Gravitational-wave observatory

A schematic diagram of a laser interferometer.

A gravitational-wave observatory (or gravitational-wave detector) is any device designed to measure gravitational waves, tiny distortions of spacetime that were first predicted by Einstein in 1916.[1] Gravitational waves are perturbations in the theoretical curvature of spacetime caused by accelerated masses. The existence of gravitational radiation is a specific prediction of general relativity, but is a feature of all theories of gravity that obey special relativity.[2] Since the 1960s, gravitational-wave detectors have been built and constantly improved. The present-day generation of resonant mass antennas and laser interferometers has reached the necessary sensitivity to detect gravitational waves from sources in the Milky Way. Gravitational-wave observatories are the primary tool of gravitational-wave astronomy.

A number of experiments have provided indirect evidence, notably the observation of binary pulsars, the orbits of which evolve precisely matching the predictions of energy loss through general relativistic gravitational-wave emission. The 1993 Nobel Prize in Physics was awarded for this work.[3]

In February 2016, the Advanced LIGO team announced that they had detected gravitational waves from a black hole merger.[4][5][6] The 2017 Nobel Prize in Physics was awarded for this work.

Complications

The direct detection of gravitational waves is complicated by the extraordinarily small effect the waves produce on a detector. The amplitude of a spherical wave falls off as the inverse of the distance from the source. Thus, even waves from extreme systems such as merging binary black holes die out to a very small amplitude by the time they reach the Earth. Astrophysicists predicted that some gravitational waves passing the Earth might produce differential motion on the order 10−18 m in a LIGO-size instrument.[7]

Resonant mass antennas

A simple device to detect the expected wave motion is called a resonant mass antenna – a large, solid body of metal isolated from outside vibrations. This type of instrument was the first type of gravitational-wave detector. Strains in space due to an incident gravitational wave excite the body's resonant frequency and could thus be amplified to detectable levels. Conceivably, a nearby supernova might be strong enough to be seen without resonant amplification. However, up to 2018, no gravitational wave observation that would have been widely accepted by the research community has been made on any type of resonant mass antenna, despite certain claims of observation by researchers operating the antennas.

There are 3 types of resonant mass antenna that have been built: the room-temperature bar antennas, the cryogenically cooled bar antennas and cryogenically cooled spherical antennas.

The earlies types of antennas were the room-temperature bar-shaped antennas called Weber bar; these were dominant in 1960s and 1970s and many were built around the world. It was claimed by Weber and some others in the late 1960s and early 1970s that these devices did observe gravitational waves; however, other experimenters failed to detect gravitational waves with these devices and thus it became consensus that these devices could not detect gravitational waves. [8]

The second generation of resonant mass antennas, developed in the 1980s and 1990s, were the cryogenic bar antennas which are also sometimes called Weber bars. There were in the 1990s 5 major cryogenic bar antennas: AURIGA (Padua, Italy), NAUTILUS (Rome, Italy), EXPLORER (CERN, Switzerland), ALLEGRO (Louisiana, USA), NIOBE (Perth, Australia). In 1997, these 5 antennas run by 4 research groups formed the International Gravitational Event Collaboration (IGEC) for collaboration. Over the years, many claims of detection of gravitational waves have been made by scientist using cryogenic bar antennas but none of these was accepted by the larger community.

In 1980s there was also a cryogenic bar antenna called ALTAIR, which along with a room-temperature bar antenna called GEOGRAV was built in Italy as a prototype for later bar antennas. GEOGRAV-detector was claimed by its operators to have seen gravitational waves coming from the supernova SN1987A (along with another room-temperature bar of Weber), but these claims were also dismissed by the wider community.

These modern cryogenic forms of the Weber bar operated with superconducting quantum interference devices to detect vibration (see for example, ALLEGRO). Some of them are still in operation, for example AURIGA, an ultracryogenic resonant cylindrical bar gravitational wave detector based at INFN in Italy. The AURIGA and LIGO teams have collaborated in joint observations.[9]

It is the current consensus that current cryogenic Weber bars are not sensitive enough to detect anything but extremely powerful gravitational waves. As of 2018, no observation of gravitational waves by cryogenic Weber bars has occurred.

In the 2000s, the third generation of resonant mass antennas, the spherical cryogenic antennas, emerged. 4 spherical antennas were proposed around year 2000 and 2 of them ended up being built (others were cancelled) as downsized versions. The proposed antennas were GRAIL (Netherlands, proposal that when downsized became MiniGRAIL), TIGA (USA, small prototypes made), SFERA (Italy), Graviton (Brasil, proposal that when downsized became Mario Schenberg).

Currently there are 2 cryogenic spherical gravitational wave antennas in the world, the MiniGRAIL and the Mario Schenberg. These antennas are actually a collaborative effort, having much in common.

MiniGRAIL is based at Leiden University, consisting of an exactingly machined 1150 kg sphere cryogenically cooled to 20 mK.[10] The spherical configuration allows for equal sensitivity in all directions, and is somewhat experimentally simpler than larger linear devices requiring high vacuum. Events are detected by measuring deformation of the detector sphere. MiniGRAIL is highly sensitive in the 2–4 kHz range, suitable for detecting gravitational waves from rotating neutron star instabilities or small black hole mergers.[11]

The Mario Schenberg antenna is located in Sao Paulo, Brazil.

Interferometers

Simplified operation of a gravitational wave observatory
Figure 1: A beamsplitter (green line) splits coherent light (from the white box) into two beams which reflect off the mirrors (cyan oblongs); only one outgoing and reflected beam in each arm is shown, and separated for clarity. The reflected beams recombine and an interference pattern is detected (purple circle).
Figure 2: A gravitational wave passing over the left arm (yellow) changes its length and thus the interference pattern.

A more sensitive detector uses laser interferometry to measure gravitational-wave induced motion between separated 'free' masses.[12] This allows the masses to be separated by large distances (increasing the signal size); a further advantage is that it is sensitive to a wide range of frequencies (not just those near a resonance as is the case for Weber bars). Ground-based interferometers are now operational. Currently, the most sensitive is LIGO – the Laser Interferometer Gravitational Wave Observatory. LIGO has three detectors: one in Livingston, Louisiana; the other two (in the same vacuum tubes) at the Hanford site in Richland, Washington. Each consists of two light storage arms which are 2 to 4 kilometers in length. These are at 90 degree angles to each other, with the light passing through 1m diameter vacuum tubes running the entire 4 kilometers. A passing gravitational wave will slightly stretch one arm as it shortens the other. This is precisely the motion to which an interferometer is most sensitive.

Even with such long arms, the strongest gravitational waves will only change the distance between the ends of the arms by at most roughly 10−18 meters. LIGO should be able to detect gravitational waves as small as . Upgrades to LIGO and other detectors such as VIRGO, GEO 600, and TAMA 300 should increase the sensitivity still further; the next generation of instruments (Advanced LIGO and Advanced Virgo) will be more than ten times more sensitive. Another highly sensitive interferometer (KAGRA) is currently in the design phase. A key point is that a ten-times increase in sensitivity (radius of "reach") increases the volume of space accessible to the instrument by one thousand. This increases the rate at which detectable signals should be seen from one per tens of years of observation, to tens per year.

Interferometric detectors are limited at high frequencies by shot noise, which occurs because the lasers produce photons randomly; one analogy is to rainfall – the rate of rainfall, like the laser intensity, is measurable, but the raindrops, like photons, fall at random times, causing fluctuations around the average value. This leads to noise at the output of the detector, much like radio static. In addition, for sufficiently high laser power, the random momentum transferred to the test masses by the laser photons shakes the mirrors, masking signals at low frequencies. Thermal noise (e.g., Brownian motion) is another limit to sensitivity. In addition to these "stationary" (constant) noise sources, all ground-based detectors are also limited at low frequencies by seismic noise and other forms of environmental vibration, and other "non-stationary" noise sources; creaks in mechanical structures, lightning or other large electrical disturbances, etc. may also create noise masking an event or may even imitate an event. All these must be taken into account and excluded by analysis before a detection may be considered a true gravitational-wave event.

Space-based interferometers, such as LISA and DECIGO, are also being developed. LISA's design calls for three test masses forming an equilateral triangle, with lasers from each spacecraft to each other spacecraft forming two independent interferometers. LISA is planned to occupy a solar orbit trailing the Earth, with each arm of the triangle being five million kilometers. This puts the detector in an excellent vacuum far from Earth-based sources of noise, though it will still be susceptible to shot noise, as well as artifacts caused by cosmic rays and solar wind.

An atomic gravitational-wave interferometric sensor (AGIS) is an alternative means to detect gravitational waves, proposed in 2008.[13][14]

Einstein@Home

In some sense, the easiest signals to detect should be constant sources. Supernovae and neutron star or black hole mergers should have larger amplitudes and be more interesting, but the waves generated will be more complicated. The waves given off by a spinning, bumpy neutron star would be "monochromatic" – like a pure tone in acoustics. It would not change very much in amplitude or frequency.

The Einstein@Home project is a distributed computing project similar to SETI@home intended to detect this type of simple gravitational wave. By taking data from LIGO and GEO, and sending it out in little pieces to thousands of volunteers for parallel analysis on their home computers, Einstein@Home can sift through the data far more quickly than would be possible otherwise.[15]

High frequency detectors

There are currently two detectors focusing on detections at the higher end of the gravitational-wave spectrum (10−7 to 105 Hz): one at University of Birmingham, England, and the other at INFN Genoa, Italy. A third is under development at Chongqing University, China. The Birmingham detector measures changes in the polarization state of a microwave beam circulating in a closed loop about one meter across. Two have been fabricated and they are currently expected to be sensitive to periodic spacetime strains of , given as an amplitude spectral density. The INFN Genoa detector is a resonant antenna consisting of two coupled spherical superconducting harmonic oscillators a few centimeters in diameter. The oscillators are designed to have (when uncoupled) almost equal resonant frequencies. The system is currently expected to have a sensitivity to periodic spacetime strains of , with an expectation to reach a sensitivity of . The Chongqing University detector is planned to detect relic high-frequency gravitational waves with the predicted typical parameters ~ 1010 Hz (10 GHz) and h ~ 10−30 to 10−31.

Pulsar timing arrays

A different approach to detecting gravitational waves is used by pulsar timing arrays, such as the European Pulsar Timing Array,[16] the North American Nanohertz Observatory for Gravitational Waves,[17] and the Parkes Pulsar Timing Array.[18] These projects propose to detect gravitational waves by looking at the effect these waves have on the incoming signals from an array of 20–50 well-known millisecond pulsars. As a gravitational wave passing through the Earth contracts space in one direction and expands space in another, the times of arrival of pulsar signals from those directions are shifted correspondingly. By studying a fixed set of pulsars across the sky, these arrays should be able to detect gravitational waves in the nanohertz range. Such signals are expected to be emitted by pairs of merging supermassive black holes.[19]

Cosmic microwave background polarization

The cosmic microwave background, radiation left over from when the Universe cooled sufficiently for the first atoms to form, can contain the imprint of gravitational waves from the very early Universe. The microwave radiation is polarized. The pattern of polarization can be split into two classes called E-modes and B-modes. This is in analogy to electrostatics where the electric field (E-field) has a vanishing curl and the magnetic field (B-field) has a vanishing divergence. The E-modes can be created by a variety of processes, but the B-modes can only be produced by gravitational lensing, gravitational waves, or scattering from dust.

On 17 March 2014, astronomers at the Harvard-Smithsonian Center for Astrophysics announced the apparent detection of the imprint gravitational waves in the cosmic microwave background, which, if confirmed, would provide strong evidence for inflation and the Big Bang.[20][21][22][23] However, on 19 June 2014, lowered confidence in confirming the findings was reported;[24][25][26] and on 19 September 2014, even more lowered confidence.[27][28] Finally, on January 30, 2015, the European Space Agency announced that the signal can be entirely attributed to dust in the Milky Way.[29]

Operational and planned gravitational-wave detectors

Noise curves for a selection of detectors as a function of frequency. The characteristic strain of potential astrophysical sources are also shown. To be detectable the characteristic strain of a signal must be above the noise curve.[30]

See also

References

  1. Clark, Stuart (17 March 2014). "What are gravitational waves?". The Guardian. Retrieved 22 May 2014.
  2. Schutz, Bernard F. (1984). "Gravitational waves on the back of an envelope". American Journal of Physics. 52 (5): 412. Bibcode:1984AmJPh..52..412S. doi:10.1119/1.13627.
  3. "Press Release: The Nobel Prize in Physics 1993". Nobel Prize. 13 October 1993. Retrieved 6 May 2014.
  4. Castelvecchi, Davide; Witze, Witze (February 11, 2016). "Einstein's gravitational waves found at last". Nature News. doi:10.1038/nature.2016.19361. Retrieved 2016-02-11.
  5. B. P. Abbott (LIGO Scientific Collaboration and Virgo Collaboration) et al. (2016). "Observation of Gravitational Waves from a Binary Black Hole Merger". Physical Review Letters. 116 (6): 061102. arXiv:1602.03837. Bibcode:2016PhRvL.116f1102A. doi:10.1103/PhysRevLett.116.061102. PMID 26918975.
  6. "Gravitational waves detected 100 years after Einstein's prediction | NSF - National Science Foundation". www.nsf.gov. Retrieved 2016-02-11.
  7. Whitcomb, S.E., "Precision Laser Interferometry in the LIGO Project", Proceedings of the International Symposium on Modern Problems in Laser Physics, August 27-September 3, 1995, Novosibirsk, LIGO Publication P950007-01-R
  8. For a review of early experiments using Weber bars, see Levine, J. (April 2004). "Early Gravity-Wave Detection Experiments, 1960-1975". Physics in Perspective. 6 (1): 42–75. Bibcode:2004PhP.....6...42L. doi:10.1007/s00016-003-0179-6.
  9. AURIGA Collaboration; LIGO Scientific Collaboration; Baggio; Cerdonio, M; De Rosa, M; Falferi, P; Fattori, S; Fortini, P; et al. (2008). "A Joint Search for Gravitational Wave Bursts with AURIGA and LIGO". Classical and Quantum Gravity. 25 (9): 095004. arXiv:0710.0497. Bibcode:2008CQGra..25i5004B. doi:10.1088/0264-9381/25/9/095004.
  10. Gravitational Radiation Antenna In Leiden
  11. de Waard, Arlette; Gottardi, Luciano; Frossati, Giorgio (2000). "Spherical Gravitational Wave Detectors: cooling and quality factor of a small CuAl6% sphere - In: Marcel Grossmann meeting on General Relativity". Rome, Italy.
  12. The idea of using laser interferometry for gravitational-wave detection was first mentioned by Gerstenstein and Pustovoit 1963 Sov. Phys.–JETP 16 433. Weber mentioned it in an unpublished laboratory notebook. Rainer Weiss first described in detail a practical solution with an analysis of realistic limitations to the technique in R. Weiss (1972). "Electromagnetically Coupled Broadband Gravitational Antenna". Quarterly Progress Report, Research Laboratory of Electronics, MIT 105: 54.
  13. Bender, Peter L. "Comment on "Atomic gravitational wave interferometric sensor"". Physical Review D. 84 (2). Bibcode:2011PhRvD..84b8101B. doi:10.1103/PhysRevD.84.028101.
  14. Johnson, David Marvin Slaughter (2011). "AGIS-LEO". Long Baseline Atom Interferometry. Stanford University. pp. 41&ndash, 98.
  15. Einstein@Home
  16. Janssen, G. H.; Stappers, B. W.; Kramer, M.; Purver, M.; Jessner, A.; Cognard, I.; Bassa, C.; Wang, Z.; Cumming, A.; Kaspi, V. M. (2008). "European Pulsar Timing Array". AIP Conference Proceedings. 983: 633–635. doi:10.1063/1.2900317.
  17. North American Nanohertz Observatory for Gravitational Waves (NANOGrav) homepage
  18. Parkes Pulsar Timing Array homepage
  19. Hobbs, G. B.; Bailes, M.; Bhat, N. D. R.; Burke-Spolaor, S.; Champion, D. J.; Coles, W.; Hotan, A.; Jenet, F.; et al. (2008). "Gravitational wave detection using pulsars: status of the Parkes Pulsar Timing Array project". Publications of the Astronomical Society of Australia. 26: 103–109. arXiv:0812.2721 [astro-ph]. Bibcode:2009PASA...26..103H. doi:10.1071/AS08023. Cite uses deprecated parameter |class= (help)
  20. Staff (17 March 2014). "BICEP2 2014 Results Release". National Science Foundation. Retrieved 18 March 2014.
  21. Clavin, Whitney (17 March 2014). "NASA Technology Views Birth of the Universe". NASA. Retrieved 17 March 2014.
  22. Overbye, Dennis (17 March 2014). "Detection of Waves in Space Buttresses Landmark Theory of Big Bang". The New York Times. Retrieved 17 March 2014.
  23. Overbye, Dennis (24 March 2014). "Ripples From the Big Bang". The New York Times. Retrieved 24 March 2014.
  24. Overbye, Dennis (19 June 2014). "Astronomers Hedge on Big Bang Detection Claim". The New York Times. Retrieved 20 June 2014.
  25. Amos, Jonathan (19 June 2014). "Cosmic inflation: Confidence lowered for Big Bang signal". BBC News. Retrieved 20 June 2014.
  26. Ade, P.A.R. (BICEP2 Collaboration); et al. (19 June 2014). "Detection of B-Mode Polarization at Degree Angular Scales by BICEP2". Physical Review Letters. 112 (24): 241101. arXiv:1403.3985. Bibcode:2014PhRvL.112x1101B. doi:10.1103/PhysRevLett.112.241101. PMID 24996078.
  27. Planck Collaboration Team (19 September 2014). "Planck intermediate results. XXX. The angular power spectrum of polarized dust emission at intermediate and high Galactic latitudes". Astronomy & Astrophysics. 586: A133. arXiv:1409.5738. Bibcode:2016A&A...586A.133P. doi:10.1051/0004-6361/201425034.
  28. Overbye, Dennis (22 September 2014). "Study Confirms Criticism of Big Bang Finding". The New York Times. Retrieved 22 September 2014.
  29. Cowen, Ron (2015-01-30). "Gravitational waves discovery now officially dead". Nature. doi:10.1038/nature.2015.16830.
  30. Moore, Christopher; Cole, Robert; Berry, Christopher (19 July 2013). "Gravitational Wave Detectors and Sources". Archived from the original on 16 April 2014. Retrieved 17 April 2014.
  31. Bhattacharya, Papiya (2016-03-25). "India's LIGO Detector Has the Money it Needs, a Site in Sight, and a Completion Date Too". The Wire. Retrieved 2016-06-16.
This article is issued from Wikipedia. The text is licensed under Creative Commons - Attribution - Sharealike. Additional terms may apply for the media files.