Betelgeuse

Source: Wikipedia, the free encyclopedia.

Betelgeuse
Map of the constellation Orion
Location of Betelgeuse (circled)
Observation data
Epoch J2000.0      Equinox J2000.0
Constellation Orion
Pronunciation
/ˈbɛtəlz, ˈbt-, -s/ BE(E)T-əl-jooz, -⁠jooss[1][2]
Right ascension 05h 55m 10.30536s[3]
Declination +07° 24′ 25.4304″[3]
Apparent magnitude (V) +0.50[4] (0.0–1.6[5])
Characteristics
Evolutionary stage Red supergiant
Spectral type M1–M2 Ia–ab[6]
Apparent magnitude (J) −3.00[7]
Apparent magnitude (K) −4.05[7]
U−B color index +2.06[4]
B−V color index +1.85[4]
Variable type SRc[8]
Distance
408-548+90
−49
 ly
(125[12]-168.1+27.5
−14.9
[11] pc)
Absolute magnitude (MV)−5.85[13]
Details
Myr
HIP 27989, SAO 113271, GC 7451, CCDM J05552+0724, AAVSO
 0549+07
Database references
SIMBADdata

Betelgeuse is a

near-infrared wavelengths. Its Bayer designation is α Orionis, Latinised
to Alpha Orionis and abbreviated Alpha Ori or α Ori.

With a radius around 640 times that of the Sun,

bow shock
over four light-years wide.

Betelgeuse became the first extrasolar star whose

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 Earth-observed angular diameter of Betelgeuse is exceeded only by those of R Doradus
and the Sun.

Starting in October 2019, Betelgeuse began to dim noticeably, and by mid-February 2020 its brightness had dropped by a factor of approximately 3, from magnitude 0.5 to 1.7. It then returned to a more normal brightness range, reaching a peak of 0.0 visual and 0.1 V-band magnitude in April 2023. Infrared observations found no significant change in

luminosity over the last 50 years, suggesting that the dimming was due to a change in extinction around the star rather than a more fundamental change. A study using the Hubble Space Telescope
suggests that occluding dust was created by a surface mass ejection; this material was cast millions of miles from the star, and then cooled to form the dust that caused the dimming.

Nomenclature

The star's designation is α Orionis (Latinised to Alpha Orionis), given by Johann Bayer in 1603.

The traditional name Betelgeuse was derived from the

bā’ (بـ) led to the European name.[22][23] In English, there are four common pronunciations of this name, depending on whether the first e is pronounced short or long and whether the s is pronounced /s/ or /z/:[1][2]

In 2016, the International Astronomical Union organized a Working Group on Star Names (WGSN)[24] to catalog and standardize proper names for stars. The WGSN's first bulletin, issued July 2016,[25] 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.[26]

Observational history

Betelgeuse and its red coloration have been noted since

Class III (orange to red) stars.[28] Three centuries before Ptolemy, in contrast, Chinese astronomers observed Betelgeuse as yellow; Such an observation, if accurate, could suggest the star was in a yellow supergiant phase around this time,[29][12] a credible possibility, given current research into these stars' complex circumstellar environment.[30]

Nascent discoveries

Aboriginal groups in South Australia have shared oral tales of the variable brightness of Betelgeuse for at least 1,000 years.[31][32]

Sir John Herschel in 1846

The variation in Betelgeuse's brightness was described in 1836 by

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.[34][35] This variability in brightness may explain why Johann Bayer, with the publication of his Uranometria in 1603, designated the star alpha, as it probably rivaled the usually brighter Rigel (beta).[36] 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").[37]

In 1920,

But limb darkening and measurement errors resulted in uncertainty about the accuracy of these measurements.

The 1950s and 1960s saw two developments that affected 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.[39][40] 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.[39]

Imaging breakthroughs

1998/9 UV HST images of Betelgeuse showing asymmetrical pulsations with corresponding spectral line profiles

Astronomers saw some major advances in astronomical imaging technology in the 1970s, beginning with

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.[41][42]
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,[43][44][45]
causing them to suspect the presence of huge gas bubbles resulting from convection.[46] However, 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 J.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.[47] The technique contributed some of the most accurate measurements of Betelgeuse while revealing bright spots on the star's photosphere.[48][49][50] 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.[51][52]

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.[53] Because ultraviolet light is absorbed by the Earth's atmosphere, observations at these wavelengths are best performed by space telescopes.[54] This image, like earlier pictures, contained a bright patch indicating a region in the southwestern quadrant 2,000 K hotter than the stellar surface.[55] 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°.[56]

2000s studies

In a study published in December 2000, the star's diameter was measured with the

mas – a figure entirely consistent with Michelson's findings eighty years earlier.[38][57]
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, an infrared interferometric study published in 2009 announced that the star had shrunk by 15% since 1993 at an increasing rate without a significant diminution in magnitude.[58][59] Subsequent observations suggest that the apparent contraction may be due to shell activity in the star's extended atmosphere.[60]

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 remains a mystery.[61]
With advances in interferometric methodologies, astronomers may be close to resolving this conundrum. Images released by the European Southern Observatory in July 2009, 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.[62] 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.[58]

2019–2020 fading

of Betelgeuse, between September 2016 and August 2023
SPHERE
images of Betelgeuse taken in January 2019 and December 2019, showing the changes in brightness and shape

A pulsating semiregular variable star, Betelgeuse is subject to multiple cycles of increasing and decreasing brightness due to changes in its size and temperature.[16] The astronomers who first noted the dimming of Betelgeuse, Villanova University astronomers Richard Wasatonic and Edward Guinan, and amateur Thomas Calderwood, theorize that a coincidence of a normal 5.9 year light-cycle minimum and a deeper-than-normal 425 day period are the driving factors.[63] Other possible causes hypothesized by late 2019 were an eruption of gas or dust or fluctuations in the star's surface brightness.[64]

By August 2020, long-term and extensive studies of Betelgeuse, primarily using ultraviolet observations by the Hubble Space Telescope, had suggested that the unexpected dimming was probably caused by an immense amount of superhot material ejected into space. The material cooled and formed a dust cloud that blocked the starlight coming from about a quarter of Betelgeuse's surface. Hubble captured signs of dense, heated material moving through the star's atmosphere in September, October and November before several telescopes observed the more marked dimming in December and the first few months of 2020.[65][66][67]

By January 2020, Betelgeuse had dimmed by a factor of approximately 2.5 from magnitude 0.5 to 1.5 and was reported still fainter in February in The Astronomer's Telegram at a record minimum of +1.614, noting that the star is currently the "least luminous and coolest" in the 25 years of their studies and also calculating a decrease in radius.[68] Astronomy magazine described it as a "bizarre dimming",[69] and popular speculation inferred that this might indicate an imminent supernova.[70][71] This dropped Betelgeuse from one of the top 10 brightest stars in the sky to outside the top 20,[63] noticeably dimmer than its near neighbor Aldebaran.[64] Mainstream media reports discussed speculation that Betelgeuse might be about to explode as a supernova,[72][73][74][75] but astronomers note that the supernova is expected to occur within approximately the next 100,000 years and is thus unlikely to be imminent.[72][74]

By 17 February 2020, Betelgeuse's brightness had remained constant for about 10 days, and the star showed signs of rebrightening.[76] On 22 February 2020, Betelgeuse may have stopped dimming altogether, all but ending the dimming episode.[77] On 24 February 2020, no significant change in the infrared over the last 50 years was detected; this seemed unrelated to the recent visual fading and suggested that an impending core collapse may be unlikely.[78] Also on 24 February 2020, further studies suggested that occluding "large-grain circumstellar dust" may be the most likely explanation for the dimming of the star.[79][80] A study that uses observations at submillimetre wavelengths rules out significant contributions from dust absorption. Instead, large starspots appear to be the cause for the dimming.[81] Followup studies, reported on 31 March 2020 in The Astronomer's Telegram, found a rapid rise in the brightness of Betelgeuse.[82]

Betelgeuse is almost unobservable from the ground between May and August because it is too close to the Sun. Before entering its 2020 conjunction with the Sun, Betelgeuse had reached a brightness of +0.4 . Observations with the STEREO-A spacecraft made in June and July 2020 showed that the star had dimmed by 0.5 since the last ground-based observation in April. This is surprising, because a maximum was expected for August/September 2020, and the next minimum should occur around April 2021. However Betelgeuse's brightness is known to vary irregularly, making predictions difficult. The fading could indicate that another dimming event might occur much earlier than expected.[83] On 30 August 2020, astronomers reported the detection of a second dust cloud emitted from Betelgeuse, and associated with recent substantial dimming (a secondary minimum on 3 August) in luminosity of the star.[84]

In June 2021, the dust was explained as possibly caused by a cool patch on its photosphere[85][86][87][88] and in August a second independent group confirmed these results.[89][90] The dust is thought to have resulted from the cooling of gas ejected from the star. An August 2022[91][92][93] study using the Hubble Space Telescope confirmed previous research and suggested the dust could have been created by a surface mass ejection. It conjectured as well that the dimming could have come from a short-term minimum coinciding with a long-term minimum producing a grand minimum, a 416-day cycle and 2010 day cycle respectively, a mechanism first suggested by astronomer L. Goldberg.[94] In April 2023, astronomers reported the star reached a peak of 0.0 visual and 0.1 V-band magnitude.[95]

Observation

Image showing Betelgeuse (top left) and the dense nebulae of the Orion molecular cloud complex (Rogelio Bernal Andreo)
Orion as seen on midday mid-June from Dome C (75 degrees South, Antarctica). Screenshot from Stellarium

As a result of its distinctive orange-red color and position within Orion, Betelgeuse is easy to find with the naked eye. It is one of three stars that make up the Winter Triangle asterism, and it marks the center of the Winter Hexagon. It can be seen rising in the east at the beginning of January of each year, just after sunset. Between mid-September and mid-March (best in mid-December), it is visible to virtually every inhabited region of the globe, except 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, centered around mid June), it is invisible to the naked eye (visible only with a telescope in daylight), except around midday low in the north in Antarctic regions between 70° and 80° south latitude (during midday twilight in polar night, when the Sun is below the horizon).

Betelgeuse is a variable star whose

Beta Crucis, themselves both slightly variable, to be the twentieth-brightest star.[35]

Betelgeuse has a B–V color index of 1.85 – a figure which points to its pronounced "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 toward and away from Betelgeuse, depending on fluctuations in the photosphere. Betelgeuse is the brightest near-infrared source in the sky with a J band magnitude of −2.99;[96] only about 13% of the star's radiant energy is emitted as visible light. If human eyes were sensitive to radiation at all wavelengths, Betelgeuse would appear as the brightest star in the night sky.[35]

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.[97][98]

Star system

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.[99]

Two spectroscopic companions to Betelgeuse have been proposed. Analysis of polarization data from 1968 through 1983 indicated a close companion with a periodic orbit of about 2.1 years, and by 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 of 273°, an orbit that would potentially place it within the star's chromosphere. The more distant companion was at 0.51″±0.01″ (≈77 AU) with a position angle of 278°.[100][101] Further studies have found no evidence for these companions or have actively refuted their existence,[102] but the possibility of a close companion contributing to the overall flux has never been fully ruled out.[103] High-resolution interferometry of Betelgeuse and its vicinity, far beyond the technology of the 1980s and 1990s, has not detected any companions.[62][104]

Distance measurements

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

mass.[105]

When the first interferometric studies were performed on the star's diameter in 1920, the assumed parallax was 0.0180

. This equated to a distance of 56 pc or roughly 180 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 1,300 ly.[105]

Before the publication of the

mas, yielding a distance of roughly 102 pc or 330 ly.[106] The second was the Hipparcos Input Catalogue (1993) with a trigonometric parallax of 5±4 mas, a distance of 200 pc or 650 ly.[107] Given this uncertainty, researchers were adopting a wide range of distance estimates, leading to significant variances in the calculation of the star's attributes.[105]

The results from the Hipparcos mission were released in 1997. The measured parallax of Betelgeuse was 7.63±1.64 mas, which equated to a distance of roughly 131 pc or 427 ly, and had a smaller reported error than previous measurements.[108] However, later evaluation of the Hipparcos parallax measurements for variable stars like Betelgeuse found that the uncertainty of these measurements had been underestimated.[109] In 2007, an improved figure of 6.55±0.83 was calculated, hence a much tighter error factor yielding a distance of roughly 152±20 pc or 500±65 ly.[3]

In 2008, measurements using the

e-Merlin gives a parallax of 4.51±0.8 mas and a distance of 222+34
−48
pc or 724+111
−156
ly.[10]

In 2020, new observational data from the space-based Solar Mass Ejection Imager aboard the Coriolis satellite and three different modeling techniques produced a refined parallax of 5.95+0.58
−0.85
mas, a radius of 764+116
−62
R, and a distance of 168.1+27.5
−14.4
pc or 548+90
−49
ly, which, if accurate, would mean Betelgeuse is nearly 25% smaller and 25% closer to Earth than previously thought.[11]

Although the

Gaia Data Release 2, which was released in 2018.[112]

Variability

of Betelgeuse (Alpha Orionis) from Dec 1988 to Aug 2002.
Orion, with Betelgeuse at its usual magnitude (left) and during the unusually deep minimum in early 2020 (right)

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.[8]

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 +1.6. Betelgeuse is listed in the General Catalogue of Variable Stars with a possible period of 2,335 days.[8] More detailed analyses have shown a main period near 400 days, a short period of 185 days,[11] and a longer secondary period around 2,100 days.[104][113] The lowest reliably-recorded V-band magnitude of +1.614 was reported in February 2020.

Radial pulsations of red supergiants are well-modelled and show that periods of a few hundred days are typically due to

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.[115] Variations in the diameter of Betelgeuse have also been measured directly.[60] First overtone pulsations of 185 days have been observed, and the ratio of the fundamental to overtone periods gives valuable information about the internal structure of the star and its age.[11]

The source of the long secondary periods is unknown, but they cannot be explained by

In addition to the discrete dominant periods, small-amplitude

granulation, similar to the same effect on the sun but on a much larger scale.[113]

Diameter

Size comparison between Arcturus, Rigel, S Doradus, Antares, and Betelgeuse
Size comparison of Betelgeuse, Mu Cephei, KY Cygni, and V354 Cephei, according to Emily Levesque[citation needed]

On 13 December 1920, Betelgeuse became the first star outside the Solar System to have the angular size of its photosphere measured.[38] 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″, 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".[38][59] Since then, other studies have produced angular diameters that range from 0.042 to 0.069″.[42][57][117] 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. 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 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;[58]
  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.[58] 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;[53][118][119]
  5. Atmospheric twinkling limits the resolution obtainable from ground-based telescopes since turbulence degrades angular resolution.[48]

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, with corrections for limb darkening and the observation wavelength.[120] For example, a measured angular diameter of 55.6 mas would correspond to a Rosseland mean diameter of 56.2 mas, while further corrections for the existence of surrounding dust and gas shells would give a diameter of 41.9 mas.[16]

To overcome these challenges, researchers have employed various solutions. Astronomical interferometry, first conceived by

space observatories like Hipparcos, Hubble and Spitzer,[53][125] and the Astronomical Multi-BEam Recombiner (AMBER), which combines the beams of three telescopes simultaneously, allowing researchers to achieve milliarcsecond spatial resolution.[126][127]

Observations in different regions of the electromagnetic spectrum—the visible, near-infrared (NIR), mid-infrared (MIR), or radio—produce very different angular measurements. In 1996, Betelgeuse was shown to have a uniform disk of 56.6±1.0 mas. In 2000, a Space Sciences Laboratory team measured a diameter of 54.7±0.3 mas, ignoring any possible contribution from hotspots, which are less noticeable in the mid-infrared.[57] 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 1,200 R, assuming the 2008 Harper distance of 197.0±45 pc,[17] a figure roughly the size of the Jovian orbit of 5.5 AU.[128][129]

In 2004, a team of astronomers working in the near-infrared announced that the more accurate photospheric measurement was 43.33±0.04 mas. 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.[118]

Infrared image of Betelgeuse, Meissa and Bellatrix with surrounding nebulae

Studies with the IOTA and VLTI published in 2009 brought strong support to the idea of dust shells and a molecular shell (MOLsphere) around Betelgeuse, and yielded diameters ranging from 42.57 to 44.28 mas with comparatively insignificant margins of error.[103][130] 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.[131] The near-infrared photospheric diameter of 43.33 mas at the Hipparcos distance of 152±20 pc equates to about 3.4 AU or 730 R.[132] 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.[133]

In 2009 it was announced that the radius of Betelgeuse had shrunk from 1993 to 2009 by 15%, with the 2008 angular measurement equal to 47.0 mas.

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.[59] The observed contraction is generally believed to be a variation in just a portion of the extended atmosphere around Betelgeuse, and observations at other wavelengths have shown an increase in diameter over a similar period.[133]

The latest models of Betelgeuse adopt a photospheric angular diameter of around 43 mas, with multiple shells out to 50-60 mas.[20] Assuming a distance of 197 pc, this means a stellar diameter of 887±203 R.[16]

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.[135]

Occultations

  • Predicted path using SOLEX
    Predicted path using
    SOLEX

Betelgeuse is too far from the ecliptic to be occulted by the major planets, but those by some asteroids (which are more wide-ranging and much more numerous) occur frequently. A partial occultation by the 19th magnitude asteroid (147857) 2005 UW381 occurred on 2 January 2012. It was partial because the angular diameter of the star was larger than that of the asteroid; the brightness of Betelgeuse dropped by only about 0.01 magnitudes.[136][137]

The 14th magnitude asteroid 319 Leona was predicted to occult on 12 December 2023, 01:12 UTC.[138] Totality was at first uncertain, and the occulation was projected to only last approximately twelve seconds (visible on a narrow path on Earth's surface, the exact width and location of which was initially uncertain due to lack of precise knowledge of the size and path of the asteroid).[139] Projections were later refined as more data were analyzed for[140] a totality ("ring of fire") of approximately five seconds and a 60 km wide path stretching from Tajikistan, Armenia, Turkey, Greece, Italy, Spain, the Atlantic Ocean, Miami, Florida and the Florida Keys to parts of Mexico.[141] (The serendiptous event would also afford detailed observations of 319 Leona itself.)[142] Among other programmes 80 amateur astronomers in Europe alone have been coordinated by astrophysicist Miguel Montargès, et al. of the Paris Observatory for the event.[143]

Physical characteristics

VV Cephei A < VY Canis Majoris

Betelgeuse is a very large, luminous but cool star classified as an M1-2 Ia-ab

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.[144]

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 126,000 L, assuming a distance of 200 pc.[145] Studies since 2001 report effective temperatures ranging from 3,250 to 3,690 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.[16] The star is also a slow rotator and the most recent velocity recorded was 5.45 km/s[20]—much slower than Antares which has a rotational velocity of 20 km/s.[146] The rotation period depends on Betelgeuse's size and orientation to Earth, but it has been calculated to take 36 years to turn on its axis, inclined at an angle of around 60° to Earth.[20]

In 2004, astronomers using computer simulations speculated that even if Betelgeuse is not rotating it might exhibit large-scale 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.

dynamo effect.[148]

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 of 9.5–21 M,[149] with values of 5 M–30 M from older studies.[150] It has been calculated that Betelgeuse began its life as a star of 15–20 M, based on a solar luminosity of 90,000–150,000.[17] 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±217 R.[149] Model fitting to evolutionary tracks give a current mass of 19.4–19.7 M, from an initial mass of 20 M.[16]

Motion

Orion OB1 association

The

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.[105][152] An observation by the Herschel Space Observatory in January 2013 revealed that the star's winds are crashing against the surrounding interstellar medium.[153]

The most likely star-formation scenario for Betelgeuse is that it is a runaway star from the

Circumstellar dynamics

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

yellow supergiant phase and then explode as a blue supergiant or Wolf–Rayet star.[30]

Artist's rendering from ESO
showing Betelgeuse with a gigantic bubble boiling on its surface and a radiant plume of gas being ejected to 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.

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.[62] 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.[158]

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.[103]

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 1,500±500 K.[103][159] 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

Another cooler region, the asymmetric gaseous envelope, 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.[103][160]

Radio-telescope images taken in 1998 confirm that Betelgeuse has a highly complex atmosphere,[161] with a temperature of 3,450±850 K, similar to that recorded on the star's surface but much lower than surrounding gas in the same region.[161][162] 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."[161] 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.[103]

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.[163] 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 5,500 K.[103][164] 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.[163] The observations have conclusively demonstrated that the warm chromospheric plasma spatially overlaps and co-exists with cool gas in Betelgeuse's gaseous envelope as well as with the dust in its circumstellar dust shells.[103][163]

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

arcseconds, or 100 to 200 AU.[165][166] 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.[165] The 1984 report of a giant asymmetric dust shell 1 pc (206,265 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.[167][168]

Although the exact size of the two outer

heliopause
is estimated at 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/s (i.e. ~6.3 AU/a) creating a

bow shock.[170][171] 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.[172] 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 one parsec wide, assuming a distance of 643 light-years.[173]

Hydrodynamic simulations of the bow shock made in 2012 indicate that it is very young—less than 30,000 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.[174] 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."[170][175] Moreover, if future research bears out this hypothesis, Betelgeuse may prove to have traveled close to 200,000 AU as a red supergiant scattering as much as M along its trajectory.

Life phases

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.

Main sequence

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

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 properties.[16][99][176] That main sequence version of Betelgeuse would have been a hot luminous star with a spectral type such as O9V.[145]

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.[176] Rapidly-rotating 20 M stars take 9.3 million years to reach the red supergiant stage, while 20 M stars with slow rotation take only 8.1 million years.[99] These are the best estimates of Betelgeuse's current age, as the time since its zero age main sequence stage is estimated to be 8.0–8.5 million years as a 20 M star with no rotation.[16]

After core hydrogen exhaustion

Betelgeuse's time spent 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.[16][99]

This is what Betelgeuse may have looked like up until about 1 million years ago, when it was a main-sequence star.

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,[16] having left the main sequence perhaps one million years ago.[176]

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.[16]

Betelgeuse's mass can also be estimated based on its position on the color‑magnitude‑diagram (CMD). Betelgeuse's color may have changed from yellow (or possibly orange) to red in the last few thousand years, based on a 2022 review of historical records. This color change combined with CMD suggest a mass of 14 M and age of 14 Myr.[12]

Celestia depiction of Orion as it might appear from Earth when Betelgeuse explodes as a supernova, which could be brighter than the supernova that exploded in 1006

All stars more massive than about 10 M are expected to end their lives when their cores collapse, typically producing a supernova explosion. Up to about 15 M, a type II-P supernova is always produced from the red supergiant stage.[176]

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 I b/c supernovae from Wolf-Rayet stars.[177] Models of rotating 20 M stars predict a peculiar type II supernova similar to SN 1987A from a blue supergiant progenitor.[176] On the other hand, non-rotating 20 M models predict a type II-P supernova from a red supergiant progenitor.[16]

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 25 M star, 550,000 years for a rotating 20 M star, and 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.

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.[152]

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, Betelgeuse as a type IIP supernova would have a peak apparent magnitude somewhere in the range −8 to −12.[178] This would be easily visible in daylight, with a possible brightness up to a significant fraction of the full moon, though likely not exceeding it. 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 sustains its brightness due to the increasing transparency of the cooling hydrogen ejected by the supernova.[179]

Media reporting

Due to misunderstandings caused by the 2009 publication of the star's 15% contraction, apparently of its outer atmosphere,[58][128] Betelgeuse has frequently been the subject of scare stories and rumors suggesting that it will explode within a year, and leading to exaggerated claims about the consequences of such an event.[180][181] The timing and prevalence of these rumors have been linked to broader misconceptions of astronomy, particularly to doomsday predictions relating to the Mayan calendrical apocalypse.[182][183] 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.[16][184]

Following the dimming of Betelgeuse in December 2019,[185][63] reports appeared in the science and mainstream media that again included speculation that the star might be about to explode as a supernova – even in the face of scientific research that a supernova is not expected for perhaps 100,000 years.[186] Some outlets reported the magnitude as faint as +1.3 as an unusual and interesting phenomenon, like Astronomy magazine,[69] the National Geographic,[72] and the Smithsonian.[187]

Some mainstream media, like The Washington Post,[73] ABC News in Australia,[74] and Popular Science,[188] reported that a supernova was possible but unlikely, whilst other outlets falsely portrayed a supernova as an imminent realistic possibility. CNN, for example, chose the headline "A giant red star is acting weird and scientists think it may be about to explode",[189] while the New York Post declared Betelgeuse as "due for explosive supernova".[75]

Phil Plait, in his Bad Astronomy blog, noting that Betelgeuse's recent behaviour, "[w]hile unusual . . . isn't unprecedented," argued that the star is not likely to explode "for a long, long time."[190] Dennis Overbye of The New York Times agreed that an explosion was not imminent but added that "astronomers are having fun thinking about it."[191]

Following the eventual supernova,

a small dense remnant will be left behind, either a neutron star or black hole. Betelgeuse does not seem to have a core massive enough for a black hole, so the remnant will probably be a neutron star of approximately 1.5 M.[16]

Ethnological attributes

Spelling and pronunciation

Betelgeuse has also been spelled Betelgeux[1] and, in German, Beteigeuze[b] (according to Bode).[192][193] Betelgeux and Betelgeuze were used until the early 20th century, when the spelling Betelgeuse became universal.[194] Consensus on its pronunciation is weak and is as varied as its spellings:

The -urz pronunciations are attempts to render the French eu sound; they only work in r-dropping accents.

Etymology

Book of Fixed Stars
. Betelgeuze is annotated as Yad al-Jauzā ("Hand of Orion"), one of the proposed etymological origins of its modern name, and also as Mankib al Jauzā' ("Shoulder of Orion").

Betelgeuse is often mistranslated as "armpit of the central one".

Jesuit priest and astronomer Giovanni Battista Riccioli had called it Bectelgeuze or Bedalgeuze.[27]

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.[198] 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.[199] Other writers have since accepted Kunitzsch's explanation.[36]

The last part of the name, "-elgeuse", comes from the Arabic الجوزاء al-Jauzā', a historical Arabic name of the constellation

Arabian legend, and of uncertain meaning. Because جوز j-w-z, the root of jauzā', means "middle", al-Jauzā' roughly means "the Central One". The modern Arabic name for Orion is الجبار al-Jabbār ("the Giant"), although the use of الجوزاء al-Jauzā' in the star's name has continued.[199] The 17th-century English translator Edmund Chilmead gave it the name Ied Algeuze ("Orion's Hand"), from Christmannus.[27] Other Arabic names recorded include Al Yad al Yamnā ("the Right Hand"), Al Dhira ("the Arm"), and Al Mankib ("the Shoulder"), all of al-Jauzā, Orion,[27]
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

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 chose 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.[202][203]

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".[204] A Hawaiian term for it was Kaulua-koko ("brilliant red star").[205] The Lacandon people of Central America knew it as chäk tulix ("red butterfly").[206]

Astronomy writer Robert Burnham Jr. proposed the term padparadaschah, which denotes a rare orange sapphire in India, for the star.[194]

Mythology

With 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.[27] 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.[33]

Aboriginal people from the Great Victoria Desert of South Australia incorporated Betelgeuse into their oral traditions as the club of Nyeeruna (Orion), which fills with fire-magic and dissipates before returning. This has been interpreted as showing that early Aboriginal observers were aware of the brightness variations of Betelgeuse.[207][208] 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.[209] In South African mythology, Betelgeuse was perceived as a lion casting a predatory gaze toward the three zebras represented by Orion's Belt.[210]

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 of 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.[33]

A Sanskrit name for Betelgeuse is ārdrā ("the moist one"), eponymous of the

Richard Hinckley Allen to Orion's stormy nature.[27] The constellations in Macedonian folklore represented agricultural items and animals, reflecting their 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 a.m. in late summer and autumn signified the time for village men to go to the fields and plough.[212] 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.[37]

The opposed locations of Orion and

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.[213]

In popular culture

As one of the brightest and best-known stars, Betelgeuse has featured in many works of fiction. The star's unusual name inspired the title of the 1988 film Beetlejuice, referring to its titular antagonist, and script writer Michael McDowell was impressed by how many people made the connection.[194] 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."[214]

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, the French supertanker Betelgeuse was moored off Whiddy Island, discharging oil when it exploded, killing 50 people in one of the worst disasters in Ireland's history.[215]

The Dave Matthews Band song "Black and Blue Bird" references the star.[216] The Blur song "Far Out" from their 1994 album Parklife mentions Betelgeuse in its lyrics.[217]

The Philip Larkin poem "The North Ship", found in the collection of the same name, references the star in the section "Above 80° N", which reads:

" 'A woman has ten claws,' /

Sang the drunken boatswain; / Farther than Betelgeuse, / More brilliant than Orion / Or the planets Venus and Mars, / The star flames on the ocean; / 'A woman has ten claws,' /

Sang the drunken boatswain."

Humbert Wolfe wrote a poem about Betelgeuse, which was set to music by Gustav Holst.[218]

Table of angular diameter estimates

This 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.

Article Year[c] Telescope
#
Spectrum λ (μm) (
mas)[d]
Radii[e] @
197±45 pc
Notes
Michelson[38] 1920 Mt. Wilson 1 Visible 0.575 47.0±4.7 3.2–6.3 AU Limb darkened +17% = 55.0
Bonneau[42] 1972 Palomar 8 Visible 0.422–0.719 52.0–69.0 3.6–9.2 AU Strong correlation of with λ
Balega[117] 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[48] 1989 WHT 4 Visible 0.633–0.710 54.0–61.0 4.0–7.9 AU Discovered asymmetries/hotspots
Wilson[102] 1991 WHT 4 Visible 0.546–0.710 49.0–57.0 3.5–7.1 AU Confirmation of hotspots
Tuthill[51] 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±3.0 3.0–5.6 AU
Gilliland[53] 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[57] 1999 ISI 2 MIR (N Band) 11.150 54.7±0.3 4.1–6.7 AU Limb darkened = 55.2±0.5
Perrin[118] 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[103] 2005 IOTA 6 NIR (H band) 1.650 44.28±0.15 3.4–5.4 AU Rosseland diameter 45.03±0.12
Hernandez[130] 2006 VLTI 2 NIR (K band) 2.099–2.198 42.57±0.02 3.2–5.2 AU High precision AMBER results.
Ohnaka[158] 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[59] 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[131] 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[105] 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.

See also

Notes

  1. ^ Stella lucida in umero dextro, quae ad rubedinem vergit.[27]
    "Bright star in right shoulder, which inclines to ruddiness."
  2. ^ Likely the result of mistaking the l for an i. Ultimately, this led to the modern "Betelgeuse".
  3. ^ The final year of observations, unless otherwise noted
  4. ^ Uniform disk measurement, unless otherwise noted
  5. ^ Radii calculations use the same methodology as outlined in Note No. 2 below Limb darkened measurement

References

  1. ^ .
  2. ^ a b c "Merriam-Webster Dictionary: Betelgeuse". Retrieved 23 April 2018.
  3. ^
    S2CID 18759600
    .
  4. ^ a b c Nicolet, B. (1978). "Catalogue of Homogeneous Data in the UBV Photoelectric Photometric System". Astronomy & Astrophysics. 34: 1–49. .
  5. ^ a b "Alpha Orionis". Variable Star Index. Retrieved 20 February 2020.
  6. S2CID 123149047
    .
  7. ^ .
  8. ^ a b c 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. 1: B/gcvs.
  9. ^ 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: 165–186.
    S2CID 17804304
    .
  10. ^ .
  11. ^ .
  12. ^ .
  13. ^ Lambert, D.L.; Brown, J.A.; Hinkle, K.H.; Johnson, H.R. (September 1984). "Carbon, nitrogen and oxygem abundances in Betelgeuse". Astrophysical Journal. 284: 223–237.
    ISSN 0004-637X
    .
  14. ^ .
  15. .
  16. ^ .
  17. ^ .
  18. ^ Lobel, Alex; Dupree, Andrea K. (2000). "Modeling the variable chromosphere of α Orionis". The Astrophysical Journal. 545 (1): 454–74. .
  19. ^ Ramírez, Solange V.; Sellgren, K.; Carr, John S.; Balachandran, Suchitra C.; Blum, Robert; Terndrup, Donald M.; Steed, Adam (July 2000). "Stellar Iron Abundances at the Galactic Center". The Astrophysical Journal. 537 (1): 205–20.
    S2CID 14713550
    .
  20. ^ a b c d e Kervella, Pierre; Decin, Leen; Richards, Anita M.S.; Harper, Graham M.; McDonald, Iain; O'Gorman, Eamon; Montargès, Miguel; Homan, Ward; Ohnaka, Keiichi (2018). "The close circumstellar environment of Betelgeuse. V. Rotation velocity and molecular envelope properties from ALMA". Astronomy and Astrophysics. 609: A67.
    S2CID 54670700
    .
  21. ^ Ridpath, Ian. "Orion: the meaning of Betelgeuse". Star Tales. Retrieved 9 September 2021.
  22. ^ a b "Definition of BETELGEUSE". www.merriam-webster.com. Retrieved 22 July 2023.
  23. ^ Lebling, James (September–October 2010). "Arabic in the Sky". Aramco World. pp. 24–33.
  24. ^ "IAU Working Group on Star Names (WGSN)". International Astronomical Union (IAU). Retrieved 22 May 2016.
  25. ^ "Bulletin of the IAU Working Group on Star Names" (PDF). International Astronomical Union (IAU). Retrieved 28 July 2016 – via University of Rochester.
  26. ^ "IAU Catalog of Star Names". IAU Division C, Working Group on Star Names (WGSN). International Astronomical Union (IAU). Retrieved 28 July 2016 – via University of Rochester.
  27. ^ .
  28. ^ Brück, H. A. (11–15 July 1978). "P. Angelo Secchi, S.J. 1818–1878". In McCarthy, M.F.; Philip, A.G.D.; Coyne, G.V. (eds.). Proceedings of the IAU Colloquium 47. Spectral Classification of the Future. Vatican City, IT (published 1979). pp. 7–20. .
  29. ^ "Ancient Chinese suggest Betelgeuse is a young star". New Scientist. Vol. 92, no. 1276. 22 October 1981. p. 238. {{cite magazine}}: Unknown parameter |agency= ignored (help)[permanent dead link]
  30. ^ .
  31. ^ Boutsalis, Kelly (10 August 2020). "Teaching indigenous star stories". The Walrus. Retrieved 6 July 2021.
  32. ^ Hamacher, Duane W. (2018). "Observations of red-giant variable stars by Aboriginal Australians".
    S2CID 119453488
    .
  33. ^ .
  34. ^ Davis, Kate (December 2000). "Variable Star of the Month: Alpha Orionis". American Association of Variable Star Observers (AAVSO). Retrieved 10 July 2010.
  35. ^ a b c .
  36. ^ a b Kaler, James B. (2002). The Hundred Greatest Stars. New York, NY: Copernicus Books. p. 33. .
  37. ^ a b 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. .
  38. ^ a b c d e
    S2CID 21969744
    . 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.
  39. ^ a b Tenn, Joseph S. (June 2009). "Martin Schwarzschild 1965". The Bruce Medalists. Astronomical Society of the Pacific (ASP). Retrieved 28 September 2010.
  40. ^ .
  41. ^ .
  42. ^ a b c Bonneau, D.; .
  43. ^ a b Sutton, E.C.; Storey, J.W.V.; Betz, A.L.; Townes, C.H.; Spears, D.L. (1977). "Spatial heterodyne tnterferometry of VY Canis Majoris, Alpha Orionis, Alpha Scorpii, and R Leonis at 11 microns". .
  44. ^ Bernat, A.P.; Lambert, D.L. (November 1975). "Observations of the circumstellar gas shells around Betelgeuse and Antares". .
  45. ^ Dyck, H.M.; Simon, T. (February 1975). "Circumstellar dust shell models for Alpha Orionis". .
  46. ^ Boesgaard, A.M.; Magnan, C. (June 1975). "The circumstellar shell of alpha Orionis from a study of the Fe II emission lines". .
  47. ^ Bernat, David (2008). "Aperture masking interferometry". Ask an Astronomer. Astronomy department. Cornell University. Retrieved 15 October 2012.
  48. ^ a b c Buscher, D.F.; Baldwin, J.E.; Warner, P.J.; Haniff, C.A. (1990). "Detection of a bright feature on the surface of Betelgeuse". .
  49. ^ Wilson, R.W.; Dhillon, V.S.; Haniff, C.A. (1997). "The changing face of Betelgeuse". .
  50. ^ Burns, D.; Baldwin, J.E.; Boysen, R.C.; Haniff, C.A.; Lawson, P.R.; MacKay, C.D.; et al. (September 1997). "The surface structure and limb-darkening profile of Betelgeuse". .
  51. ^ a b 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. .
  52. ^ a b .
  53. ^ a b c d Gilliland, Ronald L.; Dupree, Andrea K. (May 1996). "First image of the surface of a star with the Hubble Space Telescope".
    doi:10.1086/310043. 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 10×10 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.
  54. ^ Cox, A.N., ed. (2000). Allen's Astrophysical Quantities. New York, NY: Springer-Verlag. .
  55. ^ Petersen, Carolyn Collins; Brandt, John C. (1998) [1995]. Hubble Vision: Further adventures with the Hubble Space Telescope (2nd ed.). Cambridge, UK: Cambridge University Press. pp. 91–92. .
  56. ^ 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): 2501–2512.
    S2CID 117596395
    .
  57. ^ a b c d Weiner, J.; Danchi, W.C.; Hale, D.D.S.; McMahon, J.; Townes, C.H.; Monnier, J.D.; Tuthill, P.G. (December 2000). "Precision measurements of the diameters of α Orionis and ο Ceti at 11 microns". The Astrophysical Journal. 544 (2): 1097–1100. .
  58. ^ a b c d e Sanders, Robert (9 June 2009). "Red Giant Star Betelgeuse Mysteriously Shrinking". UC Berkeley News. UC Berkeley. Retrieved 18 April 2010.
  59. ^ .
  60. ^ a b Ravi, V.; Wishnow, E.; Lockwood, S.; Townes, C. (December 2011). "The many faces of Betelgeuse". Astronomical Society of the Pacific. 448: 1025. .
  61. ^ Bernat, Andrew P. (1977). "The circumstellar shells and mass-loss rates of four M supergiants". Astrophysical Journal. 213: 756–66.
    S2CID 121146305
    .
  62. ^ a b c d Kervella, P.; Verhoelst, T.; Ridgway, S.T.; Perrin, G.; Lacour, S.; Cami, J.; Haubois, X. (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.
    S2CID 14278046
    .
  63. ^ a b c Guinan, Edward F.; Wasatonic, Richard J.; Calderwood, Thomas J. (23 December 2019). "Updates on the "fainting" of Betelgeuse". The Astronomer's Telegram. ATel #13365. Retrieved 27 December 2019.
  64. ^ a b Byrd, Deborah (23 December 2019). "Betelgeuse is 'fainting' but (probably) not about to explode". Earth & Sky. Retrieved 4 January 2020.
  65. ^ Overbye, Dennis (14 August 2020). "This star looked like it would explode. Maybe it just sneezed". The New York Times. Retrieved 15 August 2020. The mysterious dimming of the red supergiant Betelgeuse is the result of a stellar exhalation, astronomers say.
  66. ^ "Hubble finds that Betegeuse's mysterious dimming is due to a traumatic outburst" (Press release). Hubble Space Telescope. 13 August 2020.
  67. ^ Dupree, Adrea K.; et al. (13 August 2020). "Spatially resolved ultraviolet spectroscopy of the great dimming of Betelgeuse".
    S2CID 221103735
    .
  68. ^ Guinan, Edward F.; Wasatonic, Richard J. (1 February 2020). "Betelgeuse Updates – 1 February 2020; 23:20 UT". The Astronomer's Telegram. ATel #13439. Retrieved 2 February 2020.
  69. ^ a b Carlson, Erika K. (27 December 2019). "Betelguese's bizarre dimming has astronomers scratching their heads". Astronomy. Retrieved 28 December 2019.
  70. ^ Griffin, Andrew (29 December 2019). "Betelgeuse: Star is behaving strangely and could be about to explode into a supernova, say astronomers". The Independent. Retrieved 30 December 2019.
  71. ^ Mack, Erick (27 December 2019). "Betelgeuse star acting like it's about to explode, even if the odds say it isn't". CNET. Retrieved 30 December 2019.
  72. ^ a b c Drake, Nadia (26 December 2019). "A giant star is acting strange, and astronomers are buzzing". National Geographic Society. Archived from the original on 26 December 2019. Retrieved 26 December 2019. The red giant Betelgeuse is the dimmest seen in years, prompting some speculation that the star is about to explode. Here's what we know.
  73. ^ a b Kaplan, Sarah (27 December 2019). "Is Betelgeuse, one of the sky's brightest stars, on the brink of a supernova?". The Washington Post. Retrieved 28 December 2019.
  74. ^ a b c Iorio, Kelsie (27 December 2019). "Is Betelgeuse, the red giant star in the constellation Orion, going to explode?". ABC News. Australia. Retrieved 28 December 2019.
  75. ^ a b Sparks, Hannah (26 December 2019). "Massive 'Betelgeuse' star in Orion constellation due for explosive supernova". New York Post. Retrieved 28 December 2019.
  76. ^ Bruce Dorminey (17 February 2020). "Betelgeuse Has Finally Stopped Dimming, Says Astronomer". Forbes. Retrieved 19 February 2020.
  77. ^ Guinan, Edward; Wasatonic, Richard; Calderwood, Thomas; Carona, Donald (22 February 2020). "The fall and rise in brightness of Betelgeuse". The Astronomer's Telegram. ATel #13512. Retrieved 22 February 2020.
  78. ^ Gehrz, R.D.; et al. (24 February 2020). "Betelgeuse remains steadfast in the infrared". The Astronomer's Telegram. ATel #13518. Retrieved 24 February 2020.
  79. ^ "Dimming Betelgeuse likely isn't cold, just dusty, new study shows".
    EurekAlert! (Press release). University of Washington
    . 6 March 2020. Retrieved 6 March 2020.
  80. ^ Levesque, Emily M.; Massey, Philip (24 February 2020). "Betelgeuse just isn't that cool: Effective temperature alone cannot explain the recent dimming of Betelgeuse". The Astrophysical Journal Letters. 891 (2): L37.
    S2CID 211296241
    .
  81. ^ Dharmawardena, Thavisha E.; Mairs, Steve; Scicluna, Peter; Bell, Graham; McDonald, Iain; Menten, Karl; Weiss, Axel; Zijlstra, Albert (29 June 2020). "Betelgeuse fainter in the submillimeter too: An analysis of JCMT and APEX monitoring during the recent optical minimum".
    S2CID 219721417
    .
  82. ^ Sigismondi, Costantino (31 March 2020). "Rapid rising of Betelgeuse's luminosity". The Astronomer's Telegram. ATel #13601. Retrieved 1 April 2020.
  83. ^ Dupree, Andrea; Guinan, Edward; Thompson, William T.; et al. (STEREO/SECCHI/HI consortium) (28 July 2020). "Photometry of Betelgeuse with the STEREO Mission while in the glare of the Sun from Earth".
    Astronomer's Telegram
    . ATel #13901. Retrieved 28 July 2020.
  84. ^ Sigismondi, Costantino; et al. (30 August 2020). "Second dust cloud on Betelgeuse". The Astronomer's Telegram. ATel #13982. Retrieved 31 August 2020.
  85. ^ Montargès M, Cannon E, Lagadec E, et al. (16 June 2021). "A dusty veil shading Betelgeuse during its Great Dimming". Nature. 594 (7863): 365–368.
    S2CID 235460928
    .
  86. ^ Levesque, E. (16 June 2021). "Great dimming of Betelgeuse explained". Nature. 594 (7863): 343–344.
    S2CID 235459976
    .
  87. ^ Montargès, M. (16 June 2021). "Imaging the great dimming of Betelgeuse". Nature.
  88. ^ Overbye, Dennis (17 June 2021). "Betelgeuse merely burped, astronomers conclude". The New York Times. Retrieved 17 June 2021. The dramatic dimming of the red supergiant in 2019 was the product of dust, not a prelude to destruction, a new study has found.
  89. ^ Alexeeva, Sofya; Zhao, Gang; Gao, Dong-Yang; Du, Junju; Li, Aigen; Li, Kai; Hu, Shaoming (5 August 2021). "Spectroscopic evidence for a large spot on the dimming Betelgeuse". Nature Communications. 12 (1): 4719.
    PMID 34354072
    .
  90. ^ Harris, Margaret (6 August 2021). "New evidence supports dark-spot theory for Betelgeuse's 'great dimming'". Physics World. Retrieved 7 August 2021.
  91. ^ Dupree, Andrea K.; Strassmeier, Klaus G.; Calderwood, Thomas; Granzer, Thomas; Weber, Michael; Kravchenko, Kateryna; et al. (2 August 2022). "The great dimming of Betelgeuse: A surface mass ejection and its consequences".
    S2CID 251280168
    .
  92. ^ Garner, Rob (13 August 2020). "Hubble finds Betelgeuse's mysterious dimming due to traumatic outburst". NASA. Retrieved 22 August 2022.
  93. ^ "How Betelgeuse blew its top and lost its rhythm". Physics World. 22 August 2022. Retrieved 22 August 2022.
  94. ^ Goldberg, L. (May 1984). "The variability of alpha Orionis". Publications of the Astronomical Society of the Pacific. 96: 366.
    S2CID 121926262
    .
  95. ^ Sigismondi, Constantino; et al. (22 April 2023). "Monitoring Betelgeuse at its brightest". The Astronomer's Telegram. Atel #16001. Retrieved 22 April 2023.
  96. ^ 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.
  97. Centre de Données astronomiques de Strasbourg
    . Retrieved 22 August 2010.
  98. .
  99. ^ .
  100. .
  101. .
  102. ^ .
  103. ^ .
  104. ^ .
  105. ^ a b c d e f g h i Harper, Graham M.; Brown, Alexander; Guinan, Edward F. (April 2008). "A New VLA-Hipparcos Distance to Betelgeuse and its Implications". The Astronomical Journal. 135 (4): 1430–40. .
  106. ^ van Altena, W. F.; Lee, J. T.; Hoffleit, D. (October 1995). "Yale Trigonometric Parallaxes Preliminary". Yale University Observatory (1991). 1174: 0. .
  107. ^ "Hipparcos Input Catalogue, Version 2 (Turon+ 1993)".
    Centre de Données astronomiques de Strasbourg
    . 1993. Retrieved 20 June 2010.
  108. ^ Perryman, M.A.C.; Lindegren, L.; Kovalevsky, J.; Hoeg, E.; Bastian, U.; Bernacca, P.L.; et al. (1997). "The Hipparcos Catalogue". Astronomy & Astrophysics. 323: L49–L52. .
  109. ^ 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. 6th Vienna Workshop in Astrophysics. ASP Conference Series. Vol. 210. Vienna, Austria: .
  110. ^ "Science Performance". European Space Agency. 19 February 2013. Retrieved 1 March 2013.
  111. S2CID 9271090
    , retrieved 21 September 2016
  112. ^ "Welcome to the Gaia Archive". European Space Agency. Retrieved 3 September 2020.
  113. ^
    S2CID 5203133
    .
  114. .
  115. .
  116. .
  117. ^ .
  118. ^ . 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
  119. ^ Young, John (24 November 2006). "Surface Imaging of Betelgeuse with COAST and the WHT". University of Cambridge. Archived from the original on 14 June 2007. 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
  120. S2CID 16674990
    .
  121. .
  122. ^ 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.
  123. .
  124. ISBN 978-1-58381-020-0. {{cite book}}: |journal= ignored (help
    )
  125. ^ "Top Five Breakthroughs From Hubble's Workhorse Camera". NASA Jet Propulsion Laboratory, California Institute of Technology. 4 May 2009. Archived from the original on 7 May 2009. Retrieved 28 August 2007.
  126. ^ 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.
  127. (PDF) on 28 July 2011. Retrieved 29 August 2007.
  128. ^
    Astronomy Magazine
    . 2009. Retrieved 14 September 2012.
  129. ^ Nemiroff, R.; Bonnell, J., eds. (6 January 2010). "The Spotty Surface of Betelgeuse". Astronomy Picture of the Day. NASA. Retrieved 18 July 2010.
  130. ^ .
  131. ^ . 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
  132. ^ .
  133. ^ .
  134. ^ Cowen, Ron (10 June 2009). "Betelgeuse Shrinks: The Red Supergiant has Lost 15 Percent of its Size". Archived from the original on 29 June 2011. Retrieved 11 June 2009. 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.
  135. S2CID 15438522
    .
  136. ^ Denissenko, Denis (3 October 2004). "Unique occultations". Archived from the original on 16 December 2012.
  137. .
  138. ^ Sigismondi, Costantino (9 December 2023). "The occultation of Betelgeuse by Leona: recovering the stellar surface brightness of a red supergiant, with a diffuse telescope, on Dec 12 1:12 UT". The Astronomer's Telegram. Archived from the original on 12 December 2023. Retrieved 11 December 2023.
  139. .
  140. ^ "IOTA-ES". www.iota-es.de. Retrieved 8 December 2023.
  141. ^ "Astronomers brace for rare eclipse as asteroid to pass in front of bright star". The Guardian. Associated Press. 8 December 2023.
  142. ^ Hernandez, Joe (10 December 2023). "A massive star called Betelgeuse will be briefly obscured by an asteroid Monday night". NPR.
  143. ^ Guenot, Marianne (7 December 2023). "Betelgeuse, one of the brightest stars in the sky, will almost disappear next week. Here's how to see it". MSN. Archived from the original on 27 December 2023.
  144. Bibcode:1993AAS...183.1710G. Archived from the original
    on 25 June 2019. Retrieved 4 February 2012.
  145. ^ .
  146. Centre de Données astronomiques de Strasbourg
    . Retrieved 7 September 2012.
  147. .
  148. ^ 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.
    S2CID 54943572
    .
  149. ^ .
  150. .
  151. ^ Maeder, André; Meynet, Georges (2003). "The Role of Rotation and Mass Loss in the Evolution of Massive Stars". Proceedings of IAU Symposium. 212: 267. .
  152. ^ .
  153. .
  154. ^ Nemiroff, R.; Bonnell, J., eds. (23 October 2010). "Orion: Head to Toe". Astronomy Picture of the Day. NASA. Retrieved 8 October 2012.
  155. . A26.
  156. ^ 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. .
  157. .
  158. ^ a b c 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.
    S2CID 17850433
    .
  159. ^ Tsuji, T. (2000). "Water on the Early M Supergiant Stars α Orionis and μ Cephei". The Astrophysical Journal. 538 (2): 801–07. .
  160. .
  161. ^ a b c Dave Finley (8 April 1998). "VLA Shows "Boiling" in Atmosphere of Betelgeuse". National Radio Astronomy Observatory. Retrieved 7 September 2010.
  162. ^ 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.
    S2CID 4431516
    .
  163. ^ a b c 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.
    S2CID 15868906
    . 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.
  164. . 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."
  165. ^ a b 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. .
  166. ^ 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. .
  167. ^ 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. .
  168. ^ David, L.; Dooling, D. (1984). "The Infrared Universe". Space World. 2: 4–7. .
  169. .
  170. ^ .
  171. .
  172. ^ "Akari Infrared Space Telescope: Latest Science Highlights". European Space Agency. 19 November 2008. Archived from the original on 17 February 2011. Retrieved 25 June 2012.
  173. . 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
  174. ^ Newton, Elizabeth (26 April 2012). "This Star Lives in Exciting Times, or, How Did Betelgeuse Make that Funny Shape?". Astrobites. Archived from the original on 30 April 2012. Retrieved 25 June 2012.
  175. S2CID 118433862
    .
  176. ^ .
  177. ^ 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.
    S2CID 84177572
    .
  178. ^ Goldberg, Jared A.; Bauer, Evan B.; Howell, D. Andrew (2020). "Apparent magnitude of Betelgeuse as a type IIP supernova". Research Notes of the AAS. 4 (3): 35.
    S2CID 216398511
    .
  179. ^ Wheeler, J. Craig (2007). Cosmic Catastrophes: Exploding stars, black holes, and mapping the universe (2nd ed.). Cambridge, UK: Cambridge University Press. pp. 115–17. .
  180. ^ Connelly, Claire (19 January 2011). "Tatooine's twin suns – coming to a planet near you just as soon as Betelgeuse explodes". News.com.au. Archived from the original on 22 September 2012. Retrieved 14 September 2012.
  181. ^ Plait, Phil (1 June 2010). "Is Betelgeuse about to blow?". Bad Astronomy. Discovery. Archived from the original on 21 April 2011. Retrieved 14 September 2012.
  182. ^ O'Neill, Ian (20 January 2011). "Don't panic! Betelgeuse won't explode in 2012!". Discovery space news. Archived from the original on 23 January 2011. Retrieved 14 September 2012.
  183. ^ Plait, Phil (21 January 2011). "Betelgeuse and 2012". Bad Astronomy. Discovery. Archived from the original on 3 November 2012. Retrieved 14 September 2012.
  184. ^ Betz, Eric (14 February 2020). "When Betelgeuse goes supernova, what will it look like from Earth?". Astronomy Magazine. Retrieved 15 June 2023.
  185. ^ Guinan, Edward F.; Wasatonic, Richard J.; Calderwood, Thomas J. (8 December 2019). "ATel #13341 – The Fainting of the Nearby Red Supergiant Betelgeuse". The Astronomer's Telegram. Retrieved 27 December 2019.
  186. ^ Plait, Phil (8 September 2014). "When will Betelgeuse explode?". Slate. Retrieved 28 December 2019.
  187. ^ Wu, Katherine J. (26 December 2019). "A giant star is dimming, which could be a sign it is about to explode". Smithsonian. Retrieved 28 December 2019.
  188. ^ Feltman, Rachel (26 December 2019). "We really don't know when Betelgeuse is going to explode". Popular Science. Retrieved 28 December 2019.
  189. ^ Prior, Ryan (26 December 2019). "A giant red star is acting weird and scientists think it may be about to explode". CNN. Retrieved 28 December 2019.
  190. ^ Plait, Phil (24 December 2019). "Don't Panic! Betelgeuse is (almost certainly) not about to explode". Syfy Wire. Retrieved 28 December 2019.
  191. ^ Overbye, Dennis (9 January 2020). "Just a fainting spell? Or is Betelgeuse about to blow? – A familiar star in the constellation Orion has dimmed noticeably since October. Astronomers wonder if its explosive finale is imminent". The New York Times. Retrieved 12 January 2020.
  192. ^ 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.
  193. ^ Bode, Johann Elert, (ed.) (1801). Uranographia: sive Astrorum Descriptio, Fridericus de Harn, Berlin, pl. XII.
  194. ^ .
  195. ^ Martha Evans Martin (1907). The friendly stars. Harper & brothers. p. 19.
  196. .
  197. .
  198. ^ Kunitzsch, Paul (1959). Arabische Sternnamen in Europa. Wiesbaden: Otto Harrassowitz.
  199. ^ .
  200. ^ "天文教育資訊網 2006 年 5 月 25 日" [Astronomy Education Information Network 25 May 2006]. aeea.nmns.edu.tw. AEEA (Activities of Exhibition and Education in Astronomy). 25 May 2006. Archived from the original on 16 July 2011. Retrieved 26 June 2012.
  201. ^ Ridpath, Ian. "Orion: Chinese associations". Star Tales. Retrieved 24 June 2012.
  202. ^ Steve Renshaw & Saori Ihara. "Yowatashi Boshi; Stars that Pass in the Night". Archived from the original on 4 June 2016. Retrieved 25 June 2012.{{cite web}}: CS1 maint: unfit URL (link) Other Versions: "Yowatashi Boshi; Stars that Pass in the Night". Griffith Observer. Vol. 63, no. 10. October 1999. pp. 2–17. and "Yowatashi Boshi; Stars that Pass in the Night". The Kyoto Journal. No. 48. July 2000.
  203. .
  204. .
  205. .
  206. .
  207. .
  208. .
  209. .
  210. .
  211. .
  212. .
  213. . Retrieved 22 September 2010.
  214. ^ Tallant, Nicolla (15 July 2007). "Survivor recalls the night an apocalypse came to Whiddy". Independent Digital. Independent News & Media PLC. Retrieved 10 June 2011.
  215. ^ "Black and Blue Bird". Dmbalmanac.com. 5 July 2015. Retrieved 30 January 2016.
  216. ^ "Blur – Far Out Lyrics". genius.com. Retrieved 7 February 2020.
  217. .

External links

  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 telescope