Stars/Vega

From Wikiversity
Jump to navigation Jump to search
This is an infrared image of the debris disc around Vega taken with the Herschel Space Observatory. Credit: Herschel Space Observatory, Steward Observatory, University of Arizona.

Vega is the fifth brightest star in the sky, and the brightest in the constellation Lyra, overhead in summer to observers in the northern hemisphere.

Vega has been extensively studied, leading it to be termed "arguably the next most important star in the sky after the Sun".[1]

Astronomy[edit | edit source]

"Vega rotates in less than a day, while the Sun's rotation period is 27 days."[2]

From Earth, Vega is observed from the direction of one of its poles.[3]

Photometric measurements of Vega during the 1930s appeared to show that the star had a low-magnitude variability on the order of ±0.03 magnitudes (around ±2.8%).[4]

This range of variability was near the limits of observational capability for that time, and so the magnitude of Vega was measured again in 1981 at the David Dunlap Observatory and showed some occasional low-amplitude pulsations associated with a Delta Scuti variable.[5]

This is a category of stars that oscillate in a coherent manner, resulting in periodic pulsations in the star's luminosity.[6]

A 2007 article surveyed these and other results, and concluded that "A conservative analysis of the foregoing results suggests that Vega is quite likely variable in the 1-2% range, with possible occasional excursions to as much as 4% from the mean".[7] Also, a 2011 article affirms on its abstract that "The long-term (year-to-year) variability of Vega was confirmed".[8]

Astrophysics[edit | edit source]

The first person to publish a star's parallax was Friedrich Georg Wilhelm von Struve, when he announced a value of 0.125 arcseconds (0.125″) for Vega.[9]

Struve's initial result was actually close to the currently accepted value of 0.129″,[10][11] as determined by the Hipparcos astrometry satellite.[12][13][14]

Stars[edit | edit source]

At one time (1932), Vega was listed as a double star.[15]

In 1963, Vega is listed as a visual double star.[16]

In 1983, it was listed as a double star.[17] This was repeated in 1994.[18] It is still listed as a double star in 1996.[19] The update in 2001 also lists it.[20] The star with Vega is 56.41 arcsec away and is designated as BD+38 3238D of unknown spectral class. That these two stars are undifferentiated between double star or binary star for some 70 years at only 25 lyrs away is remarkable.

Vega was the northern pole star around 12,000 BC and will be so again around the year 13,727, when the declination will be +86°14'.[21]

Vega is the brightest of the successive northern pole stars.[22]

Electromagnetic interactions[edit | edit source]

An extensive convection zone is not required, and any star with magnetic field strengths and geometry similar to the Sun's will possess a corona.[23]

Magnetic fields on the order of ~30 gauss have been reported for Vega (~1 gauss for the Sun), so perhaps these substantially higher average field strengths compensate for the expected reduced convective activity, resulting in surface X-ray luminosities comparable to the quiet Sun.[23]

Using spectropolarimetry, a magnetic field has been detected on the surface of Vega by a team of astronomers at the Observatoire du Pic du Midi.[24] They "report the detection of a magnetic field on Vega and argue that Vega is probably the first member of a new class of yet undetected magnetic A-type stars."[24]

The "polarization [is] a Zeeman signature [that] leads to a value of [Bl =] -0.6 ± 0.3 G for the disk-averaged line-of-sight component of the surface magnetic field."[24]

"The strength of Vega magnetic field is about 50 micro-tesla, which is close to that of the mean field on Earth and on the Sun."[2]

Spectral distributions[edit | edit source]

This is a full spectrum of Vega (instrumental responsivity corrected). Credit: Buil/AstroSurf.
This graph shows the labeled peaks for the 380.0-500.0 nm domain. Credit: Buil/Astrosurf.
The spectrum shows the annotated lines for the 500.0-600.0 nm domain. Credit: Buil/Astrosurf.
This annotated spectrum contains the 600.0-700.0 nm domain. Credit: Buil/Astrosurf.
This annotated spectrum displays the 700.0-800.0 nm domain. Credit: Buil/Astrosurf.
This annotated spectrum displays the 800.0-900.0 nm domain. Credit: Buil/Astrosurf.
This annotated spectrum displays the 900.0-1020.0 nm domain. Credit: Buil/Astrosurf.

Vega is a nearby optical spectral type A0V star at about 25 lyrs. Its polar effective temperature is near 10,000 K, while its equatorial effective temperature is 7,600 K.

Vega was the first star other than the Sun to be photographed[25] and the first to have its spectrum recorded.[26]

Each of the annotated spectra at right display portions of the spectrum at the top of this resource:

  1. the wavelength domain from 380.0 nm to 500.0 nm,
  2. the 500.0-600.0 nm domain,
  3. the 600.0-700.0 nm domain,
  4. the 700.0-800.0 nm domain,
  5. the 800.0-900.0 nm domain, and
  6. the 900.0-1020.0 nm domain.

Each of these images at right has various element lines labeled: calcium, hydrogen, iron, magnesium, nitrogen, oxygen, silicon, sodium, and titanium.

Molecular lines occur for H2O and O2 which are likely from Earth's atmosphere or the accretion disk around Vega.

The applicable astronomies are

  1. violet astronomy, blue astronomy, cyan astronomy, and green astronomy,
  2. green astronomy, yellow astronomy, and red astronomy,
  3. red astronomy and near infrared astronomy,
  4. near infrared astronomy,
  5. near infrared astronomy, and
  6. near infrared astronomy.

The principal low atomic number elements of interest, whose lines need to be identified are

  1. hydrogen,
  2. helium,
  3. lithium,
  4. beryllium,
  5. boron,
  6. carbon,
  7. nitrogen,
  8. oxygen,
  9. fluorine, and
  10. neon.

Element lines and wavelengths can be verified using these resources:

The element lines to look for and verify whether present or not are

  1. hydrogen, all high intensity lines detected and state properly labeled.
  2. helium, the two major lines at 388.86 and 706.519 nm are possible but may overlap the hydrogen line at 388.90 nm.
  3. lithium - no lines detected.
  4. beryllium - no lines detected.
  5. boron - two strong lines are possible, the 412.1933 nm as a shoulder on H I 410.17 nm line, while the other 448.705 nm overlaps the labeled 448.11 Mg II line.
  6. carbon, the line at 723.642 nm is possible but unlabeled.
  7. nitrogen, high intensity lines labeled, but the 868.028 N I is possible as an unlabeled shoulder.
  8. oxygen - the 777.42 nm O I may be mislabeled and should be 777.194 nm O I which is the stronger line.
  9. fluorine - the 685.603 nm F I line is possible.
  10. neon - not detected.

For the heavier elements:

  1. sodium - the two Na I lines are properly labeled.
  2. magnesium - all of the lines may be wrong and mislabeled.
  3. calcium - all high intensity lines present and labeled, the Ca K is a Ca II line.
  4. titanium - the Ti II line at 439.50 nm is properly labeled as are the others.
  5. silicon - lines are properly labeled.
  6. iron - all are labeled properly, because the 501.84 nm line is only Fe II, which differentiates Fe I lines from Fe II lines.

X-rays[edit | edit source]

According to SIMBAD, Vega (alpha Lyrae) (delta Sct type variable) is an X-ray source in the first Einstein catalog (1E).

The X-ray "counts observed from [...] Vega in the HRI are very likely to be due entirely to UV contamination from the photospheric emission, and hence the X-ray luminosities for [...] Vega [...] should be replaced by upper limits at least one order of magnitude lower, i.e., log LX, Vega < 26.6 [...] Thus the HRI observation of Vega is completely consistent with the upper limit obtained in the IPC, and the only remaining inconsistency concerning the X-ray emission from Vega is the rocket experiment by Topka et al. (1979), who report a detection of Vega in a 5 s pointing yielding 7 counts; however, in our opinion these authors do not convincingly rule out the possibility of UV contamination. Note in this context that the IPC's used for Topka et al.s' (1979) rocket flight and for the Einstein Observatory were not identical."[27]

"Many types of main sequence stars emit in the X-ray portion of the spectra. In massive stars, strong stellar winds ripping through the extended atmosphere of the star create X-ray photons. On lower mass stars, magnetic fields twisting through the photosphere heat it sufficiently to produce X-rays. But between these two mechanisms, in the late B to mid A classes of stars, neither of these mechanisms should be sufficient to produce X-rays. Yet when X-ray telescopes examined these stars, many were found to produce X-rays just the same."[28]

"The first exploration into the X-ray emission of this class of stars was the Einstein Observatory, launched in 1978 and deorbited in 1982. While the telescoped confirmed that these B and A stars had significantly less X-ray emission overall, seven of the 35 A type stars still had some emission. Four of these were confirmed as being in binary systems in which the secondary stars could be the source of the emission, leaving three of seven with unaccounted for X-rays."[28]

"The German ROSAT satellite found similar results, detecting 232 X-ray stars in this range. Studies explored connections with irregularities in the spectra of these stars and rotational velocities, but found no correlation with either. The suspicion was that these stars simply hid undetected, lower mass companions."[28]

Either "the main star truly is the source, or there are even more elusive, sub-arcsecond binaries skewing the data."[28]

On July 27, 1977, at 05:41:48.1 UTC, an Aerobee 350 or boosted Black Brant launched from White Sands Missile Range using Vega as a reference by its star tracker to update its position while maneuvering between X-ray targets automatically observed Vega with its X-ray telescope for 4.8 s.[23]

The quantity of detected photons (7) in the band 0.2-0.80 keV corresponds to an X-ray luminosity LX ≈ 3 x 1028 erg s-1.[23]

"The ANS 3 σ upper limit for Vega (2.5 x 1028 ergs s-1) is only slightly lower than our flux measurement."[23]

"Because the X-ray [luminosity] of Vega [is] much closer to that of the Sun than to the typical galactic X-ray sources which have been detected to date, it is natural to consider processes analogous to solar coronal activity as the explanation for the X-ray activity."[23]

"Vega is thought to be a solitary star, and therefore noncoronal X-ray-producing mechanisms seem to be excluded".[23]

"Vega is the first solitary main-sequence star beyond the Sun known to be an X-ray emitter".[23]

Vega's "computed X-ray surface luminosity [...] is comparable to that of the quiet Sun [...] Note, however, that because of our very short exposure, the average level of coronal emission may vary significantly from our single measurement."[23]

"Using estimates of the stellar [radius] derived from stelar structure calculations, we obtain [a] surface X-ray [luminosity] of ~6.4 x 104 ergs cm-2 s-1 for Vega [that falls] within the range of solar coronal X-ray emission, which can vary between ~8 x 103 ergs cm-2 s-1 in coronal holes and ~3 x 106 ergs cm-2 s-1 in active regions".[23]

Magnetic "field activity, leading to coronal heating, may account for Vega's X-ray emission because of inhomogeneous distribution of surface magnetic flux and associated coronal activity."[23]

That Vega is regarded as an X-ray source rests on one 4.8 s star-tracking observation by one sounding rocket flight carrying an X-ray detector flown on many flights that yields trustworthy results.

"Vega is a pole-on, highly oblate, rapid rotator [...] the star exhibits extreme limb darkening and a large decrement in effective temperature from pole to equator. [...] the best fittingmodel (Teff pole=10150 K, Teff eq = 7900 K, θ = 3.329 mas) has the pole inclined 5° to line of sight and rotates at 91% of the angular speed of break-up, resulting in a temperature drop of 2250 K from center to limb. [...] the total luminosity [...] is emitted in a highly non-homogeneous manner with five times more UV flux being emitted from the pole as is emitted in the equatorial plane, while the visible through near-IR flux is some 70% greater at the pole than that of the equatorial plane and 54% greater than that expected from a slow or non-rotating A0 V star."[29]

A 'polar coronal hole' is a coronal hole that occurs above one or both rotational poles of a star that has a coronal cloud around it.

"The radiant emission from coronal holes is greatly diminished relative to other coronal regions".[30]

The "emission is proportional to the integral of the square of the electron density along the line of sight [...] Data of this type are therefore heavily influenced by regions of high density along the line of sight--the low corona for disk observations, and denser structures surrounding coronal holes for limb observations."[30]

An "analysis of the northern polar region during the period 1973 June 29 to July 13 [...] can be summarized as follows. The boundary of the hole is essentially axisymmetrc about the polar axis and is nearly radial from 3 to 6 R. The boundary at these heights is located at 25° ± 5° latitude, although it is of much smaller extent (boundary ~65° latitude) as observed near the solar surface with the American Science and Engineering (AS&E) X-ray experiment on Skylab [...] the increase of the polar hole's cross sectional area from the surface to 3 R is approximately 7 times greater than for a purely radial boundary."[30]

For α Lyr, log FX/FV = -6.79 (variable X-ray source), log LX 27.6 erg s-1 (variable X-ray source).[31] Upper limits were log FX/FV = -7.4 and log LX 27.0 erg s-1.[31]

A coronal cloud is not a diffuse, homogeneous hot atmosphere, but one or more strongly structured topologically closed features dominated by magnetic confinement.

Ultraviolets[edit | edit source]

According to SIMBAD, Vega (alpha Lyrae) (delta Sct type variable) is an ultraviolet source from the CEL, EUVE, and TD1 catalogs.

Infrareds[edit | edit source]

These are infrared images of Vega. Credit: NASA/JPL-Caltech/University of Arizona.
This is a graph of infrared excesses including Vega. Credit: Herschel Space Observatory, Steward Observatory, University of Arizona.

The images at the right has a 24-micron image in blue on the left and a 70-micron infrared image on the right. "The [debris] disc extends to at least 815 astronomical units."[32] "The images are 3 arcminutes on each side."[32]

BD+38 3238D is unspecified or unlocated within the infrared images but should be within the debris disc.

According to SIMBAD there are ten objects within one arcminute of Vega including BD+38 3238D, a submillimeter source (JCMTSE J183656.4+384709), and eight infrared sources: [ MHW2003] 1-8.

"The infrared excesses are well modeled by two components, a warm belt close to the star, and a cooler belt farther out. The clear separation of the belts could be explained by the presence of planets clearing the gap."[33]

The graph at left shows the clear separation of infrared belts for Vega. This separation may "be explained by the presence of planets clearing the gap."[33]

Metallicities[edit | edit source]

Vega has an unusually low abundance of the elements with a higher atomic number than that of helium.[34]

Its metallicity [Fe/H] = −0.5[34] dex.

The metallicity of Vega's photosphere is only about 32% of the abundance of heavy elements in the Sun's atmosphere. For a metallicity of −0.5, the proportion of metals relative to the Sun is given by

.[35]

See also[edit | edit source]

References[edit | edit source]

  1. Gulliver, Austin F.; Hill, Graham; Adelman, Saul J. (1994). "Vega: A rapidly rotating pole-on star". The Astrophysical Journal 429 (2): L81–L84. doi:10.1086/187418. 
  2. 2.0 2.1 Pascal Petit (July 26, 2009). Magnetic Field On Bright Star Vega. Science Daily. http://www.sciencedaily.com/releases/2009/06/090623111947.htm. Retrieved 2014-04-06. 
  3. Peterson, D. M.; Hummel, C. A.; Pauls, T. A.; Armstrong, J. T.; Benson, J. A.; Gilbreath, G. C.; Hindsley, R. B.; Hutter, D. J. et al. (2006). "Vega is a rapidly rotating star". Nature 440 (7086): 896–899. doi:10.1038/nature04661. PMID 16612375. 
  4. Arthur N. Cox, ed (1999). Allen's Astrophysical Qualities (4th ed.). New York: Springer-Verlag. p. 382. ISBN 0-387-98746-0. 
  5. J. D. Fernie (1981). "On the variability of Vega". Publications of the Astronomical Society of the Pacific 93 (2): 333–337. doi:10.1086/130834. 
  6. A. Gautschy; H. Saio (1995). "Stellar Pulsations Across The HR Diagram: Part 1". Annual Review of Astronomy and Astrophysics 33 (1): 75–114. doi:10.1146/annurev.aa.33.090195.000451. 
  7. Raymond Gray (2007). The Problems with Vega, In: The Future of Photometric, Spectrophotometric and Polarimetric Standardization, ASP Conference Series, Proceedings of a conference held 8–11 May 2006 in Blankenberge, Belgium. 364. pp. 305-. Bibcode: 2007ASPC..364..305G. 
  8. Varvara Butkovskaya (2011). "The long-term variability of Vega". Astronomische Nachrichten 332 (9-10): 956–960. doi:10.1002/asna.201111587. http://onlinelibrary.wiley.com/doi/10.1002/asna.201111587/abstract. 
  9. Arthur Berry (1899). A Short History of Astronomy. New York: Charles Scribner's Sons. ISBN 0-486-20210-0. 
  10. Suzanne Débarbat (1988), The First Successful Attempts to Determine Stellar Parallaxes in the Light of the Bessel/Struve Correspondence, In: Mapping the Sky: Past Heritage and Future Directions, Springer, ISBN 90-277-2810-0
  11. Anonymous (2007-06-28). The First Parallax Measurements. Astroprof. http://astroprofspage.com/archives/1011. Retrieved 2007-11-12. 
  12. F. van Leeuwen (November 2007). "Validation of the new Hipparcos reduction". Astronomy and Astrophysics 474 (2): 653–664. doi:10.1051/0004-6361:20078357. 
  13. Perryman, M. A. C.; Lindegren, L.; Kovalevsky, J.; Hoeg, E.; Bastian, U.; Bernacca, P. L.; Crézé, M.; Donati, F. et al. (1997). "The Hipparcos Catalogue". Astronomy and Astrophysics 323: L49–L52. 
  14. Perryman, Michael (2010). The Making of History's Greatest Star Map. Heidelberg: Springer-Verlag. doi:10.1007/978-3-642-11602-5. 
  15. Robert Grant Aitken; Eric Doolittle (1932). New general catalogue of double stars within 120 of the North pole. Publication 417. Washington, D.C.: Carnegie institution of Washington. Bibcode: 1932ADS...C......0A. http://cdsbib.u-strasbg.fr/cgi-bin/cdsbib?1932ADS...C......0A. Retrieved 2014-04-02. 
  16. H. M. Jeffers; W. H. Van Den Bos; F. M. Greeby (1963). "Index catalogue of visual double stars, 1961.0". Publications of the Lick Observatory 21 (1). http://cdsbib.u-strasbg.fr/cgi-bin/cdsbib?1963IDS...C......0J. Retrieved 2014-04-04. 
  17. J. Dommanget (March 1983). "Un catalogue des composantes d'etoiles doubles et multiples (C.C.D.M.)". Bulletin d'Information du Centre de Donnees Stellaires (24): 83-90. http://cdsbib.u-strasbg.fr/cgi-bin/cdsbib?1983BICDS..24...83D. Retrieved 2014-04-04. 
  18. J. Dommanget; O. Nys (1994). "Catalogue des composantes d'etoiles doubles et multiples (CCDM) premiere edition - Catalogue of the components of double and multiple stars (CCDM) first edition". Com. de l'Observ. Royal de Belgique (115): 1. http://cdsbib.u-strasbg.fr/cgi-bin/cdsbib?1994CoORB.115....1D. Retrieved 2014-04-04. 
  19. C. E. Worley; G. G. Douglass (November 1997). "The Washington Double Star Catalog (WDS, 1996.0)". Astronomy & Astrophysics, Supplement Series 125 (1): 523. doi:10.1051/aas:1997239. http://cdsbib.u-strasbg.fr/cgi-bin/cdsbib?1997A%26AS..125..523W. Retrieved 2014-04-04. 
  20. Brian D. Mason; Gary L. Wycoff. William I. Hartkopf; Geoffrey G. Douglass; Charles E. Worley (December 2001). "The 2001 US Naval Observatory double star CD-ROM. I. The Washington double star catalog". Journal of Astronomy 122 (6): 3466-71. http://cdsbib.u-strasbg.fr/cgi-bin/cdsbib?2001AJ....122.3466M. Retrieved 2014-04-04. 
  21. Calculation by the Stellarium application version 0.10.2. http://www.stellarium.org. Retrieved 2009-07-28. 
  22. Richard Hinckley Allen (1963). Star Names: Their Lore and Meaning. Courier Dover Publications. ISBN 0-486-21079-0. 
  23. 23.00 23.01 23.02 23.03 23.04 23.05 23.06 23.07 23.08 23.09 23.10 K. Topka; D. Fabricant; F. R. Harnden Jr.; P. Gorenstein; R. Rosner (April 15, 1979). "Detection of Soft X-rays from α Lyrae and η Bootis with an Imaging X-ray Telescope". The Astrophysical Journal 229 (04): 661-8. doi:10.1086/157000. http://adsabs.harvard.edu/abs/1979ApJ...229..661T. Retrieved 2014-04-04. 
  24. 24.0 24.1 24.2 F. Lignières; P. Petit; T. Böhm; M. Aurière (June 2009). "First evidence of a magnetic field on Vega. Towards a new class of magnetic A-type stars". Astronomy and Astrophysics 500 (3): L41-4. doi:10.1051/0004-6361/200911996. http://adsabs.harvard.edu/abs/2009A%26A...500L..41L. Retrieved 2014-04-06. 
  25. Barger, M. Susan; White, William B. (1991). The Daguerreotype: Nineteenth-Century Technology and Modern Science. JHU Press. p. 88. ISBN 0-8018-6458-5. 
  26. Barker, George F. (1887). "On the Henry Draper Memorial Photographs of Stellar Spectra". Proceedings of the American Philosophical Society 24: 166–172. 
  27. J. H. M. M. Schmitt; L. Golub; F. R. Harnden Jr.; C. W. Maxson; R. Rosner; G. S. Vaiana (March 1, 1985). "An Einstein Observatory X-ray Survey of Main-Sequence Stars with Shallow Convection Zones". The Astrophysical Journal 290 (03): 307-20. doi:10.1086/162986. http://adsabs.harvard.edu/abs/1985ApJ...290..307S. Retrieved 2014-04-04. 
  28. 28.0 28.1 28.2 28.3 Jon Voisey (March 24, 2011). Companion Stars Could Cause Unexpected X-Rays. Universe Today. http://www.universetoday.com/84388/companion-stars-could-cause-x-rays/. Retrieved 2014-04-04. 
  29. Charles W. Engelke; Stephan D. Price; Kathleen E. Kraemer (December 2010). "Spectral Irradiance Calibration in the Infrared. XVII. Zero-Magnitude Broadband Flux Reference for Visible-to-Infrared Photometry". The Astronomical Journal 140 (6): 1919-28. doi:10.1088/0004-6256/140/6/1919. http://adsabs.harvard.edu/abs/2010AJ....140.1919E. Retrieved 2014-04-05. 
  30. 30.0 30.1 30.2 Richard H. Munro; Bernard V. Jackson (May 1, 1977). "Physical Properties of a Polar Coronal Hole from 2 to 5 R". The Astrophysical Journal 213 (05): 874-5, 877-86. doi:10.1086/155220. http://adsabs.harvard.edu/abs/1977ApJ...213..874M. Retrieved 2014-04-05. 
  31. 31.0 31.1 G. S. Vaiana; J. P. Cassinelli; G. Fabbiano; R. Giacconi; L. Golub; P. Gorenstein; B. M. Haisch; F.R. Harnden Jr. et al. (April 1, 1981). "Results from an extensive Einstein stellar survey". The Astrophysical Journal 244 (04): 163-82. doi:10.1086/158797. http://adsabs.harvard.edu/abs/1981ApJ...245..163V. Retrieved 2014-04-06. 
  32. 32.0 32.1 Sue Lavoie (January 10, 2005). PIA07218: Tiny Particles, So Far Away. Pasadena, California USA: California Institute of Technology. http://photojournal.jpl.nasa.gov/catalog/PIA07218. Retrieved 2014-04-04. 
  33. 33.0 33.1 Jessica Donaldson (January 20, 2013). Asteroid belt found in the Vega System. Astrobites. http://astrobites.com/2013/01/20/asteroid-belt-found-in-the-vega-system/figure1-8/. Retrieved 2014-04-04. 
  34. 34.0 34.1 Kinman, T.; Castelli, F. (2002). "The determination of Teff for metal-poor A-type stars using V and 2MASS J, H and K magnitudes". Astronomy and Astrophysics 391 (3): 1039–1052. doi:10.1051/0004-6361:20020806. 
  35. Francesca Matteucci (2001). The Chemical Evolution of the Galaxy. Astrophysics and Space Science Library. 253. Springer Science & Business Media. p. 7. ISBN 0792365526. https://books.google.com/books?id=PT7O1nS7CksC&pg=PA7. 

External links[edit | edit source]

{{Radiation astronomy resources}}