Gravitational microlensing

Gravitational microlensing

Gravitational microlensing is an astronomical phenomenon due to the gravitational lens effect. It can be used to detect objects ranging from the mass of a planet to the mass of a star, regardless of the light they emit. Typically, astronomers can only detect bright objects that emit lots of light (stars) or large objects that block background light (clouds of gas and dust). These objects make up only a tiny fraction of the mass of a galaxy. Microlensing allows the study of objects that emit little or no light.

When a distant star or quasar gets sufficiently aligned with a massive compact foreground object, the bending of light due to its gravitational field, as discussed by Einstein in 1915, leads to two distorted unresolved images resulting in an observable magnification. The time-scale of the transient brightening depends on the mass of the foreground object as well as on the relative proper motion between the background 'source' and the foreground 'lens' object.

Since microlensing observations do not rely on radiation received from the lens object, this effect therefore allows astronomers to study massive objects no matter how faint. It is thus an ideal technique to study the galactic population of such faint or dark objects as brown dwarfs, red dwarfs, planets, white dwarfs, neutron stars, black holes, and Massive Compact Halo Objects. Moreover, the microlensing effect is wavelength-independent, allowing study of source objects that emit any kind of electromagnetic radiation.

Microlensing by an isolated object was first detected in 1993. Since then, microlensing has been used to constrain the nature of the dark matter, detect extrasolar planets, study limb darkening in distant stars, constrain the binary star population, and constrain the structure of the Milky Way's disk. Microlensing has also been proposed as a means to find dark objects like brown dwarfs and black holes, study starspots, measure stellar rotation, and probe quasars[1][2] including their accretion disks.[3][4][5][6]

Contents

How it works

Microlensing is based on the gravitational lens effect. A massive object (the lens) will bend the light of a bright background object (the source). This can generate multiple distorted, magnified, and brightened images of the background source.[7]

Microlensing is caused by the same physical effect as strong lensing and weak lensing, but it is studied using very different observational techniques. In strong and weak lensing, the mass of the lens is large enough (mass of a galaxy or a galaxy cluster) that the displacement of light by the lens can be resolved with a high resolution telescope such as the Hubble Space Telescope. With microlensing, the lens mass is too low (mass of a planet or a star) for the displacement of light to be observed easily, but the apparent brightening of the source may still be detected. In such a situation, the lens will pass by the source in a reasonable amount of time, seconds to years instead of millions of years. As the alignment changes, the source's apparent brightness changes, and this can be monitored to detect and study the event. Thus, unlike with strong and weak gravitational lenses, a microlensing event is a transient phenomenon from a human timescale perspective.[8]

Unlike with strong and weak lensing, no single observation can establish that microlensing is occurring. Instead the rise and fall of the source brightness must be monitored over time using photometry. This function of brightness versus time is known as a light curve. A typical microlensing light curve is shown below:

Typical light curve of gravitational microlensing event (OGLE-2005-BLG-006) with its model fitted (red)

A typical microlensing event like this one has a very simple shape, and only one physical parameter can be extracted: the time scale, which is related to the lens mass, distance, and velocity. There are several effects, however, that contribute to the shape of more atypical lensing events:

  • Lens mass distribution. If the lens mass is not concentrated in a single point, the light curve can be dramatically different, particularly with caustic-crossing events, which may exhibit strong spikes in the light curve. In microlensing, this can be seen when the lens is a binary star or a planetary system.
  • Finite source size. In extremely bright or quickly-changing microlensing events, like caustic-crossing events, the source star cannot be treated as an infinitesimally small point of light: the size of the star's disk and even limb darkening can mollify extreme features.
  • Parallax. For events lasting for months, the motion of the Earth around the Sun can cause the alignment to change slightly, affecting the light curve.

Most focus is currently on the more unusual microlensing events, especially those that might lead to the discovery of extrasolar planets. Although it has not yet been observed, another way to get more information from microlensing events that may soon be feasible involves measuring the astrometric shifts in the source position during the course of the event[9] and even resolving the separate images with interferometry.[10]

Observing microlensing

In practice, because the alignment needed is so precise and difficult to predict, microlensing is very rare. Events, therefore, are generally found with surveys, which photometrically monitor tens of millions of potential source stars, every few days for several years. Dense background fields suitable for such surveys are nearby galaxies, such as the Magellanic Clouds and the Andromeda galaxy, and the Milky Way bulge. In each case, the lens population studied comprises the objects between Earth and the source field: for the bulge, the lens population is the Milky Way disk stars, and for external galaxies, the lens population is the Milky Way halo, as well as objects in the other galaxy itself. The density, mass, and location of the objects in these lens populations determines the frequency of microlensing along that line of sight, which is characterized by a value known as the optical depth due to microlensing. (This is not to be confused with the more common meaning of optical depth, although it shares some properties.) The optical depth is, roughly speaking, the average fraction of source stars undergoing microlensing at a given time, or equivalently the probability that a given source star is undergoing lensing at a given time. The MACHO project found the optical depth toward the LMC to be 1.2×10−7 or about 1 in 8,000,000,[11] and the optical depth toward the bulge to be 2.43×10−6 or about 1 in 400,000.[12]

Complicating the search is the fact that for every star undergoing microlensing, there are thousands of stars changing in brightness for other reasons (about 2% of the stars in a typical source field are naturally variable stars) and other transient events (such as novae and supernovae), and these must be weeded out to find true microlensing events. After a microlensing event in progress has been identified, the monitoring program that detects it often alerts the community to its discovery, so that other specialized programs may follow the event more intensively, hoping to find interesting deviations from the typical light curve. This is because these deviations – particularly ones due to exoplanets – require hourly monitoring to be identified, which the survey programs are unable to provide while still searching for new events. The question of how to prioritize events in progress for detailed followup with limited observing resources is very important for microlensing researchers today.

History

In 1704 Isaac Newton suggested that a light ray could be deflected by gravity. In 1801 Johann Georg von Soldner calculated the amount of deflection of a light ray from a star under Newtonian gravity. In 1915 Einstein correctly predicted the amount of deflection under General Relativity, which was twice the amount predicted by von Soldner. Einstein's prediction was validated by a 1919 expedition led by Arthur Eddington, which was a great early success for General Relativity.[13] In 1924 Orest Chwolson found that lensing could produce multiple images of the star. A correct prediction of the concomitant brightening of the source, the basis for microlensing, was published in 1936 by Einstein.[14] Because of the unlikely alignment required, he concluded that "there is no great chance of observing this phenomenon". Gravitational lensing's modern theoretical framework was established with works by Yu Klimov (1963), Sidney Liebes (1964), and Sjur Refsdal (1964).[1]

Gravitational lensing was first observed in 1979, in the form of a quasar lensed by a foreground galaxy. That same year Kyongae Chang and Sjur Refsdal showed that individual stars in the lens galaxy could act as smaller lenses within the main lens, causing the source quasar's images to fluctuate on a timescale of months.[15] Bohdan Paczyński first used the term "microlensing" to describe this phenomenon. This type of microlensing is difficult to identify because of the intrinsic variability of quasars, but in 1989 Mike Irwin et al. published detection of microlensing in Huchra's Lens.

In 1986, Paczyński proposed using microlensing to look for dark matter in the form of massive compact halo objects (MACHOs) in the Galactic halo, by observing background stars in a nearby galaxy. Two groups of particle physicists working on dark matter heard his talks and joined with astronomers to form the Anglo-Australian MACHO collaboration[16] and the French EROS[17] collaboration. In 1991 Paczyński suggested that microlensing might be used to find planets, and in 1992 he founded the OGLE microlensing experiment,[18] which began searching for events in the direction of the Galactic bulge.

The first two microlensing events in the direction of the Large Magellanic Cloud that might be caused by dark matter were reported in back to back Nature papers by MACHO[19] and EROS[20] in 1993, and in the following years, events continued to be detected. The MACHO collaboration ended in 1999. Their data refuted the hypothesis that 100% of the dark halo comprises MACHOs, but they found a significant unexplained excess of roughly 20% of the halo mass, which might be due to MACHOs or to lenses within the Large Magellanic Cloud itself.[21] EROS subsequently published even stronger upper limits on MACHOs,[22] and it is currently uncertain as to whether there is any halo microlensing excess that could be due to dark matter at all. The SuperMACHO project[23] currently underway seeks to locate the lenses responsible for MACHO's results.

Despite not solving the dark matter problem, microlensing has been shown to be a useful tool for many applications. Hundreds of microlensing events are detected per year toward the Galactic bulge, where the microlensing optical depth (due to stars in the Galactic disk) is about 20 times greater than through the Galactic halo. In 2007, the OGLE project identified 611 event candidates, and the MOA project (a Japan-New Zealand collaboration)[24] identified 488 (although not all candidates turn out to be microlensing events, and there is a significant overlap between the two projects). In addition to these surveys, followup projects are underway to study in detail potentially interesting events in progress, primarily with the aim of detecting extrasolar planets. These include PLANET,[25] RoboNet[26] and MicroFUN.[27]

Mathematics

The mathematics of microlensing, along with modern notation, are described by Gould[28] and we use his notation in this section, though other authors have used other notation. The Einstein radius, also called the Einstein angle, is the angular radius of the Einstein ring in the event of perfect alignment. It depends on the lens mass M, the distance of the lens dL, and the distance of the source dS:

\theta_E = \sqrt{\frac{4GM}{c^2} \frac{d_S - d_L}{d_S d_L}} (in radians)

For M equal to the mass of the Sun, dL = 4000 parsecs, and dS = 8000 parsecs (typical for a Bulge microlensing event), the Einstein radius is 0.001 arcseconds (1 milliarcsecond). By comparison, ideal Earth-based observations have angular resolution around 0.4 arcseconds, 400 times greater. Since θE is so small, it is not generally observed for a typical microlensing event, but it can be observed in some extreme events as described below.

Although there is no clear beginning or end of a microlensing event, by convention the event is said to last while the angular separation between the source and lens is less than θE. Thus the event duration is determined by the time it takes the apparent motion of the lens in the sky to cover an angular distance θE. The Einstein radius is also the same order of magnitude as the angular separation between the two lensed images, and the astrometric shift of the image positions throughout the course of the microlensing event.

During a microlensing event, the brightness of the source is amplified by an amplification factor A. This factor depends only on the closeness of the alignment between observer, lens, and source. The unitless number u is defined as the angular separation of the lens and the source, divided by θE. The amplification factor is given in terms of this value:

A(u) = \frac{u^2 + 2}{u \sqrt{u^2 + 4}}

This function has several important properties. A(u) is always greater than 1, so microlensing can only increase the brightness of the source star, not decrease it. A(u) always decreases as u increases, so the closer the alignment, the brighter the source becomes. As u approaches infinity, A(u) approaches 1, so that at wide separations, microlensing has no effect. Finally, as u approaches 0, A(u) approaches infinity as the images approach an Einstein ring. For perfect alignment (u = 0), A(u) is theoretically infinite. In practice, finite source size effects will set a limit to how large an amplification can occur for very close alignment, but some microlensing events can cause a brightening by a factor of hundreds.

Unlike gravitational macrolensing where the lens is a galaxy or cluster of galaxies, in microlensing u changes significantly in a short period of time. The relevant time scale is called the Einstein time tE, and it's given by the time it takes the lens to traverse an angular distance θE relative to the source in the sky. For typical microlensing events, tE is on the order of a few days to a few months. The function u(t) is simply determined by the Pythagorean theorem:

u(t) = \sqrt{u_{min}^2 + \left ( \frac{t-t_0}{t_E} \right )^2}

The minimum value of u, called umin, determines the peak brightness of the event.

In a typical microlensing event, the light curve is well fit by assuming that the source is a point, the lens is a single point mass, and the lens is moving in a straight line: the point source-point lens approximation. In these events, the only physically significant parameter that can be measured is the Einstein timescale tE. Since this observable is a degenerate function of the lens mass, distance, and velocity, we cannot determine these physical parameters from a single event.

However, in some extreme events, θE may be measurable while other extreme events can probe an additional parameter: the size of the Einstein ring in the plane of the observer, known as the Projected Einstein radius: \tilde{r}_E. This parameter describes how the event will appear to be different from two observers at different locations, such as a satellite observer. The projected Einstein radius is related to the physical parameters of the lens and source by

\tilde{r}_E = \sqrt{\frac{4GM}{c^2} \frac{d_S d_L}{d_S - d_L}}.

It is mathematically convenient to use the inverses of some of these quantities. These are the Einstein proper motion

\vec{\mu}_E = {t_E}^{-1}

and the Einstein parallax

\vec{\pi}_E = {\tilde{r}_E}^{-1}.

These vector quantities point in the direction of the relative motion of the lens with respect to the source. Some extreme microlensing events can only constrain one component of these vector quantities. Should these additional parameters be fully measured, the physical parameters of the lens can be solved yielding the lens mass, parallax, and proper motion as

M=\frac{c^2}{4G}\theta_E \tilde{r}_E

πL = πEθE + πS

μL = μEθE + μS

Extreme microlensing events

In a typical microlensing event, the light curve is well fit by assuming that the source is a point, the lens is a single point mass, and the lens is moving in a straight line: the point source-point lens approximation. In these events, the only physically significant parameter that can be measured is the Einstein timescale tE. However, in some cases, events can be analyzed to yield the additional parameters of the Einstein angle and parallax: θE and πE. These include very high magnification events, binary lenses, parallax, and xallarap events, and events where the lens is visible.

Events yielding the Einstein angle

Although the Einstein angle is too small to be directly visible from a ground-based telescope, several techniques have been proposed to observe it.

If the lens passes directly in front of the source star, then the finite size of the source star becomes an important parameter. The source star must be treated as a disk on the sky, not a point, breaking the point-source approximation, and causing a deviation from the traditional microlensing curve that lasts as long as the time for the lens to cross the source, known as a finite source light curve. The length of this deviation can be used to determine the time needed for the lens to cross the disk of the source star tS. If the angular size of the source θS is known, the Einstein angle can be determined as

\theta_E = \theta_S \frac{t_E}{t_S} .

These measurements are rare, since they require an extreme alignment between source and lens. They are more likely when θS / θE is (relatively) large, i.e., for nearby giant sources with slow-moving low-mass lenses close to the source.

In finite source events, different parts of the source star are magnified at different rates at different times during the event. These events can thus be used to study the limb-darkening of the source star.

Binary lenses

If the lens is a binary star with separation of roughly the Einstein radius, the magnification pattern is more complex than in the single star lenses. In this case, there are typically three images when the lens is distant from the source, but there is a range of alignments where two additional images are created. These alignments are known as caustics. At these alignments, the magnification of the source is formally infinite under the point-source approximation.

Caustic crossings in binary lenses can happen with a wider range of lens geometries than in a single lens. Like a single lens source caustic, it takes a finite time for the source to cross the caustic. If this caustic-crossing time tS can be measured, and if the angular radius of the source is known, then again the Einstein angle can be determined.

As in the single lens case when the source magnification is formally infinite, caustic crossing binary lenses will magnify different portions of the source star at different times. They can thus probe the structure of the source and its limb darkening.

An animation of a binary lens event can be found at this YouTube video.

Events yielding the Einstein parallax

In principle, the Einstein parallax can be measured by having two observers simultaneously observe the event from different locations, e.g., from the earth and from a distant spacecraft.[29] The difference in amplification observed by the two observers yields the component of \vec{\pi}_E perpendicular to the motion of the lens while the difference in the time of peak amplification yields the component parallel to the motion of the lens. This direct measurement was recently reported[30] using the Spitzer Space Telescope. In extreme cases, the differences may even be measurable from small differences seen from telescopes at different locations on the earth.[31]

More typically, the Einstein parallax is measured from the non-linear motion of the observer caused by the rotation of the earth about the sun. It was first reported in 1995 [32] and has been reported in a handful of events since. Parallax in point-lens events can best be measured in long-timescale events with a large πE-- from slow-moving, low mass lenses which are close to the observer.

If the source star is a binary star, then it too will have a non-linear motion which can also cause slight, but detectable changes in the light curve. This effect is known as Xallarap (parallax spelled backwards).

Detection of extrasolar planets

Gravitational microlensing of an extrasolar planet

If the lensing object is a star with a planet orbiting it, this is an extreme example of a binary lens event. If the source crosses a caustic, the deviations from a standard event can be large even for low mass planets. These deviations allow us to infer the existence and determine the mass and separation of the planet around the lens. Deviations typically last a few hours or a few days. Because the signal is strongest when the event itself is strongest, high-magnification events are the most promising candidates for detailed study. Typically, a survey team notifies the community when they discover a high-magnification event in progress. Followup groups then intensively monitor the ongoing event, hoping to get good coverage of the deviation if it occurs. When the event is over, the light curve is compared to theoretical models to find the physical parameters of the system. The parameters that can be determined directly from this comparison are the mass ratio of the planet to the star, and the ratio of the star-planet angular separation to the Einstein angle. From these ratios, along with assumptions about the lens star, the mass of the planet and its orbital distance can be estimated.

Exoplanets discovered using microlensing, by year, through 2010-01-13.

The first success of this technique was made in 2003 by both OGLE and MOA of the microlensing event OGLE 2003–BLG–235 (or MOA 2003–BLG–53). Combining their data, they found the most likely planet mass to be 1.5 times the mass of Jupiter.[33] As of January 2011, eleven exoplanets have been detected by this method, including OGLE-2005-BLG-071Lb,[34] OGLE-2005-BLG-390Lb,[35] OGLE-2005-BLG-169Lb,[36] two exoplanets around OGLE-2006-BLG-109L,[37] and MOA-2007-BLG-192Lb.[38] Notably, at the time of its announcement in January 2006, the planet OGLE-2005-BLG-390Lb probably had the lowest mass of any known exoplanet orbiting a regular star, with a median at 5.5 times the mass of the Earth and roughly a factor two uncertainty. This record was contested in 2007 by Gliese 581 c with a minimal mass of 5 Earth masses, and since 2009 Gliese 581 e is the lightest known "regular" exoplanet, with minimum 1.9 Earth masses.

Comparing this method of detecting extrasolar planets with other techniques such as the transit method, one advantage is that the intensity of the planetary deviation does not depend on the planet mass as strongly as effects in other techniques do. This makes microlensing well suited to finding low-mass planets. One disadvantage is that followup of the lens system is very difficult after the event has ended, because it takes a long time for the lens and the source to be sufficiently separated to resolve them separately.

Microlensing experiments

There are two basic types of microlensing experiments. "Search" groups use large-field images to find new microlensing events. "Follow-up" groups often coordinate telescopes around the world to provide intensive coverage of select events. The initial experiments all had somewhat risqué names until the formation of the PLANET group. There are current proposals to build new specialized microlensing satellites, or to use other satellites to study microlensing.

Search collaborations

  • Alard; Mao; Guibert (1995). "Object DUO 2: A New Binary Lens Candidate". arXiv:astro-ph/9506101 [astro-ph].  Photographic plate search of bulge. Remarkable for largely being the work of a single graduate student, Christophe Alard, for his Ph.D. Thesis.
  • Experience de Recherche des Objets Sombres (EROS) (1993–2002) Largely French collaboration. EROS1: Photographic plate search of LMC: EROS2: CCD search of LMC, SMC, Bulge & spiral arms.
  • MACHO (1993–1999) Australia & US collaboration. CCD search of bulge and LMC.
  • Optical Gravitational Lensing Experiment (OGLE) ( 1992 – ), Polish collaboration established by Paczynski and Udalski. Dedicated 1.3m telescope in Chile run by the University of Warsaw. Targets on bulge and Magellanic Clouds.
  • Microlensing Observations in Astrophysics (MOA) (1998 – ), Japanese-New Zealand collaboration. Dedicated 1.8m telescope in New Zealand. Targets on bulge and Magellanic Clouds.
  • SuperMACHO (2001 – ), successor to the MACHO collaboration used 4 m CTIO telescope to study faint LMC microlenses.

Follow-up collaborations

Andromeda galaxy pixel lensing collaborations

Proposed satellite experiments

References

  1. ^ a b Wambsganss (2006). "Gravitational Microlensing". arXiv:astro-ph/0604278 [astro-ph]. doi:10.1007/978-3-540-30310-7_4. 
  2. ^ Kochanek, C. S. (2004). "Quantitative Interpretation of Quasar Microlensing Light Curves". The Astrophysical Journal 605: 58. arXiv:astro-ph/03074223. Bibcode 2004ApJ...605...58K. doi:10.1086/382180. 
  3. ^ Poindexter et al., Shawn; Morgan, Nicholas; Kochanek, Christopher S. (2008). "The Spatial Structure of An Accretion Disk". The Astrophysical Journal, 673: 34. arXiv:0707.0003. Bibcode 2008ApJ...673...34P. doi:10.1086/524190. 
  4. ^ Eigenbrod et al., A.; Courbin, F.; Meylan, G.; Agol, E.; Anguita, T.; Schmidt, R. W.; Wambsganss, J. (2008). "Microlensing variability in the gravitationally lensed quasar QSO 2237+0305 = the Einstein Cross. II. Energy profile of the accretion disk". Astronomy & Astrophysics 490 (3): 933. arXiv:0810.0011. Bibcode 2008A&A...490..933E. doi:10.1051/0004-6361:200810729. 
  5. ^ Mosquera et al., A. M.; Muñoz, J. A.; Mediavilla, E. (2009). "Detection of chromatic microlensing in Q 2237+0305 A". The Astrophysical Journal, 691 (2): 1292. arXiv:0810.1626. Bibcode 2009ApJ...691.1292M. doi:10.1088/0004-637X/691/2/1292. 
  6. ^ Floyd et al., David J. E.; Bate, N. F.; Webster, R. L. (2009). "The accretion disc in the quasar SDSS J0924+0219". Monthly Notices of the Royal Astronomical Society 398: 233–239. arXiv:0905.2651. Bibcode 2009MNRAS.398..233F. doi:10.1111/j.1365-2966.2009.15045.x. 
  7. ^ Refsdal, S. (1964). "The gravitational lens effect". Monthly Notices of the Royal Astronomical Society 128: 295. Bibcode 1964MNRAS.128..295R. 
  8. ^ Paczynski, B. (1986). "Gravitational microlensing by the galactic halo". Astrophysical Journal 304: 1. Bibcode 1986ApJ...304....1P. doi:10.1086/164140. 
  9. ^ Boden, A. F.; Shao, M.; van Buren, D. (1998). "Astrometric Observation of MACHO Gravitational Microlensing". Astrophysical Journal 502 (2): 538. arXiv:astro-ph/9802179. Bibcode 1998ApJ...502..538B. doi:10.1086/305913. 
  10. ^ Delplancke, F.; Górski, K. M.; Richichi, A. (2001). "Resolving gravitational microlensing events with long-baseline optical interferometry". Astronomy and Astrophysics 375 (2): 701–710. arXiv:astro-ph/0108178. Bibcode 2001A&A...375..701D. doi:10.1051/0004-6361:20010783. 
  11. ^ The MACHO collaboration; Alcock; Allsman; Alves; Axelrod; Becker; Bennett; Cook et al. (2000). "The MACHO Project: Microlensing Results from 5.7 Years of LMC Observations". Astrophys.J. 542 (2000) 281–307 542: 281–307. arXiv:astro-ph/0001272. Bibcode 2000ApJ...542..281A. doi:10.1086/309512. 
  12. ^ Alcock; Allsman; Alves; Axelrod; Becker; Bennett; Cook; Drake et al. (2000). "The MACHO project: Microlensing Optical Depth towards the Galactic Bulge from Difference Image Analysis". Astrophysical Journal 541 (2): 734–766. arXiv:astro-ph/0002510. Bibcode 2000ApJ...541..734A. doi:10.1086/309484. 
  13. ^ Schneider, Ehlers, and Falco. Gravitational Lenses. 1992.
  14. ^ Einstein, A. (1936). "Lens-Like Action of a Star by the Deviation of Light in the Gravitational Field". Science 84 (2188): 506. Bibcode 1936Sci....84..506E. doi:10.1126/science.84.2188.506. PMID 17769014. 
  15. ^ Chang, K.; Refsdal, S. (1979). "Flux variations of QSO 0957 + 561 A, B and image splitting by stars near the light path". Nature 282 (5739): 561. Bibcode 1979Natur.282..561C. doi:10.1038/282561a0. 
  16. ^ mcmaster.ca
  17. ^ eros.in2p3.fr
  18. ^ OGLE homepage at ogle.astrouw.edu.pl
  19. ^ Alcock, C.; Akerlof, C. W.; Allsman, R. A.; Axelrod, T. S.; Bennett, D. P.; Chan, S.; Cook, K. H.; Freeman, K. C. et al. (1993). "Possible gravitational microlensing of a star in the Large Magellanic Cloud". Nature 365 (6447): 621–623. arXiv:astro-ph/9309052. Bibcode 1993Natur.365..621A. doi:10.1038/365621a0. 
  20. ^ Aubourg, E.; Bareyre, P.; Bréhin, S.; Gros, M.; Lachièze-Rey, M.; Laurent, B.; Lesquoy, E.; Magneville, C. et al. (1993). "Evidence for gravitational microlensing by dark objects in the Galactic halo". Nature 365 (6447): 623–625. Bibcode 1993Natur.365..623A. doi:10.1038/365623a0. 
  21. ^ Alcock, C.; Allsman, R. A.; Alves, D. R.; Axelrod, T. S.; Becker, A. C.; Bennett, D. P.; Cook, K. H.; Dalal, N. et al. (2000). "The MACHO Project: Microlensing Results from 5.7 Years of Large Magellanic Cloud Observations". The Astrophysical Journal 542: 281. arXiv:astro-ph/0001272. Bibcode 2000ApJ...542..281A. doi:10.1086/309512. 
  22. ^ Tisserand, P.; Le Guillou, L.; Afonso, C.; Albert, J. N.; Andersen, J.; Ansari, R.; Aubourg, É.; Bareyre, P. et al. (2007). "Limits on the Macho content of the Galactic Halo from the EROS-2 Survey of the Magellanic Clouds". Astronomy and Astrophysics 469 (2): 387–404. arXiv:astro-ph/0607207. Bibcode 2007A&A...469..387T. doi:10.1051/0004-6361:20066017. 
  23. ^ An NOAO Long Term Survey with the MOSAIC Imager on the Blanco 4 meter telescope. Ctio.noao.edu (2005-01-03). Retrieved on 2011-05-22.
  24. ^ Microlensing Observations in Astrophysics
  25. ^ μFUN-PLANET collaboration
  26. ^ RoboNet-II
  27. ^ Microlensing Follow-up Network
  28. ^ Gould, Andrew (2000). "A Natural Formalism for Microlensing". The Astrophysical Journal 542 (2): 785. arXiv:astro-ph/0001421. Bibcode 2000ApJ...542..785G. doi:10.1086/317037. 
  29. ^ Gould, Andrew (1994). "MACHO velocities from satellite-based parallaxes". The Astrophysical Journal 421: L75. Bibcode 1994ApJ...421L..75G. doi:10.1086/187191. 
  30. ^ Dong, Subo; Udalski, A.; Gould, A.; Reach, W. T.; Christie, G. W.; Boden, A. F.; Bennett, D. P.; Fazio, G. et al. (2007). "First Space‐Based Microlens Parallax Measurement:SpitzerObservations of OGLE‐2005‐SMC‐001". The Astrophysical Journal 664 (2): 862. arXiv:astro-ph/0702240. Bibcode 2007ApJ...664..862D. doi:10.1086/518536. 
  31. ^ Hardy, S. J.; Walker, M. A. (1995). "Parallax effects in binary microlensing events". Monthly Notices of the Royal Astronomical Society 276: L79. Bibcode 1995MNRAS.276L..79H. 
  32. ^ Alcock,, C.; Allsman,, R. A.; Alves,, D.; Axelrod,, T. S.; Bennett,, D. P.; Cook,, K. H.; Freeman,, K. C.; Griest,, K. et al. (1995). "First Observation of Parallax in a Gravitational Microlensing Event". The Astrophysical Journal 454 (2). arXiv:astro-ph/9506114. Bibcode 1995ApJ...454L.125A. doi:10.1086/309783. 
  33. ^ Bond; Udalski; Jaroszynski; Rattenbury; Paczynski; Soszynski; Wyrzykowski; Szymanski et al. (2004). "OGLE 2003-BLG-235/MOA 2003-BLG-53: A planetary microlensing event". Astrophys.J.606:L155-L158,2004 606 (2): L155–L158. arXiv:astro-ph/0404309. Bibcode 2004ApJ...606L.155B. doi:10.1086/420928. 
  34. ^ Udalski; Jaroszynski; Paczynski; Kubiak; Szymanski; Soszynski; Pietrzynski; Ulaczyk et al. (2005). "A Jovian-mass Planet in Microlensing Event OGLE-2005-BLG-071". The Astrophysical Journal 628 (2): L109–L112. arXiv:astro-ph/0505451. Bibcode 2005ApJ...628L.109U. doi:10.1086/432795. 
  35. ^ OGLE website
  36. ^ Gould; Udalski; An; Bennett; Zhou; Dong; Rattenbury; Gaudi et al. (2006). "Microlens OGLE-2005-BLG-169 Implies Cool Neptune-Like Planets are Common". Astrophys.J. 644 (2006) L37-L40 644: L37–L40. arXiv:astro-ph/0603276. Bibcode 2006ApJ...644L..37G. doi:10.1086/505421. 
  37. ^ Gaudi; Bennett; Udalski; Gould; Christie; Maoz; Dong; McCormick et al. (2008). "Discovery of a Jupiter/Saturn Analog with Gravitational Microlensing". Science 319 (5865): 927–930. arXiv:0802.1920. Bibcode 2008Sci...319..927G. doi:10.1126/science.1151947. PMID 18276883. 
  38. ^ Paul Rincon, Tiniest extrasolar planet found, BBC, 2 June 2008

External links


Wikimedia Foundation. 2010.

Игры ⚽ Поможем сделать НИР

Look at other dictionaries:

  • Microlensing Observations in Astrophysics — The Microlensing Observations in Astrophysics (MOA) telescope dome at the top of Mount John Microlensing Observations in Astrophysics (MOA) is a collaborative project between researchers in New Zealand and Japan, led by Professor Yasushi Muraki… …   Wikipedia

  • Gravitational lens — Gravitational Lensing Formalism Strong lensing …   Wikipedia

  • Microlensing Observations in Astrophysics — Pour les articles homonymes, voir MOA. Le dome du télescope MOA au sommet du Mount John Microlensing Observations in Astrophysics (MOA) est une collaboration entre des chercheurs …   Wikipédia en Français

  • microlensing — noun gravitational lensing …   Wiktionary

  • Optical Gravitational Lensing Experiment — OGLE telescope at the Las Campanas Observatory …   Wikipedia

  • Optical Gravitational Lensing Experiment — (OGLE, буквально Оптический Эксперимент по Гравитационному Линзированию)  польско американский астрономический проект по изучению темной материи при помощи метода гравитационного микролинзирования. Проект был основан в 1992, под руководством …   Википедия

  • Optical Gravitational Lensing Experiment — 29° 00′ 36″ S 70° 41′ 56″ W / 29.010, 70.699 …   Wikipédia en Français

  • Methods of detecting extrasolar planets — Any planet is an extremely faint light source compared to its parent star. In addition to the intrinsic difficulty of detecting such a faint light source, the light from the parent star causes a glare that washes it out. For those reasons, only a …   Wikipedia

  • OGLE-2005-BLG-390Lb — Extrasolar planet List of extrasolar planets An artist s conception of the planet OGLE 2005 BLG 390Lb Parent star Star …   Wikipedia

  • Exoplanetology — is the Art and Science of Exoplanets and the search for Life on other Worlds. It is formally defined as the study of the characteristics, formation and fate of planets outside our solar system. The discovery of extrasolar planets or exoplanets in …   Wikipedia

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”