Void (astronomy)

Last updated
Matter distribution in a cubic section of the universe. The blue fiber structures represent the matter (primarily dark matter) and the empty regions in between represent the cosmic voids. Structure of the Universe.jpg
Matter distribution in a cubic section of the universe. The blue fiber structures represent the matter (primarily dark matter) and the empty regions in between represent the cosmic voids.

Cosmic voids (also known as dark space) are vast spaces between filaments (the largest-scale structures in the universe), which contain very few or no galaxies. Most galaxies are not located in voids, despite their size, due to most galaxies being gravitationally bound together, creating huge cosmic structures known as galaxy filaments. The cosmological evolution of the void regions differs drastically from the evolution of the Universe as a whole: there is a long stage when the curvature term dominates, which prevents the formation of galaxy clusters and massive galaxies. Hence, although even the emptiest regions of voids contain more than ~15% of the average matter density of the Universe, the voids look almost empty to an observer. [1]

Contents

Voids typically have a diameter of 10 to 100 megaparsecs (30 to 300 million light-years); particularly large voids, defined by the absence of rich superclusters, are sometimes called supervoids. They were first discovered in 1978 in a pioneering study by Stephen Gregory and Laird A. Thompson at the Kitt Peak National Observatory. [2]

Voids are believed to have been formed by baryon acoustic oscillations in the Big Bang, collapses of mass followed by implosions of the compressed baryonic matter. Starting from initially small anisotropies from quantum fluctuations in the early universe, the anisotropies grew larger in scale over time. Regions of higher density collapsed more rapidly under gravity, eventually resulting in the large-scale, foam-like structure or "cosmic web" of voids and galaxy filaments seen today. Voids located in high-density environments are smaller than voids situated in low-density spaces of the universe. [3]

Voids appear to correlate with the observed temperature of the cosmic microwave background (CMB) because of the Sachs–Wolfe effect. Colder regions correlate with voids, and hotter regions correlate with filaments because of gravitational redshifting. As the Sachs–Wolfe effect is only significant if the universe is dominated by radiation or dark energy, the existence of voids is significant in providing physical evidence for dark energy. [4] [5]

Large-scale structure

A map of galaxy voids Galaxy superclusters and galaxy voids.png
A map of galaxy voids

The structure of the Universe can be broken down into components that can help describe the characteristics of individual regions of the cosmos. These are the main structural components of the cosmic web:

Voids have a mean density less than a tenth of the average density of the universe. This serves as a working definition even though there is no single agreed-upon definition of what constitutes a void. The matter density value used for describing the cosmic mean density is usually based on a ratio of the number of galaxies per unit volume rather than the total mass of the matter contained in a unit volume. [9]

Discovery

Study of cosmic voids within the discipline of astrophysics began in the mid-1970s when redshift surveys led two separate teams of astrophysicists in 1978 to identify superclusters and voids in the distribution of galaxies and Abell clusters. [10] [11] The new redshift surveys revolutionized the field of astronomy by adding depth to the two-dimensional maps of cosmological structure, which were often densely packed and overlapping, [7] allowing for the first three-dimensional mapping of the universe. Through redshift surveys, their depth was calculated from the individual redshifts of the galaxies due to the expansion of the universe according to Hubble's law. [12]

Timeline

A summarized timeline of important events in the field of cosmic voids from its beginning to recent times is as follows:

Methods for finding

There exist a number of ways for finding voids with the results of large-scale surveys of the universe. Of the many different algorithms, virtually all fall into one of three general categories. [27] The first class consists of void finders that try to find empty regions of space based on local galaxy density. [28] The second class are those which try to find voids via the geometrical structures in the dark matter distribution as suggested by the galaxies. [29] The third class is made up of those finders which identify structures dynamically by using gravitationally unstable points in the distribution of dark matter. [30] The three most popular methods through the study of cosmic voids are listed below:

VoidFinder algorithm

This first-class method uses each galaxy in a catalog as its target and then uses the Nearest Neighbor Approximation to calculate the cosmic density in the region contained in a spherical radius determined by the distance to the third-closest galaxy. [31] El Ad & Piran introduced this method in 1997 to allow a quick and effective method for standardizing the cataloging of voids. Once the spherical cells are mined from all of the structure data, each cell is expanded until the underdensity returns to average expected wall density values. [32] One of the helpful features of void regions is that their boundaries are very distinct and defined, with a cosmic mean density that starts at 10% in the body and quickly rises to 20% at the edge and then to 100% in the walls directly outside the edges. The remaining walls and overlapping void regions are then gridded into, respectively, distinct and intertwining zones of filaments, clusters, and near-empty voids. Any overlap of more than 10% with already known voids are considered to be subregions within those known voids. All voids admitted to the catalog had a minimum radius of 10 Mpc in order to ensure all identified voids were not accidentally cataloged due to sampling errors. [31]

Zone bordering on voidness (ZOBOV) algorithm

This particular second-class algorithm uses a Voronoi tessellation technique and mock border particles in order to categorize regions based on a high-density contrasting border with a very low amount of bias. [33] Neyrinck introduced this algorithm in 2008 with the purpose of introducing a method that did not contain free parameters or presumed shape tessellations. Therefore, this technique can create more accurately shaped and sized void regions. Although this algorithm has some advantages in shape and size, it has been criticized often for sometimes providing loosely defined results. Since it has no free parameters, it mostly finds small and trivial voids, although the algorithm places a statistical significance on each void it finds. A physical significance parameter can be applied in order to reduce the number of trivial voids by including a minimum density to average density ratio of at least 1:5. Subvoids are also identified using this process which raises more philosophical questions on what qualifies as a void. [34] Void finders such as VIDE [35] are based on ZOBOV.

Dynamical void analysis (DIVA) algorithm

This third-class method is drastically different from the previous two algorithms listed. The most striking aspect is that it requires a different definition of what it means to be a void. Instead of the general notion that a void is a region of space with a low cosmic mean density; a hole in the distribution of galaxies, it defines voids to be regions in which matter is escaping; which corresponds to the dark energy equation of state, w. Void centers are then considered to be the maximal source of the displacement field denoted as Sψ. The purpose for this change in definitions was presented by Lavaux and Wandelt in 2009 as a way to yield cosmic voids such that exact analytical calculations can be made on their dynamical and geometrical properties. This allows DIVA to heavily explore the ellipticity of voids and how they evolve in the large-scale structure, subsequently leading to the classification of three distinct types of voids. These three morphological classes are True voids, Pancake voids, and Filament voids. Another notable quality is that even though DIVA also contains selection function bias just as first-class methods do, DIVA is devised such that this bias can be precisely calibrated, leading to much more reliable results. Multiple shortfalls of this Lagrangian-Eulerian hybrid approach exist. One example is that the resulting voids from this method are intrinsically different than those found by other methods, which makes an all-data points inclusive comparison between results of differing algorithms very difficult. [27]

Significance

Voids have contributed significantly to the modern understanding of the cosmos, with applications ranging from shedding light on the current understanding of dark energy, to refining and constraining cosmological evolution models. The Milky Way Galaxy is in a cosmic void named the KBC Void. [36] Some popular applications are mentioned in detail below.

Dark energy

The simultaneous existence of the largest-known voids and galaxy clusters requires about 70% dark energy in the universe today, consistent with the latest data from the cosmic microwave background. [5] Voids act as bubbles in the universe that are sensitive to background cosmological changes. This means that the evolution of a void's shape is in part the result of the expansion of the universe. Since this acceleration is believed to be caused by dark energy, studying the changes of a void's shape over a period of time can be used to constrain the standard ΛCDM model, [37] [38] or further refine the Quintessence + Cold Dark Matter (QCDM) model and provide a more accurate dark energy equation of state. [39] Additionally the abundance of voids is a promising way to constrain the dark energy equation of state. [40] [41]

Neutrinos

Neutrinos, due to their very small mass and extremely weak interaction with other matter, will free-stream in and out of voids which are smaller than the mean-free path of neutrinos. This has an effect on the size and depth distribution of voids, and is expected to make it possible with future astronomical surveys (e.g. the Euclid satellite) to measure the sum of the masses of all neutrino species by comparing the statistical properties of void samples to theoretical predictions. [41]

Galactic formation and evolution models

A 43x43x43-megaparsec cube shows the evolution of the large-scale structure over a logarithmic period starting from a redshift of 30 and ending at redshift 0. The model makes it clear to see how the matter-dense regions contract under the collective gravitational force while simultaneously aiding in the expansion of cosmic voids as the matter flees to the walls and filaments. Large-scale structure formation.gif
A 43×43×43-megaparsec cube shows the evolution of the large-scale structure over a logarithmic period starting from a redshift of 30 and ending at redshift 0. The model makes it clear to see how the matter-dense regions contract under the collective gravitational force while simultaneously aiding in the expansion of cosmic voids as the matter flees to the walls and filaments.

Cosmic voids contain a mix of galaxies and matter that is slightly different than other regions in the universe. This unique mix supports the biased galaxy formation picture predicted in Gaussian adiabatic cold dark matter models. This phenomenon provides an opportunity to modify the morphology-density correlation that holds discrepancies with these voids. Such observations like the morphology-density correlation can help uncover new facets about how galaxies form and evolve on the large scale. [42] On a more local scale, galaxies that reside in voids have differing morphological and spectral properties than those that are located in the walls. One feature that has been found is that voids have been shown to contain a significantly higher fraction of starburst galaxies of young, hot stars when compared to samples of galaxies in walls. [43]

Voids offer opportunities to study the strength of intergalactic magnetic fields. For example, a 2015 study concluded, based on the deflection of blazar gamma-ray emissions that travel through voids, that intergalactic space contains a magnetic field of strength at least 10-17 G. The specific large-scale magnetic structure of the universe suggests primordial "magnetogenesis", which in turn could have played a role in the formation of magnetic fields within galaxies, and could also change estimates of the timeline of recombination in the early universe. [44] [45]

Anomalies in anisotropies

Cold spots in the cosmic microwave background, such as the WMAP cold spot found by Wilkinson Microwave Anisotropy Probe, could possibly be explained by an extremely large cosmic void that has a radius of ~120 Mpc, as long as the late integrated Sachs–Wolfe effect was accounted for in the possible solution. Anomalies in CMB screenings are now being potentially explained through the existence of large voids located down the line-of-sight in which the cold spots lie. [46]

CMB screening of the universe WMAP white.png
CMB screening of the universe

Expansion

Although dark energy is currently the most popular explanation for the acceleration in the expansion of the universe, another theory elaborates on the possibility of our galaxy being part of a very large, not-so-underdense, cosmic void. According to this theory, such an environment could naively lead to the demand for dark energy to solve the problem with the observed acceleration. As more data has been released on this topic the chances of it being a realistic solution in place of the current ΛCDM interpretation has been largely diminished but not all together abandoned. [47]

Gravitational theories

The abundance of voids, particularly when combined with the abundance of clusters of galaxies, is a promising method for precision tests of deviations from general relativity on large scales and in low-density regions. [48] [49]

The insides of voids often seem to adhere to cosmological parameters which differ from those of the known universe[ citation needed ]. It is because of this unique feature that cosmic voids are useful laboratories to study the effects that gravitational clustering and growth rates have on local galaxies and structure when the cosmological parameters have different values from the outside universe. Due to the observation that larger voids predominantly remain in a linear regime, with most structures within exhibiting spherical symmetry in the underdense environment; that is, the underdensity leads to near-negligible particle-particle gravitational interactions that would otherwise occur in a region of normal galactic density. Testing models for voids can be performed with very high accuracy. The cosmological parameters that differ in these voids are Ωm, ΩΛ, and H0. [50]

See also

Related Research Articles

<span class="mw-page-title-main">Copernican principle</span> Principle that humans are not privileged observers of the universe

In physical cosmology, the Copernican principle states that humans, on the Earth or in the Solar System, are not privileged observers of the universe, that observations from the Earth are representative of observations from the average position in the universe. Named for Copernican heliocentrism, it is a working assumption that arises from a modified cosmological extension of Copernicus' argument of a moving Earth.

In astronomy, dark matter is a hypothetical form of matter that appears not to interact with light or the electromagnetic field. Dark matter is implied by gravitational effects which cannot be explained by general relativity unless more matter is present than can be seen. Such effects occur in the context of formation and evolution of galaxies, gravitational lensing, the observable universe's current structure, mass position in galactic collisions, the motion of galaxies within galaxy clusters, and cosmic microwave background anisotropies.

<span class="mw-page-title-main">Supercluster</span> Large group of smaller galaxy clusters or galaxy groups

A supercluster is a large group of smaller galaxy clusters or galaxy groups; they are among the largest known structures in the universe. The Milky Way is part of the Local Group galaxy group, which in turn is part of the Virgo Supercluster, which is part of the Laniakea Supercluster, which is part of the Pisces–Cetus Supercluster Complex. The large size and low density of superclusters means that they, unlike clusters, expand with the Hubble expansion. The number of superclusters in the observable universe is estimated to be 10 million.

<span class="mw-page-title-main">Virgo Supercluster</span> Galactic supercluster containing the Virgo Cluster

The Virgo Supercluster or the Local Supercluster is a mass concentration of galaxies containing the Virgo Cluster and Local Group, which itself contains the Milky Way and Andromeda galaxies, as well as others. At least 100 galaxy groups and clusters are located within its diameter of 33 megaparsecs. The Virgo SC is one of about 10 million superclusters in the observable universe and is in the Pisces–Cetus Supercluster Complex, a galaxy filament.

<span class="mw-page-title-main">Great Attractor</span> Region of overdensity of galaxies within the local supercluster

The Great Attractor is a region of gravitational attraction in intergalactic space and the apparent central gravitational point of the Laniakea Supercluster of galaxies that includes the Milky Way galaxy, as well as about 100,000 other galaxies.

<span class="mw-page-title-main">Observable universe</span> All of space observable from the Earth at the present

The observable universe is a ball-shaped region of the universe comprising all matter that can be observed from Earth or its space-based telescopes and exploratory probes at the present time; the electromagnetic radiation from these objects has had time to reach the Solar System and Earth since the beginning of the cosmological expansion. Initially, it was estimated that there may be 2 trillion galaxies in the observable universe. That number was reduced in 2021 to only several hundred billion based on data from New Horizons. Assuming the universe is isotropic, the distance to the edge of the observable universe is roughly the same in every direction. That is, the observable universe is a spherical region centered on the observer. Every location in the universe has its own observable universe, which may or may not overlap with the one centered on Earth.

The Sunyaev–Zeldovich effect is the spectral distortion of the cosmic microwave background (CMB) through inverse Compton scattering by high-energy electrons in galaxy clusters, in which the low-energy CMB photons receive an average energy boost during collision with the high-energy cluster electrons. Observed distortions of the cosmic microwave background spectrum are used to detect the disturbance of density in the universe. Using the Sunyaev–Zeldovich effect, dense clusters of galaxies have been observed.

The Sachs–Wolfe effect, named after Rainer K. Sachs and Arthur M. Wolfe, is a property of the cosmic microwave background radiation (CMB), in which photons from the CMB are gravitationally redshifted, causing the CMB spectrum to appear uneven. This effect is the predominant source of fluctuations in the CMB for angular scales larger than about ten degrees.

<span class="mw-page-title-main">Reionization</span> Process that caused matter to reionize early in the history of the Universe

In the fields of Big Bang theory and cosmology, reionization is the process that caused electrically neutral atoms in the universe to reionize after the lapse of the "dark ages".

The Lambda-CDM, Lambda cold dark matter, or ΛCDM model is a mathematical model of the Big Bang theory with three major components:

  1. a cosmological constant denoted by lambda (Λ) associated with dark energy
  2. the postulated cold dark matter
  3. ordinary matter

Redshift quantization, also referred to as redshift periodicity, redshift discretization, preferred redshifts and redshift-magnitude bands, is the hypothesis that the redshifts of cosmologically distant objects tend to cluster around multiples of some particular value.

<span class="mw-page-title-main">CMB cold spot</span> Region in space

The CMB Cold Spot or WMAP Cold Spot is a region of the sky seen in microwaves that has been found to be unusually large and cold relative to the expected properties of the cosmic microwave background radiation (CMBR). The "Cold Spot" is approximately 70 µK (0.00007 K) colder than the average CMB temperature, whereas the root mean square of typical temperature variations is only 18 µK. At some points, the "cold spot" is 140 µK colder than the average CMB temperature.

<span class="mw-page-title-main">Galaxy filament</span> Largest structures in the universe, made of galaxies

In cosmology, galaxy filaments are the largest known structures in the universe, consisting of walls of galactic superclusters. These massive, thread-like formations can commonly reach 50/h to 80/h Megaparsecs — with the largest found to date being the Hercules-Corona Borealis Great Wall at around 3 gigaparsecs (9.8 Gly) in length — and form the boundaries between voids. Due to the accelerating expansion of the universe, the individual clusters of gravitationally bound galaxies that make up galaxy filaments are moving away from each other at an accelerated rate; in the far future they will dissolve.

<span class="mw-page-title-main">Weak gravitational lensing</span>

While the presence of any mass bends the path of light passing near it, this effect rarely produces the giant arcs and multiple images associated with strong gravitational lensing. Most lines of sight in the universe are thoroughly in the weak lensing regime, in which the deflection is impossible to detect in a single background source. However, even in these cases, the presence of the foreground mass can be detected, by way of a systematic alignment of background sources around the lensing mass. Weak gravitational lensing is thus an intrinsically statistical measurement, but it provides a way to measure the masses of astronomical objects without requiring assumptions about their composition or dynamical state.

In cosmology, baryon acoustic oscillations (BAO) are fluctuations in the density of the visible baryonic matter of the universe, caused by acoustic density waves in the primordial plasma of the early universe. In the same way that supernovae provide a "standard candle" for astronomical observations, BAO matter clustering provides a "standard ruler" for length scale in cosmology. The length of this standard ruler is given by the maximum distance the acoustic waves could travel in the primordial plasma before the plasma cooled to the point where it became neutral atoms, which stopped the expansion of the plasma density waves, "freezing" them into place. The length of this standard ruler can be measured by looking at the large scale structure of matter using astronomical surveys. BAO measurements help cosmologists understand more about the nature of dark energy by constraining cosmological parameters.

<span class="mw-page-title-main">Hubble bubble (astronomy)</span> Variation in the Hubble constant

In astronomy, a Hubble bubble would be "a departure of the local value of the Hubble constant from its globally averaged value," or, more technically, "a local monopole in the peculiar velocity field, perhaps caused by a local void in the mass density."

<span class="mw-page-title-main">Huge-LQG</span> Possible astronomical structure

The Huge Large Quasar Group, is a possible structure or pseudo-structure of 73 quasars, referred to as a large quasar group, that measures about 4 billion light-years across. At its discovery, it was identified as the largest and the most massive known structure in the observable universe, though it has been superseded by the Hercules–Corona Borealis Great Wall at 10 billion light-years. There are also issues about its structure.

In cosmology, intensity mapping is an observational technique for surveying the large-scale structure of the universe by using the integrated radio emission from unresolved gas clouds.

References

  1. Baushev, A. N. (2021). "The central region of a void: an analytical solution". Monthly Notices of the Royal Astronomical Society: Letters . 504 (1): L56–L60. arXiv: 2104.01359 . Bibcode:2021MNRAS.504L..56B. doi:10.1093/mnrasl/slab036.
  2. Freedman, R. A., & Kaufmann III, W. J. (2008). Stars and galaxies: Universe. New York City: W.H. Freeman and Company.
  3. U. Lindner; J. Einasto; M. Einasto; W. Freudling; K. Fricke; E. Tago (1995). "The structure of supervoids. I. Void hierarchy in the Northern Local Supervoid". Astron. Astrophys. 301: 329. arXiv: astro-ph/9503044 . Bibcode:1995A&A...301..329L.
  4. Granett, B. R.; Neyrinck, M. C.; Szapudi, I. (2008). "An Imprint of Superstructures on the Microwave Background due to the Integrated Sachs-Wolfe Effect". Astrophysical Journal . 683 (2): L99–L102. arXiv: 0805.3695 . Bibcode:2008ApJ...683L..99G. doi:10.1086/591670. S2CID   15976818.
  5. 1 2 Sahlén, Martin; Zubeldía, Íñigo; Silk, Joseph (2016). "Cluster–Void Degeneracy Breaking: Dark Energy, Planck, and the Largest Cluster and Void". The Astrophysical Journal Letters. 820 (1): L7. arXiv: 1511.04075 . Bibcode:2016ApJ...820L...7S. doi: 10.3847/2041-8205/820/1/L7 . ISSN   2041-8205. S2CID   119286482.
  6. Ryden, Barbara Sue; Peterson, Bradley M. (2010-01-01). Foundations of Astrophysics (International ed.). Addison-Wesley. p. 522. ISBN   9780321595584.
  7. 1 2 Carroll, Bradley W.; Ostlie, Dale A. (2013-07-23). An Introduction to Modern Astrophysics (International ed.). Pearson. p. 1171. ISBN   9781292022932.
  8. Pan, Danny C.; Michael S. Vogeley; Fiona Hoyle; Yun-Young Choi; Changbom Park (23 Mar 2011). "Cosmic Voids in Sloan Digital Sky Survey Data Release 7". Monthly Notices of the Royal Astronomical Society. 421 (2): 926–934. arXiv: 1103.4156 . Bibcode:2012MNRAS.421..926P. doi:10.1111/j.1365-2966.2011.20197.x. S2CID   119182772.
  9. Neyrinck, Mark C. (29 Feb 2008). "ZOBOV: a parameter-free void-finding algorithm". Monthly Notices of the Royal Astronomical Society. 386 (4): 2101–2109. arXiv: 0712.3049 . Bibcode:2008MNRAS.386.2101N. doi:10.1111/j.1365-2966.2008.13180.x. S2CID   5670329.
  10. 1 2 Gregory, S. A.; Thompson, L. A. (1978). "The Coma/A1367 supercluster and its environs". The Astrophysical Journal. 222: 784. Bibcode:1978ApJ...222..784G. doi:10.1086/156198. ISSN   0004-637X.
  11. Jõeveer, M.; Einasto, J. (1978). Longair, M. S.; Einasto, J. (eds.). The Large Scale Structure of the Universe. Dordrecht: Reidel. p. 241.{{cite book}}: CS1 maint: location missing publisher (link)
  12. Rex, Andrew F.; Bennett, Jeffrey O.; Donahue, Megan; Schneider, Nicholas; Voit, Mark (1998-12-01). The Cosmic Perspective. Pearson College Division. p. 602. ISBN   978-0-201-47399-5 . Retrieved 4 May 2014.
  13. Abell, George O. (1961). "Evidence regarding second-order clustering of galaxies and interactions between clusters of galaxies". The Astronomical Journal. 66: 607. Bibcode:1961AJ.....66..607A. doi:10.1086/108472. ISSN   0004-6256.
  14. Joeveer, Einasto and Tago 1978, Dordrecht, N/A, 241.
  15. Kirshner, R. P.; Oemler, A. Jr.; Schechter, P. L.; Shectman, S. A. (1981). "A million cubic megaparsec void in Bootes". The Astrophysical Journal. 248: L57. Bibcode:1981ApJ...248L..57K. doi:10.1086/183623. ISSN   0004-637X.
  16. Kirshner, Robert P.; Oemler, Augustus Jr.; Schechter, Paul L.; Shectman, Stephen A. (1987). "A survey of the Bootes void". The Astrophysical Journal. 314: 493. Bibcode:1987ApJ...314..493K. doi:10.1086/165080. ISSN   0004-637X. S2CID   118385803.
  17. Merlott, A. L. (November 1983). "Clustering velocities in the adiabatic picture of galaxy formation". Monthly Notices of the Royal Astronomical Society. 205 (3): 637–641. Bibcode:1983MNRAS.205..637M. doi: 10.1093/mnras/205.3.637 . ISSN   0035-8711.
  18. Frenk, C. S.; White, S. D. M.; Davis, M. (1983). "Nonlinear evolution of large-scale structure in the universe". The Astrophysical Journal. 271: 417. Bibcode:1983ApJ...271..417F. doi:10.1086/161209. ISSN   0004-637X.
  19. Giovanelli, R.; Haynes, M. P. (1985). "A 21 CM survey of the Pisces-Perseus supercluster. I – The declination zone +27.5 to +33.5 degrees". The Astronomical Journal. 90: 2445. Bibcode:1985AJ.....90.2445G. doi: 10.1086/113949 . ISSN   0004-6256.
  20. Geller, M. J.; Huchra, J. P. (1989). "Mapping the Universe". Science. 246 (4932): 897–903. Bibcode:1989Sci...246..897G. doi:10.1126/science.246.4932.897. ISSN   0036-8075. PMID   17812575. S2CID   31328798.
  21. Kirshner, 1991, Physical Cosmology, 2, 595.
  22. Fisher, Karl; Huchra, John; Strauss, Michael; Davis, Marc; Yahil, Amos; Schlegel, David (1995). "The IRAS 1.2 Jy Survey: Redshift Data". The Astrophysical Journal Supplement Series. 100: 69. arXiv: astro-ph/9502101 . Bibcode:1995ApJS..100...69F. doi:10.1086/192208. S2CID   13605316.
  23. Colless, Matthew; Dalton, G. B.; Maddox, S. J.; Sutherland, W. J.; Norberg, P.; Cole, S.; Bland-Hawthorn, J.; Bridges, T. J.; Cannon, R. D.; Collins, C. A.; J Couch, W.; Cross, N. G. J.; Deeley, K.; DePropris, R.; Driver, S. P.; Efstathiou, G.; Ellis, R. S.; Frenk, C. S.; Glazebrook, K.; Jackson, C. A.; Lahav, O.; Lewis, I. J.; Lumsden, S. L.; Madgwick, D. S.; Peacock, J. A.; Peterson, B. A.; Price, I. A.; Seaborne, M.; Taylor, K. (2001). "The 2dF Galaxy Redshift Survey: Spectra and redshifts". Monthly Notices of the Royal Astronomical Society. 328 (4): 1039–1063. arXiv: astro-ph/0106498 . Bibcode:2001MNRAS.328.1039C. doi:10.1046/j.1365-8711.2001.04902.x. S2CID   40393799.
  24. Abazajian, K.; for the Sloan Digital Sky Survey; Agüeros, Marcel A.; Allam, Sahar S.; Prieto, Carlos Allende; An, Deokkeun; Anderson, Kurt S. J.; Anderson, Scott F.; Annis, James; Bahcall, Neta A.; Bailer-Jones, C. A. L.; Barentine, J. C.; Bassett, Bruce A.; Becker, Andrew C.; Beers, Timothy C.; Bell, Eric F.; Belokurov, Vasily; Berlind, Andreas A.; Berman, Eileen F.; Bernardi, Mariangela; Bickerton, Steven J.; Bizyaev, Dmitry; Blakeslee, John P.; Blanton, Michael R.; Bochanski, John J.; Boroski, William N.; Brewington, Howard J.; Brinchmann, Jarle; Brinkmann, J.; et al. (2009). "The Seventh Data Release of the Sloan Digital Sky Survey". The Astrophysical Journal Supplement Series. 182 (2): 543–558. arXiv: 0812.0649 . Bibcode:2009ApJS..182..543A. doi:10.1088/0067-0049/182/2/543. S2CID   14376651.
  25. Thompson, Laird A.; Gregory, Stephen A. (2011). "An Historical View: The Discovery of Voids in the Galaxy Distribution". arXiv: 1109.1268 [physics.hist-ph].
  26. Mao, Qingqing; Berlind, Andreas A.; Scherrer, Robert J.; Neyrinck, Mark C.; Scoccimarro, Román; Tinker, Jeremy L.; McBride, Cameron K.; Schneider, Donald P.; Pan, Kaike (2017). "A Cosmic Void Catalog of SDSS DR12 BOSS Galaxies". The Astrophysical Journal. 835 (2): 161. arXiv: 1602.02771 . Bibcode:2017ApJ...835..161M. doi: 10.3847/1538-4357/835/2/161 . ISSN   0004-637X. S2CID   119098071.
  27. 1 2 Lavaux, Guilhem; Wandelt, Benjamin D. (2010). "Precision cosmology with voids: Definition, methods, dynamics". Monthly Notices of the Royal Astronomical Society. 403 (3): 403–1408. arXiv: 0906.4101 . Bibcode:2010MNRAS.403.1392L. doi:10.1111/j.1365-2966.2010.16197.x. S2CID   15294193.
  28. Hoyle, Fiona; Vogeley, Michael S. (2002). "Voids in the PSCz Survey and the Updated Zwicky Catalog". The Astrophysical Journal. 566 (2): 641–651. arXiv: astro-ph/0109357 . Bibcode:2002ApJ...566..641H. doi:10.1086/338340. S2CID   5822042.
  29. Colberg, Joerg M.; Sheth, Ravi K.; Diaferio, Antonaldo; Gao, Liang; Yoshida, Naoki (2005). "Voids in a [Lambda] CDM Universe". Monthly Notices of the Royal Astronomical Society. 360 (1): 216–226. arXiv: astro-ph/0409162v2 . Bibcode:2005MNRAS.360..216C. doi:10.1111/j.1365-2966.2005.09064.x. S2CID   18912038.
  30. Hahn, Oliver; Porciani, Cristiano; Marcella Carollo, C.; Dekel, Avishai (2007). "Properties of Dark Matter Haloes in Clusters, Filaments, Sheets and Voids". Monthly Notices of the Royal Astronomical Society. 375 (2): 489–499. arXiv: astro-ph/0610280 . Bibcode:2007MNRAS.375..489H. doi:10.1111/j.1365-2966.2006.11318.x. S2CID   14225529.
  31. 1 2 Pan, Danny C.; Vogeley, Michael S.; Hoyle, Fiona; Choi, Yun-Young; Park, Changbom (2011). "Cosmic Voids in Sloan Digital Sky Survey Data Release 7". Monthly Notices of the Royal Astronomical Society. 421 (2): 926–934. arXiv: 1103.4156 . Bibcode:2012MNRAS.421..926P. doi:10.1111/j.1365-2966.2011.20197.x. S2CID   119182772.
  32. El-Ad, Hagai; Piran, Tsvi (1997). "Voids in the Large-Scale Structure". The Astrophysical Journal. 491 (2): 421–435. arXiv: astro-ph/9702135 . Bibcode:1997ApJ...491..421E. doi:10.1086/304973. S2CID   16336543.
  33. Sutter, P. M.; Lavaux, Guilhem; Wandelt, Benjamin D.; Weinberg, David H. (2013). "A response to arXiv:1310.2791: A self-consistent public catalogue of voids and superclusters in the SDSS Data Release 7 galaxy surveys". arXiv: 1310.5067 [astro-ph.CO].
  34. Neyrinck, Mark C. (2008). "ZOBOV: A parameter-free void-finding algorithm". Monthly Notices of the Royal Astronomical Society. 386 (4): 2101–2109. arXiv: 0712.3049 . Bibcode:2008MNRAS.386.2101N. doi:10.1111/j.1365-2966.2008.13180.x. S2CID   5670329.
  35. Sutter, P. M. (2015). "VIDE: The Void IDentification and Examination toolkit". Astronomy and Computing. 9: 1–9. arXiv: 1406.1191 . Bibcode:2015A&C.....9....1S. doi:10.1016/j.ascom.2014.10.002. S2CID   62620511.
  36. Howell, Elizabeth (2017-06-14). "We Live in a Cosmic Void, Another Study Confirms". Space.com. Retrieved 2023-11-26.
  37. Lavaux, Guilhem; Wandelt, Benjamin D. (1 August 2012). "Precision Cosmography with Stacked Voids". The Astrophysical Journal. 754 (2): 109. arXiv: 1110.0345 . Bibcode:2012ApJ...754..109L. doi: 10.1088/0004-637X/754/2/109 .
  38. Mao, Qingqing; Berlind, Andreas A.; Scherrer, Robert J.; Neyrinck, Mark C.; Scoccimarro, Román; Tinker, Jeremy L.; McBride, Cameron K.; Schneider, Donald P. (25 January 2017). "Cosmic Voids in the SDSS DR12 BOSS Galaxy Sample: the Alcock–Paczyński test". The Astrophysical Journal. 835 (2): 160. arXiv: 1602.06306 . Bibcode:2017ApJ...835..160M. doi: 10.3847/1538-4357/835/2/160 . S2CID   119276823.
  39. Lee, Jounghun; Park, Daeseong (2007). "Constraining the Dark Energy Equation of State with Cosmic Voids". The Astrophysical Journal. 696 (1): L10–L12. arXiv: 0704.0881 . Bibcode:2009ApJ...696L..10L. doi:10.1088/0004-637X/696/1/L10. S2CID   18219268.
  40. Pisani, Alice; Sutter, P. M.; Hamaus, Nico; Alizadeh, Esfandiar; Biswas, Rahul; Wandelt, Benjamin D.; Hirata, Christopher M. (2015). "Counting voids to probe dark energy". Physical Review D. 92 (8): 083531. arXiv: 1503.07690 . Bibcode:2015PhRvD..92h3531P. doi:10.1103/PhysRevD.92.083531. S2CID   119253930.
  41. 1 2 Sahlén, Martin (2019-03-22). "Cluster-void degeneracy breaking: Neutrino properties and dark energy". Physical Review D. 99 (6): 063525. arXiv: 1807.02470 . Bibcode:2019PhRvD..99f3525S. doi:10.1103/PhysRevD.99.063525. ISSN   2470-0010. S2CID   85530907.
  42. Peebles, P. J. E. (2001). "The Void Phenomenon". The Astrophysical Journal. 557 (2): 495–504. arXiv: astro-ph/0101127 . Bibcode:2001ApJ...557..495P. doi:10.1086/322254. S2CID   2138259.
  43. Constantin, Anca; Hoyle, Fiona; Vogeley, Michael S. (2007). "Active Galactic Nuclei in Void Regions". The Astrophysical Journal. 673 (2): 715–729. arXiv: 0710.1631 . Bibcode:2008ApJ...673..715C. doi:10.1086/524310. S2CID   15383038.
  44. Wolchover, Natalie (2 July 2020). "The Hidden Magnetic Universe Begins to Come Into View". Quanta Magazine. Retrieved 7 July 2020.
  45. Chen, Wenlei; Buckley, James H.; Ferrer, Francesc (16 November 2015). "Search for GeV γ-Ray Pair Halos Around Low Redshift Blazars". Physical Review Letters. 115 (21): 211103. arXiv: 1410.7717 . Bibcode:2015PhRvL.115u1103C. doi: 10.1103/PhysRevLett.115.211103 . PMID   26636838. S2CID   32638647.
  46. Rudnick, Lawrence; Brown, Shea; Williams, Liliya R. (2007). "Extragalactic Radio Sources and the WMAP Cold Spot". The Astrophysical Journal. 671 (1): 40–44. arXiv: 0704.0908 . Bibcode:2007ApJ...671...40R. doi:10.1086/522222. S2CID   14316362.
  47. Alexander, Stephon; Biswas, Tirthabir; Notari, Alessio; Vaid, Deepak (2009). "Local Void vs Dark Energy: Confrontation with WMAP and Type Ia Supernovae". Journal of Cosmology and Astroparticle Physics. 2009 (9): 025. arXiv: 0712.0370 . Bibcode:2009JCAP...09..025A. doi:10.1088/1475-7516/2009/09/025. S2CID   119259755.
  48. Sahlén, Martin; Silk, Joseph (2018-05-03). "Cluster-void degeneracy breaking: Modified gravity in the balance". Physical Review D. 97 (10): 103504. arXiv: 1612.06595 . Bibcode:2018PhRvD..97j3504S. doi:10.1103/PhysRevD.97.103504. S2CID   73621033.
  49. Nan, Yue; Yamamoto, Kazuhiro (2018-08-28). "Gravitational redshift in the void-galaxy cross-correlation function in redshift space". Physical Review D. 98 (4): 043527. arXiv: 1805.05708 . Bibcode:2018PhRvD..98d3527N. doi:10.1103/PhysRevD.98.043527. S2CID   119351761.
  50. Goldberg, David M.; Vogeley, Michael S. (2004). "Simulating Voids". The Astrophysical Journal. 605 (1): 1–6. arXiv: astro-ph/0307191 . Bibcode:2004ApJ...605....1G. doi:10.1086/382143. S2CID   13242401.

Further reading