. Values outside this range have previously been reported, and much of the variation is believed to be real, due to pulsations in the atmosphere.[11] The star is also a slow rotator and the most recent velocity recorded was 5 km/s—[15] much slower than Antares which has a rotational velocity of 20 km/s.[93] The rotation period depends on Betelgeuse's size and orientation to Earth, but it has been calculated to take 8.4 years to turn on its axis.[11]

In 2002, astronomers using computer simulations speculated that Betelgeuse might exhibit magnetic activity in its extended atmosphere, a factor where even moderately strong fields could have a meaningful influence over the star's dust, wind and mass-loss properties.[94] A series of spectropolarimetric observations obtained in 2010 with the Bernard Lyot Telescope at Pic du Midi Observatory revealed the presence of a weak magnetic field at the surface of Betelgeuse, suggesting that the giant convective motions of supergiant stars are able to trigger the onset of a small-scale dynamo effect.[95]

Mass

Betelgeuse has no known orbital companions, so its mass cannot be calculated by that direct method. Modern mass estimates from theoretical modelling have produced values from 9.5 - 21 M,[10] with values ranging from 5 M to 30 M from older studies.[96] It has been calculated that Betelgeuse began its life as a star of 15 to 20 M, based on a solar luminosity of 90000150000.[12] A novel method of determining the supergiant's mass was proposed in 2011, arguing for a current stellar mass of 11.6 M with an upper limit of 16.6 and lower of 7.7 M, based on observations of the star's intensity profile from narrow H-band interferometry and using a photospheric measurement of roughly 4.3 AU or 955 R.[10] Model fitting to evolutionary tracks give a current mass of 19.4 - 19.7 M, from an initial mass of 20 M.[11]

Motion

Orion OB1 Association

The kinematics of Betelgeuse are complex. The age of Class M supergiants with an initial mass of 20 M is roughly 10 million years.[9][97] Starting from its present position and motion a projection back in time would place Betelgeuse around 290 parsecs farther from the galactic plane—an implausible location, as there is no star formation region there. Moreover, Betelgeuse's projected pathway does not appear to intersect with the 25 Ori subassociation or the far younger Orion Nebula Cluster (ONC, also known as Ori OB1d), particularly since Very Long Baseline Array astrometry yields a distance from Betelgeuse to the ONC of between 389 and 414 parsecs. Consequently, it is likely that Betelgeuse has not always had its current motion through space but has changed course at one time or another, possibly the result of a nearby stellar explosion.[9][98] An observation by the Herschel Space Observatory in January 2013 revealed that the star's winds are crashing against the surrounding interstellar medium.[99]

The most likely star-formation scenario for Betelgeuse is that it is a runaway star from the Orion OB1 Association. Originally a member of a high-mass multiple system within Ori OB1a, Betelgeuse was probably formed about 10–12 million years ago,[100] but has evolved rapidly due to its high mass.[9]

Like many young stars in Orion whose mass is greater than 10 M, Betelgeuse will use its fuel quickly and not live long. On the Hertzsprung-Russell diagram, Betelgeuse has moved off the main sequence and has swelled and cooled to become a red supergiant. Although young, Betelgeuse has exhausted the hydrogen in its core, causing the core to contract under the force of gravity into a hotter and denser state. As a result, it has begun to fuse helium into carbon and oxygen and has ignited a hydrogen shell outside the core. The hydrogen-burning shell and the contracting core cause the outer envelope to expand and cool. Its mass is such that the star will eventually fuse higher elements through neon, magnesium, and silicon all the way to iron, at which point it will collapse and explode, probably as a type II supernova.[101][102]

Density

Wembley Stadium - the center circle (9.15 m radius) is a close analogy for the Earth's orbit around the Sun, while the air in the stadium is far denser than Betelgeuse.

As an early M-type supergiant, Betelgeuse is one of the largest, most luminous and yet one of the most ethereal stars known. A radius of 5.5 AU is roughly 1180 times the radius of the Sun—able to contain over 2 quadrillion Earths (2.15 × 1015) or more than 1.6 billion (1.65 × 109) Suns. That is the equivalent of Betelgeuse being a football stadium like Wembley Stadium in London with the Earth a tiny pearl, 1 millimeter in diameter, orbiting a Sun the size of a mango.[note 2] Moreover, observations from 2009 of Betelgeuse exhibiting a 15% contraction in angular diameter would equate to a reduction of the star's radius from about 5.5 to 4.6 AU, assuming that the photosphere is a perfect sphere. A reduction of this magnitude would correspond to a diminution in photospheric volume of about 41%.[note 3]

Not only is the photosphere enormous, but the star is surrounded by a complex circumstellar environment where light could take over three years to escape.[103] In the outer reaches of the photosphere the density is extremely low, yet the total mass of the star is believed to be no more than 20 M. Consequently, the average density is less than twelve parts per billion (1.119×10−8) that of the Sun. Such star matter is so tenuous that Betelgeuse has often been called a "red-hot vacuum".[25][26]

Circumstellar dynamics

Image from ESO's Very Large Telescope showing the stellar disk and an extended atmosphere with a previously unknown plume of surrounding gas

In the late phase of stellar evolution, massive stars like Betelgeuse exhibit high rates of mass loss, possibly as much as 1 M every 10000 years, resulting in a complex circumstellar environment that is constantly in flux. In a 2009 paper, stellar mass loss was cited as the "key to understanding the evolution of the universe from the earliest cosmological times to the current epoch, and of planet formation and the formation of life itself".[104] However, the physical mechanism is not well understood.[85] When Schwarzschild first proposed his theory of huge convection cells, he argued it was the likely cause of mass loss in evolved supergiants like Betelgeuse.[43] Recent work has corroborated this hypothesis, yet there are still uncertainties about the structure of their convection, the mechanism of their mass loss, the way dust forms in their extended atmosphere, and the conditions which precipitate their dramatic finale as a type II supernova.[85] In 2001, Graham Harper estimated a stellar wind at 0.03 M every 10000 years,[105] but research since 2009 has provided evidence of episodic mass loss making any total figure for Betelgeuse uncertain.[106] Current observations suggest that a star like Betelgeuse may spend a portion of its lifetime as a red supergiant, but then cross back across the H-R diagram, pass once again through a brief yellow supergiant phase and then explode as a blue supergiant or Wolf-Rayet star.[23]

Artist's rendering from ESO showing Betelgeuse with a gigantic bubble boiling on its surface and a radiant plume of gas being ejected to at least six photospheric radii or roughly the orbit of Neptune

Astronomers may be close to solving this mystery. They noticed a large plume of gas extending at least six times its stellar radius indicating that Betelgeuse is not shedding matter evenly in all directions.[15] The plume's presence implies that the spherical symmetry of the star's photosphere, often observed in the infrared, is not preserved in its close environment. Asymmetries on the stellar disk had been reported at different wavelengths. However, due to the refined capabilities of the NACO adaptive optics on the VLT, these asymmetries have come into focus. The two mechanisms that could cause such asymmetrical mass loss, were large-scale convection cells or polar mass loss, possibly due to rotation.[15] Probing deeper with ESO's AMBER, gas in the supergiant's extended atmosphere has been observed vigorously moving up and down, creating bubbles as large as the supergiant itself, leading his team to conclude that such stellar upheaval is behind the massive plume ejection observed by Kervella.[106]

Asymmetric shells

In addition to the photosphere, six other components of Betelgeuse's atmosphere have now been identified. They are a molecular environment otherwise known as the MOLsphere, a gaseous envelope, a chromosphere, a dust environment and two outer shells (S1 and S2) composed of carbon monoxide (CO). Some of these elements are known to be asymmetric while others overlap.[60]

Exterior view of ESO's Very Large Telescope (VLT) in Paranal, Chile

At about 0.45 stellar radii (~2–3 AU) above the photosphere there may lie a molecular layer known as the MOLsphere or molecular environment. Studies show it to be composed of water vapor and carbon monoxide with an effective temperature of about 1500±500 K.[60][107] Water vapor had been originally detected in the supergiant's spectrum in the 1960s with the two Stratoscope projects but had been ignored for decades. The MOLsphere may also contain SiO and Al2O3—molecules which could explain the formation of dust particles.

Interior view of one of the four 8.2-meter Unit Telescopes at ESO's VLT

The asymmetric gaseous envelope, another cooler region, extends for several radii (~10–40 AU) from the photosphere. It is enriched in oxygen and especially in nitrogen relative to carbon. These composition anomalies are likely caused by contamination by CNO-processed material from the inside of Betelgeuse.[60][108]

Radio-telescope images taken in 1998 confirm that Betelgeuse has a highly complex atmosphere,[109] with a temperature of 3450±850 K, similar to that recorded on the star's surface but much lower than surrounding gas in the same region.[109][110] The VLA images also show this lower-temperature gas progressively cools as it extends outward. Although unexpected, it turns out to be the most abundant constituent of Betelgeuse's atmosphere. "This alters our basic understanding of red-supergiant star atmospheres", explained Jeremy Lim, the team's leader. "Instead of the star's atmosphere expanding uniformly due to gas heated to high temperatures near its surface, it now appears that several giant convection cells propel gas from the star's surface into its atmosphere."[109] This is the same region in which Kervella's 2009 finding of a bright plume, possibly containing carbon and nitrogen and extending at least six photospheric radii in the southwest direction of the star, is believed to exist.[60]

The chromosphere was directly imaged by the Faint Object Camera on board the Hubble Space Telescope in ultraviolet wavelengths. The images also revealed a bright area in the southwest quadrant of the disk.[111] The average radius of the chromosphere in 1996 was about 2.2 times the optical disk (~10 AU) and was reported to have a temperature no higher than 5500 K.[60][112] However, in 2004 observations with the STIS, Hubble's high-precision spectrometer, pointed to the existence of warm chromospheric plasma at least one arcsecond away from the star. At a distance of 197 pc, the size of the chromosphere could be up to 200 AU.[111] The observations have conclusively demonstrated that the warm chromospheric plasma spatially overlaps and coexists with cool gas in Betelgeuse's gaseous envelope as well as with the dust in its circumstellar dust shells (see below).[60][111]

This infrared image from the ESO's VLT shows complex shells of gas and dust around Betelgeuse - the tiny red circle in the middle is the size of the photosphere.

The first claim of a dust shell surrounding Betelgeuse was put forth in 1977 when it was noted that dust shells around mature stars often emit large amounts of radiation in excess of the photospheric contribution. Using heterodyne interferometry, it was concluded that the red supergiant emits most of its excess radiation from positions beyond 12 stellar radii or roughly the distance of the Kuiper belt at 50 to 60 AU, which depends on the assumed stellar radius.[34][60] Since then, there have been studies done of this dust envelope at varying wavelengths yielding decidedly different results. Studies from the 1990s have estimated the inner radius of the dust shell anywhere from 0.5 to 1.0 arcseconds, or 100 to 200 AU.[113][114] These studies point out that the dust environment surrounding Betelgeuse is not static. In 1994, it was reported that Betelgeuse undergoes sporadic decades long dust production, followed by inactivity. In 1997, significant changes in the dust shell's morphology in one year were noted, suggesting that the shell is asymmetrically illuminated by a stellar radiation field strongly affected by the existence of photospheric hotspots.[113] The 1984 report of a giant asymmetric dust shell 1 pc (206265 AU) has not been corroborated by recent studies, although another published the same year said that three dust shells were found extending four light-years from one side of the decaying star, suggesting that Betelgeuse sheds its outer layers as it moves.[103][115]

Although the exact size of the two outer CO shells remains elusive, preliminary estimates suggest that one shell extends from about 1.5 to 4.0 arcseconds and the other expands as far as 7.0 arcseconds.[116] Assuming the Jovian orbit of 5.5 AU as the star radius, the inner shell would extend roughly 50 to 150 stellar radii (~300 to 800 AU) with the outer one as far as 250 stellar radii (~1400 AU). The Sun's heliopause is estimated at about 100 AU, so the size of this outer shell would be almost fourteen times the size of the Solar System.

Supersonic bow shock

Betelgeuse is travelling supersonically through the interstellar medium at a speed of 30 km per second (i.e. ~6.3 AU per year) creating a bow shock.[117][118] The shock is not created by the star, but by its powerful stellar wind as it ejects vast amounts of gas into the interstellar medium at a speed of 17 km/s, heating the material surrounding the star, thereby making it visible in infrared light.[119] Because Betelgeuse is so bright, it was only in 1997 that the bow shock was first imaged. The cometary structure is estimated to be at least 1 parsec wide, assuming a distance of 643 light-years.[120]

Hydrodynamic simulations of the bow shock made in 2012 indicate that it is very young—less than 30000 years old—suggesting two possibilities: that Betelgeuse moved into a region of the interstellar medium with different properties only recently or that Betelgeuse has undergone a significant transformation producing a changed stellar wind.[121] A 2012 paper, proposed that this phenomenon was caused by Betelgeuse transitioning from a blue supergiant (BSG) to a red supergiant (RSG). There is evidence that in the late evolutionary stage of a star like Betelgeuse, such stars "may undergo rapid transitions from red to blue and vice versa on the Hertzsprung-Russell diagram, with accompanying rapid changes to their stellar winds and bow shocks."[117][122] Moreover, if future research bears out this hypothesis, Betelgeuse may prove to have traveled close to 200000 AU as a red supergiant scattering as much as 3 M along its trajectory.

Evolution

Hertzsprung–Russell diagram identifying supergiants like Betelgeuse that have moved off the main sequence

Betelgeuse is a red supergiant that has evolved from an O-type main sequence star. Its core will eventually collapse, producing a supernova explosion and leaving behind a compact remnant. The details depend on the exact initial mass and other physical properties of that main sequence star.

So far

The initial mass of Betelgeuse can only be estimated by testing different stellar evolutionary models to match its current observed properties. The unknowns of both the models and the current properties mean that there is considerable uncertainty in Betelgeuse's initial appearance, but its mass is usually estimated to have been in the range of 10-25 M, with modern models finding values of 15-20 M. Its chemical makeup can be reasonably assumed to have been around 70% hydrogen, 28% helium, and 2.4% heavy elements, slightly more metal-rich than the sun but otherwise similar. The initial rotation rate is more uncertain, but models with slow to moderate initial rotation rates produce the best matches to Betelgeuse's current proprties.[11][56][123] That main sequence version of Betelgeuse would have been a hot luminous star with a spectral type such as O9V.[92]

A 15 M star would take between 11.5 and 15 million years to reach the red supergiant stage, with more rapidly rotating stars taking the longest.[123] Rapidly-rotating 20 M stars take only 9.3 million years to reach the red supergiant stage, while 20 M stars with slow rotation take only 8.1 million years.[56] These form the best estimates of Betelgeuse's current age, with a preferred age since the zero age main sequence of 8.0 - 8.5 million years for a 20 M star with no rotation.[11]

The time spent so far as a red supergiant can be estimated by comparing mass loss rates to the observed circumstellar material, as well as the abundances of heavy elements at the surface. Estimates range from 20,000 years to a maximum of 140,000 years. Betelgeuse appears to undergo short periods of heavy mass loss and is a runaway star moving rapidly through space, so comparisons of its current mass loss to the total lost mass are difficult.[11][56] The surface of Betelgeuse shows enhancement of nitrogen, relatively low levels of carbon, and a high proportion of 13C relative to 12C, all indicative of a star that has experienced the first dredge-up. However the first dredge-up occurs soon after a star reaches the red supergiant phase and so this only means that Betelgeuse has been a red supergiant for at least a few thousand years. The best prediction is that Betelgeuse has already spent around 40,000 years as a red supergiant,[11] having left the main sequence perhaps one million years ago.[123]

The current mass can be estimated from evolutionary models from the initial mass and the expected mass lost so far. For Betelgeuse, the total mass lost is predicted to be no more than about one M, giving a current mass of 19.4-19.7 M, considerably higher than estimated by other means such as pulsational properties or limb-darkening models.[11]

Approaching supernova

Celestia depiction of Orion as it might appear from Earth when Betelgeuse explodes as a supernova

All stars more massive than about 10 M are expected to end their lives when their core collapses, typically producing a supernova explosion. Up to about 15 M, a type II-P supernova is always produced from the red supergiant stage.[123] More massive stars can lose mass quickly enough that they evolve towards higher temperatures before their cores can collapse, particularly for rotating stars and models with especially high mass loss rates. These stars can produce type II-L or type IIb supernovae from yellow or blue supergiants, or type Ib/c supernovae from Wolf-Rayet stars.[124] Models of rotating 20 M stars predict a peculiar type II supernova similar to SN 1987A from a blue supergiant progenitor.[123] On the other hand, non-rotating 20 M models predict a type II-P supernova from a red supergiant progenitor.[11]

The time until Betelgeuse explodes depends on the predicted initial conditions and on the estimate of the time already spent as a red supergiant. The total lifetime from the start of the red supergiant phase to core collapse varies from about 300,000 years for a rotating 20 M star, 550,000 years for a rotating 20 M star, up to a million years for a non-rotating 15 M star. Given the estimated time since Betelgeuse became a red supergiant, estimates of its remaining lifetime range from a "best guess" of under 100,000 years for a non-rotating 20 M model to far longer for rotating models or lower mass stars.[11][123] Betelgeuse's suspected birthplace in the Orion OB1 Association is the location of several previous supernovae. It is believed that runaway stars may be caused by supernovae, and there is strong evidence that OB stars μ Columbae, AE Aurigae and 53 Arietis all originated from such explosions in Ori OB1 2.2, 2.7 and 4.9 million years ago.[98]

A typical type II-P supernova emits 2×1046 J of neutrinos and produces an explosion with a kinetic energy of 2×1044 J. As seen from Earth, it would have a peak apparent magnitude of about 12.4.[11] It may outshine the full moon and would be easily visible in daylight. This type of supernova would remain at roughly constant brightness for 2–3 months before rapidly dimming. The visible light is produced mainly by the radioactive decay of cobalt, and maintains its brightness due to the increasing transparency of the cooling hydrogen ejected by the supernova.[125]

Due to misunderstandings caused by the 2009 publication of the star's 15% contraction, apparently of its outer atmosphere,[49][81] Betelgeuse has frequently been the subject of scare stories and rumors suggesting that it will explode within a year, leading to exaggerated claims about the consequences of such an event.[126][127] The timing and prevalence of these rumors have been linked to broader misconceptions of astronomy, particularly to doomsday predictions relating to the Mayan calendar.[128][129] Betelgeuse is not likely to produce a gamma-ray burst and is not close enough for its x-rays, ultraviolet radiation, or ejected material to cause significant effects on Earth.[11]

Remnant

Following Betelgeuse's supernova, a small dense remnant will be left behind, either a neutron star or black hole. This is predicted to be a neutron star of approximately 1.5 M.[11] Only more massive stars, or low metallicity stars with lower mass loss, would produce a black hole.[124]

Ethnological attributes

Spelling and pronunciation

Betelgeuse has been known as Betelgeux,[1] and in German Beteigeuze[130] (according to Bode).[131][132] Betelgeux and Betelgeuze were used until the early 20th century, when the spelling Betelgeuse became universal.[133] There is no consensus for the correct pronunciation of the name,[134] and pronunciations for the star are as varied as its spellings:

Etymology

Betelgeuse is often mistranslated as "armpit of the central one".[136] In his 1899 work Star-Names and Their Meanings, American amateur naturalist Richard Hinckley Allen stated the derivation was from the ابط الجوزاء Ibṭ al-Jauzah, which he claimed degenerated into a number of forms including Bed Elgueze, Beit Algueze, Bet El-gueze, Beteigeuze and more, to the forms Betelgeuse, Betelguese, Betelgueze and Betelgeux. The star was named Beldengeuze in the Alfonsine Tables,[137] and Italian Jesuit priest and astronomer Giovanni Battista Riccioli had called it Bectelgeuze or Bedalgeuze.[19] Paul Kunitzsch, Professor of Arabic Studies at the University of Munich, refuted Allen's derivation and instead proposed that the full name is a corruption of the Arabic يد الجوزاء Yad al-Jauzā' meaning "the Hand of al-Jauzā'", i.e., Orion.[138] European mistransliteration into medieval Latin led to the first character y (, with two dots underneath) being misread as a b (, with only one dot underneath). During the Renaissance, the star's name was written as بيت الجوزاء Bait al-Jauzā' ("house of Orion") or بط الجوزاء Baţ al-Jauzā', incorrectly thought to mean "armpit of Orion" (a true translation of "armpit" would be ابط, transliterated as Ibţ). This led to the modern rendering as Betelgeuse.[139] Other writers have since accepted Kunitzsch's explanation.[102]

The last part of the name, "-elgeuse", comes from the Arabic الجوزاء al-Jauzā', a historical Arabic name of the constellation Orion, a feminine name in old Arabian legend, and of uncertain meaning. Because جوز j-w-z, the root of jauzā', means "middle", al-Jauzā' roughly means "the Central One". Later, al-Jauzā' was also designated as the scientific Arabic name for Orion and for Gemini. The modern Arabic name for Orion is الجبار al-Jabbār ("the Giant"), although the use of الجوزاء al-Jauzā' in the name of the star has continued.[139] The 17th-century English translator Edmund Chilmead gave it the name Ied Algeuze ("Orion's Hand"), from Christmannus.[19] Other Arabic names recorded include Al Yad al Yamnā ("the Right Hand"), Al Dhira ("the Arm"), and Al Mankib ("the Shoulder"), all appended to "of the giant",[19] as منكب الجوزاء Mankib al Jauzā'.

Dunhuang Star Chart, circa AD 700, showing 参宿四 Shēnxiùsì (Betelgeuse), the Fourth Star of the constellation of Three Stars

Other names

Other names for Betelgeuse included the Persian Bašn "the Arm", and Coptic Klaria "an Armlet".[19] Bahu was its Sanskrit name, as part of a Hindu understanding of the constellation as a running antelope or stag.[19] In traditional Chinese astronomy, Betelgeuse was known as 参宿四 (Shēnxiùsì, the Fourth Star of the constellation of Three Stars)[140] as the Chinese constellation 参宿 originally referred to the three stars in the girdle of Orion. This constellation was ultimately expanded to ten stars, but the earlier name stuck.[141] In Japan, the Taira or Heike clan adopted Betelgeuse and its red color as its symbol, calling the star Heike-boshi, (平家星), while the Minamoto or Genji clan had chosen Rigel and its white color. The two powerful families fought a legendary war in Japanese history, the stars seen as facing each other off and only kept apart by the Belt.[142][143]

In Tahitian lore, Betelgeuse was one of the pillars propping up the sky, known as Anâ-varu, the pillar to sit by. It was also called Ta'urua-nui-o-Mere "Great festivity in parental yearnings".[144] A Hawaiian term for it was Kaulua-koko "brilliant red star".[145] The Lacandon people of Central America knew it as chäk tulix "red butterfly".[146]

Mythology

With the history of astronomy intimately associated with mythology and astrology before the scientific revolution, the red star, like the planet Mars that derives its name from a Roman war god, has been closely associated with the martial archetype of conquest for millennia, and by extension, the motif of death and rebirth.[19] Other cultures have produced different myths. Stephen R. Wilk has proposed the constellation of Orion could have represented the Greek mythological figure Pelops, who had an artificial shoulder of ivory made for him, with Betelgeuse as the shoulder, its color reminiscent of the reddish yellow sheen of ivory.[24]

In the Americas, Betelgeuse signifies a severed limb of a man-figure (Orion)—the Taulipang of Brazil know the constellation as Zililkawai, a hero whose leg was cut off by his wife, with the variable light from Betelgeuse linked to the severing of the limb. Similarly, the Lakota people of North America see it as a chief whose arm has been severed.[24] The Wardaman people of northern Australia knew the star as Ya-jungin "Owl Eyes Flicking", its variable light signifying its intermittent watching of ceremonies led by the Red Kangaroo Leader Rigel.[147] In South African mythology, Betelgeuse was perceived as a lion casting a predatory gaze toward the three zebras represented by Orion's Belt.[148]

A Sanskrit name for Betelgeuse is ārdrā "the moist one", eponymous of the Ardra lunar mansion in Hindu astrology.[149] The Rigvedic God of storms Rudra presided over the star; this association was linked by 19th-century star enthusiast Richard Hinckley Allen to Orion's stormy nature.[19] The constellations in Macedonian folklore represented agricultural items and animals, reflecting their village way of life. To them, Betelgeuse was Orach "the ploughman", alongside the rest of Orion which depicted a plough with oxen. The rising of Betelgeuse at around 3 am in late summer and autumn signified the time for village men to go to the fields and plough.[150] To the Inuit, the appearance of Betelgeuse and Bellatrix high in the southern sky after sunset marked the beginning of spring and lengthening days in late February and early March. The two stars were known as Akuttujuuk "those (two) placed far apart", referring to the distance between them, mainly to people from North Baffin Island and Melville Peninsula.[28]

The opposed locations of Orion and Scorpius, with their corresponding bright variable red stars Betelgeuse and Antares, were noted by ancient cultures around the world. The setting of Orion and rising of Scorpius signify the death of Orion by the scorpion. In China they signify brothers and rivals Shen and Shang.[24] The Batak of Sumatra marked their New Year with the first new moon after the sinking of Orion's Belt below the horizon, at which point Betelgeuse remained "like the tail of a rooster". The positions of Betelgeuse and Antares at opposite ends of the celestial sky were considered significant and their constellations were seen as a pair of scorpions. Scorpion days marked as nights that both constellations could be seen.[151]

The star's unusual name inspired the title of the 1988 film Beetlejuice, and script writer Michael McDowell was impressed by how many people made the connection. He added that they had received a suggestion the sequel be named Sanduleak-69 202 after the former star of SN 1987A.[133] In August Derleth's short story "The Dweller in the Darkness" set in H. P. Lovecraft's Cthulhu Mythos, Betelgeuse is the home of the "benign" Elder Gods.[152] The identity of the red star Borgil mentioned in Lord of the Rings was much debated; Aldebaran, Betelgeuse and the planet Mars were touted as candidates. Professor Kristine Larsen has concluded the evidence points to it being Aldebaran as it precedes Menelvagor (Orion).[153] Astronomy writer Robert Burnham, Jr. proposed the term padparadaschah which denotes a rare orange sapphire in India, for the star.[133] In the popular science fiction series The Hitchhiker's Guide to the Galaxy by Douglas Adams, Ford Prefect was from "a small planet somewhere in the vicinity of Betelgeuse."[152] In the poetic work Betelguese, a Trip Through Hell by Jean Louis De Esque, hell is on Betelgeuse because De Esque believed that it was "a celestial pariah, an outcast, the largest of all known comets or outlawed suns in the universe.".[154] In his 1953 story "Tony and the Beetles" Philip K. Dick describes a planet system around Betelguese and the neighboring stars occupied by the Terra (Earth) invaders. The main character incorrectly states the name of the star is of Jewish origin.

Two American navy ships were named after the star, both of them World War II vessels, the USS Betelgeuse (AKA-11) launched in 1939 and USS Betelgeuse (AK-260) launched in 1944. In 1979, a French supertanker named Betelgeuse was moored off Whiddy Island discharging oil when it exploded, killing 50 people in one of the worst disasters in Ireland's history.[155]

The Dave Matthews Band song "Black and Blue Bird" references the star.[156]

Humbert Wolfe wrote a poem about Betelgeuse, which was set to music by Gustav Holst.[157]

Notes

Article Year1 Telescope # Spectrum λ (μm) (mas)2 Radii3 @
197±45 pc
Notes
Michelson[29] 1920 Mt-Wilson 1 Visible 0.575 47.0 ± 4.7 3.2–6.3 AU Limb darkened +17% = 55.0
Bonneau[33] 1972 Palomar 8 Visible 0.422–0.719 52.0–69.0 3.6–9.2 AU Strong correlation of with λ
Balega[71] 1978 ESO 3 Visible 0.405–0.715 45.0–67.0 3.1–8.6 AU No correlation of with λ
1979 SAO 4 Visible 0.575–0.773 50.0–62.0 3.5–8.0 AU
Buscher[39] 1989 WHT 4 Visible 0.633–0.710 54.0–61.0 4.0–7.9 AU Discovered asymmetries/hotspots
Wilson[59] 1991 WHT 4 Visible 0.546–0.710 49.0–57.0 3.5–7.1 AU Confirmation of hotspots
Tuthill[42] 1993 WHT 8 Visible 0.633–0.710 43.5–54.2 3.2–7.0 AU Study of hotspots on 3 stars
1992 WHT 1 NIR 0.902 42.6 ± 0:03 3.0–5.6 AU
Gilliland[44] 1995 HST UV 0.24–0.27 104–112 10.3–11.1 FWHM diameters
0.265–0.295 92–100 9.1–9.8
Weiner[48] 1999 ISI 2 MIR (N Band) 11.150 54.7 ± 0.3 4.1–6.7 AU Limb darkened = 55.2 ± 0.5
Perrin[72] 1997 IOTA 7 NIR (K band) 2.200 43.33 ± 0.04 3.3–5.2 AU K and L bands, 11.5 μm data contrast
Haubois[60] 2005 IOTA 6 NIR (H band) 1.650 44.28 ± 0.15 3.4–5.4 AU Rosseland diameter 45.03 ± 0.12
Hernandez[83] 2006 VLTI 2 NIR (K band) 2.099–2.198 42:57 ± 0:02 3.2–5.2 AU High precision AMBER results.
Ohnaka[106] 2008 VLTI 3 NIR (K band) 2.280–2.310 43.19 ± 0.03 3.3–5.2 AU Limb darkened 43.56 ± 0.06
Townes[50] 1993 ISI 17 MIR (N band) 11.150 56.00 ± 1.00 4.2–6.8 AU Systematic study involving 17 measurements at the same wavelength from 1993 to 2009
2008 ISI MIR (N band) 11.150 47.00 ± 2.00 3.6–5.7 AU
2009 ISI MIR (N band) 11.150 48.00 ± 1.00 3.6–5.8 AU
Ohnaka[84] 2011 VLTI 3 NIR (K band) 2.280–2.310 42.05 ± 0.05 3.2–5.2 AU Limb darkened 42.49 ± 0.06
Harper[9] 2008 VLA Also noteworthy, Harper et al. in the conclusion of their paper make the following remark: "In a sense, the derived distance of 200 pc is a balance between the 131 pc (425 ly) Hipparcos distance and the radio which tends towards 250 pc (815 ly)"—hence establishing ± 815 ly as the outside distance for the star.
  1. 1 2 The above table provides a non-exhaustive list of angular measurements conducted since 1920. Also included is a column providing a current range of radii for each study based on Betelgeuse's most recent distance estimate (Harper et al.) of 197 ± 45 pc.

References

  1. 1 2 3 4 Simpson, J.; Weiner, E., eds. (1989). "Betelgeuse". Oxford English Dictionary (2nd ed.). Oxford: Clarendon Press. p. 130. ISBN 0-19-861186-2.
  2. 1 2 Merriam-Webster Dictionary
  3. 1 2 3 van Leeuwen, F (November 2007). "Hipparcos, the New Reduction". Astronomy and Astrophysics. VizieR: Centre de Données astronomiques de Strasbourg. 474 (2): 653. arXiv:0708.1752Freely accessible. Bibcode:2007A&A...474..653V. doi:10.1051/0004-6361:20078357.
  4. Keenan, Philip C.; McNeil, Raymond C. (1989). "The Perkins catalog of revised MK types for the cooler stars". Astrophysical Journal Supplement Series. 71: 245. Bibcode:1989ApJS...71..245K. doi:10.1086/191373.
  5. 1 2 3 Nicolet, B. (1978). "Catalogue of Homogeneous Data in the UBV Photoelectric Photometric System". Astronomy & Astrophysics. 34: 1–49. Bibcode:1978A&AS...34....1N.
  6. 1 2 3 4 Samus, N. N.; Durlevich, O. V.; et al. (2009). "VizieR Online Data Catalog: General Catalogue of Variable Stars (Samus+ 2007-2013)". VizieR On-line Data Catalog: B/gcvs. Originally published in: 2009yCat....102025S. 1. Bibcode:2009yCat....102025S.
  7. 1 2 Ducati, J. R. (2002). "VizieR Online Data Catalog: Catalogue of Stellar Photometry in Johnson's 11-color system". CDS/ADC Collection of Electronic Catalogues. 2237. Bibcode:2002yCat.2237....0D.
  8. Famaey, B.; Jorissen, A.; Luri, X.; Mayor, M.; Udry, S.; Dejonghe, H.; Turon, C. (2005). "Local kinematics of K and M giants from CORAVEL/Hipparcos/Tycho-2 data. Revisiting the concept of superclusters". Astronomy and Astrophysics. 430: 165. arXiv:astro-ph/0409579Freely accessible. Bibcode:2005A&A...430..165F. doi:10.1051/0004-6361:20041272.
  9. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 Harper, Graham M.; Brown, Alexander; Guinan, Edward F. (April 2008). "A New VLA-Hipparcos Distance to Betelgeuse and its Implications" (PDF). The Astronomical Journal. 135 (4): 1430–40. Bibcode:2008AJ....135.1430H. doi:10.1088/0004-6256/135/4/1430. Retrieved 10 July 2010.
  10. 1 2 3 4 Neilson, H. R.; Lester, J. B.; Haubois, X. (December 2011). "Weighing Betelgeuse: Measuring the Mass of α Orionis from Stellar Limb-darkening". Astronomical Society of the Pacific. 9th Pacific Rim Conference on Stellar Astrophysics. Proceedings of a conference held at Lijiang, China in 14–20 April 2011. ASP Conference Series, Vol. 451: 117. arXiv:1109.4562Freely accessible. Bibcode:2010ASPC..425..103L.
  11. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 Dolan, Michelle M.; Mathews, Grant J.; Lam, Doan Duc; Lan, Nguyen Quynh; Herczeg, Gregory J.; Dearborn, David S. P. (2016). "Evolutionary Tracks for Betelgeuse". The Astrophysical Journal. 819: 7. arXiv:1406.3143v2Freely accessible. Bibcode:2016ApJ...819....7D. doi:10.3847/0004-637X/819/1/7.
  12. 1 2 3 Smith, Nathan; Hinkle, Kenneth H.; Ryde, Nils (March 2009). "Red Supergiants as Potential Type IIn Supernova Progenitors: Spatially Resolved 4.6 μm CO Emission Around VY CMa and Betelgeuse". The Astronomical Journal. 137 (3): 3558–3573. arXiv:0811.3037Freely accessible. Bibcode:2009AJ....137.3558S. doi:10.1088/0004-6256/137/3/3558.
  13. Lobel, Alex; Dupree, Andrea K. (2000). "Modeling the Variable Chromosphere of α Orionis" (PDF). The Astrophysical Journal. 545 (1): 454–74. Bibcode:2000ApJ...545..454L. doi:10.1086/317784. Retrieved 10 July 2010.
  14. Ramírez, Solange V.; Sellgren, K.; Carr, John S.; Balachandran, Suchitra C.; et al. (July 2000). "Stellar Iron Abundances at the Galactic Center" (PDF). The Astrophysical Journal. 537 (1): 205–20. arXiv:astro-ph/0002062Freely accessible. Bibcode:2000ApJ...537..205R. doi:10.1086/309022. Retrieved 9 July 2010.
  15. 1 2 3 4 5 6 Kervella, P.; Verhoelst, T.; Ridgway, S. T.; Perrin, G.; Lacour, S.; et al. (September 2009). "The Close Circumstellar Environment of Betelgeuse. Adaptive Optics Spectro-imaging in the Near-IR with VLT/NACO". Astronomy and Astrophysics. 504 (1): 115–25. arXiv:0907.1843Freely accessible. Bibcode:2009A&A...504..115K. doi:10.1051/0004-6361/200912521.
  16. "IAU Working Group on Star Names (WGSN)". Retrieved 22 May 2016.
  17. "Bulletin of the IAU Working Group on Star Names, No. 1" (PDF). Retrieved 28 July 2016.
  18. "IAU Catalog of Star Names". Retrieved 28 July 2016.
  19. 1 2 3 4 5 6 7 8 Allen, Richard Hinckley (1963) [1899]. Star Names: Their Lore and Meaning (rep. ed.). New York, NY: Dover Publications Inc. pp. 310–12. ISBN 0-486-21079-0.
  20. Stella lucida in umero dextro, quae ad rubedinem vergit. "Bright star in right shoulder, which inclines to ruddiness."
  21. Brück, H. A. (11–15 July 1978). M. F. McCarthy, A. G. D. Philip, and G. V. Coyne, eds. P. Angelo Secchi, S. J. 1818–1878. Spectral Classification of the Future, Proceedings of the IAU Colloq. 47. Vatican City (published 1979). pp. 7–20. Bibcode:1979RA......9....7B.
  22. Information, Reed Business (22 October 1981). "Ancient Chinese Suggest Betelgeuse is a Young Star". New Scientist. 92 (1276): 238.
  23. 1 2 Levesque, E. M. (June 2010). "The Physical Properties of Red Supergiants". Astronomical Society of the Pacific. 425 Hot and Cool: Bridging Gaps in Massive Star Evolution ASP Conference Series: 103. arXiv:0911.4720Freely accessible. Bibcode:2010ASPC..425..103L.
  24. 1 2 3 4 Wilk, Stephen R. (1999). "Further Mythological Evidence for Ancient Knowledge of Variable Stars". The Journal of the American Association of Variable Star Observers. 27 (2): 171–74. Bibcode:1999JAVSO..27..171W.
  25. 1 2 Davis, Kate (December 2000). "Variable Star of the Month: Alpha Orionis". American Association of Variable Star Observers (AAVSO). Retrieved 10 July 2010.
  26. 1 2 3 Burnham, Robert (1978). Burnham's Celestial Handbook: An Observer's Guide to the Universe Beyond the Solar System, Volume 2. New York: Courier Dover Publications. p. 1290. ISBN 0-486-23568-8.
  27. Kaler, James B. (2002). The Hundred Greatest Stars. New York: Copernicus Books. p. 33. ISBN 0-387-95436-8.
  28. 1 2 MacDonald, John (1998). The Arctic sky: Inuit astronomy, star lore, and legend. Toronto, Ontario/Iqaluit, NWT: Royal Ontario Museum/Nunavut Research Institute. pp. 52–54, 119. ISBN 978-0-88854-427-8.
  29. 1 2 3 4 5 Michelson, Albert Abraham; Pease, Francis G. (1921). "Measurement of the diameter of Alpha Orionis with the interferometer". Astrophysical Journal. 53: 249–59. Bibcode:1921ApJ....53..249M. doi:10.1086/142603. The 0.047 arcsecond measurement was for a uniform disk. In the article Michelson notes that limb darkening would increase the angular diameter by about 17%, hence 0.055 arcseconds
  30. 1 2 Tenn, Joseph S. (June 2009). "The Bruce Medalists". Martin Schwarzschild 1965. Astronomical Society of the Pacific (ASP). Retrieved 28 September 2010.
  31. Schwarzschild, Martin (1958). Structure and Evolution of the Stars. Princeton University Press. Bibcode:1958ses..book.....S. ISBN 0-486-61479-4.
  32. Labeyrie, A. (May 1970). "Attainment of Diffraction Limited Resolution in Large Telescopes by Fourier Analysing Speckle Patterns in Star Images" (PDF). Astronomy and Astrophysics. 6: 85. Bibcode:1970A&A.....6...85L. Retrieved 12 October 2012.
  33. 1 2 3 Bonneau, D.; Labeyrie, A. (1973). "Speckle Interferometry: Color-Dependent Limb Darkening Evidenced on Alpha Orionis and Omicron Ceti". Astrophysical Journal. 181: L1. Bibcode:1973ApJ...181L...1B. doi:10.1086/181171.
  34. 1 2 Sutton, E. C.; Storey, J. W. V.; Betz, A. L.; Townes, C. H.; Spears, D. L. (1977). "Spatial Heterodyne Interferometry of VY Canis Majoris, Alpha Orionis, Alpha Scorpii, and R Leonis at 11 Microns" (PDF). Astrophysical Journal Letters. 217: L97–L100. Bibcode:1977ApJ...217L..97S. doi:10.1086/182547.
  35. Bernat, A. P.; Lambert, D. L. (November 1975). "Observations of the circumstellar gas shells around Betelgeuse and Antares" (PDF). Astrophysical Journal. 201: L153–L156. Bibcode:1975ApJ...201L.153B. doi:10.1086/181964.
  36. Dyck, H. M.; Simon, T. (February 1975). "Circumstellar dust shell models for Alpha Orionis" (PDF). Astrophysical Journal. 195: 689–693. Bibcode:1975ApJ...195..689D. doi:10.1086/153369.
  37. Boesgaard, A. M.; Magnan, C. (June 1975). "The circumstellar shell of alpha Orionis from a study of the Fe II emission lines" (PDF). Astrophysical Journal. 198 (1): 369–371, 373–378. Bibcode:1975ApJ...198..369B. doi:10.1086/153612.
  38. Bernat, David (2008). "Aperture Masking Interferometry". Ask An Astronomer. Cornell University Astronomy. Retrieved 15 October 2012.
  39. 1 2 3 Buscher, D. F.; Baldwin, J. E.; Warner, P. J.; Haniff, C. A. (1990). "Detection of a bright feature on the surface of Betelgeuse" (PDF). Monthly Notices of the Royal Astronomical Society. 245: 7. Bibcode:1990MNRAS.245P...7B.
  40. Wilson, R. W.; Dhillon, V. S.; Haniff, C. A. (1997). "The changing face of Betelgeuse". Monthly Notices of the Royal Astronomical Society. 291 (4): 819. Bibcode:1997MNRAS.291..819W. doi:10.1093/mnras/291.4.819.
  41. Burns, D.; Baldwin, J. E.; Boysen, R. C.; Haniff, C. A.; Lawson, P. R.; et al. (September 1997). "The surface structure and limb-darkening profile of Betelgeuse". Monthly Notices of the Royal Astronomical Society. 290 (1): L11–L16. Bibcode:1997MNRAS.290L..11B. doi:10.1093/mnras/290.1.l11.
  42. 1 2 Tuthill, P. G.; Haniff, C. A.; Baldwin, J. E. (March 1997). "Hotspots on late-type supergiants". Monthly Notices of the Royal Astronomical Society. 285 (3): 529–39. Bibcode:1997MNRAS.285..529T. doi:10.1093/mnras/285.3.529.
  43. 1 2 Schwarzschild, Martin (1975). "On the Scale of Photospheric Convection in Red Giants and Supergiants". Astrophysical Journal. 195 (1): 137–44. Bibcode:1975ApJ...195..137S. doi:10.1086/153313.
  44. 1 2 3 4 Gilliland, Ronald L.; Dupree, Andrea K. (May 1996). "First Image of the Surface of a Star with the Hubble Space Telescope" (PDF). Astrophysical Journal Letters. 463 (1): L29. Bibcode:1996ApJ...463L..29G. doi:10.1086/310043. Retrieved 1 August 2010. The yellow/red "image" or "photo" of Betelgeuse commonly seen is not a picture of the red supergiant, but a mathematically generated image based on the photograph. The photograph was of much lower resolution: The entire Betelgeuse image fit within a 10x10 pixel area on the Hubble Space Telescopes Faint Object Camera. The images were oversampled by a factor of 5 with bicubic spline interpolation, then deconvolved.
  45. A. N. Cox, editor (2000). Allen's Astrophysical Quantities. New York: Springer-Verlag. ISBN 0-387-98746-0.
  46. Petersen, Carolyn Collins; Brandt, John C. (1998) [1995]. Hubble Vision: Further Adventures with the Hubble Space Telescope (2nd ed.). Cambridge, England: Cambridge University Press. pp. 91–92. ISBN 0-521-59291-7.
  47. Uitenbroek, Han; Dupree, Andrea K.; Gilliland, Ronald L. (1998). "Spatially Resolved Hubble Space Telescope Spectra of the Chromosphere of α Orionis". The Astronomical Journal. 116 (5): 2501–12. Bibcode:1998AJ....116.2501U. doi:10.1086/300596. Retrieved 20 June 2007.
  48. 1 2 3 4 Weiner, J.; Danchi, W. C.; Hale, D. D. S.; McMahon, J.; et al. (December 2000). "Precision Measurements of the Diameters of α Orionis and ο Ceti at 11 Microns" (PDF). The Astrophysical Journal. 544 (2): 1097–1100. Bibcode:2000ApJ...544.1097W. doi:10.1086/317264. Retrieved 23 June 2007.
  49. 1 2 3 4 5 Sanders, Robert (9 June 2009). "Red Giant Star Betelgeuse Mysteriously Shrinking". UC Berkeley News. UC Berkeley. Retrieved 18 April 2010.
  50. 1 2 3 4 5 Townes, C. H.; Wishnow, E. H.; Hale, D. D. S.; Walp, B. (2009). "A Systematic Change with Time in the Size of Betelgeuse" (PDF). The Astrophysical Journal Letters. 697 (2): L127–28. Bibcode:2009ApJ...697L.127T. doi:10.1088/0004-637X/697/2/L127.
  51. 1 2 3 Ravi, V.; Wishnow, E.; Lockwood, S.; Townes, C. (December 2011). "The Many Faces of Betelgeuse". Astronomical Society of the Pacific. 448: 1025. arXiv:1012.0377Freely accessible. Bibcode:2011ASPC..448.1025R.
  52. Bernat, Andrew P. (1977). "The Circumstellar Shells and Mass Loss Rates of Four M Supergiants". Astrophysical Journal. 213: 756–66. Bibcode:1977ApJ...213..756B. doi:10.1086/155205.
  53. Cutri, R.; Skrutskie. M. (7 September 2009). "Very Bright Stars in the 2MASS Point Source Catalog (PSC)". The Two Micron All Sky Survey at IPAC. Retrieved 28 December 2011.
  54. "CCDM (Catalog of Components of Double & Multiple stars (Dommanget+ 2002)". VizieR. Centre de Données astronomiques de Strasbourg. Retrieved 22 August 2010.
  55. Mason, Brian D.; Wycoff, Gary L.; Hartkopf, William I.; Douglass, Geoffrey G.; Worley, Charles E. (2001). "The 2001 US Naval Observatory Double Star CD-ROM. I. The Washington Double Star Catalog". The Astronomical Journal. 122 (6): 3466. Bibcode:2001AJ....122.3466M. doi:10.1086/323920.
  56. 1 2 3 4 Van Loon, J. Th. (2013). "Betelgeuse and the Red Supergiants". Betelgeuse Workshop 2012. Edited by P. Kervella. 60: 307. arXiv:1303.0321Freely accessible. Bibcode:2013EAS....60..307V. doi:10.1051/eas/1360036.
  57. Karovska, M.; Noyes, R. W.; Roddier, F.; Nisenson, P.; Stachnik, R. V. (1985). "On a Possible Close Companion to α Ori". Bulletin of the American Astronomical Society. 17: 598. Bibcode:1985BAAS...17..598K.
  58. Karovska, M.; Nisenson, P.; Noyes, R. (1986). "On the alpha Orionis triple system". Astrophysical Journal. 308: 675–85. Bibcode:1986ApJ...308..260K. doi:10.1086/164497.
  59. 1 2 Wilson, R. W.; Baldwin, J. E.; Buscher, D. F.; Warner, P. J. (1992). "High-resolution imaging of Betelgeuse and Mira". Monthly Notices of the Royal Astronomical Society. 257 (3): 369–76. Bibcode:1992MNRAS.257..369W. doi:10.1093/mnras/257.3.369.
  60. 1 2 3 4 5 6 7 8 9 10 11 Haubois, X.; Perrin, G.; Lacour, S.; Verhoelst, T.; Meimon, S.; et al. (2009). "Imaging the Spotty Surface of Betelgeuse in the H Band". Astronomy & Astrophysics. 508 (2): 923–32. arXiv:0910.4167Freely accessible. Bibcode:2009A&A...508..923H. doi:10.1051/0004-6361/200912927.
  61. 1 2 3 Montargès, M.; Kervella, P.; Perrin, G.; Chiavassa, A.; Le Bouquin, J.-B.; Aurière, M.; López Ariste, A.; Mathias, P.; Ridgway, S. T.; Lacour, S.; Haubois, X.; Berger, J.-P. (2016). "The close circumstellar environment of Betelgeuse. IV. VLTI/PIONIER interferometric monitoring of the photosphere". Astronomy & Astrophysics. 588: A130. arXiv:1602.05108Freely accessible. Bibcode:2016A&A...588A.130M. doi:10.1051/0004-6361/201527028.
  62. van Altena, W. F.; Lee, J. T.; Hoffleit, D. (October 1995). "Yale Trigonometric Parallaxes Preliminary". Yale University Observatory (1991). 1174: 0. Bibcode:1995yCat.1174....0V.
  63. "Hipparcos Input Catalogue, Version 2 (Turon+ 1993)". VizieR. Centre de Données astronomiques de Strasbourg. 1993. Retrieved 20 June 2010.
  64. Perryman, M. A. C.; et al. (1997). "The Hipparcos Catalogue". Astronomy & Astrophysics. 323: L49–L52. Bibcode:1997A&A...323L..49P.
  65. Eyer, L.; Grenon, M. (2000). "Problems Encountered in the Hipparcos Variable Stars Analysis". Delta Scuti and Related Stars, Reference Handbook and Proceedings of the 6th Vienna Workshop in Astrophysics. ASP Conference Series. 210: 482. arXiv:astro-ph/0002235Freely accessible. Bibcode:2000ASPC..210..482E. ISBN 1-58381-041-2.
  66. "Science Performance". European Space Agency. 19 February 2013. Retrieved 1 March 2013.
  67. 1 2 3 4 Kiss, L. L.; Szabó, Gy. M.; Bedding, T. R. (2006). "Variability in red supergiant stars: Pulsations, long secondary periods and convection noise". Monthly Notices of the Royal Astronomical Society. 372 (4): 1721. arXiv:astro-ph/0608438Freely accessible. Bibcode:2006MNRAS.372.1721K. doi:10.1111/j.1365-2966.2006.10973.x.
  68. Guo, J. H.; Li, Y. (2002). "Evolution and Pulsation of Red Supergiants at Different Metallicities". The Astrophysical Journal. 565: 559. Bibcode:2002ApJ...565..559G. doi:10.1086/324295.
  69. Goldberg, L. (1984). "The variability of alpha Orionis". Astronomical Society of the Pacific. 96: 366. Bibcode:1984PASP...96..366G. doi:10.1086/131347.
  70. Wood, P. R.; Olivier, E. A.; Kawaler, S. D. (2004). "Long Secondary Periods in Pulsating Asymptotic Giant Branch Stars: An Investigation of their Origin". The Astrophysical Journal. 604 (2): 800. Bibcode:2004ApJ...604..800W. doi:10.1086/382123.
  71. 1 2 Balega, Iu.; Blazit, A.; Bonneau, D.; Koechlin, L.; Labeyrie, A.; Foy, R.. (November 1982). "The angular diameter of Betelgeuse". Astronomy and Astrophysics. 115 (2): 253–56. Bibcode:1982A&A...115..253B.
  72. 1 2 3 4 Perrin, G.; Ridgway, S. T.; Coudé du Foresto, V.; Mennesson, B.; Traub, W. A.; Lacasse, M. G. (2004). "Interferometric Observations of the Supergiant Stars α Orionis and α Herculis with FLUOR at IOTA". Astronomy and Astrophysics. 418 (2): 675–85. arXiv:astro-ph/0402099Freely accessible. Bibcode:2004A&A...418..675P. doi:10.1051/0004-6361:20040052. Assuming a distance of 197 ± 45 pc, an angular distance of 43.33 ± 0.04 mas would equate to a radius of 4.3 AU or 920 R
  73. Young, John (24 November 2006). "Surface Imaging of Betelgeuse with COAST and the WHT". University of Cambridge. Retrieved 21 June 2007. Images of hotspots on the surface of Betelgeuse taken at visible and infra-red wavelengths using high resolution ground-based interferometers
  74. Perrin, Guy; Malbet, Fabien (2003). "Observing with the VLTI". EAS Publications Series. 6: 3. Bibcode:2003EAS.....6D...3P. doi:10.1051/eas/20030601.
  75. Nemiroff, R.; Bonnell, J., eds. (21 April 2012). "3 ATs". Astronomy Picture of the Day. NASA. Retrieved 17 August 2012. Photograph showing three of the four enclosures which house 1.8 meter Auxiliary Telescopes (ATs) at the Paranal Observatory in the Atacama Desert region of Chile.
  76. Worden, S. (1978). "Speckle Interferometry". New Scientist. 78: 238–40. Bibcode:1978NewSc..78..238W.
  77. Roddier, F. (1999). "Ground-Based Interferometry with Adaptive Optics". Working on the Fringe: Optical and IR Interferometry from Ground and Space. Proceedings from ASP Conference. 194: 318. Bibcode:1999ASPC..194..318R. ISBN 1-58381-020-X.
  78. "Top Five Breakthroughs From Hubble's Workhorse Camera". NASA Jet Propulsion Laboratory, California Institute of Technology. 4 May 2009. Retrieved 28 August 2007.
  79. Melnick, J.; Petrov R.; Malbet, F. (23 February 2007). "The Sky Through Three Giant Eyes, AMBER Instrument on VLT Delivers a Wealth of Results". European Southern Observatory. Retrieved 29 August 2007.
  80. Wittkowski, M. (23 February 2007). "MIDI and AMBER from the User's Point of View" (PDF). European Southern Observatory VLTI. Retrieved 29 August 2007.
  81. 1 2 "Red Giant Star Betelgeuse in the Constellation Orion is Mysteriously Shrinking". Astronomy Magazine. 2009. Retrieved 14 September 2012.
  82. Nemiroff, R.; Bonnell, J., eds. (6 January 2010). "The Spotty Surface of Betelgeuse". Astronomy Picture of the Day. NASA. Retrieved 18 July 2010.
  83. 1 2 Hernandez Utrera, O.; Chelli, A (2009). "Accurate Diameter Measurement of Betelgeuse Using the VLTI/AMBER Instrument" (PDF). Revista Mexicana de Astronomía y Astrofísica (Serie de Conferencias). 37: 179–80. Bibcode:2009RMxAC..37..179H.
  84. 1 2 Ohnaka, K.; Weigelt, G.; Millour, F.; Hofmann, K.-H.; Driebe, T.; Schertl, D.; Chelli, A.; Massi, F.; Petrov, R.; Stee, Ph. (May 2011). "Imaging the Dynamical Atmosphere of the Red Supergiant Betelgeuse in the CO First Overtone Lines with VLTI/AMBER". Astronomy & Astrophysics. 529, id.A163: A163. arXiv:1104.0958Freely accessible. Bibcode:2011A&A...529A.163O. doi:10.1051/0004-6361/201016279. We derive a uniform-disk diameter of 42.05 ± 0.05 mas and a power-law-type limb-darkened disk diameter of 42.49 ± 0.06 mas and a limb-darkening parameter of (9.7 ± 0.5) × 10−2
  85. 1 2 3 Kervella, P.; Perrin, G.; Chiavassa, A.; Ridgway, S. T.; Cami, J.; Haubois, X.; Verhoelst, T. (2011). "The Close Circumstellar Environment of Betelgeuse. II. Diffraction-limited Spectro-imaging from 7.76 to 19.50 μm with VLT/VISIR". Astronomy & Astrophysics. 531, id.A117: A117. arXiv:1106.5041Freely accessible. Bibcode:2011A&A...531A.117K. doi:10.1051/0004-6361/201116962.
  86. Montargès, M.; Kervella, P.; Perrin, G.; Ohnaka, K.; Chiavassa, A.; Ridgway, S. T.; Lacour, S. (2014). "Properties of the CO and H2O MOLsphere of the red supergiant Betelgeuse from VLTI/AMBER observations". Astronomy & Astrophysics. 572: id.A17. arXiv:1408.2994Freely accessible. Bibcode:2014A&A...572A..17M. doi:10.1051/0004-6361/201423538.
  87. Cowen, Ron (10 June 2009). "Betelgeuse Shrinks: The Red Supergiant has Lost 15 Percent of its Size". The shrinkage corresponds to the star contracting by a distance equal to that between Venus and the Sun, researchers reported June 9 at an American Astronomical Society meeting and in the June 1 Astrophysical Journal Letters.
  88. Courtland, Rachel (2009). "Betelgeuse: The incredible Shrinking Star?". New Scientist. Reed Business Information Ltd. Retrieved 25 September 2010.
  89. Bedding, T. R.; Zijlstra, A. A.; Von Der Luhe, O.; Robertson, J. G.; et al. (1997). "The Angular Diameter of R Doradus: a Nearby Mira-like Star". Monthly Notices of the Royal Astronomical Society. 286 (4): 957–62. arXiv:astro-ph/9701021Freely accessible. Bibcode:1997MNRAS.286..957B. doi:10.1093/mnras/286.4.957.
  90. Dyck, H. M.; Van Belle, G. T.; Thompson, R. R. (1998). "Radii and Effective Temperatures for K and M Giants and Supergiants. II". The Astronomical Journal. 116 (2): 981. doi:10.1086/300453.
  91. Garrison, R. F. (1993). "Anchor Points for the MK System of Spectral Classification". Bulletin of the American Astronomical Society. 25: 1319. Bibcode:1993AAS...183.1710G. Retrieved 4 February 2012.
  92. 1 2 Le Bertre, T.; Matthews, L. D.; Gérard, E.; Libert, Y. (2012). "Discovery of a detached H I gas shell surrounding α Orionis". Monthly Notices of the Royal Astronomical Society. 422 (4): 3433. arXiv:1203.0255Freely accessible. Bibcode:2012MNRAS.422.3433L. doi:10.1111/j.1365-2966.2012.20853.x.
  93. "Bright Star Catalogue, 5th Revised Ed. (Hoffleit+, 1991)". VizieR. Centre de Données astronomiques de Strasbourg. Retrieved 7 September 2012.
  94. Dorch, S. B. F. (2004). "Magnetic Activity in Late-type Giant Stars: Numerical MHD Simulations of Non-linear Dynamo Action in Betelgeuse" (PDF). Astronomy & Astrophysics. 423 (3): 1101–07. arXiv:astro-ph/0403321Freely accessible. Bibcode:2004A&A...423.1101D. doi:10.1051/0004-6361:20040435.
  95. Aurière, M; Donati, J.-F.; Konstantinova-Antova, R.; Perrin, G.; Petit, P.; Roudier, T. (2010). "The Magnetic Field of Betelgeuse : a Local Dynamo from Giant Convection Cells?". Astronomy & Astrophysics. 516: L2. arXiv:1005.4845Freely accessible. Bibcode:2010A&A...516L...2A. doi:10.1051/0004-6361/201014925.
  96. Posson-Brown, Jennifer; Kashyap, Vinay L.; Pease, Deron O.; Drake, Jeremy J. (2006). "Dark Supergiant: Chandra's Limits on X-rays from Betelgeuse": arXiv:astro–ph/0606387. arXiv:astro-ph/0606387Freely accessible. Bibcode:2006astro.ph..6387P.
  97. Maeder, André; Meynet, Georges (2003). "The Role of Rotation and Mass Loss in the Evolution of Massive Stars". Proceedings of IAU Symposium. 212: 267. Bibcode:2003IAUS..212..267M.
  98. 1 2 Reynolds, R. J.; Ogden, P. M. (1979). "Optical evidence for a very large, expanding shell associated with the I Orion OB association, Barnard's loop, and the high galactic latitude H-alpha filaments in Eridanus". The Astrophysical Journal. 229: 942. Bibcode:1979ApJ...229..942R. doi:10.1086/157028.
  99. Decin, L.; Cox, N. L. J.; Royer, P.; Van Marle, A. J.; Vandenbussche, B.; Ladjal, D.; Kerschbaum, F.; Ottensamer, R.; Barlow, M. J.; Blommaert, J. A. D. L.; Gomez, H. L.; Groenewegen, M. A. T.; Lim, T.; Swinyard, B. M.; Waelkens, C.; Tielens, A. G. G. M. (2012). "The enigmatic nature of the circumstellar envelope and bow shock surrounding Betelgeuse as revealed by Herschel. I. Evidence of clumps, multiple arcs, and a linear bar-like structure". Astronomy & Astrophysics. 548: A113. arXiv:1212.4870Freely accessible. Bibcode:2012A&A...548A.113D. doi:10.1051/0004-6361/201219792.
  100. Nemiroff, R.; Bonnell, J., eds. (23 October 2010). "Orion: Head to Toe". Astronomy Picture of the Day. NASA. Retrieved 8 October 2012.
  101. SolStation. "Betelgeuse; Release No.: 04-03". Sol Company. Retrieved 20 July 2010.
  102. 1 2 Kaler, James B. "Betelgeuse (Alpha Orionis)". Stars website. University of Illinois. Retrieved 19 July 2009.
  103. 1 2 Baud, B.; Waters, R.; De Vries, J.; Van Albada, G. D.; et al. (January 1984). "A Giant Asymmetric Dust Shell around Betelgeuse". Bulletin of the American Astronomical Society. 16: 405. Bibcode:1984BAAS...16..405B.
  104. Ridgway, Stephen; Aufdenberg, Jason; Creech-Eakman, Michelle; Elias, Nicholas; et al. (2009). "Quantifying Stellar Mass Loss with High Angular Resolution Imaging". Astronomy & Astrophysics. 247: 247. arXiv:0902.3008Freely accessible. Bibcode:2009astro2010S.247R.
  105. Harper, Graham M.; Brown, Alexander; Lim, Jeremy (April 2001). "A Spatially Resolved, Semiempirical Model for the Extended Atmosphere of α Orionis (M2 Iab)". The Astrophysical Journal. 551 (2): 1073–98. Bibcode:2001ApJ...551.1073H. doi:10.1086/320215.
  106. 1 2 3 A. P. Ohnaka, K.; Hofmann, K.-H.; Benisty, M.; Chelli, A.; et al. (2009). "Spatially Resolving the Inhomogeneous Structure of the Dynamical Atmosphere of Betelgeuse with VLTI/AMBER". Astronomy & Astrophysics. 503 (1): 183–95. arXiv:0906.4792Freely accessible. Bibcode:2009A&A...503..183O. doi:10.1051/0004-6361/200912247.
  107. Tsuji, T. (2000). "Water on the Early M Supergiant Stars α Orionis and μ Cephei" (PDF). The Astrophysical Journal. 538 (2): 801–07. Bibcode:2000ApJ...538..801T. doi:10.1086/309185.
  108. Lambert, D. L.; Brown, J. A.; Hinkle, K. H.; Johnson, H. R. (1984). "Carbon, Nitrogen, and Oxygen Abundances in Betelgeuse". Astrophysical Journal. 284: 223–37. Bibcode:1984ApJ...284..223L. doi:10.1086/162401.
  109. 1 2 3 Dave Finley (8 April 1998). "VLA Shows "Boiling" in Atmosphere of Betelgeuse". National Radio Astronomy Observatory. Retrieved 7 September 2010.
  110. Lim, Jeremy; Carilli, Chris L.; White, Stephen M.; Beasley, Anthony J.; Marson, Ralph G. (1998). "Large Convection Cells as the Source of Betelgeuse's Extended Atmosphere". Nature. 392 (6676): 575–77. Bibcode:1998Natur.392..575L. doi:10.1038/33352.
  111. 1 2 3 Lobel, A.; Aufdenberg, J.; Dupree, A. K.; Kurucz, R. L.; Stefanik, R. P.; Torres, G. (2004). "Spatially Resolved STIS Spectroscopy of Betelgeuse's Outer Atmosphere". Proceedings of the 219th symposium of the IAU. 219: 641. arXiv:astro-ph/0312076Freely accessible. Bibcode:2004IAUS..219..641L. In the article, Lobel et al. equate 1 arcsecond to approximately 40 stellar radii, a calculation which in 2004 likely assumed a Hipparcos distance of 131 pc (430 ly) and a photospheric diameter of 0.0552" from Weiner et al.
  112. Dupree, Andrea K.; Gilliland, Ronald L. (December 1995). "HST Direct Image of Betelgeuse". Bulletin of the American Astronomical Society. 27: 1328. Bibcode:1995AAS...187.3201D. Such a major single feature is distinctly different from scattered smaller regions of activity typically found on the Sun although the strong ultraviolet flux enhancement is characteristic of stellar magnetic activity. This inhomogeneity may be caused by a large scale convection cell or result from global pulsations and shock structures that heat the chromosphere."
  113. 1 2 Skinner, C. J.; Dougherty, S. M.; Meixner, M.; Bode, M. F.; Davis, R. J.; et al. (1997). "Circumstellar Environments – V. The Asymmetric Chromosphere and Dust Shell of Alpha Orionis". Monthly Notices of the Royal Astronomical Society. 288 (2): 295–306. Bibcode:1997MNRAS.288..295S. doi:10.1093/mnras/288.2.295.
  114. Danchi, W. C.; Bester, M.; Degiacomi, C. G.; Greenhill, L. J.; Townes, C. H. (1994). "Characteristics of Dust Shells around 13 Late-type Stars". The Astronomical Journal. 107 (4): 1469–1513. Bibcode:1994AJ....107.1469D. doi:10.1086/116960.
  115. David, L.; Dooling, D. (1984). "The Infrared Universe". Space World. 2: 4–7. Bibcode:1984SpWd....2....4D.
  116. Harper, Graham M.; Carpenter, Kenneth G.; Ryde, Nils; Smith, Nathan; Brown, Joanna; et al. (2009). "UV, IR, and mm Studies of CO Surrounding the Red Supergiant α Orionis (M2 Iab)". AIP Conference Proceedings. 1094: 868–71. Bibcode:2009AIPC.1094..868H. doi:10.1063/1.3099254.
  117. 1 2 Mohamed, S.; Mackey, J.; Langer, N. (2012). "3D Simulations of Betelgeuse's Bow Shock". Astronomy & Astrophysics. 541, id.A1: A1. arXiv:1109.1555v2Freely accessible. Bibcode:2012A&A...541A...1M. doi:10.1051/0004-6361/201118002.
  118. Lamers, Henny J. G. L. M. & Cassinelli, Joseph P. (June 1999). Introduction to Stellar Winds. Cambridge, UK: Cambridge University Press. Bibcode:1999isw..book.....L. ISBN 978-0-521-59565-0.
  119. "Akari Infrared Space Telescope: Latest Science Highlights". European Space Agency. 19 November 2008. Archived from the original on 2011-02-17. Retrieved 25 June 2012.
  120. Noriega-Crespo, Alberto; van Buren, Dave; Cao, Yu; Dgani, Ruth (1997). "A Parsec-Size Bow Shock around Betelgeuse" (PDF). Astronomical Journal. 114: 837–40. Bibcode:1997AJ....114..837N. doi:10.1086/118517. Retrieved 25 June 2012. Noriega in 1997 estimated the size to be 0.8 parsecs, having assumed the earlier distance estimate of 400 ly. With a current distance estimate of 643 ly, the bow shock would measure ~1.28 parsecs or over 4 ly
  121. Newton, Elizabeth (26 April 2012). "This Star Lives in Exciting Times, or, How Did Betelgeuse Make that Funny Shape?". Astrobites. Retrieved 25 June 2012.
  122. MacKey, Jonathan; Mohamed, Shazrene; Neilson, Hilding R.; Langer, Norbert; Meyer, Dominique M.-A. (2012). "Double Bow Shocks Around Young, Runaway Red Supergiants: Application to Betelgeuse". The Astrophysical Journal. 751: L10. Bibcode:2012ApJ...751L..10M. doi:10.1088/2041-8205/751/1/L10.
  123. 1 2 3 4 5 6 Meynet, G.; Haemmerlé, L.; Ekström, S.; Georgy, C.; Groh, J.; Maeder, A. (2013). "The past and future evolution of a star like Betelgeuse". Betelgeuse Workshop 2012. Edited by P. Kervella. 60: 17. arXiv:1303.1339Freely accessible. Bibcode:2013EAS....60...17M. doi:10.1051/eas/1360002.
  124. 1 2 Groh, Jose H.; Meynet, Georges; Georgy, Cyril; Ekstrom, Sylvia (2013). "Fundamental properties of core-collapse Supernova and GRB progenitors: Predicting the look of massive stars before death". Astronomy & Astrophysics. 558: A131. arXiv:1308.4681v1Freely accessible. doi:10.1051/0004-6361/201321906.
  125. Wheeler, J. Craig (2007). Cosmic Catastrophes: Exploding Stars, Black Holes, and Mapping the Universe (2nd ed.). Cambridge, UK: Cambridge University Press. pp. 115–17. ISBN 0-521-85714-7.
  126. Connelly, Claire (19 January 2011). "Tatooine's twin suns – coming to a planet near you just as soon as Betelgeuse explodes". News.com.au. Retrieved 14 September 2012.
  127. Plait, Phil (1 June 2010). "Is Betelgeuse about to blow?". Bad Astronomy. Discovery. Retrieved 14 September 2012.
  128. O'Neill, Ian (20 January 2011). "Don't panic! Betelgeuse won't explode in 2012!". Discovery space news. Archived from the original on 2011-01-23. Retrieved 14 September 2012.
  129. Plait, Phil (21 January 2011). "Betelgeuse and 2012". Bad Astronomy. Discovery. Retrieved 14 September 2012.
  130. Likely the result of mistaking the l for an i. Ultimately, this led to the modern Betelgeuse.
  131. Bode, Johann Elert, (ed.). (1782) Vorstellung der Gestirne: auf XXXIV Kupfertafeln nach der Parisier Ausgabe des Flamsteadschen Himmelsatlas, Gottlieb August Lange, Berlin / Stralsund, pl. XXIV.
  132. Bode, Johann Elert, (ed.) (1801). Uranographia: sive Astrorum Descriptio, Fridericus de Harn, Berlin, pl. XII.
  133. 1 2 3 Schaaf, Fred (2008). "Betelgeuse". The Brightest Stars. Hoboken, New Jersey: Wiley. pp. 174–82. ISBN 0-471-70410-5.
  134. Dibon-Smith, Richard. "Alpha Orionis (Betelgeuse )". The Constellations Web Page. Retrieved 23 January 2010.
  135. Kanipe, Jeff (30 June 2005). "SpaceWatch – A Star by Any Other Name". Archived from the original on 22 May 2009. Retrieved 23 October 2009.
  136. Ridpath, Ian (2006). The Monthly Sky Guide (7th ed.). Cambridge University Press. p. 8. ISBN 0-521-68435-8.
  137. Kunitzsch, Paul (1986). "The Star Catalogue Commonly Appended to the Alfonsine Tables". Journal for the History of Astronomy. 17 (49): 89–98. Bibcode:1986JHA....17...89K.
  138. Kunitzsch, Paul (1959). Arabische Sternnamen in Europa. Wiesbaden: Otto Harrassowitz.
  139. 1 2 Kunitzsch, Paul; Smart, Tim (2006). A Dictionary of Modern star Names: A Short Guide to 254 Star Names and Their Derivations (2nd rev. ed.). Cambridge, MA: Sky Pub. p. 45. ISBN 978-1-931559-44-7.
  140. (Chinese) AEEA (Activities of Exhibition and Education in Astronomy) 天文教育資訊網 2006 年 5 月 25 日
  141. Ridpath, Ian. "Orion: Chinese associations". Star Tales. Retrieved 24 June 2012.
  142. Steve Renshaw & Saori Ihara (October 1999). "Yowatashi Boshi; Stars that Pass in the Night". Griffith Observer. pp. 2–17. Retrieved 25 June 2012.
  143. Hōei Nojiri"Shin seiza jyunrei"p.19 ISBN 978-4-12-204128-8
  144. Henry, Teuira (1907). "Tahitian Astronomy: Birth of Heavenly Bodies". The Journal of the Polynesian Society. 16 (2): 101–04. JSTOR 20700813.
  145. Brosch, Noah (2008). Sirius Matters. Springer. p. 46. ISBN 1-4020-8318-1.
  146. Milbrath, Susan (1999). Star Gods of the Maya: Astronomy in Art, Folklore, and Calendars. Austin, Texas: University of Texas Press. p. 39. ISBN 0-292-75226-1.
  147. Harney, Bill Yidumduma; Cairns, Hugh C. (2004) [2003]. Dark Sparklers (Revised ed.). Merimbula, New South Wales: Hugh C. Cairns. pp. 139–40. ISBN 0-9750908-0-1.
  148. Littleton, C. Scott (2005). Gods, goddesses, and mythology. 1. Marshall Cavendish. p. 1056. ISBN 0-7614-7559-1.
  149. Motz, Lloyd; Nathanson, Carol (1991). The Constellations: An Enthusiast's Guide to the Night Sky. London, United Kingdom: Aurum Press. p. 85. ISBN 1-85410-088-2.
  150. Cenev, Gjore (2008). "Macedonian Folk Constellations". Publications of the Astronomical Observatory of Belgrade. 85: 97–109. Bibcode:2008POBeo..85...97C.
  151. Kelley, David H.; Milone, Eugene F.; Aveni, A.F. (2011). Exploring Ancient Skies: A Survey of Ancient and Cultural Astronomy. New York, New York: Springer. p. 307. ISBN 1-4419-7623-X.
  152. 1 2 Conley, Craig (2008). Magic Words: A Dictionary. Weiser. p. 121. ISBN 1-57863-434-2. Retrieved 22 September 2010.
  153. Larsen, Kristine (2005). "A Definitive Identification of Tolkien's "Borgil": An Astronomical and Literary Approach". Tolkien Studies. 2 (1): 161–70. doi:10.1353/tks.2005.0023.
  154. De Esque, Jean Louis (1908). Wikisource link to Betelguese, a trip through hell. Connoisseur's Press. Wikisource. pp. 7.
  155. Tallant, Nicolla (15 July 2007). "Survivor recalls the night an apocalypse came to Whiddy". Independent Digital. Independent News & Media PLC. Retrieved 10 June 2011.
  156. "DMBAlmanac.com". Dmbalmanac.com. Retrieved 2016-01-30.
  157. Ford, Andrew (2012). "Holst, the Mystic". Try Whistling This: Writings on Music. Collingwood, Victoria: Black Incorporated. ISBN 9781921870682.
Wikimedia Commons has media related to Betelgeuse.
  1. Mars and Orion Over Monument Valley Skyscape showing the relative brightness of Betelgeuse and Rigel.
  2. Orion: Head to Toe Breathtaking vista the Orion Molecular Cloud Complex from Rogelio Bernal Andreo.
  3. The Spotty Surface of Betelgeuse A reconstructed image showing two hotspots, possibly convection cells.
  4. Simulated Supergiant Star Freytag's "Star in a Box" illustrating the nature of Betelgeuse's "monster granules".
  5. Why Stars Twinkle Image of Betelgeuse showing the effect of atmospheric twinkling in a microscope.

This article is issued from Wikipedia - version of the 11/30/2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.