ART

In cosmology, the cosmological constant (usually denoted by the Greek capital letter lambda: Λ) is the energy density of space, or vacuum energy, that arises in Albert Einstein's field equations of general relativity. It is closely associated to the concepts of dark energy and quintessence.[1]

Einstein originally introduced the concept in 1917[2] to counterbalance the effects of gravity and achieve a static universe, a notion which was the accepted view at the time. Einstein abandoned the concept in 1931 after Hubble's confirmation of the expanding universe.[3] From the 1930s until the late 1990s, most physicists assumed the cosmological constant to be equal to zero.[4] That changed with the surprising discovery in 1998 that the expansion of the universe is accelerating, implying the possibility of a positive nonzero value for the cosmological constant.[5]

Since the 1990s, studies have shown that around 68% of the mass–energy density of the universe can be attributed to so-called dark energy.[6] The cosmological constant Λ is the simplest possible explanation for dark energy, and is used in the current standard model of cosmology known as the ΛCDM model.

According to quantum field theory (QFT) which underlies modern particle physics, empty space is defined by the vacuum state which is a collection of quantum fields. All these quantum fields exhibit fluctuations in their ground state (lowest energy density) arising from the zero-point energy present everywhere in space. These zero-point fluctuations should act as a contribution to the cosmological constant Λ, but when calculations are performed these fluctuations give rise to an enormous vacuum energy.[7] The discrepancy between theorized vacuum energy from quantum field theory and observed vacuum energy from cosmology is a source of major contention, with the values predicted exceeding observation by some 120 orders of magnitude, a discrepancy that has been called "the worst theoretical prediction in the history of physics!".[8] This issue is called the cosmological constant problem and it is one of the greatest mysteries in science with many physicists believing that "the vacuum holds the key to a full understanding of nature".[9]

History

Einstein included the cosmological constant as a term in his field equations for general relativity because he was dissatisfied that otherwise his equations did not allow, apparently, for a static universe: gravity would cause a universe that was initially at dynamic equilibrium to contract. To counteract this possibility, Einstein added the cosmological constant.[3] However, soon after Einstein developed his static theory, observations by Edwin Hubble indicated that the universe appears to be expanding; this was consistent with a cosmological solution to the original general relativity equations that had been found by the mathematician Friedmann, working on the Einstein equations of general relativity. Einstein reportedly referred to his failure to accept the validation of his equations—when they had predicted the expansion of the universe in theory, before it was demonstrated in observation of the cosmological redshift—as his "biggest blunder".[10]

In fact, adding the cosmological constant to Einstein's equations does not lead to a static universe at equilibrium because the equilibrium is unstable: if the universe expands slightly, then the expansion releases vacuum energy, which causes yet more expansion. Likewise, a universe that contracts slightly will continue contracting.[11]

However, the cosmological constant remained a subject of theoretical and empirical interest. Empirically, the onslaught of cosmological data in the past decades strongly suggests that our universe has a positive cosmological constant.[5] The explanation of this small but positive value is an outstanding theoretical challenge, the so-called cosmological constant problem.

Some early generalizations of Einstein's gravitational theory, known as classical unified field theories, either introduced a cosmological constant on theoretical grounds or found that it arose naturally from the mathematics. For example, Sir Arthur Stanley Eddington claimed that the cosmological constant version of the vacuum field equation expressed the "epistemological" property that the universe is "self-gauging", and Erwin Schrödinger's pure-affine theory using a simple variational principle produced the field equation with a cosmological term.
Equation
Estimated ratios of dark matter and dark energy (which may be the cosmological constant[1]) in the universe. According to current theories of physics, dark energy now dominates as the largest source of energy of the universe, in contrast to earlier epochs when it was insignificant.

The cosmological constant \( \Lambda \) appears in Einstein's field equation in the form

\( {\displaystyle R_{\mu \nu }-{\tfrac {1}{2}}Rg_{\mu \nu }+\Lambda g_{\mu \nu }={8\pi G \over c^{4}}T_{\mu \nu },}} \)

where the Ricci tensor/scalar R and the metric tensor g describe the structure of spacetime, the stress–energy tensor T describes the energy and momentum density and flux of the matter in that point in spacetime, and the universal constants G and c are conversion factors that arise from using traditional units of measurement. When Λ is zero, this reduces to the field equation of general relativity usually used in the mid-20th century. When T is zero, the field equation describes empty space (the vacuum).

The cosmological constant has the same effect as an intrinsic energy density of the vacuum, ρvac (and an associated pressure). In this context, it is commonly moved onto the right-hand side of the equation, and defined with a proportionality factor of 8π: Λ = 8πρvac, where unit conventions of general relativity are used (otherwise factors of G and c would also appear, i.e. Λ = 8π(G/c2)ρvac = κρvac, where κ is the Einstein gravitational constant). It is common to quote values of energy density directly, though still using the name "cosmological constant", with convention 8πG = 1. The true dimension of Λ is a length−2.

Given the Planck (2018) values of ΩΛ = 0.6889±0.0056 and H0 = 67.66±0.42 (km/s)/Mpc = (2.1927664±0.0136)×10−18 s−1, Λ has the value of

\( {\displaystyle {\begin{aligned}\Lambda =3\left({\frac {H_{0}}{c}}\right)^{2}\Omega _{\Lambda }&=1.1056\times 10^{-52}\ {\text{m}}^{-2}\\&=2.888\times 10^{-122}\ l_{\text{P}}^{-2}\end{aligned}}} \)

where \( {\displaystyle l_{\text{P}}} \) is the Planck length. A positive vacuum energy density resulting from a cosmological constant implies a negative pressure, and vice versa. If the energy density is positive, the associated negative pressure will drive an accelerated expansion of the universe, as observed. (See dark energy and cosmic inflation for details.)
ΩΛ (Omega Lambda)

Instead of the cosmological constant itself, cosmologists often refer to the ratio between the energy density due to the cosmological constant and the critical density of the universe, the tipping point for a sufficient density to stop the universe from expanding forever. This ratio is usually denoted ΩΛ, and is estimated to be 0.6889±0.0056, according to results published by the Planck Collaboration in 2018.[12]

In a flat universe, ΩΛ is the fraction of the energy of the universe due to the cosmological constant, i.e., what we would intuitively call the fraction of the universe that is made up of dark energy. Note that this value changes over time: the critical density changes with cosmological time, but the energy density due to the cosmological constant remains unchanged throughout the history of the universe: the amount of dark energy increases as the universe grows, while the amount of matter does not.
Equation of state

Another ratio that is used by scientists is the equation of state, usually denoted w, which is the ratio of pressure that dark energy puts on the universe to the energy per unit volume.[13] This ratio is w = −1 for a true cosmological constant, and is generally different for alternative time-varying forms of vacuum energy such as quintessence. The Planck Collaboration (2018) has measured w = −1.028±0.032, consistent with −1, assuming no evolution in w over cosmic time.

Positive value
Lambda-CDM, accelerated expansion of the universe. The time-line in this schematic diagram extends from the Big Bang/inflation era 13.7 Byr ago to the present cosmological time.

Observations announced in 1998 of distance–redshift relation for Type Ia supernovae[5] indicated that the expansion of the universe is accelerating. When combined with measurements of the cosmic microwave background radiation these implied a value of ΩΛ ≈ 0.7,[14] a result which has been supported and refined by more recent measurements.[15] There are other possible causes of an accelerating universe, such as quintessence, but the cosmological constant is in most respects the simplest solution. Thus, the current standard model of cosmology, the Lambda-CDM model, includes the cosmological constant, which is measured to be on the order of 10−52 m−2, in metric units. It is often expressed as 10−35 s−2 (by multiplication with c2, i.e. ≈1017 m2⋅s−2) or as 10−122[16] (by multiplication with square Planck length, i.e. ≈10−70 m2). The value is based on recent measurements of vacuum energy density, ρ vacuum = \( {\displaystyle \rho _{\text{vacuum}}=5.96\times 10^{-27}{\text{ kg/m}}^{3}} \) .[17]

As was only recently seen, by works of 't Hooft, Susskind and others, a positive cosmological constant has surprising consequences, such as a finite maximum entropy of the observable universe (see the holographic principle).[18]
Predictions
Quantum field theory
See also: Cosmological constant problem
Question, Web Fundamentals.svg Unsolved problem in physics:
Why does the zero-point energy of the quantum vacuum not cause a large cosmological constant? What cancels it out?
(more unsolved problems in physics)

A major outstanding problem is that most quantum field theories predict a huge value for the quantum vacuum. A common assumption is that the quantum vacuum is equivalent to the cosmological constant. Although no theory exists that supports this assumption, arguments can be made in its favor.[19]

Such arguments are usually based on dimensional analysis and effective field theory. If the universe is described by an effective local quantum field theory down to the Planck scale, then we would expect a cosmological constant of the order of \( {\displaystyle M_{\rm {pl}}^{2}} \) ( 1 in reduced Planck units). As noted above, the measured cosmological constant is smaller than this by a factor of ~10−120. This discrepancy has been called "the worst theoretical prediction in the history of physics!".[8]

Some supersymmetric theories require a cosmological constant that is exactly zero, which further complicates things. This is the cosmological constant problem, the worst problem of fine-tuning in physics: there is no known natural way to derive the tiny cosmological constant used in cosmology from particle physics.

No vacuum in the string theory landscape is known to support a metastable, positive cosmological constant, and in 2018 a group of four physicists advanced a controversial conjecture which would imply that no such universe exists.[20]

Anthropic principle

One possible explanation for the small but non-zero value was noted by Steven Weinberg in 1987 following the anthropic principle.[21] Weinberg explains that if the vacuum energy took different values in different domains of the universe, then observers would necessarily measure values similar to that which is observed: the formation of life-supporting structures would be suppressed in domains where the vacuum energy is much larger. Specifically, if the vacuum energy is negative and its absolute value is substantially larger than it appears to be in the observed universe (say, a factor of 10 larger), holding all other variables (e.g. matter density) constant, that would mean that the universe is closed; furthermore, its lifetime would be shorter than the age of our universe, possibly too short for intelligent life to form. On the other hand, a universe with a large positive cosmological constant would expand too fast, preventing galaxy formation. According to Weinberg, domains where the vacuum energy is compatible with life would be comparatively rare. Using this argument, Weinberg predicted that the cosmological constant would have a value of less than a hundred times the currently accepted value.[22] In 1992, Weinberg refined this prediction of the cosmological constant to 5 to 10 times the matter density.[23]

This argument depends on a lack of a variation of the distribution (spatial or otherwise) in the vacuum energy density, as would be expected if dark energy were the cosmological constant. There is no evidence that the vacuum energy does vary, but it may be the case if, for example, the vacuum energy is (even in part) the potential of a scalar field such as the residual inflaton (also see quintessence). Another theoretical approach that deals with the issue is that of multiverse theories, which predict a large number of "parallel" universes with different laws of physics and/or values of fundamental constants. Again, the anthropic principle states that we can only live in one of the universes that is compatible with some form of intelligent life. Critics claim that these theories, when used as an explanation for fine-tuning, commit the inverse gambler's fallacy.

In 1995, Weinberg's argument was refined by Alexander Vilenkin to predict a value for the cosmological constant that was only ten times the matter density,[24] i.e. about three times the current value since determined.
Failure to detect dark energy

An attempt to directly observe dark energy in a laboratory failed to detect a new force.[25]
See also

Big Rip
Higgs mechanism
Lambdavacuum solution
Naturalness (physics)
Quantum electrodynamics
de Sitter relativity
Unruh effect

References
Footnotes

It may well be that dark energy is explained by a static cosmological constant, or that this mysterious energy is not constant at all and has changed over time, as in the case with quintessence, see for example:

"Physics invites the idea that space contains energy whose gravitational effect approximates that of Einstein's cosmological constant, Λ; nowadays the concept is termed dark energy or quintessence." Peebles & Ratra (2003), p. 1
"It would then appear that the cosmological fluid is dominated by some sort of fantastic energy density, which has negative pressure, and has just begun to play an important role today. No convincing theory has yet been constructed to explain this state of affairs, although cosmological models based on a dark energy component, such as the cosmological constant (Λ) or quintessence (Q),are leading candidates." Caldwell (2002), p. 2

Einstein (1917)
Rugh & Zinkernagel (2001), p. 3
On the Cosmological Constant being thought to have zero value see for example:

"Since the cosmological upper bound on \( {\displaystyle \left|\left\langle \rho \right\rangle +\lambda /8\pi G\right|} \) was vastly less than any value expected from particle theory, most particle theorists simply assumed that for some unknown reason this quantity was zero." Weinberg (1989), p. 3
"An epochal astronomical discovery would be to establish by convincing observation that Λ is nonzero." Carroll, Press & Turner (1992), p. 500
"Before 1998, there was no direct astronomical evidence for Λ and the observational upper bound was so strong ( Λ < 10−120 Planck units) that many particle physicists suspected that some fundamental principle must force its value to be precisely zero." Barrow & Shaw (2011), p. 1
"The only other natural value is Λ = 0. If Λ really is tiny but not zero, it adds a most stimulating though enigmatic clue to physics to be discovered." Peebles & Ratra (2003), p. 333

See for example:

"This is the independent result of two teams. Supernova Cosmology Project (Perlmutter et al. (1999); also see Perlmutter et al. (1998)) and the High-Z Supernova Search Team (Riess et al. (1998); also see Schmidt et al. (1998))" Weinberg (2015), p. 376

Redd (2013)
Rugh & Zinkernagel (2001), p. 1
See for example:

"This gives an answer about 120 orders of magnitude higher than the upper limits on Λ set by cosmological observations. This is probably the worst theoretical prediction in the history of physics!" Hobson, Efstathiou & Lasenby (2006), p. 187
"This, as we will see later, is approximately 120 orders of magnitude larger than what is allowed by observation." Carroll, Press & Turner (1992), p. 503
"Theoretical expectations for the cosmological constant exceed observational limits by some 120 orders of magnitude." Weinberg (1989), p. 1

See for example:

"the vacuum holds the key to a full understanding of nature" Davies (1985), p. 104
"The theoretical problem of explaining the cosmological constant is one of the greatest challenges of theoretical physics. It is most likely that we require a fully developed theory of quantum gravity (perhaps superstring theory) before we can understand Λ." Hobson, Efstathiou & Lasenby (2006), p. 188

There is some debate over whether Einstein labelled the cosmological constant his "biggest blunder", with all references being traced back to a single person: George Gamow. (See Gamow (1956, 1970).) For example:

"Astrophysicist and author Mario Livio can find no documentation that puts those words into Einstein's mouth (or, for that matter, his pen). Instead, all references eventually lead back to one man—physicist George Gamow—who reported Einstein's use of the phrase in two sources: His posthumously published autobiography My World Line (1970) and a Scientific American article from September 1956." Rosen (2013)
" We also find it quite plausible that Einstein made such a statement to Gamow in particular. We conclude that there is little doubt that Einstein came to view the introduction of the cosmological constant a serious error, and that it is very plausible that he labelled the term his "biggest blunder" on at least one occasion". O'Raifeartaigh & Mitton (2018), p. 1

Ryden (2003), p. 59
Planck Collaboration (2018)
Brumfiel (2007), p. 246
See e.g. Baker et al. (1999)
See for example Table 9 in The Planck Collaboration (2015a), p. 27
Barrow & Shaw (2011)
Calculated based on the Hubble constant and Ω Λ {\displaystyle \Omega _{\Lambda }} \Omega _{\Lambda } from The Planck Collaboration (2015b)
Dyson, Kleban & Susskind (2002)
Rugh & Zinkernagel (2001), p. ?
Wolchover, Natalie (9 August 2018). "Dark Energy May Be Incompatible With String Theory". Quanta Magazine. Simons Foundation. Retrieved 2 April 2020.
Weinberg (1987)
Vilenkin (2006), pp. 138–139
Weinberg (1992), p. 182
Vilenkin (2006), p. 146

D. O. Sabulsky; I. Dutta; E. A. Hinds; B. Elder; C. Burrage; E. J. Copeland (2019). "Experiment to Detect Dark Energy Forces Using Atom Interferometry". Physical Review Letters. 123 (6): 061102. arXiv:1812.08244. Bibcode:2019PhRvL.123f1102S. doi:10.1103/PhysRevLett.123.061102. PMID 31491160.

Bibliography
Primary literature

Baker, J. C.; Grainge, K.; Hobson, M.P.; Jones, M.E.; Kneissl, R.; Lasenby, A.N.; O'Sullivan, C.M. M.; Pooley, G.; Rocha, G.; Saunders, R.; Scott, P.F.; Waldram, E.M.; et al. (1999). "Detection of cosmic microwave background structure in a second field with the Cosmic Anisotropy Telescope". Monthly Notices of the Royal Astronomical Society. 308 (4): 1173–1178.arXiv:astro-ph/9904415. Bibcode:1999MNRAS.308.1173B. doi:10.1046/j.1365-8711.1999.02829.x. ISSN 0035-8711.
Dyson, L.; Kleban, M.; Susskind, L. (2002). "Disturbing Implications of a Cosmological Constant". Journal of High Energy Physics. 2002 (10): 011. arXiv:hep-th/0208013. Bibcode:2002JHEP...10..011D. doi:10.1088/1126-6708/2002/10/011. ISSN 1029-8479.
Einstein, A. (1917). "Kosmologische Betrachtungen zur allgemeinen Relativitätstheorie". Sitzungsberichte der Königlich Preußischen Akademie der Wissenschaften. Berlin, DE. part 1: 142–152. Bibcode:1917SPAW.......142E. Archived from the original on 2019-03-21. Retrieved 2014-11-15.
Gamow, G. (1956). "The evolutionary universe" . Scientific American. 195 (3): 136–156. Bibcode:1956SciAm.195c.136G. doi:10.1038/scientificamerican0956-136. JSTOR 24941749.
Gamow, G. (1970). My World Line: An informal autobiography. New York, NY: Viking Press. ISBN 978-0-670-50376-6. LCCN 79094855. OCLC 70097.
Perlmutter, S.; Aldering, G.; Valle, M. Della; Deustua, S.; Ellis, R.S.; Fabbro, S.; Fruchter, A.; Goldhaber, G.; Groom, D.E.; Hook, I.M.; Kim, A.G.; Kim, M.Y.; Knop, R. A.; Lidman, C.; McMahon, R.G.; Nugent, P.; Pain, R.; Panagia, N.; Pennypacker, C. R.; Ruiz-Lapuente, P.; Schaefer, B.; Walton, N. (1998). "Discovery of a supernova explosion at half the age of the Universe". Nature. 391 (6662): 51–54.arXiv:astro-ph/9712212. Bibcode:1998Natur.391...51P. doi:10.1038/34124. ISSN 0028-0836.
Perlmutter, S.; Aldering, G.; Goldhaber, G.; Knop, R.A.; Nugent, P.; Castro, P.G.; Deustua, S.; Fabbro, S.; Goobar, A.; Groom, D.E.; Hook, I.M.; Kim, A.G.; Kim, M.Y.; Lee, J.C.; Nunes, N.J.; Pain, R.; Pennypacker, C.R.; Quimby, R.; Lidman, C.; Ellis, R.S.; Irwin, M.; McMahon, R.G.; Ruiz-Lapuente, P.; Walton, N.; Schaefer, B.; Boyle, B.J.; Filippenko, A.V.; Matheson, T.; Fruchter, A.S.; Panagia, N.; Newberg, H.J.M.; Couch, W.J.; The Supernova Cosmology Project (1999). "Measurements of Ω and Λ from 42 high-redshift supernovae". The Astrophysical Journal. 517 (2): 565–586. arXiv:astro-ph/9812133. Bibcode:1999ApJ...517..565P. doi:10.1086/307221. ISSN 0004-637X.
Riess, A.G.; Filippenko, A.V.; Challis, P.; Clocchiatti, A.; Diercks, A.; Garnavich, P.M.; Gilliland, R.L.; Hogan, C.J.; Jha, S.; Kirshner, R.P.; Leibundgut, B.; Phillips, M.M.; Reiss, D.; Schmidt, B.P.; Schommer, R.A.; Smith, R.C.; Spyromilio, J.; Stubbs, C.; Suntzeff, N.B.; Tonry, J. (1998). "Observational Evidence from Supernovae for an Accelerating Universe and a Cosmological Constant". The Astronomical Journal. 116 (3): 1009–1038.arXiv:astro-ph/9805201. Bibcode:1998AJ....116.1009R. doi:10.1086/300499. ISSN 0004-6256.
Schmidt, B.P.; Suntzeff, N.B.; Phillips, M.M.; Schommer, R.A.; Clocchiatti, A.; Kirshner, R.P.; Garnavich, P.; Challis, P.; Leibundgut, B.; Spyromilio, J.; Riess, A.G.; Filippenko, A.V.; Hamuy, M.; Smith, R. C.; Hogan, C.; Stubbs, C.; Diercks, A.; Reiss, D.; Gilliland, R.; Tonry, J.; Maza, J.; Dressler, A.; Walsh, J.; Ciardullo, R. (1998). "The High-Z Supernova Search: Measuring Cosmic Deceleration and Global Curvature of the Universe Using Type Ia Supernovae". The Astrophysical Journal. 507 (1): 46–63. arXiv:astro-ph/9805200. Bibcode:1998ApJ...507...46S. doi:10.1086/306308. ISSN 0004-637X.
The Planck Collaboration (2016). "Planck 2015 results I. Overview of products and scientific results". Astronomy and Astrophysics. 594: A1. arXiv:1502.01582. Bibcode:2016A&A...594A...1P. doi:10.1051/0004-6361/201527101.
Planck Collaboration (2016). "Planck 2015 results. XIII. Cosmological parameters". Astronomy and Astrophysics. 594: A13.arXiv:1502.01589. Bibcode:2016A&A...594A..13P. doi:10.1051/0004-6361/201525830. ISSN 0004-6361.
The Planck Collaboration (2018). "Planck 2018 results. VI. Cosmological parameters". arXiv:1807.06209. Bibcode:2018arXiv180706209P.
Weinberg, S. (1987). "Anthropic Bound on the Cosmological Constant". Physical Review Letters. 59 (22): 2607–2610. Bibcode:1987PhRvL..59.2607W. doi:10.1103/PhysRevLett.59.2607. PMID 10035596.

Secondary literature: news, popular science articles & books

Abbott, Larry (1988). "The Mystery of the Cosmological Constant" . Scientific American. 258 (5): 106–113. Bibcode:1988SciAm.258e.106A. doi:10.1038/scientificamerican0588-106. ISSN 0036-8733. S2CID 30023659.
Barrow, J. D.; Webb, J. K. (2005). "Inconstant Constants" (PDF) . Scientific American. 292 (6): 56–63. Bibcode:2005SciAm.292f..56B. doi:10.1038/scientificamerican0605-56. ISSN 0036-8733. PMID 15934653.
Brumfiel, G. (2007). "A constant problem" (PDF). Nature. 448 (7151): 245–248. Bibcode:2007Natur.448..245B. doi:10.1038/448245a. ISSN 0028-0836. PMID 17637631.
Davies, P. C. W. (1985). Superforce: The Search for a Grand Unified Theory of Nature. New York: Simon and Schuster. ISBN 978-0-671-47685-4. LCCN 84005473. OCLC 12397205.
Hogan, J. (2007). "Welcome to the dark side" (PDF). Nature. 448 (7151): 240–245. Bibcode:2007Natur.448..240H. doi:10.1038/448240a. ISSN 0028-0836. PMID 17637630.
O'Raifeartaigh, C.; Mitton, S. (2018). "Einstein's "biggest blunder" – interrogating the legend". Physics in Perspective. 20 (4): 318–341.arXiv:1804.06768. doi:10.1007/s00016-018-0228-9.
Redd, N. T. (2013). "What is Dark Energy?". space.com. Archived from the original on 19 May 2016. Retrieved 28 October 2018.
Rosen, R. J. (2013). "Einstein Likely Never Said One of His Most Oft-Quoted Phrases". theatlantic.com. The Atlantic. Archived from the original on 10 August 2013. Retrieved 6 March 2017.

Secondary literature: review articles, monographs and textbooks

Barrow, J. D.; Shaw, D. J. (2011). "The value of the cosmological constant". General Relativity and Gravitation. 43 (10): 2555–2560. arXiv:1105.3105. Bibcode:2011GReGr..43.2555B. doi:10.1007/s10714-011-1199-1. ISSN 0001-7701.
Caldwell, R. R. (2002). "A phantom menace? Cosmological consequences of a dark energy component with super-negative equation of state". Physics Letters B. 545 (1–2): 23–29. arXiv:astro-ph/9908168. Bibcode:2002PhLB..545...23C. doi:10.1016/S0370-2693(02)02589-3. ISSN 0370-2693.
Carroll, S. M.; Press, W. H.; Turner, E. L. (1992). "The Cosmological Constant" (PDF). Annual Review of Astronomy and Astrophysics. 30 (1): 499–542. Bibcode:1992ARA&A..30..499C. doi:10.1146/annurev.aa.30.090192.002435. ISSN 0066-4146.
Hobson, M. P.; Efstathiou, G. P.; Lasenby, A. N. (2006). General Relativity: An Introduction for Physicists (2014 ed.). Cambridge: Cambridge University Press. ISBN 978-0-521-82951-9. LCCN 2006277059. OCLC 903178203.
Joyce, A.; Jain, B.; Khoury, J.; Trodden, M. (2015). "Beyond the cosmological standard model". Physics Reports. 568: 1–98. arXiv:1407.0059. Bibcode:2015PhR...568....1J. doi:10.1016/j.physrep.2014.12.002. ISSN 0370-1573.
Peebles, P. J. E.; Ratra, B. (2003). "The Cosmological Constant and Dark Energy". Reviews of Modern Physics. 75 (2): 559–606. arXiv:astro-ph/0207347. Bibcode:2003RvMP...75..559P. doi:10.1103/RevModPhys.75.559. ISSN 0034-6861.
Rugh, S; Zinkernagel, H. (2001). "The Quantum Vacuum and the Cosmological Constant Problem". Studies in History and Philosophy of Modern Physics. 33 (4): 663–705.arXiv:hep-th/0012253. Bibcode:2002SHPMP..33..663R. doi:10.1016/S1355-2198(02)00033-3.
Ryden, B. S. (2003). Introduction to Cosmology. San Francisco: Addison-Wesley. ISBN 978-0-8053-8912-8. LCCN 2002013176. OCLC 50478401.
Vilenkin, A. (2006). Many worlds in one: The Search For Other Universes. New York: Hill and Wang. ISBN 978-0-8090-9523-0. LCCN 2005027057. OCLC 799428013.
Weinberg, S. (1989). "The Cosmological Constant Problem" (PDF). Reviews of Modern Physics. 61 (1): 1–23. Bibcode:1989RvMP...61....1W. doi:10.1103/RevModPhys.61.1. hdl:2152/61094. ISSN 0034-6861.
Weinberg, S. (1992). Dreams of a Final Theory: The Scientist's Search for the Ultimate Laws of Nature. New York: Pantheon Books. ISBN 978-0-679-74408-5. LCCN 93030534. OCLC 319776354.
Weinberg, S. (2015). Lectures on Quantum Mechanics (2nd ed.). Cambridge: Cambridge University Press. ISBN 978-1-107-11166-0. LCCN 2015021123. OCLC 910664598.

Physics Encyclopedia

World

Index

Hellenica World - Scientific Library

Retrieved from "http://en.wikipedia.org/"
All text is available under the terms of the GNU Free Documentation License