Virgo interferometer

Last updated

The Virgo experiment
Formation1993
TypeInternational scientific collaboration
Purpose Gravitational wave detection
Headquarters European Gravitational Observatory
Location
Coordinates 43°37′53″N10°30′16″E / 43.6313°N 10.5045°E / 43.6313; 10.5045
Region
Italy
FieldsBasic research
Spokesperson
Gianluca Gemme
AffiliationsLVK (LIGO-Virgo-KAGRA collaboration)
Budget
About ten million euros per year
Staff
Around 850 people participate in the Virgo Collaboration
Website www.virgo-gw.eu

The Virgo interferometer is a large Michelson interferometer designed to detect the gravitational waves predicted by general relativity. It is located in Santo Stefano a Macerata, near the city of Pisa, Italy. The instrument's two arms are three kilometres long, housing its mirrors and instrumentation inside an ultra-high vacuum.

Contents

Virgo is hosted by the European Gravitational Observatory (EGO), a consortium founded by the French CNRS and Italian INFN. [1] The Virgo Collaboration operates the detector and defines the strategy and policy for its use and upgrades. It is composed of several hundreds of members across 16 different countries. [2] Other interferometers similar to Virgo have the same goal of detecting gravitational waves, including the two LIGO interferometers in the United States (at the Hanford Site and in Livingston, Louisiana) and the Japanese interferometer KAGRA. Since 2007, Virgo and LIGO have agreed to share and jointly analyze the data recorded by their detectors and to jointly publish their results; this agreement was joined by KAGRA in 2019. [3] Because the interferometric detectors are not directional (they survey the whole sky) and are looking for signals which are weak and infrequent, simultaneous detection of a gravitational wave by multiple instruments is crucial for improving confidence in the signal validity and deducing the location of its source.

The interferometer is named after the Virgo Cluster, a cluster of about 1,500 galaxies in the Virgo constellation, about 50 million light-years from Earth. Founded at a time when gravitational waves were only a prediction by general relativity, it has now participated in detecting multiple gravitational wave events; the detector is still being periodically improved to increase its sensitivity and scientific output.

Organization

The Virgo experiment is managed by the European Gravitational Observatory (EGO) consortium, created in December 2000 by the CNRS and INFN. [4] The Dutch Institute for Nuclear and High-Energy Physics, Nikhef, later joined as an observer and eventually became a full member. EGO is responsible for the Virgo site, in charge of the construction, maintenance, and operation of the detector, as well as its upgrades. One of the goals of EGO is also to promote research on and studies of gravitation in Europe. [1]

In addition, the Virgo Collaboration consolidates all the researchers working on various aspects of the detector. As of May 2023, around 850 members, representing 142 institutions in 16 different countries, are part of the collaboration. [2] [5] This includes institutions from France, Italy, the Netherlands, Poland, Spain, Belgium, Germany, Hungary, Portugal, Greece, Czechia, Denmark, Ireland, Monaco, China, and Japan. [6]

The Virgo Collaboration is also part of the larger LIGO-Virgo-KAGRA (LVK) Collaboration, which gathers scientists from the other major gravitational waves experiment, for the purpose of carrying out joint analysis of the data which is crucial for gravitational wave detections. [7] LVK first started in 2007 [3] as the LIGO-Virgo Collaboration, and was expanded when KAGRA joined in 2019. [8] [9]

History

The Virgo project was approved in 1992 by the French CNRS and in 1993 by the Italian INFN, the two institutes at the origin of the experiment. The construction of the detector started in 1996 at the Cascina site near Pisa, Italy, and was completed in 2003. After several observation runs without detection, the interferometer was shut down in 2011 to allow for significant upgrades as part of the Advanced Virgo project. It started making observations again in 2017, quickly making its first detections along with the LIGO detectors.

Conception

Although the concept of gravitational waves is more than 100 years old, having been predicted by Einstein in 1916, [10] it was not before the 1970s that serious projects for detecting them started to appear. The first were the so-called Weber bars, invented by Joseph Weber, [11] which could in principle detect gravitational waves. This triggered a number of projects, and while none of them succeeded, they did spark the creation of many research groups dedicated to the detection of gravitational waves. [12]

The idea of a large interferometric detector began to gain credibility in the early 1980s, and in 1985, the Virgo project was conceptualized by the Italian researcher Adalberto Giazotto and the French researcher Alain Brillet after they met in Rome. One of the key ideas that set Virgo apart from other projects was targeting low frequencies (around 10 Hz), whereas most projects focused on higher frequencies (around 500 Hz); many believed at the time that this was not doable, and only France and Italy started working on the project, [13] which was first presented in 1987. [14] After being approved by the CNRS and the INFN, the construction of the interferometer began in 1996, with the aim of beginning observations by the year 2000. [15]

The first goal of Virgo was to directly observe gravitational waves. The study of the binary pulsar 1913+16 over three decades, whose discoverers were awarded the 1993 Nobel Prize in Physics, had already led to indirect evidence of the existence of gravitational waves. The observed decrease over time of this binary pulsar's orbital period was in agreement with the hypothesis that the system was losing energy by emitting gravitational waves. [16]

Initial Virgo detector

In the 2000s, the Virgo detector was first built, commissioned, and operated. The instrument successfully reached its planned design sensitivity to gravitational wave signals. This initial endeavor was used to validate the Virgo technical design choices; it also demonstrated that giant interferometers were promising devices for detecting gravitational waves in a wide frequency band. [17] [18] This original detector is generally referred to as the "initial Virgo" or "original Virgo".

The construction of the initial Virgo detector was completed in June 2003, [19] and several data collection periods ("science runs") followed between 2007 and 2011. [20] [21] Some of these runs were done simultaneously with the two LIGO detectors. There was a shut-down of a few months in 2010 to allow for a major upgrade of the Virgo suspension system: the original steel suspension wires were replaced by glass fibers in order to reduce the thermal noise. [22]

However, the initial Virgo detector was not sensitive enough to detect gravitational waves. After several months of data collection with the upgraded suspension system, the initial Virgo detector was shut down in September 2011 to begin the installation of Advanced Virgo. [23]

Advanced Virgo detector

First direct detection of a gravitational wave by Virgo, GW170814 GW170814 signal.png
First direct detection of a gravitational wave by Virgo, GW170814

The Advanced Virgo detector aimed to increase the sensitivity (and thus the distance at which a signal can be detected) by a factor of 10, allowing it to probe a volume of the Universe 1,000 times larger, making detection of gravitational waves more likely. [13] [24] It benefited from the experience gained with the initial detector and subsequent technological advances.

The Advanced Virgo detector kept the same vacuum infrastructure as the initial Virgo, with four additional cryotraps located at both ends of both the three-kilometre-long arms to trap residual particles coming from the mirror towers, but the remainder of the interferometer was significantly upgraded. The new mirrors were larger (350 mm in diameter, with a weight of 40 kg), and their optical performance was improved. The critical optical elements used to control the interferometer are under vacuum on suspended benches. A system of adaptive optics was to be installed to correct the mirror aberrations in-situ. In the final Advanced Virgo configuration, the laser power is expected to be 200 W. [25]

Advanced Virgo started the commissioning process in 2016, joining the two advanced LIGO detectors ("aLIGO") on 1 August 2017, during the "O2" observation period. On 14 August 2017, LIGO and Virgo detected a signal, GW170814, which was reported on 27 September 2017. It was the first binary black hole merger detected by both LIGO and Virgo (and the first one for Virgo). [26] [27]

Just a few days later, GW170817 was detected by LIGO and Virgo on 17 August 2017. The signal was produced by the last minutes of two neutron stars spiralling closer to each other and finally merging, and represents both the first binary neutron star merger observed and the first gravitational wave observation which was confirmed by non-gravitational means. Indeed, the resulting gamma-ray burst was also detected, and optical telescopes later discovered a kilonova corresponding to the merger. [28] [29]

After further upgrades, Virgo started the third observation run ("O3") in April 2019, planned to last one year, followed by further upgrades. [30] On 27 March 2020, the O3 run was suspended because of the COVID-19 pandemic. [31]

The upgrades currently underway are part of the "Advanced Virgo +" program, divided in two phases, the first one preceding the O4 run and the second one preceding the O5 run. The first phase focuses on the reduction of quantum noise by introducing a more powerful laser, improving the squeezing introduced in O3, and implementing a new technique called signal recycling; seismic sensors will also be installed around the mirrors. The second phase will then try to reduce the mirror thermal noise, by changing the geometry of the laser beam to increase its size on the mirrors (spreading the energy on a larger area and thus reducing the temperature), and by improving the coating of the mirrors; the end mirrors will also be significantly larger, requiring improvements to the suspension. Further improvements for quantum noise reduction are also expected in the second phase, building upon the changes from the first phase. [32]

The fourth observation run ("O4") was scheduled to start in May 2023, and is planned to last for a total of 20 months, including a commissioning break of up to two months. [33] However, on 11 May 2023, Virgo announced that it would not join at the beginning of O4, as the interferometer was not stable enough to reach the expected sensitivity and needs to undergo the replacement of one of the mirrors, requiring several weeks of work. [34] Virgo has not joined the O4 run during the first part of the run ("O4a"), which ended on 16 January 2024, as it only managed to reach a peak sensitivity of 45 Mpc instead of the 80 to 115 Mpc initially expected; it joined the second part of the run ("O4b") which began on 10 April 2024, [33] with a sensitivity of 50 to 55 Mpc. [35]

Future

Following the O4 run, the detector will once again be shut down to undergo upgrades, including an improvement in the coating of the mirrors. A fifth observing run (O5) is currently planned for the beginning of 2027; the target sensitivity for Virgo, which was originally set to be 150–260 Mpc, is currently being redefined in light of the performance during O4; plans to enter the O5 run are expected to be known before the end of 2024. [33]

No official plans have been announced for the future of the Virgo installations following the O5 period, although projects for further improving the detectors have been suggested; the current plans of the collaboration are referred to as the Virgo_nEXT project. [36]

Science case

Numerical simulations of the gravitational waves emitted by the inspiral and merger of two black holes. MergingBlackHoles V2.jpg
Numerical simulations of the gravitational waves emitted by the inspiral and merger of two black holes.

Virgo is designed to look for gravitational waves emitted by astrophysical sources across the universe, which can be broadly classified into three types: [37]

Typical "chirp" of a gravitational wave signal, from the GW170817 event. The x-axis represents the time, and y-axis is the frequency; the way the frequency rises with the time is typical of gravitational waves from compact object binaries, and its exact shape is mainly determined by the masses of the objects. GW170817 Gravitational Wave Chirp Spectrogram.jpg
Typical "chirp" of a gravitational wave signal, from the GW170817 event. The x-axis represents the time, and y-axis is the frequency; the way the frequency rises with the time is typical of gravitational waves from compact object binaries, and its exact shape is mainly determined by the masses of the objects.

The detection of these sources gives a new way to observe this type of objects (often carrying different informations than more classical ways, e.g. using telescopes), but also to probe fundamental properties of gravity, such as the polarization of gravitational waves, [38] possible gravitational lensing, [39] or more generally whether the observed signals are correctly described by general relativity. [40] It also provides a way to measure the Hubble constant. [41]

Instrument

Principle

Aerial view of the site of the Virgo experiment showing the central building, the Mode-Cleaner building, the full 3 km-long west arm and the beginning of the north arm (on the right). The other buildings include offices, workshops, the local computing center and the interferometer control room. When this picture was shot, the building hosting the project management and the canteen had not been built yet. Virgo aerial view 01.jpg
Aerial view of the site of the Virgo experiment showing the central building, the Mode-Cleaner building, the full 3 km-long west arm and the beginning of the north arm (on the right). The other buildings include offices, workshops, the local computing center and the interferometer control room. When this picture was shot, the building hosting the project management and the canteen had not been built yet.

In general relativity, a gravitational wave is a space-time perturbation which propagates at the speed of light. It thus slightly curves space-time, which locally changes the light path. Concretely, it can be detected using a Michelson interferometer design, where a laser is divided in two beams travelling in orthogonal directions, bouncing on a mirror located at the end of each arm. As the gravitational wave passes, it alters the path of the two beams in a different manner; the two beams are recombined, and the resulting interferometric pattern is measured using a photodiode. As the induced deformation is extremely small, the design requires an extremely high precision in the position of the mirrors, the stability of the laser, the measurements, and a very good isolation from the outside world to reduce the amount of noise. [42]

Animation displaying the principle of gravitational wave detection with an interferometer such as Virgo. Mirror displacements and phase difference are widely exaggerated ; time is also slowed down by more than a factor 10.

Laser and injection system

Layout of the Virgo interferometer during the O4 run (2023-2024). It includes the signal recycling mirror and the filter cavity, not present during the previous run. All laser power estimates are indicative as they can fluctuate rapidly. Virgo Interferometer O4 diagram.png
Layout of the Virgo interferometer during the O4 run (2023-2024). It includes the signal recycling mirror and the filter cavity, not present during the previous run. All laser power estimates are indicative as they can fluctuate rapidly.

The laser is the light source of the experiment. It must be powerful, while extremely stable in frequency and amplitude. [43] To meet all these (somewhat opposing) specifications, the beam starts from a very low power, yet very stable, laser. [44] The light from this laser passes through several amplifiers which enhance its power by a factor of 100. A 50 W output power was achieved for the last configuration of the initial Virgo detector, and later reached 100 W during the O3 run, following the Advanced Virgo upgrades; it is expected to be upgraded to 130 W at the beginning of the O4 run. [32] The original Virgo detector used a master-slave laser system, where a "master" laser is used to stabilize a high-powered "slave" laser; the master laser was a Nd:YAG laser, and the slave laser a Nd:YVO4 laser. [19] The retained solution for Advanced Virgo is to have a fiber laser with an amplification stage made of fibers as well, to improve the robustness of the system; in its final configuration, it is planned to coherently combine the light of two lasers in ordered to achieve the required power. [25] [45] The wavelength of the laser is 1064 nanometres, in both the original and Advanced Virgo configurations. [32]

This laser is then sent into the interferometer after passing through the injection system, which further ensures the stability of the beam, adjusts its shape and power, and positions it correctly for entering the interferometer. Key components of the injection system include the input mode cleaner (a 140-metre-long cavity made for improving the beam quality, by stabilizing the frequency, removing light propagating in an unwanted way and reduce the effect of misalignment of the laser), a Faraday isolator preventing any light from returning to the laser, and a mode matching telescope, which adapts the size and position of the beam right before it enters the interferometer. [25]

Mirrors

The large mirrors of the arm cavities are the most critical optics of the interferometer. They include the two end mirrors, located at the ends of the 3-km interferometer arms, and the two input mirrors, located near the beginning of the arms. Together, those mirrors make a resonant optical cavity in each arm, where the light bounces thousands of times before returning to the beam splitter, maximizing the effect of the signal on the laser path. [46] It also allows to increase the power of the light circulating in the arms. These mirrors have been specifically designed for Virgo and are made from state-of-the-art technologies. They are cylinders 35 cm in diameter and 20 cm thick, [25] made from the purest glass in the world. [47] The mirrors are polished to the atomic level in order to not diffuse (and hence lose) any light. [48] Finally, a reflective coating (a Bragg reflector made with ion beam sputtering) is added. The mirrors located at the end of the arms reflect almost all incoming light; less than 0.002% of the light is lost at each reflection. [49]

One of the mirrors from the initial Virgo detector, now used as an exposition model at the Virgo site. Initial Virgo mirror.jpg
One of the mirrors from the initial Virgo detector, now used as an exposition model at the Virgo site.

In addition, two other mirrors are present in the final design:

Superattenuators

Any Virgo mirror is supported, under vacuum, by a mechanical structure enormously damping seismic vibrations. A "superattenuator" consists of a chain of pendula, hanging from an upper platform, supported by three long flexible legs clamped to ground, forming an inverted pendulum. In this way seismic vibrations above 10 Hz are reduced by more than 10 times and the position of the mirror is very carefully controlled. Virgo3 1.jpg
Any Virgo mirror is supported, under vacuum, by a mechanical structure enormously damping seismic vibrations. A "superattenuator" consists of a chain of pendula, hanging from an upper platform, supported by three long flexible legs clamped to ground, forming an inverted pendulum. In this way seismic vibrations above 10 Hz are reduced by more than 10 times and the position of the mirror is very carefully controlled.

In order to mitigate the seismic noise which could propagate up to the mirrors, shaking them and hence obscuring potential gravitational wave signals, the large mirrors are suspended by a complex system. All of the main mirrors are suspended by four thin fibers made of silica [51] which are attached to a series of attenuators. This chain of suspension, called the "superattenuator", is close to 8 meters high and is also under vacuum. [52] The superattenuators do not only limit the disturbances on the mirrors, they also allow the mirror position and orientation to be precisely steered. The optical table where the injection optics used to shape the laser beam are located, such as the benches used for the light detection, are also suspended and under vacuum, in order to limit the seismic and acoustic noises. In the Advanced Virgo configuration, the whole instrumentation used to detect gravitational waves signals and to steer the interferometer (photodiodes, cameras, and the associated electronics) is also installed on several suspended benches, and under vacuum. [25]

The design of the superattenuators is mainly based on the passive attenuation of the seismic noise, which is achieved by chaining several pendula, each acting as an harmonic oscillator. They are characterized by a resonance frequency (which diminishes with the length of the pendulum) above which the noise will be dampened; chaining several pendula allows to reduce the noise by twelve orders of magnitude, at the cost of introducing multiple, collecitve resonance frequencies, which are at a higher frequency than a single pendulum. [53] In the current design, the highest resonance frequency is around 2 Hz, providing a meaningful noise reduction starting at 4 Hz, [25] and reaching the level needed for detecting gravitational waves around 10 Hz. A limit of the system is that the noise in the resonance frequency band (below 2 Hz) is not filtered and can generate large oscillations; this is mitigated by an active damping system, including sensors measuring the seismic noise and actuators controlling the superattenuator to counteract the noise. [53]

Detection system

Part of the light circulating in the arm cavities is sent towards the detection system by the beam splitter. In its optimal configuration, the interferometer works close to the "dark fringe", meaning that very little light is sent towards the output (most of it is sent back to the input, to be collected by the power recycling mirror). A fraction of this light is reflected back by the signal recycling mirror, and the rest is collected by the detection system. It first passes through the output mode cleaner, which allows to filter the so-called "high-order modes" (light propagating in an unwanted way, typically introduced by small defects in the mirrors, and susceptible to degrade the measurement [54] ), before reaching the photodiodes, which measure the light intensity. Both the output mode cleaner and the photodiodes are suspended and under vacuum. [24]

Detection bench of the Virgo interferometer before being installed in April 2015. It is 88 cm wide and hosts the output mode cleaner; the photodiode measuring the signal is placed on another bench. VirgoDetectionBench2015.jpg
Detection bench of the Virgo interferometer before being installed in April 2015. It is 88 cm wide and hosts the output mode cleaner; the photodiode measuring the signal is placed on another bench.

Starting with the O3 run, a squeezed vacuum source was introduced in order to reduce the quantum noise, which is one of the main limitations to sensitivity. When replacing the standard vacuum by a squeezed vacuum, the fluctuations of a quantity, either the amplitude or the phase of the light, is decreased, at the expense of increasing the fluctuations of the other quantity due to Heisenberg's uncertainty principle. In the case of Virgo, the two quantities are the amplitude and the phase. The idea of using squeezed vacuum was first proposed in 1981 by Carlton Caves, during the infancy of gravitational wave detectors. [55]

During the O3 run, frequency-independent squeezing was implemented, meaning that the squeezing is identical at all frequencies; it was used to reduce the shot noise (at high frequencies) and increase the radiation pressure noise (at low frequencies), as the latter was not limiting the instrument's sensitivity. [56] Due to the addition of the squeezed vacuum injection, the quantum noise was reduced by 3.2 dB at high frequencies, resulting in an increase of the range of the detector by 5–8%. [57]

Currently, more sophisticated squeezed states are produced [58] by combining the technology from O3 with a new 285 m long cavity, known as the filter cavity. This technology is known as frequency-dependent squeezing, and helps reduce the shot noise at high frequencies (where radiation pressure noise is not relevant), and reduce the radiation pressure noise at low frequencies (where shot noise is low). [59] [60]

Infrastructure

Seen from the air, the Virgo detector has a characteristic "L" shape with its two 3-km-long perpendicular arms. The arm "tunnels" house vacuum pipes in which the laser beams are travelling under an ultra-high vacuum.

Virgo is the largest ultra-high vacuum installation in Europe, with a total volume of 6,800 cubic meters. [61] The two 3-km arms are made of a long steel pipe 1.2m in diameter in which the target residual pressure is about 1 thousandth of a billionth of an atmosphere (improving by a factor of 100 from the original Virgo level). Thus, the residual gas molecules (mainly hydrogen and water) have a limited impact on the path of the laser beams. [25] Large gate valves are located at both ends of the arms so that work can be done in the mirror vacuum towers without breaking an arm's ultra-high vacuum. The towers containing the mirrors and attenuators are themselves split in two sections with different pressures. [62] The tubes undergo a process called baking, where they are heated at 150°C in order to remove unwanted particles stuck on the surfaces; while the towers were also baked-out in the initial Virgo design, cryogenic traps are now used to prevent contamination. [25]

Due to the high power in the interferometer, the mirrors are susceptible to thermal effects due to the heating induced by the laser (despite having an extremely low absorption). These thermal effects can take the shape of a deformation of the surface due to dilation, or a change in the refractive index of the substrate; this results in power escaping from the interferometer and in perturbations of the signal. These two effects are accounted for by the thermal compensation system (TCS), which includes sensors called Hartmann wavefront sensors [63] (HWS), used to measure the optical aberration through an auxiliary light source, and two actuators: CO2 lasers, which selectively heat parts of the mirror to correct the defects, and ring heaters, which precisely adjust the radius of curvature of the mirror. The system also corrects the "cold defects", which are permanent defects introduced during the mirror manufacturing. [64] [25] During the O3 run, the TCS was able to increase the power circulating inside the interferometer by 15%, and decrease the power leaving the interferometer by a factor of 2. [65]

One of the Newtonian calibrators ("NCal") before it was installed at the detector. Several of them are installed near one of the end mirror ; the movement of the rotor generates a varying gravitational force on the mirror, allowing to move it in a controlled manner. Newtonian Calibrator Virgo.png
One of the Newtonian calibrators ("NCal") before it was installed at the detector. Several of them are installed near one of the end mirror ; the movement of the rotor generates a varying gravitational force on the mirror, allowing to move it in a controlled manner.

Another important component is the system for controlling stray light, which refers to any light leaving the designated path of the interferometer, either by scattering on a surface or from unwanted reflection. The recombination of this stray light with the main beam of the interferometer can be a significant source of noise, and is often hard to track and to model. Most of the efforts to mitigate stray light are based on absorbing plates called "baffles", placed near the optics as well as within the tubes; additional precautions are needed to prevent the baffles from having an effect on the interferometer operation. [66] [67] [61]

In order to properly estimate the response of the detector to gravitational waves and thus correctly reconstruct the signal, a calibration step is required, which involves moving the mirrors in a controlled way and measuring the result. During the initial Virgo era, this was primarily achieved by agitating one of the pendulum to which the mirror is suspended using coils to generate a magnetic field interacting with magnets fixed to the pendulum. [68] This technique was employed until O2. For O3, the main calibration method became the photon calibration ("PCal") which had until then been used as a secondary method to validate the results; it uses an auxiliary laser to displace the mirror via the radiation pressure. [69] [70] In addition, a new method called Newtonian calibration ("NCal") has been introduced at the end of O2 and is now used to validate the PCal; it relies on gravity to move the mirror, by placing a rotating mass at a specific distance of the mirror. [71] [70]

Finally, the instrument requires an efficient data acquisition system. This system is in charge of managing the data measured at the output of the interferometer and from the many sensors present on the site, writing it in files, and distributing the files for data analysis. To this end, dedicated hardware and software have been developed in order to accommodate the specific needs of Virgo. [72]

Noise and sensitivity

Noise sources

Visualization of a gravitational wave "koi fish" glitch, from LIGO Hanford data taken in 2015. The top part represents the output of the detector ("strain") as a function of time, while the bottom part displays the frequency distribution of the power as a function of time. This type of glitch is of unknown origin, and covers a broad frequency range, with characteristic "fins" at lower frequencies in the time-frequency plot. Gravitational wave "Koi fish" glitch.png
Visualization of a gravitational wave "koi fish" glitch, from LIGO Hanford data taken in 2015. The top part represents the output of the detector ("strain") as a function of time, while the bottom part displays the frequency distribution of the power as a function of time. This type of glitch is of unknown origin, and covers a broad frequency range, with characteristic "fins" at lower frequencies in the time-frequency plot.

Due to the precision required in the measurement, the Virgo detector is sensitive to a number of sources of noise which limit the precision of the measurement. Some of these sources correspond to large frequency ranges and limit the overall sensitivity of the detector, [74] [61] such as:

  • seismic noise (any ground motion from numerous sources, such as waves in the Mediterranean Sea, wind, or human activity like traffic), generally in the low frequencies up to about 10 Hertz (Hz)
  • thermal noise of the mirrors and their suspension wires, corresponding to the agitation of the mirror/suspension from its own temperature, from a few tens to a few hundreds of Hz
  • quantum noise, which includes the laser shot noise, corresponding to the fluctuation of the power received by the detectors and relevant above a few hundreds of Hz, and the radiation pressure noise, corresponding to the pressure applied by the laser on the mirror, which is relevant at low frequency
  • Newtonian noise, caused by the variation of the gravity field which affects the position of the mirror, relevant below 20 Hz

In addition to these broad noise sources, a number of peaks are visible in the noise spectrum, related to specific noise sources. These notably include a line at 50 Hz (as well as harmonics at 100, 150, and 200 Hz), corresponding to the frequency of the European power grid; so-called "violin modes" at 300 Hz (and a number of harmonics), corresponding to the resonance frequency of the suspension fibers (which can vibrate at a specific frequency just as the strings of a violin do); and calibration lines, appearing when mirrors are moved for calibration. [75] [76]

Additional noise sources may also have a short-term impact—bad weather or earthquakes may temporarily increase the noise level. [61]

Finally, a number of short-lived artifacts may appear in the data due to many possible instrumental issues; these are usually referred to as 'glitches'. It is estimated that about 20% of the detected events are impacted by glitches, requiring specific data processing methods to mitigate their impact. [77]

Detector sensitivity

A sensitivity curve from the Virgo detector in the frequency band [10 Hz; 10 kHz], computed in August 2011. Its shape is typical: the thermal noise of the mirror suspension pendulum dominates at low frequency while the increase at high frequency is due to the laser shot noise. In between, one can see resonances (for instance, the suspension wire violin modes) and contributions from various instrumental noises (among which the 50 Hz frequency from the power grid and its harmonics) which one is trying to reduce continuously. BestVirgoSensitivityCurveVSR4.png
A sensitivity curve from the Virgo detector in the frequency band [10 Hz; 10 kHz], computed in August 2011. Its shape is typical: the thermal noise of the mirror suspension pendulum dominates at low frequency while the increase at high frequency is due to the laser shot noise. In between, one can see resonances (for instance, the suspension wire violin modes) and contributions from various instrumental noises (among which the 50 Hz frequency from the power grid and its harmonics) which one is trying to reduce continuously.

A detector like Virgo is characterized by its sensitivity, a figure of merit providing information about the tiniest signal the instrument could detect—the smaller the value of the sensitivity, the better the detector. The sensitivity varies with frequency as each noise has its own frequency range.

The most common measure for the sensitivity of a gravitational wave detector is the "horizon distance", defined as the distance at which a binary neutron star with masses 1.4 M–1.4 M (where M is the solar mass) produces a signal-to-noise ratio of 8 in the detector. It is generally expressed in megaparsecs. [79] For instance, the range for Virgo during the O3 run was between 40 and 50 Mpc. [33] This range is only an indicator and does not represent a maximal range for the detector; signals from more massive sources will have a larger amplitude, and can thus be detected from further away.

Virgo is a wide band detector whose sensitivity ranges from a few Hz up to 10 kHz. Mathematically speaking, its sensitivity is characterized by its power spectrum which is computed in real time using the data recorded by the detector. The image attached shows an example of Virgo amplitude spectrum density (the square root of the power spectrum) from 2011, plotted using a log-log scale.

Calculations show that the detector sensitivity roughly scales as , where is the arm cavity length and the laser power on the beam splitter. To improve it, these two quantities must be increased. This is achieved by having long arms, using optical cavities inside the arm to maximize the exposition to the signal, and implementing power recycling to increase the power in the arms. [74] [80]

Data analysis

An important part of the Virgo collaboration resources is dedicated to the development and deployment of data analysis software designed to process the output of the detector. Apart from the data acquisition software and the tools for distributing the data, this effort is mostly shared with members of the LIGO and KAGRA collaborations, as part of the LIGO-Virgo-KAGRA (LVK) collaboration. [81]

The data from the detector is initially only available to LVK members; segments of data around detected events are released at the time of publication of the related paper, and the full data is released after a proprietary period, currently lasting 18 months. During the third observing run (O3), this resulted in two separated data releases (O3a and O3b), corresponding to the first six months and last six months of the run respectively. [82] The data is then available for anyone on the Gravitational Wave Open Science Center (GWOSC) platform. [83] [84]

The analysis of the data requires a variety of different techniques, targetting the different type of sources. The major part of the effort is dedicated to the detection and analysis of mergers of compact objects, the only type of source detected up until now. Several different analysis software are running on the data searching for this event, and a dedicated infrastructure is used to emit alerts to the online community. Other efforts are carried out after the data taking period ("offline"), including searches for continuous sources or for a stochastic background, as well as deeper analysis of the detected events.

Scientific results

Sky localization of the GW170814 event, both with the two LIGO detectors and with the full network. The addition of Virgo allows for a much more precise localization. GW170814.png
Sky localization of the GW170814 event, both with the two LIGO detectors and with the full network. The addition of Virgo allows for a much more precise localization.

The first detection of a gravitational signal by Virgo took place at the beginning of the second observing run (O2), as Virgo was absent from the first observing run. The event, named GW170814, was a coalescence between two black holes, and also the first event to be detected by three different detectors, allowing for its localization to be greatly improved compared to the events from the first observing run. It also allowed for the first conclusive measure of gravitational wave polarizations, providing evidence against the existence of polarizations other than the ones predicted by general relativity. [26]

It was soon followed by the more famous GW170817, first merger of two neutron stars detected by the gravitational wave network, and as of January 2023 the only event with a confirmed detection of an electromagnetic counterpart, both in gamma rays and in optical telescopes, and later in the radio and x-ray domains. While no signal was observed in Virgo, this absence was crucial to put tighter constraints on the localization of the event. [28] This event had tremendous repercussions in the astronomical community, involving more than 4000 astronomers, [85] improving the understanding of neutron star mergers, [86] and putting very tight constraints on the speed of gravity. [87]

Several searches for continuous gravitational waves have been performed on data from the past runs. On the O3 run, these include an all-sky search, [88] targeted searches toward Scorpius X-1 [89] and a number of known pulsars (including the Crab and Vela pulsars), [90] [91] and directed search towards the supernova remnants Cassiopeia A and Vela Jr. [92] and the Galactic Center. [93] While none of the sources managed to identify a signal, this allowed upper limits to be set on some parameters; in particular, it was found that the deviation from perfect spinning balls for close known pulsars is at most of the order of 1 mm. [88]

Virgo was included in the latest search for a gravitational wave background along with LIGO, combining the results of O3 with the ones from the O1 and O2 runs (which only used LIGO data). No stochastic background was observed, improving previous constraints on the energy of the background by an order of magnitude. [94]

Constraints on the Hubble constant have also been obtained; the current best estimate is 68+12
-8
km s−1 Mpc−1, combining results from binary black holes and from the GW170817 event. This result is coherent with other estimates of the constant, but not precise enough to resolve the tension regarding its exact value. [95]

Outreach

The Virgo collaboration participates in a number of activities promoting communication and education on gravitational waves towards the general public. [96] This includes a wide variety of activities, such as:

Related Research Articles

<span class="mw-page-title-main">LIGO</span> Gravitational wave detector

The Laser Interferometer Gravitational-Wave Observatory (LIGO) is a large-scale physics experiment and observatory designed to detect cosmic gravitational waves and to develop gravitational-wave observations as an astronomical tool. Two large observatories were built in the United States with the aim of detecting gravitational waves by laser interferometry. These observatories use mirrors spaced four kilometers apart which are capable of detecting a change of less than one ten-thousandth the charge diameter of a proton.

<span class="mw-page-title-main">Laser Interferometer Space Antenna</span> European space mission to measure gravitational waves

The Laser Interferometer Space Antenna (LISA) is a planned space probe to detect and accurately measure gravitational waves—tiny ripples in the fabric of spacetime—from astronomical sources. LISA will be the first dedicated space-based gravitational-wave observatory. It aims to measure gravitational waves directly by using laser interferometry. The LISA concept has a constellation of three spacecraft arranged in an equilateral triangle with sides 2.5 million kilometres long, flying along an Earth-like heliocentric orbit. The distance between the satellites is precisely monitored to detect a passing gravitational wave.

<span class="mw-page-title-main">Max Planck Institute for Gravitational Physics</span>

The Max Planck Institute for Gravitational Physics is a Max Planck Institute whose research is aimed at investigating Einstein's theory of relativity and beyond: Mathematics, quantum gravity, astrophysical relativity, and gravitational-wave astronomy. The institute was founded in 1995 and is located in the Potsdam Science Park in Golm, Potsdam and in Hannover where it closely collaborates with the Leibniz University Hannover. Both the Potsdam and the Hannover parts of the institute are organized in three research departments and host a number of independent research groups.

<span class="mw-page-title-main">GEO600</span> Gravitational wave detector in Germany

GEO600 is a gravitational wave detector located near Sarstedt, a town 20 km to the south of Hanover, Germany. It is designed and operated by scientists from the Max Planck Institute for Gravitational Physics, Max Planck Institute of Quantum Optics and the Leibniz Universität Hannover, along with University of Glasgow, University of Birmingham and Cardiff University in the United Kingdom, and is funded by the Max Planck Society and the Science and Technology Facilities Council (STFC). GEO600 is capable of detecting gravitational waves in the frequency range 50 Hz to 1.5 kHz, and is part of a worldwide network of gravitational wave detectors. This instrument, and its sister interferometric detectors, when operational, are some of the most sensitive gravitational wave detectors ever designed. They are designed to detect relative changes in distance of the order of 10−21, about the size of a single atom compared to the distance from the Sun to the Earth. Construction on the project began in 1995.

<span class="mw-page-title-main">KAGRA</span> Japanese underground gravitational wave detector

The Kamioka Gravitational Wave Detector (KAGRA) is a large interferometer designed to detect gravitational waves predicted by the general theory of relativity. KAGRA is a Michelson interferometer that is isolated from external disturbances: its mirrors and instrumentation are suspended and its laser beam operates in a vacuum. The instrument's two arms are three kilometres long and located underground at the Kamioka Observatory which is near the Kamioka section of the city of Hida in Gifu Prefecture, Japan.

<span class="mw-page-title-main">Gravitational wave</span> Propagating spacetime ripple

Gravitational waves are waves of the intensity of gravity that are generated by the accelerated masses of binary stars and other motions of gravitating masses, and propagate as waves outward from their source at the speed of light. They were first proposed by Oliver Heaviside in 1893 and then later by Henri Poincaré in 1905 as the gravitational equivalent of electromagnetic waves. Gravitational waves are sometimes called gravity waves, but gravity waves typically refer to displacement waves in fluids. In 1916 Albert Einstein demonstrated that gravitational waves result from his general theory of relativity as ripples in spacetime.

<span class="mw-page-title-main">Gravitational-wave observatory</span> Device used to measure gravitational waves

A gravitational-wave detector is any device designed to measure tiny distortions of spacetime called gravitational waves. Since the 1960s, various kinds of gravitational-wave detectors have been built and constantly improved. The present-day generation of laser interferometers has reached the necessary sensitivity to detect gravitational waves from astronomical sources, thus forming the primary tool of gravitational-wave astronomy.

<span class="mw-page-title-main">Gravitational-wave astronomy</span> Branch of astronomy using gravitational waves

Gravitational-wave astronomy is a subfield of astronomy concerned with the detection and study of gravitational waves emitted by astrophysical sources.

<span class="mw-page-title-main">Carlton M. Caves</span> American physicist

Carlton Morris Caves is an American theoretical physicist. He is currently professor emeritus and research professor of physics and astronomy at the University of New Mexico. Caves works in the areas of physics of information; information, entropy, and complexity; quantum information theory; quantum chaos, quantum optics; the theory of non-classical light; the theory of quantum noise; and the quantum theory of measurement. He is a Fellow of the American Physical Society and of the American Association for the Advancement of Science and is a member of the US National Academy of Sciences.

Allegro was a ground-based, cryogenic resonant Weber bar, gravitational-wave detector run by Warren Johnson, et al. at Louisiana State University in Baton Rouge, Louisiana. The detector was commissioned in the early 1990s, and was decommissioned in 2008.

INDIGO or IndIGO is a consortium of Indian gravitational wave physicists. It is an initiative to set up advanced experimental facilities for a multi-institutional observatory project in gravitational-wave astronomy to be located near Aundha Nagnath, Hingoli District, Maharashtra, India. Predicted date of commission is in 2030.

<span class="mw-page-title-main">First observation of gravitational waves</span> 2015 direct detection of gravitational waves by the LIGO and VIRGO interferometers

The first direct observation of gravitational waves was made on 14 September 2015 and was announced by the LIGO and Virgo collaborations on 11 February 2016. Previously, gravitational waves had been inferred only indirectly, via their effect on the timing of pulsars in binary star systems. The waveform, detected by both LIGO observatories, matched the predictions of general relativity for a gravitational wave emanating from the inward spiral and merger of a pair of black holes of around 36 and 29 solar masses and the subsequent "ringdown" of the single resulting black hole. The signal was named GW150914. It was also the first observation of a binary black hole merger, demonstrating both the existence of binary stellar-mass black hole systems and the fact that such mergers could occur within the current age of the universe.

<span class="mw-page-title-main">GW170817</span> Gravitational-wave signal detected in 2017

GW 170817 was a gravitational wave (GW) signal observed by the LIGO and Virgo detectors on 17 August 2017, originating from the shell elliptical galaxy NGC 4993. The signal was produced by the last moments of the inspiral process of a binary pair of neutron stars, ending with their merger. It is the first GW observation that has been confirmed by non-gravitational means. Unlike the five previous GW detections—which were of merging black holes and thus not expected to produce a detectable electromagnetic signal—the aftermath of this merger was seen across the electromagnetic spectrum by 70 observatories on 7 continents and in space, marking a significant breakthrough for multi-messenger astronomy. The discovery and subsequent observations of GW 170817 were given the Breakthrough of the Year award for 2017 by the journal Science.

In quantum physics, light is in a squeezed state if its electric field strength Ԑ for some phases has a quantum uncertainty smaller than that of a coherent state. The term squeezing thus refers to a reduced quantum uncertainty. To obey Heisenberg's uncertainty relation, a squeezed state must also have phases at which the electric field uncertainty is anti-squeezed, i.e. larger than that of a coherent state. Since 2019, the gravitational-wave observatories LIGO and Virgo employ squeezed laser light, which has significantly increased the rate of observed gravitational-wave events.

GW 190412 was a gravitational wave (GW) signal observed by the LIGO and Virgo detectors on 12 April 2019. In April 2020, it was announced as the first time a collision of a pair of very differently sized black holes has been detected. As a result of this asymmetry, the signal included two measurable harmonics with frequencies approximately a factor 1.5 apart.

<span class="mw-page-title-main">Rana X. Adhikari</span> American experimental physicist (born 1974)

Rana X. Adhikari is an American experimental physicist. He is a professor of physics at the California Institute of Technology (Caltech) and an associate faculty member of the International Centre for Theoretical Sciences of Tata Institute of Fundamental Research (ICTS-TIFR).

Lisa Barsotti is a research scientist at the Massachusetts Institute of Technology Kavli Institute.

<span class="mw-page-title-main">C.S. Unnikrishnan</span> Indian physicist and professor (born 1962)

C. S. Unnikrishnan is an Indian physicist and professor known for his contributions in multiple areas of experimental and theoretical physics. He has been a professor at the Tata Institute of Fundamental Research Mumbai and is currently a professor in the School of Quantum Technology at the Defence Institute of Advanced Technology in Pune. He has made significant contributions in foundational issues in gravity and quantum physics and has published over 250 research papers and articles. Unnikrishnan is also a key member of the LIGO-India project and a member of the global LIGO Scientific Collaboration

Ground-based interferometric gravitational-wave search refers to methods and devices used to search and detect gravitational waves based on interferometers built on the ground. Most of current gravitational wave observations have been made using these techniques; the first one was made in 2015 by the two LIGO detectors. The current major detectors are the two LIGO in the United States, Virgo in Italy and KAGRA in Japan, which are all part of the second generation of detectors; future projects include LIGO-India as part of the second generation, and the Einstein Telescope and Cosmic Explorer forming a third generation.

References

  1. 1 2 "Our mission". www.ego-gw.it. European Gravitational Observatory. Retrieved 11 October 2023.
  2. 1 2 "The Virgo Collaboration". virgo-gw.eu. The Virgo Collaboration. 18 February 2021. Retrieved 11 October 2023.
  3. 1 2 "LIGO-M060038-v5: Memorandum of Understanding (MoU) Between VIRGO and LIGO". dcc.ligo.org. Retrieved 4 July 2023.
  4. "Communique de presse – Le CNRS signe l'accord franco-italien de création du consortium EGO European Gravitational Observatory" [Press release - The CNRS signs the franco-italian agreement on the creation of the EGO (European Gravitational Observatory) consortium.]. Cnrs.fr (in French). Archived from the original on 5 March 2016. Retrieved 11 February 2016.
  5. "Gravitational wave detectors prepare for next observing run – Virgo". www.virgo-gw.eu. Retrieved 4 May 2023.
  6. "The Virgo Institutions". virgo-gw.eu. The Virgo Collaboration. Retrieved 11 October 2023.
  7. "Scientific Collaboration – Virgo". www.virgo-gw.eu. Retrieved 31 March 2023.
  8. "LIGO Scientific Collaboration - Learn about the LSC". www.ligo.org. Retrieved 31 March 2023.
  9. "KAGRA to Join LIGO and Virgo in Hunt for Gravitational Waves". LIGO Lab | Caltech. Retrieved 4 July 2023.
  10. Einstein, Albert (1 January 1916). "Näherungsweise Integration der Feldgleichungen der Gravitation" [Approximative Integration of the Field Equations of Gravitation]. Sitzungsberichte der Königlich Preussischen Akademie der Wissenschaften (Minutes of the Royal Prussian Academy of Sciences) (in German): 688–696. Bibcode:1916SPAW.......688E.
  11. Weber, J. (3 June 1968). "Gravitational-Wave-Detector Events". Physical Review Letters . 20 (23): 1307–1308. Bibcode:1968PhRvL..20.1307W. doi:10.1103/PhysRevLett.20.1307.
  12. 1 2 Bersanetti, Diego; Patricelli, Barbara; Piccinni, Ornella Juliana; Piergiovanni, Francesco; Salemi, Francesco; Sequino, Valeria (August 2021). "Advanced Virgo: Status of the Detector, Latest Results and Future Prospects". Universe . 7 (9): 322. Bibcode:2021Univ....7..322B. doi: 10.3390/universe7090322 . hdl: 11568/1161730 . ISSN   2218-1997.
  13. 1 2 Giazotto, Adalberto (2018). La musica nascosta dell'universo: La mia vita a caccia delle onde gravitazionali[The hidden music of the Universe : my life of running after gravitational waves] (in Italian). Turin: Einaudi. ASIN   B07FY52PGV. Bibcode:2018lmnd.book.....G.
  14. Giazotto, Adalberto; Milano, Leopoldo; Bordoni, Franco; Brillet, Alain; Tourrenc, Philippe (12 May 1987). Proposta di Antenna interferometrica a grande base per la ricerca di Onde Gravitazionali [Proposition for an interferometric antenna with long arms for searching gravitational waves](PDF). ego-gw.it (Technical report) (in Italian).
  15. Caron, B.; Dominjon, A.; Drezen, C.; Flaminio, R.; Grave, X.; Marion, F.; Massonnet, L.; Mehmel, C.; Morand, R.; Mours, B.; Yvert, M.; Babusci, D.; Giordano, G.; Matone, G.; Mackowski, J. -M. (1 May 1996). "Status of the VIRGO experiment". Nuclear Physics B - Proceedings Supplements. Proceedings of the Fourth International Workshop on Theoretical and Phenomenological Aspects of Underground Physics. 48 (1): 107–109. Bibcode:1996NuPhS..48..107C. doi:10.1016/0920-5632(96)00220-4. ISSN   0920-5632.
  16. J.M. Weisberg and J.H. Taylor (2004). "Relativistic Binary Pulsar B1913+16: Thirty Years of Observations and Analysis". ASP Conference Series. 328: 25. arXiv: astro-ph/0407149 . Bibcode:2005ASPC..328...25W.
  17. Riles, K. (2013). "Gravitational Waves: Sources, Detectors and Searches". Progress in Particle and Nuclear Physics. 68: 1–54. arXiv: 1209.0667 . Bibcode:2013PrPNP..68....1R. doi:10.1016/j.ppnp.2012.08.001. S2CID   56407863.
  18. Sathyaprakash and, B.S.; Schutz, Bernard F. (2009). "Physics, Astrophysics and Cosmology with Gravitational Waves". Living Reviews in Relativity. 12 (1): 2. arXiv: 0903.0338 . Bibcode:2009LRR....12....2S. doi:10.12942/lrr-2009-2. PMC   5255530 . PMID   28163611. Archived from the original on 4 March 2016. Retrieved 11 February 2016.
  19. 1 2 Acernese, F.; Amico, P.; Al-Shourbagy, M.; Aoudia, S.; Avino, S.; et al. (August 2004). "The status of VIRGO". 5th Rencontres du Vietnam Particle Physics and Astrophysics. Hanoi, Vietnam: 1–6 via HAL.
  20. "Ondes gravitationnelles : Virgo entre dans sa phase d'exploitation scientifique – Communiqués et dossiers de presse" [Gravitational waves : Virgo enters in its scientific exploitation phase - Press releases and communications](PDF). Cnrs.fr (in French). Retrieved 21 February 2024.
  21. Accadia, T.; Acernese, F.; Alshourbagy, M.; Amico, P.; Antonucci, F.; Aoudia, S.; Arnaud, N.; Arnault, C.; Arun, K. G.; Astone, P.; Avino, S.; Babusci, D.; Ballardin, G.; Barone, F.; Barrand, G.; Barsotti, L.; Barsuglia, M.; Basti, A.; Bauer, Th S.; Beauville, F.; Bebronne, M.; Bejger, M.; Beker, M. G.; Bellachia, F.; Belletoile, A.; Beney, J. L.; Bernardini, M.; Bigotta, S.; Bilhaut, R.; et al. (29 March 2012). "Virgo: a laser interferometer to detect gravitational waves". Journal of Instrumentation . 7 (3): P03012. Bibcode:2012JInst...7.3012A. doi: 10.1088/1748-0221/7/03/P03012 .
  22. Lorenzini, Matteo (April 2010). "The monolithic suspension for the Virgo interferometer". Classical and Quantum Gravity . 27 (8): 084021. Bibcode:2010CQGra..27h4021L. doi:10.1088/0264-9381/27/8/084021. S2CID   123269358.
  23. The Virgo Collaboration (2011). "Status of the Virgo project" (PDF). Classical and Quantum Gravity. 28 (11): 114002. Bibcode:2011CQGra..28k4002A. doi:10.1088/0264-9381/28/11/114002. S2CID   59369141.
  24. 1 2 Acernese, F.; Agathos, M.; Agatsuma, K.; Aisa, D.; Allemandou, N.; Allocca, A.; Amarni, J.; Astone, P.; Balestri, G.; Ballardin, G.; Barone, F.; Baronick, J-P; Barsuglia, M.; Basti, A.; Basti, F.; Bauer, Th S.; Bavigadda, V.; Bejger, M.; Beker, M. G.; Belczynski, C.; Bersanetti, D.; Bertolini, A.; Bitossi, M.; Bizouard, M. A.; Bloemen, S.; Blom, M.; Boer, M.; Bogaert, G.; Bondi, D.; et al. (2015). "Advanced Virgo: A second-generation interferometric gravitational wave detector". Classical and Quantum Gravity. 32 (2): 024001. arXiv: 1408.3978 . Bibcode:2015CQGra..32b4001A. doi:10.1088/0264-9381/32/2/024001. S2CID   20640558.
  25. 1 2 3 4 5 6 7 8 9 Many authors of the Virgo Collaboration (13 April 2012). Advanced Virgo Technical Design Report VIR–0128A–12 (PDF).
  26. 1 2 Abbott, B. P.; Abbott, R.; Abbott, T. D.; Acernese, F.; Ackley, K.; Adams, C.; Adams, T.; Addesso, P.; Adhikari, R. X.; Adya, V. B.; Affeldt, C.; Afrough, M.; Agarwal, B.; Agathos, M.; Agatsuma, K. (6 October 2017). "GW170814: A Three-Detector Observation of Gravitational Waves from a Binary Black Hole Coalescence". Physical Review Letters. 119 (14): 141101. arXiv: 1709.09660 . Bibcode:2017PhRvL.119n1101A. doi:10.1103/PhysRevLett.119.141101. ISSN   0031-9007. PMID   29053306. S2CID   46829350.
  27. Gibney, Elizabeth (27 September 2017). "European detector spots its first gravitational wave". Nature. Retrieved 21 February 2024.
  28. 1 2 Abbott, B. P.; Abbott, R.; Abbott, T. D.; Acernese, F.; Ackley, K.; Adams, C.; Adams, T.; Addesso, P.; Adhikari, R. X.; Adya, V. B.; Affeldt, C.; Afrough, M.; Agarwal, B.; Agathos, M.; Agatsuma, K. (16 October 2017). "Multi-messenger Observations of a Binary Neutron Star Merger". The Astrophysical Journal . 848 (2): L12. arXiv: 1710.05833 . Bibcode:2017ApJ...848L..12A. doi: 10.3847/2041-8213/aa91c9 . ISSN   2041-8213. S2CID   217162243.
  29. Abbott, B. P.; Abbott, R.; Abbott, T. D.; Acernese, F.; Ackley, K.; Adams, C.; Adams, T.; Addesso, P.; Adhikari, R. X.; Adya, V. B.; Affeldt, C.; Afrough, M.; Agarwal, B.; Agathos, M. (28 February 2018). "GW170817: Implications for the Stochastic Gravitational-Wave Background from Compact Binary Coalescences". Physical Review Letters. 120 (9): 091101. arXiv: 1710.05837 . Bibcode:2018PhRvL.120i1101A. doi:10.1103/PhysRevLett.120.091101. PMID   29547330. S2CID   3889124.
  30. Bersanetti, Diego (13 July 2019). "Status of the Virgo gravitational-wave detector and the O3 Observing Run - EPS-HEP2019". cern.ch. Retrieved 29 February 2024.
  31. "LIGO Suspends Third Observing Run (O3)". LIGO Lab | Caltech. Retrieved 16 April 2023.
  32. 1 2 3 4 Flaminio, Raffaele (13 December 2020). "Status and plans of the Virgo gravitational wave detector". In Marshall, Heather K.; Spyromilio, Jason; Usuda, Tomonori (eds.). Ground-based and Airborne Telescopes VIII (PDF). SPIE Conference Series. Vol. 11445. SPIE. pp. 205–214. Bibcode:2020SPIE11445E..11F. doi:10.1117/12.2565418. ISBN   9781510636774. S2CID   230549331.
  33. 1 2 3 4 "IGWN | Observing Plans". observing.docs.ligo.org. Retrieved 16 January 2024.
  34. "Virgo postpones entry into O4 observing run – Virgo". www.virgo-gw.eu. Retrieved 13 May 2023.
  35. "IGWN". Observing Plans. 10 April 2024. Retrieved 23 April 2024.
  36. The Virgo Collaboration (31 May 2022). Virgo nEXT: beyond the AdV+ project - A concept study (PDF). ego-gw.it (Technical report).
  37. "Astrophysical Sources of Gravitational Waves". Virgo. Retrieved 17 May 2024.
  38. Eardley, Douglas M.; Lee, David L.; Lightman, Alan P.; Wagoner, Robert V.; Will, Clifford M. (30 April 1973). "Gravitational-Wave Observations as a Tool for Testing Relativistic Gravity". Physical Review Letters. 30 (18): 884–886. Bibcode:1973PhRvL..30..884E. doi:10.1103/PhysRevLett.30.884. hdl: 2060/19730012613 . S2CID   120335306.
  39. Abbott, R.; et al. (2021). "Search for Lensing Signatures in the Gravitational-Wave Observations from the First Half of LIGO–Virgo's Third Observing Run". The Astrophysical Journal. 923 (1): 14. arXiv: 2105.06384 . Bibcode:2021ApJ...923...14A. doi: 10.3847/1538-4357/ac23db . S2CID   234482851.
  40. Van Den Broeck, Chris (2014), Ashtekar, Abhay; Petkov, Vesselin (eds.), "Probing Dynamical Spacetimes with Gravitational Waves", Springer Handbook of Spacetime, Springer Handbooks, Berlin, Heidelberg: Springer, pp. 589–613, arXiv: 1301.7291 , Bibcode:2014shst.book..589V, doi:10.1007/978-3-642-41992-8_27, ISBN   978-3-642-41992-8, S2CID   119242493 , retrieved 23 April 2023
  41. The LIGO Scientific Collaboration; the Virgo Collaboration; the KAGRA Collaboration; Abbott, R.; Abe, H.; Acernese, F.; Ackley, K.; Adhikari, N.; Adhikari, R. X.; Adkins, V. K.; Adya, V. B.; Affeldt, C.; Agarwal, D.; Agathos, M.; Agatsuma, K. (2023). "Constraints on the Cosmic Expansion History from GWTC–3". The Astrophysical Journal. 949 (2): 76. arXiv: 2111.03604 . Bibcode:2023ApJ...949...76A. doi: 10.3847/1538-4357/ac74bb . S2CID   243832919.
  42. Vinet, Jean-Yves; The Virgo Collaboration (2006). The VIRGO physics book Vol. II (PDF). p. 19.
  43. F. Bondu; et al. (1996). "Ultrahigh-spectral-purity laser for the VIRGO experiment". Optics Letters . 21 (8): 582–4. Bibcode:1996OptL...21..582B. doi:10.1364/OL.21.000582. PMID   19876090.
  44. F. Bondu; et al. (2002). "The VIRGO injection system" (PDF). Classical and Quantum Gravity. 19 (7): 1829–1833. Bibcode:2002CQGra..19.1829B. doi:10.1088/0264-9381/19/7/381. S2CID   250902832.
  45. Wei, Li-Wei (3 December 2015). High-power laser system for Advanced Virgo gravitational wave detector : coherently combined master oscillator fiber power amplifiers (PhD thesis). Université Nice Sophia Antipolis.
  46. "Optical Layout – Virgo". www.virgo-gw.eu. Retrieved 5 March 2023.
  47. J. Degallaix (2015). "Silicon, the test mass substrate of tomorrow?" (PDF). The Next Detectors for Gravitational Wave Astronomy. Archived from the original (PDF) on 8 December 2015. Retrieved 16 December 2015.
  48. R. Bonnand (2012). The Advanced Virgo Gravitational Wave Detector/ Study of the optical design and development of the mirrors (PhD) (in French). Université Claude Bernard – Lyon I.
  49. R Flaminio; et al. (2010). "A study of coating mechanical and optical losses in view of reducing mirror thermal noise in gravitational wave detectors" (PDF). Classical and Quantum Gravity. 27 (8): 084030. Bibcode:2010CQGra..27h4030F. doi:10.1088/0264-9381/27/8/084030. S2CID   122750664.
  50. Boschi, Valerio (1 March 2019). "Seismic isolation in advanced Virgo gravitational wave detector". Journal of the Acoustical Society of America . 145 (3_Supplement): 1668. Bibcode:2019ASAJ..145.1668B. doi:10.1121/1.5101119. ISSN   0001-4966. S2CID   150337668.
  51. M. Lorenzini & Virgo Collaboration (2010). "The monolithic suspension for the virgo interferometer". Classical and Quantum Gravity. 27 (8): 084021. Bibcode:2010CQGra..27h4021L. doi:10.1088/0264-9381/27/8/084021. S2CID   123269358.
  52. Braccini, S.; Barsotti, L.; Bradaschia, C.; Cella, G.; Virgilio, A. Di; Ferrante, I.; Fidecaro, F.; Fiori, I.; Frasconi, F.; Gennai, A.; Giazotto, A.; Paoletti, F.; Passaquieti, R.; Passuello, D.; Poggiani, R. (1 July 2005). "Measurement of the seismic attenuation performance of the VIRGO Superattenuator". Astroparticle Physics. 23 (6): 557–565. Bibcode:2005APh....23..557B. doi:10.1016/j.astropartphys.2005.04.002. ISSN   0927-6505.
  53. 1 2 Beker, M. G.; Blom, M.; van den Brand, J. F. J.; Bulten, H. J.; Hennes, E.; Rabeling, D. S. (1 January 2012). "Seismic Attenuation Technology for the Advanced Virgo Gravitational Wave Detector". Physics Procedia. Proceedings of the 2nd International Conference on Technology and Instrumentation in Particle Physics (TIPP 2011). 37: 1389–1397. Bibcode:2012PhPro..37.1389B. doi: 10.1016/j.phpro.2012.03.741 . ISSN   1875-3892.
  54. Beauville, F; Buskulic, D; Derome, L; Dominjon, A; Flaminio, R; Hermel, R; Marion, F; Masserot, A; Massonnet, L; Mours, B; Moreau, F; Mugnier, P; Ramonet, J; Tournefier, E; Verkindt, D (7 May 2006). "Improvement in the shot noise of a laser interferometer gravitational wave detector by means of an output mode-cleaner". Classical and Quantum Gravity. 23 (9): 3235–3250. Bibcode:2006CQGra..23.3235B. doi:10.1088/0264-9381/23/9/030. ISSN   0264-9381. S2CID   123072147.
  55. Caves, Carlton M. (15 April 1981). "Quantum-mechanical noise in an interferometer". Physical Review D. 23 (8): 1693–1708. Bibcode:1981PhRvD..23.1693C. doi:10.1103/PhysRevD.23.1693.
  56. The Virgo Collaboration; Acernese, F.; Agathos, M.; Aiello, L.; Ain, A.; Allocca, A.; Amato, A.; Ansoldi, S.; Antier, S.; Arène, M.; Arnaud, N.; Ascenzi, S.; Astone, P.; Aubin, F.; Babak, S. (22 September 2020). "Quantum Backaction on kg-Scale Mirrors: Observation of Radiation Pressure Noise in the Advanced Virgo Detector". Physical Review Letters. 125 (13): 131101. Bibcode:2020PhRvL.125m1101A. doi: 10.1103/PhysRevLett.125.131101 . hdl: 11390/1193696 . PMID   33034506. S2CID   222235425.
  57. Virgo Collaboration; Acernese, F.; Agathos, M.; Aiello, L.; Allocca, A.; Amato, A.; Ansoldi, S.; Antier, S.; Arène, M.; Arnaud, N.; Ascenzi, S.; Astone, P.; Aubin, F.; Babak, S.; Bacon, P. (5 December 2019). "Increasing the Astrophysical Reach of the Advanced Virgo Detector via the Application of Squeezed Vacuum States of Light". Physical Review Letters. 123 (23): 231108. Bibcode:2019PhRvL.123w1108A. doi: 10.1103/PhysRevLett.123.231108 . hdl: 11585/709335 . PMID   31868444. S2CID   209446443.
  58. Virgo Collaboration; Acernese, F.; Agathos, M.; Ain, A.; Albanesi, S.; Alléné, C.; Allocca, A.; Amato, A.; Amra, C.; Andia, M.; Andrade, T.; Andres, N.; Andrés-Carcasona, M.; Andrić, T.; Ansoldi, S. (25 July 2023). "Frequency-Dependent Squeezed Vacuum Source for the Advanced Virgo Gravitational-Wave Detector". Physical Review Letters. 131 (4): 041403. Bibcode:2023PhRvL.131d1403A. doi: 10.1103/PhysRevLett.131.041403 . hdl: 11568/1196710 . PMID   37566847. S2CID   260185660.
  59. Zhao, Yuhang; Aritomi, Naoki; Capocasa, Eleonora; Leonardi, Matteo; Eisenmann, Marc; Guo, Yuefan; Polini, Eleonora; Tomura, Akihiro; Arai, Koji; Aso, Yoichi; Huang, Yao-Chin; Lee, Ray-Kuang; Lück, Harald; Miyakawa, Osamu; Prat, Pierre (28 April 2020). "Frequency-Dependent Squeezed Vacuum Source for Broadband Quantum Noise Reduction in Advanced Gravitational-Wave Detectors". Physical Review Letters. 124 (17): 171101. arXiv: 2003.10672 . Bibcode:2020PhRvL.124q1101Z. doi:10.1103/PhysRevLett.124.171101. PMID   32412296. S2CID   214623227.
  60. Polini, E (1 August 2021). "Broadband quantum noise reduction via frequency dependent squeezing for Advanced Virgo Plus". Physica Scripta . 96 (8): 084003. Bibcode:2021PhyS...96h4003P. doi:10.1088/1402-4896/abfef0. ISSN   0031-8949. S2CID   235285860.
  61. 1 2 3 4 "Fighting Noises – Virgo". www.virgo-gw.eu. Retrieved 21 February 2023.
  62. VIRGO Vacuum System Overview, A.Pasqualetti https://workarea.ego-gw.it/ego2/virgo/advanced-virgo/vac/varies/Virgo_Vacuum_system_Overview_r2.pdf
  63. Kelly, Thu-Lan; Veitch, Peter J.; Brooks, Aidan F.; Munch, Jesper (20 February 2007). "Accurate and precise optical testing with a differential Hartmann wavefront sensor". Applied Optics . 46 (6): 861–866. Bibcode:2007ApOpt..46..861K. doi:10.1364/AO.46.000861. hdl: 2440/43095 . ISSN   2155-3165. PMID   17279130.
  64. Rocchi, A; Coccia, E; Fafone, V; Malvezzi, V; Minenkov, Y; Sperandio, L (1 June 2012). "Thermal effects and their compensation in Advanced Virgo". Journal of Physics: Conference Series . 363 (1): 012016. Bibcode:2012JPhCS.363a2016R. doi: 10.1088/1742-6596/363/1/012016 . ISSN   1742-6596. S2CID   122763506.
  65. Nardecchia, Ilaria (2022). "Detecting Gravitational Waves with Advanced Virgo". Galaxies. 10 (1): 28. Bibcode:2022Galax..10...28N. doi: 10.3390/galaxies10010028 . ISSN   2075-4434.
  66. Vinet, Jean-Yves; Brisson, Violette; Braccini, Stefano; Ferrante, Isidoro; Pinard, Laurent; Bondu, François; Tournié, Eric (15 November 1997). "Scattered light noise in gravitational wave interferometric detectors: A statistical approach". Physical Review D. 56 (10): 6085–6095. Bibcode:1997PhRvD..56.6085V. doi:10.1103/PhysRevD.56.6085.
  67. Vinet, Jean-Yves; Brisson, Violette; Braccini, Stefano (15 July 1996). "Scattered light noise in gravitational wave interferometric detectors: Coherent effects". Physical Review D. 54 (2): 1276–1286. Bibcode:1996PhRvD..54.1276V. doi:10.1103/PhysRevD.54.1276. PMID   10020804.
  68. Accadia, T; Acernese, F; Antonucci, F; Astone, P; Ballardin, G; Barone, F; Barsuglia, M; Basti, A; Bauer, Th S; Beker, M G; Belletoile, A; Birindelli, S; Bitossi, M; Bizouard, M A; Blom, M (21 January 2011). "Calibration and sensitivity of the Virgo detector during its second science run". Classical and Quantum Gravity. 28 (2): 025005. arXiv: 1009.5190 . Bibcode:2011CQGra..28b5005A. doi:10.1088/0264-9381/28/2/025005. ISSN   0264-9381. S2CID   118586058.
  69. Estevez, D; Lagabbe, P; Masserot, A; Rolland, L; Seglar-Arroyo, M; Verkindt, D (25 February 2021). "The Advanced Virgo photon calibrators". Classical and Quantum Gravity. 38 (7): 075007. arXiv: 2009.08103 . Bibcode:2021CQGra..38g5007E. doi:10.1088/1361-6382/abe2db. ISSN   0264-9381. S2CID   221761337.
  70. 1 2 Acernese, F; Agathos, M; Ain, A; Albanesi, S; Allocca, A; Amato, A; Andrade, T; Andres, N; Andrić, T; Ansoldi, S; Antier, S; Arène, M; Arnaud, N; Assiduo, M; Astone, P (21 January 2022). "Calibration of advanced Virgo and reconstruction of the detector strain h(t) during the observing run O3". Classical and Quantum Gravity. 39 (4): 045006. arXiv: 2107.03294 . Bibcode:2022CQGra..39d5006A. doi:10.1088/1361-6382/ac3c8e. ISSN   0264-9381. S2CID   238634092.
  71. Estevez, D; Lieunard, B; Marion, F; Mours, B; Rolland, L; Verkindt, D (9 November 2018). "First tests of a Newtonian calibrator on an interferometric gravitational wave detector". Classical and Quantum Gravity. 35 (23): 235009. arXiv: 1806.06572 . Bibcode:2018CQGra..35w5009E. doi:10.1088/1361-6382/aae95f. ISSN   0264-9381. S2CID   119192600.
  72. Acernese, F.; Amico, P.; Alshourbagy, M.; Antonucci, F.; Aoudia, S.; Astone, P.; Avino, S.; Babusci, D.; Ballardin, G.; Barone, F.; Barsotti, L.; Barsuglia, M.; Bauer, Th. S.; Beauville, F.; Bigotta, S. (April 2007). "Data Acquisition System of the Virgo Gravitational Waves Interferometric Detector". 2007 15th IEEE-NPSS Real-Time Conference. pp. 1–8. doi:10.1109/RTC.2007.4382842. ISBN   978-1-4244-0866-5. S2CID   140107498.
  73. Glanzer, J.; Banagiri, S.; Coughlin, S. B.; Soni, S.; Zevin, M.; Berry, C. P. L.; Patane, O.; Bahaadini, S.; Rohani, N.; Crowston, K.; Kalogera, V.; Østerlund, C.; Katsaggelos, A. (16 March 2023). "Data quality up to the third observing run of Advanced LIGO: Gravity Spy glitch classifications". Classical and Quantum Gravity. 40 (6): 065004. arXiv: 2208.12849 . Bibcode:2023CQGra..40f5004G. doi:10.1088/1361-6382/acb633. ISSN   0264-9381. S2CID   251903127.
  74. 1 2 G. Vajente (2008). Analysis of sensitivity and noise sources for the Virgo gravitational wave interferometer (PDF).
  75. "O2 Instrumental Lines". www.gw-openscience.org. Retrieved 24 March 2023.
  76. "Virgo Logbook - Detector Characterisation (Spectral lines)". logbook.virgo-gw.eu. Retrieved 24 March 2023.
  77. Davis, D; Littenberg, T B; Romero-Shaw, I M; Millhouse, M; McIver, J; Di Renzo, F; Ashton, G (15 December 2022). "Subtracting glitches from gravitational-wave detector data during the third LIGO-Virgo observing run". Classical and Quantum Gravity. 39 (24): 245013. arXiv: 2207.03429 . Bibcode:2022CQGra..39x5013D. doi:10.1088/1361-6382/aca238. ISSN   0264-9381. S2CID   250334515.
  78. "Virgo Sensitivity Curves". 2011. Archived from the original on 1 December 2015. Retrieved 15 December 2015.
  79. Chen, Hsin-Yu; Holz, Daniel E; Miller, John; Evans, Matthew; Vitale, Salvatore; Creighton, Jolien (4 March 2021). "Distance measures in gravitational-wave astrophysics and cosmology". Classical and Quantum Gravity. 38 (5): 055010. arXiv: 1709.08079 . Bibcode:2021CQGra..38e5010C. doi:10.1088/1361-6382/abd594. ISSN   0264-9381. S2CID   119057584.
  80. Hello, Patrice (1997). Détection des ondes gravitationnelles - Ecole Joliot Curie [Detection of gravitational waves - Joliot Curie School](PDF) (Report) (in French). Retrieved 20 April 2023.
  81. "Our Collaborations". LIGO Lab | Caltech. Retrieved 26 February 2023.
  82. "LIGO-M1000066-v27: LIGO Data Management Plan". dcc.ligo.org. Retrieved 26 February 2023.
  83. "GWOSC". www.gw-openscience.org. Retrieved 5 March 2023.
  84. The LIGO Scientific Collaboration; the Virgo Collaboration; the KAGRA Collaboration; Abbott, R.; Abe, H.; Acernese, F.; Ackley, K.; Adhicary, S.; Adhikari, N.; Adhikari, R. X.; Adkins, V. K.; Adya, V. B.; Affeldt, C.; Agarwal, D.; Agathos, M. (7 February 2023). "Open Data from the Third Observing Run of LIGO, Virgo, KAGRA, and GEO". The Astrophysical Journal Supplement Series. 267 (2): 29. arXiv: 2302.03676 . Bibcode:2023ApJS..267...29A. doi: 10.3847/1538-4365/acdc9f . S2CID   256627681.
  85. "Astronomers Catch Gravitational Waves from Colliding Neutron Stars". Sky & Telescope. 16 October 2017. Retrieved 20 February 2023.
  86. Watson, Darach; Hansen, Camilla J.; Selsing, Jonatan; Koch, Andreas; Malesani, Daniele B.; Andersen, Anja C.; Fynbo, Johan P. U.; Arcones, Almudena; Bauswein, Andreas; Covino, Stefano; Grado, Aniello; Heintz, Kasper E.; Hunt, Leslie; Kouveliotou, Chryssa; Leloudas, Giorgos (October 2019). "Identification of strontium in the merger of two neutron stars". Nature. 574 (7779): 497–500. arXiv: 1910.10510 . Bibcode:2019Natur.574..497W. doi:10.1038/s41586-019-1676-3. ISSN   1476-4687. PMID   31645733. S2CID   204837882.
  87. Abbott, B. P.; Abbott, R.; Abbott, T. D.; Acernese, F.; Ackley, K.; Adams, C.; Adams, T.; Addesso, P.; Adhikari, R. X.; Adya, V. B.; Affeldt, C.; Afrough, M.; Agarwal, B.; Agathos, M.; Agatsuma, K. (16 October 2017). "Gravitational Waves and Gamma-Rays from a Binary Neutron Star Merger: GW170817 and GRB 170817A". The Astrophysical Journal. 848 (2): L13. arXiv: 1710.05834 . Bibcode:2017ApJ...848L..13A. doi: 10.3847/2041-8213/aa920c . ISSN   2041-8213. S2CID   126310483.
  88. 1 2 LIGO Scientific Collaboration, Virgo Collaboration, and KAGRA Collaboration; Abbott, R.; Abe, H.; Acernese, F.; Ackley, K.; Adhikari, N.; Adhikari, R. X.; Adkins, V. K.; Adya, V. B.; Affeldt, C.; Agarwal, D.; Agathos, M.; Agatsuma, K.; Aggarwal, N.; Aguiar, O. D. (28 November 2022). "All-sky search for continuous gravitational waves from isolated neutron stars using Advanced LIGO and Advanced Virgo O3 data". Physical Review D. 106 (10): 102008. arXiv: 2201.00697 . Bibcode:2022PhRvD.106j2008A. doi:10.1103/PhysRevD.106.102008. hdl:1854/LU-01GXN8M856WCY1YG62A5ACCPTN. S2CID   245650351.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  89. Whelan, John T.; Sundaresan, Santosh; Zhang, Yuanhao; Peiris, Prabath (20 May 2015). "Model-based cross-correlation search for gravitational waves from Scorpius X-1". Physical Review D. 91 (10): 102005. arXiv: 1504.05890 . Bibcode:2015PhRvD..91j2005W. doi:10.1103/PhysRevD.91.102005. S2CID   59360101.
  90. Abbott, R.; Abe, H.; Acernese, F.; Ackley, K.; Adhikari, N.; Adhikari, R. X.; Adkins, V. K.; Adya, V. B.; Affeldt, C.; Agarwal, D.; Agathos, M.; Agatsuma, K.; Aggarwal, N.; Aguiar, O. D.; Aiello, L. (25 May 2022). "Searches for Gravitational Waves from Known Pulsars at Two Harmonics in the Second and Third LIGO-Virgo Observing Runs". The Astrophysical Journal. 935 (1): 1. arXiv: 2111.13106 . Bibcode:2022ApJ...935....1A. doi: 10.3847/1538-4357/ac6acf . ISSN   0004-637X. S2CID   244709285.
  91. "LIGO Scientific Collaboration - The science of LSC research". www.ligo.org. Retrieved 29 March 2023.
  92. LIGO Scientific Collaboration and Virgo Collaboration; Abbott, R.; Abbott, T. D.; Acernese, F.; Ackley, K.; Adams, C.; Adhikari, N.; Adhikari, R. X.; Adya, V. B.; Affeldt, C.; Agarwal, D.; Agathos, M.; Agatsuma, K.; Aggarwal, N.; Aguiar, O. D. (28 April 2022). "Search of the early O3 LIGO data for continuous gravitational waves from the Cassiopeia A and Vela Jr. supernova remnants". Physical Review D. 105 (8): 082005. arXiv: 2111.15116 . Bibcode:2022PhRvD.105h2005A. doi:10.1103/PhysRevD.105.082005. S2CID   244729269.
  93. LIGO Scientific Collaboration, Virgo Collaboration, and KAGRA Collaboration; Abbott, R.; Abe, H.; Acernese, F.; Ackley, K.; Adhikari, N.; Adhikari, R. X.; Adkins, V. K.; Adya, V. B.; Affeldt, C.; Agarwal, D.; Agathos, M.; Agatsuma, K.; Aggarwal, N.; Aguiar, O. D. (9 August 2022). "Search for continuous gravitational wave emission from the Milky Way center in O3 LIGO-Virgo data". Physical Review D. 106 (4): 042003. arXiv: 2204.04523 . Bibcode:2022PhRvD.106d2003A. doi:10.1103/PhysRevD.106.042003. S2CID   248085352.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  94. Abbott, R.; Abbott, T. D.; Abraham, S.; Acernese, F.; Ackley, K.; Adams, A.; Adams, C.; Adhikari, R. X.; Adya, V. B.; Affeldt, C.; Agarwal, D.; Agathos, M.; Agatsuma, K.; Aggarwal, N.; Aguiar, O. D. (23 July 2021). "Upper limits on the isotropic gravitational-wave background from Advanced LIGO and Advanced Virgo's third observing run". Physical Review D. 104 (2): 022004. arXiv: 2101.12130 . Bibcode:2021PhRvD.104b2004A. doi:10.1103/PhysRevD.104.022004. ISSN   2470-0010. S2CID   231719405.
  95. The LIGO Scientific Collaboration; the Virgo Collaboration; the KAGRA Collaboration; Abbott, R.; Abe, H.; Acernese, F.; Ackley, K.; Adhikari, N.; Adhikari, R. X.; Adkins, V. K.; Adya, V. B.; Affeldt, C.; Agarwal, D.; Agathos, M.; Agatsuma, K. (2023). "Constraints on the Cosmic Expansion History from GWTC–3". The Astrophysical Journal. 949 (2): 76. arXiv: 2111.03604 . Bibcode:2023ApJ...949...76A. doi: 10.3847/1538-4357/ac74bb . S2CID   243832919.
  96. 1 2 "Outreach – Virgo". www.virgo-gw.eu. Retrieved 8 May 2023.
  97. "Festa di Scienza e Filosofia" [Science and Philosophy Festival]. Festa di Scienza e Filosofia (in Italian). Retrieved 8 May 2023.
  98. "Gravitional[sic] waves: Windows to the universe". Athens Science Festival. Retrieved 8 May 2023.
  99. Rossi, Giada (23 November 2022). "Black Hole: a new interactive installation by EGO and INFN at Città della Scienza in Naples". EGO - European Gravitational Observatory. Retrieved 8 May 2023.
  100. "Home page". Il Ritmo Dello Spazio (The Rhythm of Space). Retrieved 26 February 2023.
  101. "On Air". Studio Tomás Sarceno. 13 October 2018. Retrieved 26 February 2023.
  102. "Guided Tour". EGO - European Gravitational Observatory. Retrieved 26 February 2023.
  103. "International Day of Women and Girls in Science 2023 – Virgo". www.virgo-gw.eu. Retrieved 26 February 2023.