Flatness problem

Flatness problem
The local geometry of the universe is determined by whether the relative density Ω is less than, equal to or greater than 1. From top to bottom: a spherical universe with greater than critical density (Ω>1, k>0); a hyperbolic, underdense universe (Ω<1, k<0); and a flat universe with exactly the critical density (Ω=1, k=0). Our universe, unlike the diagrams, is three-dimensional.

The flatness problem is a cosmological fine-tuning problem within the Big Bang model of the universe. Such problems arise from the observation that some of the initial conditions of the universe appear to be fine-tuned to very 'special' values, and that a small deviation from these values would have had massive effects on the nature of the universe at the current time.

In the case of the flatness problem, the parameter which appears fine-tuned is the density of matter and energy in the universe. This value affects the curvature of space-time, with a very specific critical value being required for a flat universe. The current density of the universe is observed to be very close to this critical value. Since the total density departs rapidly from the critical value over cosmic time,[1] the early universe must have had a density even closer to the critical density, departing from it by one part in 1062 or less. This leads cosmologists to question how the initial density came to be so closely fine-tuned to this 'special' value.

The problem was first mentioned by Robert Dicke in 1969. The most commonly accepted solution among cosmologists is cosmic inflation, the idea that the universe went through a brief period of extremely rapid expansion in the first fraction of a second after the Big Bang; along with the monopole problem and the horizon problem, the flatness problem is one of the three primary motivations for inflationary theory.[2]

Contents

Energy density and the Friedmann equation

According to Einstein's field equations of general relativity, the structure of spacetime is affected by the presence of matter and energy. On small scales space appears flat – as does the surface of the Earth if one looks at a small area. On large scales however, space is bent by the gravitational effect of matter. Since relativity indicates that matter and energy are equivalent, this effect is also produced by the presence of energy (such as light and other electromagnetic radiation) in addition to matter. The amount of bending (or curvature) of the universe depends on the density of matter/energy present.

This relationship can be expressed by the first Friedmann equation. In a universe without a cosmological constant, this is:

H^2 = \frac{8 \pi G}{3} \rho - \frac{kc^2}{a^2}

Here H is the Hubble parameter, a measure of the rate at which the universe is expanding. ρ is the total density of mass and energy in the universe, a is the scale factor (essentially the 'size' of the universe), and k is the curvature parameter — that is, a measure of how curved spacetime is. A positive, zero or negative value of k corresponds to a respectively closed, flat or open universe. The constants G and c are Newton's gravitational constant and the speed of light, respectively.

Cosmologists often simplify this equation by defining a critical density, ρc. For a given value of H, this is defined as the density required for a flat universe, i.e. k = 0. Thus the above equation implies

\rho_c = \frac{3H^2}{8\pi G}.

Since the constant G is known and the expansion rate H can be measured by observing the speed at which distant galaxies are receding from us, ρc can be determined. Its value is currently around 10−26 kg m−3. The ratio of the actual density to this critical value is called Ω, and its difference from 1 determines the geometry of the universe: Ω > 1 corresponds to a greater than critical density, ρ > ρc, and hence a closed universe. Ω < 1 gives a low density open universe, and Ω equal to exactly 1 gives a flat universe.

The Friedmann equation above can now be rearranged as follows:

\frac{3a^2}{8\pi G}H^2 = \rho a^2 - \frac{3kc^2}{8 \pi G}
\rho_c a^2 - \rho a^2 = - \frac{3kc^2}{8 \pi G}
(\Omega^{-1} - 1)\rho a^2 = \frac{-3kc^2}{8 \pi G}.[3]

The right hand side of this expression contains only constants, and therefore the left hand side must remain constant throughout the evolution of the universe.

As the universe expands the scale factor a increases, but the density ρ decreases as matter (or energy) becomes spread out. For the standard model of the universe which contains mainly matter and radiation for most of its history, ρ decreases more quickly than a2 increases, and so the factor ρa2 will decrease. Since the time of the Planck era, shortly after the Big Bang, this term has decreased by a factor of around 1060,[3] and so − 1 − 1) must have increased by a similar amount to retain the constant value of their product.

Current value of Ω

The relative density Ω against cosmic time t (neither axis to scale). Each curve represents a possible universe: note that Ω diverges rapidly from 1. The blue curve is a universe similar to our own, which at the present time (right of the graph) has a small |Ω − 1| and therefore must have begun with Ω very close to 1 indeed. The red curve is a hypothetical different universe in which the initial value of Ω differed slightly too much from 1: by the present day it has diverged massively and would not be able to support galaxies, stars or planets.

Measurement

The value of Ω at the present time is denoted Ω0. This value can be deduced by measuring the curvature of spacetime (since Ω=1, or ρ = ρc, is defined as the density for which the curvature k=0). The curvature can be inferred from a number of observations.

One such observation is that of anisotropies (that is, variations with direction - see below) in the Cosmic Microwave Background (CMB) radiation. The CMB is electromagnetic radiation which fills the universe, left over from an early stage in its history when it was filled with photons and a hot, dense plasma. This plasma cooled as the universe expanded, and when it cooled enough to form stable atoms it no longer absorbed the photons. The photons present at that stage have been propagating ever since, growing fainter and less energetic as they spread through the ever-expanding universe.

The temperature of this radiation is almost the same at all points on the sky, but there is a slight variation (around one part in 100,000) between the temperature received from different directions. The angular scale of these fluctuations - the typical angle between a hot patch and a cold patch on the sky[nb 1] - depends on the curvature of the universe which in turn depends on its density as described above. Thus, measurements of this angular scale allow an estimation of Ω0.[4][nb 2]

Another probe of Ω0 is the frequency of Type-Ia supernovae at different distances from Earth.[5][6] These supernovae, the explosions of massive stars, are a type of standard candle; this means that the processes governing their intrinsic brightness are well understood so that a measure of apparent brightness when seen from Earth can be used to derive accurate distance measures for them (the apparent brightness decreasing in proportion to the square of the distance - see luminosity distance). Comparing this distance to the redshift of the supernovae gives a measure of the rate at which the universe has been expanding at different points in history. Since the expansion rate evolves differently over time in cosmologies with different total densities, Ω0 can be inferred from the supernovae data.

Data from the Wilkinson Microwave Anisotropy Probe (measuring CMB anisotropies) combined with that from the Sloan Digital Sky Survey and observations of type-Ia supernovae constrain Ω0 to be 1 within 1%.[7] In other words the term |Ω − 1| is currently less than 0.01, and therefore must have been less than 10−62 at the Planck era.

Implication

This tiny value is the crux of the flatness problem. If the initial density of the universe could take any value, it would seem extremely surprising to find it so 'finely tuned' to the critical value ρc. Indeed, a very small departure of Ω from 1 in the early universe would have been magnified during billions of years of expansion to create a current density very far from critical. In the case of an overdensity (ρ > ρc) this would lead to a universe so dense it would cease expanding and collapse into a Big Crunch (an opposite to the Big Bang in which all matter and energy falls back into an incredibly dense state) in a few years or less; in the case of an underdensity (ρ < ρc) it would expand so quickly and become so sparse it would soon seem essentially empty, and gravity would not be strong enough by comparison to cause matter to collapse and form galaxies. In either case the universe would contain no complex structures such as galaxies, stars, planets and people.[8]

This problem with the Big Bang model was first pointed out by Robert Dicke in 1969,[9] and it motivated a search for some reason the density should take such a specific value.

Solutions to the problem

Some cosmologists agreed with Dicke that the flatness problem was a serious one, in need of a fundamental reason for the closeness of the density to criticality. But there was also a school of thought which denied that there was a problem to solve, arguing instead that since the universe must have some density it may as well have one close to ρcrit as far from it, and that speculating on a reason for any particular value was "beyond the domain of science".[9] Enough cosmologists saw the problem as a real one, however, for various solutions to be proposed.

Anthropic principle

One solution to the problem is to invoke the anthropic principle, which states that humans should take into account the conditions necessary for them to exist when speculating about causes of the universe's properties. If two types of universe seem equally likely but only one is suitable for the evolution of intelligent life, the anthropic principle suggests that finding ourselves in that universe is no surprise: if the other universe had existed instead, there would be no observers to notice the fact.

The principle can be applied to solve the flatness problem in two somewhat different ways. The first (an application of the 'strong anthropic principle') was suggested by C. B. Collins and Stephen Hawking,[10][11] who in 1973 considered the existence of an infinite number of universes such that every possible combination of initial properties was held by some universe. In such a situation, they argued, only those universes with exactly the correct density for forming galaxies and stars would give rise to intelligent observers such as humans: therefore, the fact that we observe Ω to be so close to 1 would be "simply a reflection of our own existence."[10]

An alternative approach, which makes use of the 'weak anthropic principle', is to suppose that the universe is infinite in size, but with the density varying in different places (i.e. an inhomogeneous universe). Thus some regions will be over-dense (Ω > 1) and some under-dense (Ω < 1). These regions may be extremely far apart - perhaps so far that light has not had time to travel from one to another during the age of the universe (that is, they lie outside one another's cosmological horizons). Therefore each such region would behave essentially as a separate universe: if we happened to live in a large patch of almost-critical density we would have no way of knowing of the existence of far-off under- or over-dense patches since no light or other signal has reached us from them. An appeal to the anthropic principle can then be made, arguing that intelligent life would only arise in those patches with Ω very close to 1, and that therefore our living in such a patch is unsurprising.[12]

This latter argument makes use of a version of the anthropic principle which is 'weaker' in the sense that it requires no speculation on multiple universes, or on the probabilities of various different universes existing instead of the current one. It requires only a single universe which is infinite - or merely large enough that many disconnected patches can form - and that the density varies in different regions (which is certainly the case on smaller scales, giving rise to galactic clusters and voids).

However, the Anthropic Principle has been criticised by many scientists.[11] For example, in 1979 Bernard Carr and Martin Rees argued that the principle “is entirely post hoc: it has not yet been used to predict any feature of the Universe.”[11][13] Others have taken objection to its philosophical basis, with Ernan McMullin writing in 1994 that "the weak Anthropic principle is trivial ... and the strong Anthropic principle is indefensible." Since many physicists and philosophers of science do not consider the principle to be compatible with the scientific method,[11] another explanation for the flatness problem was needed.

Inflation

The standard solution to the Flatness Problem invokes cosmic inflation, a process whereby the universe expands exponentially quickly (i.e. a grows as eλt with time t, for some constant λ) during a short period in its early history. The theory of inflation was first proposed in 1979, and published in 1981, by Alan Guth.[14][15] His two main motivations for doing so were the flatness problem and the horizon problem, another fine-tuning problem of physical cosmology.

The proposed cause of inflation is a field which permeates space and drives the expansion. The field contains a certain energy density, but unlike the density of the matter or radiation present in the late universe, which decrease over time, the density of the inflationary field remains roughly constant as space expands. Therefore the term ρa2 increases extremely rapidly as the scale factor a grows exponentially. Recalling the Friedman Equation

(\Omega^{-1} - 1)\rho a^2 = \frac{-3kc^2}{8\pi G},

and the fact that the right-hand side of this expression is constant, the term | Ω − 1 − 1 | must therefore decrease with time.

Thus if | Ω − 1 − 1 | initially takes any arbitrary value, a period of inflation can force it down towards 0 and leave it extremely small - around 10 − 62 as required above, for example. Subsequent evolution of the universe will cause the value to grow, bringing it to the currently observed value of around 0.01. Thus the sensitive dependence on the initial value of Ω has been removed: a large and therefore 'unsurprising' starting value need not become massively amplified and lead to a very curved universe with no opportunity to form galaxies and other structures.

This success in solving the flatness problem is considered one of the major motivations for inflationary theory.[2][16]

Post inflation

Although inflationary theory is regarded as having had much success, and the evidence for it as compelling, it is not universally accepted: cosmologists recognise that there are still gaps in the theory and are open to the possibility that future observations will disprove it.[17][18] In particular, in the absence of any firm evidence for what the field driving inflation should be, many different versions of the theory have been proposed.[19] Many of these contain parameters or initial conditions which themselves require fine-tuning[19] in much the way that the early density does without inflation.

For these reasons work is still being done on alternative solutions to the flatness problem. These have included non-standard interpretations of the effect of dark energy[20] and gravity,[21] particle production in an oscillating universe,[22] and use of a Bayesian statistical approach to argue that the problem is non-existent. The latter argument, suggested for example by Evrard and Coles,[23] maintains that the idea that Ω being close to 1 is 'unlikely' is based on assumptions about the likely distribution of the parameter which are not necessarily justified.

Despite this ongoing work, inflation remains by far the dominant explanation for the flatness problem.[1][2]

Notes

  1. ^ Since there are fluctuations on many scales, not a single angular separation between hot and cold spots, the necessary measure is the angular scale of the first peak in the anisotropies' power spectrum. See Cosmic Microwave Background#Primary anisotropy.
  2. ^ Liddle[4] uses an alternative notation in which Ω0 is the current density of matter alone, excluding any contribution from dark energy; his Ω0Λ corresponds to Ω0 in this article.

References

  1. ^ a b Peacock, J. A. (1998). Cosmological Physics. Cambridge: Cambridge University Press. ISBN 978-0521422703. 
  2. ^ a b c Barbara Ryden (2002). Introduction to Cosmology. San Francisco: Addison Wesley. ISBN 0805389121. 
  3. ^ a b Peter Coles and Francesco Lucchin (1997). Cosmology. Chichester: Wiley. ISBN 047195473X. 
  4. ^ a b Liddle, Andrew. An Introduction to Modern Cosmology. Wiley. p. 157. ISBN 978-0-490-84835-7. 
  5. ^ Ryden p. 168
  6. ^ Stompor, Radek; et al. (2001). "Cosmological Implications of the MAXIMA-1 High-Resolution Cosmic Microwave Background Anisotropy Measurement". The Astrophysical Journal 561 (1): L7–L10. arXiv:astro-ph/0105062. Bibcode 2001ApJ...561L...7S. doi:10.1086/324438. 
  7. ^ D. N. Spergel et al. (June 2007). "Wilkinson Microwave Anisotropy Probe (WMAP) Three Year Results: Implications for Cosmology". ApJS 170 (2): 337–408. arXiv:astro-ph/0603449. Bibcode 2007ApJS..170..377S. doi:10.1086/513700. 
  8. ^ Ryden p. 193
  9. ^ a b Agazzi, Evandro; Massimo Pauri (2000). The Reality of the Unobservable: Observability, Unobservability and Their Impact on the Issue of Scientific Realism. Springer. p. 226. ISBN 9780792363118. http://books.google.com/?id=OIG0F37QrmQC&pg=PT237&lpg=PT237&dq=%22flatness+problem+was%22. 
  10. ^ a b Collins, C. B.; Hawking, S. (1973). "Why is the Universe Isotropic?". Astrophysical Journal 180: 317–334. Bibcode 1973ApJ...180..317C. doi:10.1086/151965. 
  11. ^ a b c d Mosterín, Jesús (2003). "Anthropic Explanations in Cosmolgy". http://philsci-archive.pitt.edu/archive/00001658/. Retrieved 2008-08-01. 
  12. ^ Barrow, John D.; Tiple, Frank J. (1986). The Anthropic Cosmological Principle. Oxford: Clarendon Press. p. 411. ISBN 0-19-851949-4. 
  13. ^ Carr, Bernard J.; Rees, Martin (April 1979). "The anthropic principle and the structure of the physical world". Nature 278 (5705): 605–612. Bibcode 1979Natur.278..605C. doi:10.1038/278605a0. 
  14. ^ Castelvecchi, Davide (1981). "The Growth of Inflation". Physical Review D 23 (2): 347. Bibcode 1981PhRvD..23..347G. doi:10.1103/PhysRevD.23.347. 
  15. ^ Guth, Alan (January 1981). "Inflationary universe: A possible solution to the horizon and flatness problems". Physical Review D 23 (2): 347–356. Bibcode 1981PhRvD..23..347G. doi:10.1103/PhysRevD.23.347. 
  16. ^ Coles, Peter; Ellis, George F. R. (1997). Is the Universe Open or Closed? The Density of Matter in the Universe. Cambridge: Cambridge University Press. ISBN 0-521-56689-4. 
  17. ^ Albrecht, Andreas (August 2000). Proceedings of the NATO Advanced Study Institute on Structure Formation in the Universe, Cambridge 1999. arXiv:astro-ph/0007247. Bibcode 2001sfu..conf...17A. ISBN 140200155X. 
  18. ^ Guth, Alan (1997). "Was Cosmic Inflation the 'Bang' of the Big Bang?". The Beamline 27. http://nedwww.ipac.caltech.edu/level5/Guth/Guth_contents.html. Retrieved 2008-09-07. 
  19. ^ a b Bird, Simeon; Peiris, Hiranya V.; Easther, Richard (July 2008). "Fine-tuning criteria for inflation and the search for primordial gravitational waves". Physical Review D 78 (8): 083518. Bibcode 2008PhRvD..78h3518B. doi:10.1103/PhysRevD.78.083518. 
  20. ^ Chernin, Arthur D. (January 2003). "Cosmic vacuum and the 'flatness problem' in the concordant model". New Astronomy 8 (1): 79–83. arXiv:astro-ph/0211489. Bibcode 2003NewA....8...79C. doi:10.1016/S1384-1076(02)00180-X. 
  21. ^ Nikolic, Hrvoje (August 1999). "Some Remarks on a Nongeometrical Interpretation of Gravity and the Flatness Problem". General Relativity and Gravitation 31 (8): 1211. arXiv:gr-qc/9901057. Bibcode 1999GReGr..31.1211N. doi:10.1023/A:1026760304901. 
  22. ^ Anderson, P. R.; Schokman; Zaramensky; R. Schokman, M. Zaramensky (May 1997). "A Solution to the Flatness Problem via Particle Production in an Oscillating Universe". Bulletin of the American Astronomical Society 29: 828. Bibcode 1997AAS...190.3806A. 
  23. ^ Evrard, G; P. Coles (October 1995). "Getting the measure of the flatness problem". Classical Quantum Gravity 12 (10): L93–L97. arXiv:astro-ph/9507020. Bibcode 1995CQGra..12L..93E. doi:10.1088/0264-9381/12/10/001. .

Wikimedia Foundation. 2010.

Игры ⚽ Нужен реферат?

Look at other dictionaries:

  • Horizon problem — The horizon problem is a problem with the standard cosmological model of the Big Bang which was identified in the 1970s. It points out that different regions of the universe have not contacted each other due to the great distances between them,… …   Wikipedia

  • Inflation (cosmology) — Inflation model and Inflation theory redirect here. For a general rise in the price level, see Inflation. For other uses, see Inflation (disambiguation). Physical cosmology …   Wikipedia

  • Big Bang — This article is about the cosmological model. For the sitcom, see The Big Bang Theory. For other uses, see Big Bang (disambiguation) …   Wikipedia

  • Cosmic inflation — In physical cosmology, cosmic inflation is the idea that the universe passed through a phase of exponential expansion that was driven by a negative pressure vacuum energy density. [Liddle and Lyth (2000) and Mukhanov (2005) are recent cosmology… …   Wikipedia

  • Alan Guth — Infobox Scientist name = Alan Harvey Guth caption = Alan Guth at Harvard University birth date = Birth date and age|1947|2|27|df=y birth place = New Brunswick, New Jersey death date = death place = residence = United States nationality = American …   Wikipedia

  • Self-creation cosmology — (SCC) theories are gravitational theories in which the mass of the universe is created out of its self contained gravitational and scalar fields, as opposed to the theory of continuous creation cosmology or the steady state theory which depend on …   Wikipedia

  • Timeline of cosmology — This timeline of cosmological theories and discoveries is a chronological catalog of the evolution of humankind s understanding of the cosmos over the last two plus millennia. Modern cosmological conceptions follow the development of the… …   Wikipedia

  • General relativity — For a generally accessible and less technical introduction to the topic, see Introduction to general relativity. General relativity Introduction Mathematical formulation Resources …   Wikipedia

  • Non-standard cosmology — Physical cosmology Universe · Big Bang …   Wikipedia

  • Fine-tuning — In theoretical physics, fine tuning refers to circumstances when the parameters of a model must be adjusted very precisely in order to agree with observations. Theories requiring fine tuning are regarded as problematic in the absence of a known… …   Wikipedia

Share the article and excerpts

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