Flatness problem

Last updated
The local geometry of the universe is determined by whether the relative density O is less than, equal to or greater than 1. From top to bottom: a spherical universe with greater than critical density (O>1, k>0); a hyperbolic, underdense universe (O<1, k<0); and a flat universe with exactly the critical density (O=1, k=0). The spacetime of the universe is, unlike the diagrams, four-dimensional. End of universe.jpg
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). The spacetime of the universe is, unlike the diagrams, four-dimensional.

The flatness problem (also known as the oldness 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 small deviations from these values would have extreme effects on the appearance of the universe at the current time.

Contents

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 any departure of the total density from the critical value would increase rapidly 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. [2] :62, [3] :61 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. [4]

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:

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

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

.

Since the constant is known and the expansion rate can be measured by observing the speed at which distant galaxies are receding from us, can be determined. Its value is currently around 1026 kg m3. 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, , 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,

can be re-arranged into

which after factoring , and using , leads to

[5]

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

As the universe expands the scale factor increases, but the density 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, decreases more quickly than increases, and so the factor will decrease. Since the time of the Planck era, shortly after the Big Bang, this term has decreased by a factor of around [5] and so must have increased by a similar amount to retain the constant value of their product.

Current value of Ω

The relative density O against cosmic time t (neither axis to scale). Each curve represents a possible universe: note that O 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 |O - 1| and therefore must have begun with O very close to 1 indeed. The red curve is a hypothetical different universe in which the initial value of O differed slightly too much from 1: by the present day it has diverged extremely and would not be able to support galaxies, stars or planets. Flatness problem density graph.svg
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 extremely 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 , 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. [6] [nb 2]

Another probe of Ω0 is the frequency of Type-Ia supernovae at different distances from Earth. [7] [8] These supernovae, the explosions of degenerate white dwarf 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 (WMAP, 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%. [9] In other words, the term 1| is currently less than 0.01, and therefore must have been less than 1062 at the Planck era. The cosmological parameters measured by Planck spacecraft mission reaffirmed previous results by WMAP. [10] [11] [12]

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 . 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 () 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 extremely dense state) in a few years or less; in the case of an underdensity () 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 resulting in a big freeze. In either case the universe would contain no complex structures such as galaxies, stars, planets and any form of life. [13]

This problem with the Big Bang model was first pointed out by Robert Dicke in 1969, [14] 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 as far from it, and that speculating on a reason for any particular value was "beyond the domain of science". [14] That, however, is a minority viewpoint, even among those sceptical of the existence of the flatness problem. Several cosmologists have argued that, for a variety of reasons, the flatness problem is based on a misunderstanding, [15] but that seems to be widely ignored by many. 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, [16] 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." [16]

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 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. [17]

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. [18] 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.” [18] [19] 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, [18] 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. grows as with time , for some constant ) during a short period in its early history. The theory of inflation was first proposed in 1979, and published in 1981, by Alan Guth. [20] [21] His two main motivations for doing so were the flatness problem and the horizon problem, another fine-tuning problem of physical cosmology. However, “In December, 1980 when Guth was developing his inflation model, he was not trying to solve either the flatness or horizon problems. Indeed, at that time, he knew nothing of the horizon problem and had never quantitatively calculated the flatness problem”. [22] He was a particle physicist trying to solve the magnetic monopole problem.”

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 increases extremely rapidly as the scale factor grows exponentially. Recalling the Friedmann Equation

,

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

Thus if initially takes any arbitrary value, a period of inflation can force it down towards 0 and leave it extremely small - around 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 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. [4] [23]

Post inflation

Although inflationary theory is regarded as having had much success, and the evidence for it is compelling, it is not universally accepted: cosmologists recognize that there are still gaps in the theory and are open to the possibility that future observations will disprove it. [24] [25] 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. [26] Many of these contain parameters or initial conditions which themselves require fine-tuning [26] 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 [27] and gravity, [28] particle production in an oscillating universe, [29] 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, 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. [30] Despite this ongoing work, inflation remains by far the dominant explanation for the flatness problem. [1] [4] The question arises, however, whether it is still the dominant explanation because it is the best explanation, or because the community is unaware of progress on this problem. [31] In particular, in addition to the idea that Ω is not a suitable parameter in this context, other arguments against the flatness problem have been presented: if the universe collapses in the future, then the flatness problem "exists", but only for a relatively short time, so a typical observer would not expect to measure Ω appreciably different from 1; [32] in the case of a universe which expands forever with a positive cosmological constant, fine-tuning is needed not to achieve a (nearly) flat universe, but also to avoid it. [33]

Einstein–Cartan theory

The flatness problem is naturally solved by the Einstein–Cartan–Sciama–Kibble theory of gravity, without an exotic form of matter required in inflationary theory. [34] [35] This theory extends general relativity by removing a constraint of the symmetry of the affine connection and regarding its antisymmetric part, the torsion tensor, as a dynamical variable. It has no free parameters. Including torsion gives the correct conservation law for the total (orbital plus intrinsic) angular momentum of matter in the presence of gravity. The minimal coupling between torsion and Dirac spinors obeying the nonlinear Dirac equation generates a spin-spin interaction which is significant in fermionic matter at extremely high densities. Such an interaction averts the unphysical big bang singularity, replacing it with a bounce at a finite minimum scale factor, before which the Universe was contracting. The rapid expansion immediately after the big bounce explains why the present Universe at largest scales appears spatially flat, homogeneous and isotropic. As the density of the Universe decreases, the effects of torsion weaken and the Universe smoothly enters the radiation-dominated era.

See also

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 [6] 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.

Related Research Articles

<span class="mw-page-title-main">Big Bang</span> Theory on how the universe expanded

The Big Bang event is a physical theory that describes how the universe expanded from an initial state of high density and temperature. It was first proposed in 1927 by Roman Catholic priest and physicist Georges Lemaître. Various cosmological models of the Big Bang explain the evolution of the observable universe from the earliest known periods through its subsequent large-scale form. These models offer a comprehensive explanation for a broad range of observed phenomena, including the abundance of light elements, the cosmic microwave background (CMB) radiation, and large-scale structure. The overall uniformity of the Universe, known as the flatness problem, is explained through cosmic inflation: a sudden and very rapid expansion of space during the earliest moments. However, physics currently lacks a widely accepted theory of quantum gravity that can successfully model the earliest conditions of the Big Bang.

<span class="mw-page-title-main">Physical cosmology</span> Branch of cosmology which studies mathematical models of the universe

Physical cosmology is a branch of cosmology concerned with the study of cosmological models. A cosmological model, or simply cosmology, provides a description of the largest-scale structures and dynamics of the universe and allows study of fundamental questions about its origin, structure, evolution, and ultimate fate. Cosmology as a science originated with the Copernican principle, which implies that celestial bodies obey identical physical laws to those on Earth, and Newtonian mechanics, which first allowed those physical laws to be understood.

In physical cosmology, cosmic inflation, cosmological inflation, or just inflation, is a theory of exponential expansion of space in the early universe. The inflationary epoch is believed to have lasted from 10−36 seconds to between 10−33 and 10−32 seconds after the Big Bang. Following the inflationary period, the universe continued to expand, but at a slower rate. The acceleration of this expansion due to dark energy began after the universe was already over 7.7 billion years old.

<span class="mw-page-title-main">Cosmic microwave background</span> Trace radiation from the early universe

The cosmic microwave background is microwave radiation that fills all space in the observable universe. It is a remnant that provides an important source of data on the primordial universe. With a standard optical telescope, the background space between stars and galaxies is almost completely dark. However, a sufficiently sensitive radio telescope detects a faint background glow that is almost uniform and is not associated with any star, galaxy, or other object. This glow is strongest in the microwave region of the radio spectrum. The accidental discovery of the CMB in 1965 by American radio astronomers Arno Penzias and Robert Wilson was the culmination of work initiated in the 1940s.

<span class="mw-page-title-main">Cosmological constant</span> Constant representing stress–energy density of the vacuum

In cosmology, the cosmological constant, alternatively called Einstein's cosmological constant, is the constant coefficient of a term that Albert Einstein temporarily added to his field equations of general relativity. He later removed it. Much later it was revived and reinterpreted as the energy density of space, or vacuum energy, that arises in quantum mechanics. It is closely associated with the concept of dark energy.

<span class="mw-page-title-main">Accelerating expansion of the universe</span> Cosmological phenomenon

Observations show that the expansion of the universe is accelerating, such that the velocity at which a distant galaxy recedes from the observer is continuously increasing with time. The accelerated expansion of the universe was discovered in 1998 by two independent projects, the Supernova Cosmology Project and the High-Z Supernova Search Team, which used distant type Ia supernovae to measure the acceleration. The idea was that as type Ia supernovae have almost the same intrinsic brightness, and since objects that are farther away appear dimmer, the observed brightness of these supernovae can be used to measure the distance to them. The distance can then be compared to the supernovae's cosmological redshift, which measures how much the universe has expanded since the supernova occurred; the Hubble law established that the farther away that an object is, the faster it is receding. The unexpected result was that objects in the universe are moving away from one another at an accelerating rate. Cosmologists at the time expected that recession velocity would always be decelerating, due to the gravitational attraction of the matter in the universe. Three members of these two groups have subsequently been awarded Nobel Prizes for their discovery. Confirmatory evidence has been found in baryon acoustic oscillations, and in analyses of the clustering of galaxies.

<span class="mw-page-title-main">Hubble's law</span> Observation in physical cosmology

Hubble's law, also known as the Hubble–Lemaître law, is the observation in physical cosmology that galaxies are moving away from Earth at speeds proportional to their distance. In other words, the farther they are, the faster they are moving away from Earth. The velocity of the galaxies has been determined by their redshift, a shift of the light they emit toward the red end of the visible spectrum.

The ultimate fate of the universe is a topic in physical cosmology, whose theoretical restrictions allow possible scenarios for the evolution and ultimate fate of the universe to be described and evaluated. Based on available observational evidence, deciding the fate and evolution of the universe has become a valid cosmological question, being beyond the mostly untestable constraints of mythological or theological beliefs. Several possible futures have been predicted by different scientific hypotheses, including that the universe might have existed for a finite and infinite duration, or towards explaining the manner and circumstances of its beginning.

The particle horizon is the maximum distance from which light from particles could have traveled to the observer in the age of the universe. Much like the concept of a terrestrial horizon, it represents the boundary between the observable and the unobservable regions of the universe, so its distance at the present epoch defines the size of the observable universe. Due to the expansion of the universe, it is not simply the age of the universe times the speed of light, but rather the speed of light times the conformal time. The existence, properties, and significance of a cosmological horizon depend on the particular cosmological model.

The Friedmann–Lemaître–Robertson–Walker metric is a metric based on the exact solution of the Einstein field equations of general relativity. The metric describes a homogeneous, isotropic, expanding universe that is path-connected, but not necessarily simply connected. The general form of the metric follows from the geometric properties of homogeneity and isotropy; Einstein's field equations are only needed to derive the scale factor of the universe as a function of time. Depending on geographical or historical preferences, the set of the four scientists – Alexander Friedmann, Georges Lemaître, Howard P. Robertson and Arthur Geoffrey Walker – are variously grouped as Friedmann, Friedmann–Robertson–Walker (FRW), Robertson–Walker (RW), or Friedmann–Lemaître (FL). This model is sometimes called the Standard Model of modern cosmology, although such a description is also associated with the further developed Lambda-CDM model. The FLRW model was developed independently by the named authors in the 1920s and 1930s.

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, and
  3. ordinary matter.
<span class="mw-page-title-main">Friedmann equations</span> Equations in physical cosmology

The Friedmann equations are a set of equations in physical cosmology that govern the expansion of space in homogeneous and isotropic models of the universe within the context of general relativity. They were first derived by Alexander Friedmann in 1922 from Einstein's field equations of gravitation for the Friedmann–Lemaître–Robertson–Walker metric and a perfect fluid with a given mass density ρ and pressure p. The equations for negative spatial curvature were given by Friedmann in 1924.

Primordial fluctuations are density variations in the early universe which are considered the seeds of all structure in the universe. Currently, the most widely accepted explanation for their origin is in the context of cosmic inflation. According to the inflationary paradigm, the exponential growth of the scale factor during inflation caused quantum fluctuations of the inflaton field to be stretched to macroscopic scales, and, upon leaving the horizon, to "freeze in". At the later stages of radiation- and matter-domination, these fluctuations re-entered the horizon, and thus set the initial conditions for structure formation.

In cosmology, the equation of state of a perfect fluid is characterized by a dimensionless number , equal to the ratio of its pressure to its energy density :

In physical cosmology, structure formation is the formation of galaxies, galaxy clusters and larger structures from small early density fluctuations. The universe, as is now known from observations of the cosmic microwave background radiation, began in a hot, dense, nearly uniform state approximately 13.8 billion years ago. However, looking at the night sky today, structures on all scales can be seen, from stars and planets to galaxies. On even larger scales, galaxy clusters and sheet-like structures of galaxies are separated by enormous voids containing few galaxies. Structure formation attempts to model how these structures were formed by gravitational instability of small early ripples in spacetime density or another emergence.

In physical cosmology, cosmological perturbation theory is the theory by which the evolution of structure is understood in the Big Bang model. Cosmological perturbation theory may be broken into two categories: Newtonian or general relativistic. Each case uses its governing equations to compute gravitational and pressure forces which cause small perturbations to grow and eventually seed the formation of stars, quasars, galaxies and clusters. Both cases apply only to situations where the universe is predominantly homogeneous, such as during cosmic inflation and large parts of the Big Bang. The universe is believed to still be homogeneous enough that the theory is a good approximation on the largest scales, but on smaller scales more involved techniques, such as N-body simulations, must be used. When deciding whether to use general relativity for perturbation theory, note that Newtonian physics is only applicable in some cases such as for scales smaller than the Hubble horizon, where spacetime is sufficiently flat, and for which speeds are non-relativistic.

The expansion of the universe is the increase in distance between gravitationally unbound parts of the observable universe with time. It is an intrinsic expansion; the universe does not expand "into" anything and does not require space to exist "outside" it. To any observer in the universe, it appears that all but the nearest galaxies recede at speeds that are proportional to their distance from the observer, on average. While objects cannot move faster than light, this limitation only applies with respect to local reference frames and does not limit the recession rates of cosmologically distant objects.

The deceleration parameter in cosmology is a dimensionless measure of the cosmic acceleration of the expansion of space in a Friedmann–Lemaître–Robertson–Walker universe. It is defined by:

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">Cosmic microwave background spectral distortions</span> Fluctuations in the energy spectrum of the microwave background

CMB spectral distortions are tiny departures of the average cosmic microwave background (CMB) frequency spectrum from the predictions given by a perfect black body. They can be produced by a number of standard and non-standard processes occurring at the early stages of cosmic history, and therefore allow us to probe the standard picture of cosmology. Importantly, the CMB frequency spectrum and its distortions should not be confused with the CMB anisotropy power spectrum, which relates to spatial fluctuations of the CMB temperature in different directions of the sky.

References

  1. 1 2 Peacock, J. A. (1998). Cosmological Physics . Cambridge: Cambridge University Press. ISBN   978-0-521-42270-3.
  2. Robert H. Dicke (1970). Gravitation and the Universe: Jayne Lectures for 1969. American Philosophical Society. ISBN   978-0871690784.
  3. Alan P. Lightman (1 January 1993). Ancient Light: Our Changing View of the Universe. Harvard University Press. ISBN   978-0-674-03363-4.
  4. 1 2 3 Ryden, Barbara (2002). Introduction to Cosmology. San Francisco: Addison Wesley. ISBN   978-0-8053-8912-8.
  5. 1 2 Peter Coles; Francesco Lucchin (1997). Cosmology. Chichester: Wiley. ISBN   978-0-471-95473-6.
  6. 1 2 Liddle, Andrew (2007). An Introduction to Modern Cosmology (2nd ed.). Chichester; Hoboken, NJ: Wiley. p.  157. ISBN   978-0-470-84835-7.
  7. Ryden p. 168
  8. 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. S2CID   119352299.
  9. D. N. Spergel, et al. (June 2007). "Wilkinson Microwave Anisotropy Probe (WMAP) Three Year Results: Implications for Cosmology". Astrophysical Journal Supplement Series. 170 (2): 337–408. arXiv: astro-ph/0603449 . Bibcode:2007ApJS..170..377S. doi:10.1086/513700. S2CID   1386346.
  10. Cain, Fraser; Today, Universe. "How do we know the universe is flat? Discovering the topology of the universe". phys.org. Retrieved 2023-03-26.
  11. darkmatterdarkenergy (2015-03-06). "Planck Mission Full Results Confirm Canonical Cosmology Model". Dark Matter, Dark Energy, Dark Gravity. Retrieved 2023-03-26.
  12. Planck Collaboration; Aghanim, N.; Akrami, Y.; Ashdown, M.; Aumont, J.; Baccigalupi, C.; Ballardini, M.; Banday, A. J.; Barreiro, R. B.; Bartolo, N.; Basak, S.; Battye, R.; Benabed, K.; Bernard, J.-P.; Bersanelli, M. (August 2021). "Planck 2018 results: VI. Cosmological parameters (Corrigendum)". Astronomy & Astrophysics. 652: C4. Bibcode:2021A&A...652C...4P. doi:10.1051/0004-6361/201833910e. hdl: 10902/24951 . ISSN   0004-6361.
  13. Ryden p. 193
  14. 1 2 Agazzi, Evandro; Massimo Pauri (2000). The Reality of the Unobservable: Observability, Unobservability and Their Impact on the Issue of Scientific Realism. Springer. p. 226. Bibcode:2000ruou.book.....A. ISBN   978-0-7923-6311-8.
  15. Helbig, Phillip (December 2021). "Arguments against the flatness problem in classical cosmology: a review" (PDF). European Physical Journal H. 46 (1): 10. Bibcode:2021EPJH...46...10H. doi:10.1140/epjh/s13129-021-00006-9. S2CID   233403196.
  16. 1 2 Collins, C. B.; Hawking, S. (1973). "Why is the Universe Isotropic?". Astrophysical Journal. 180: 317–334. Bibcode:1973ApJ...180..317C. doi: 10.1086/151965 .
  17. Barrow, John D.; Tipler, Frank J. (1986). The Anthropic Cosmological Principle . Oxford: Clarendon Press. p.  411. ISBN   978-0-19-851949-2.
  18. 1 2 3 Mosterín, Jesús (2003). "Anthropic Explanations in Cosmology" . Retrieved 2008-08-01.
  19. 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. S2CID   4363262.
  20. Castelvecchi, Davide (1981). "The Growth of Inflation". Physical Review D . 23 (2): 347. Bibcode:1981PhRvD..23..347G. doi: 10.1103/PhysRevD.23.347 .
  21. 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 .
  22. Brawer, Roberta (February 1996). "Inflationary Cosmology and the Horizon and Flatness Problems: The Mutual Constitution of Explanation and Questions".
  23. 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   978-0-521-56689-6.
  24. Albrecht, Andreas (August 2000). Proceedings of the NATO Advanced Study Institute on Structure Formation in the Universe, Cambridge 1999. Vol. 565. p. 17. arXiv: astro-ph/0007247 . Bibcode:2001ASIC..565...17A. ISBN   978-1-4020-0155-0.
  25. Guth, Alan (1997). "Was Cosmic Inflation the 'Bang' of the Big Bang?". The Beamline. 27. Retrieved 2008-09-07.
  26. 1 2 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. arXiv: 0807.3745 . Bibcode:2008PhRvD..78h3518B. doi:10.1103/PhysRevD.78.083518. S2CID   118432957.
  27. 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. S2CID   15885200.
  28. 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. S2CID   1113031.
  29. Anderson, P. R.; 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.
  30. Evrard, G; P. Coles (October 1995). "Getting the measure of the flatness problem". Classical and Quantum Gravity. 12 (10): L93–L97. arXiv: astro-ph/9507020 . Bibcode:1995CQGra..12L..93E. doi:10.1088/0264-9381/12/10/001. S2CID   14096945..
  31. Holman, Marc (November 2018). "How Problematic is the Near-Euclidean Spatial Geometry of the Large-Scale Universe?". Foundations of Physics. 48 (11): 1617–1647. arXiv: 1803.05148 . Bibcode:2018FoPh...48.1617H. doi:10.1007/s10701-018-0218-4. S2CID   119066780.
  32. Helbig, Phillip (March 2012). "Is there a flatness problem in classical cosmology?". Monthly Notices of the Royal Astronomical Society. 421 (1): 561–569. arXiv: 1112.1666 . Bibcode:2012MNRAS.421..561H. doi:10.1111/j.1365-2966.2011.20334.x. S2CID   85526633.
  33. Lake, Kayll (May 2005). "The Flatness Problem and Λ". Physical Review Letters. 94 (20): 201102. arXiv: astro-ph/0404319 . Bibcode:2005PhRvL..94t1102L. doi:10.1103/PhysRevLett.94.201102. PMID   16090234. S2CID   40500958.
  34. Poplawski, N. J. (2010). "Cosmology with torsion: An alternative to cosmic inflation". Phys. Lett. B. 694 (3): 181–185. arXiv: 1007.0587 . Bibcode:2010PhLB..694..181P. doi:10.1016/j.physletb.2010.09.056.
  35. Poplawski, N. (2012). "Nonsingular, big-bounce cosmology from spinor-torsion coupling". Phys. Rev. D. 85 (10): 107502. arXiv: 1111.4595 . Bibcode:2012PhRvD..85j7502P. doi:10.1103/PhysRevD.85.107502. S2CID   118434253.