Cosmological perturbation theory

Last updated

In physical cosmology, cosmological perturbation theory [1] [2] [3] [4] [5] 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.

Contents

Because of the gauge invariance of general relativity, the correct formulation of cosmological perturbation theory is subtle. In particular, when describing an inhomogeneous spacetime, there is often not a preferred coordinate choice. There are currently two distinct approaches to perturbation theory in classical general relativity:

Newtonian perturbation theory

In this section, we will focus on the effect of matter on structure formation in the hydrodynamical fluid regime. This regime is useful because dark matter has dominated structure growth for most of the universe's history. In this regime, we are on sub-Hubble scales (where is the Hubble parameter) so we can take spacetime to be flat, and ignore general relativistic corrections. But these scales are above a cut-off, such that perturbations in pressure and density are sufficiently linear Next we assume low pressure so that we can ignore radiative effects and low speeds so we are in the non-relativistic regime.

The first governing equation follows from matter conservation – the continuity equation [6]

where is the scale factor and is the peculiar velocity. Although we don't explicitly write it, all variables are evaluated at time and the divergence is in comoving coordinates. Second, momentum conservation gives us the Euler equation

where is the gravitational potential. Lastly, we know that for Newtonian gravity, the potential obeys the Poisson equation

So far, our equations are fully nonlinear, and can be hard to interpret intuitively. It's therefore useful to consider a perturbative expansion and examine each order separately. We use the following decomposition

where is a comoving coordinate.

At linear order, the continuity equation becomes

where is the velocity divergence. And the linear Euler equation is

By combining the linear continuity, Euler, and Poisson equations, we arrive at a simple master equation governing evolution

where we defined a sound speed to give us a closure relation. This master equation admits wave solutions in which tell us how matter fluctuations grow over time due to a combination of competing effects – the fluctuation's self-gravity, pressure forces, the universe's expansion, and the background gravitational field.

Gauge-invariant perturbation theory

The gauge-invariant perturbation theory is based on developments by Bardeen (1980), [7] Kodama and Sasaki (1984) [8] building on the work of Lifshitz (1946). [9] This is the standard approach to perturbation theory of general relativity for cosmology. [10] This approach is widely used for the computation of anisotropies in the cosmic microwave background radiation [11] as part of the physical cosmology program and focuses on predictions arising from linearisations that preserve gauge invariance with respect to Friedmann-Lemaître-Robertson-Walker (FLRW) models. This approach draws heavily on the use of Newtonian like analogue and usually has as it starting point the FRW background around which perturbations are developed. The approach is non-local and coordinate dependent but gauge invariant as the resulting linear framework is built from a specified family of background hyper-surfaces which are linked by gauge preserving mappings to foliate the space-time. Although intuitive this approach does not deal well with the nonlinearities natural to general relativity.

1+3 covariant gauge-invariant perturbation theory

In relativistic cosmology using the Lagrangian threading dynamics of Ehlers (1971) [12] and Ellis (1971) [13] it is usual to use the gauge-invariant covariant perturbation theory developed by Hawking (1966) [14] and Ellis and Bruni (1989). [15] Here rather than starting with a background and perturbing away from that background one starts with full general relativity and systematically reduces the theory down to one that is linear around a particular background. [16] The approach is local and both covariant as well as gauge invariant but can be non-linear because the approach is built around the local comoving observer frame (see frame bundle) which is used to thread the entire space-time. This approach to perturbation theory produces differential equations that are of just the right order needed to describe the true physical degrees of freedom and as such no non-physical gauge modes exist. It is usual to express the theory in a coordinate free manner. For applications of kinetic theory, because one is required to use the full tangent bundle, it becomes convenient to use the tetrad formulation of relativistic cosmology. The application of this approach to the computation of anisotropies in cosmic microwave background radiation [17] requires the linearization of the full relativistic kinetic theory developed by Thorne (1980) [18] and Ellis, Matravers and Treciokas (1983). [19]

Gauge freedom and frame fixing

In relativistic cosmology there is a freedom associated with the choice of threading frame; this frame choice is distinct from the choice associated with coordinates. Picking this frame is equivalent to fixing the choice of timelike world lines mapped into each other. This reduces the gauge freedom; it does not fix the gauge but the theory remains gauge invariant under the remaining gauge freedoms. In order to fix the gauge a specification of correspondences between the time surfaces in the real universe (perturbed) and the background universe are required along with the correspondences between points on the initial spacelike surfaces in the background and in the real universe. This is the link between the gauge-invariant perturbation theory and the gauge-invariant covariant perturbation theory. Gauge invariance is only guaranteed if the choice of frame coincides exactly with that of the background; usually this is trivial to ensure because physical frames have this property.

Newtonian-like equations

Newtonian-like equations emerge from perturbative general relativity with the choice of the Newtonian gauge; the Newtonian gauge provides the direct link between the variables typically used in the gauge-invariant perturbation theory and those arising from the more general gauge-invariant covariant perturbation theory.

See also

Related Research Articles

<span class="mw-page-title-main">Stress–energy tensor</span> Tensor describing energy momentum density in spacetime

The stress–energy tensor, sometimes called the stress–energy–momentum tensor or the energy–momentum tensor, is a tensor physical quantity that describes the density and flux of energy and momentum in spacetime, generalizing the stress tensor of Newtonian physics. It is an attribute of matter, radiation, and non-gravitational force fields. This density and flux of energy and momentum are the sources of the gravitational field in the Einstein field equations of general relativity, just as mass density is the source of such a field in Newtonian gravity.

In the general theory of relativity, the Einstein field equations relate the geometry of spacetime to the distribution of matter within it.

In relativistic physics, Lorentz symmetry or Lorentz invariance, named after the Dutch physicist Hendrik Lorentz, is an equivalence of observation or observational symmetry due to special relativity implying that the laws of physics stay the same for all observers that are moving with respect to one another within an inertial frame. It has also been described as "the feature of nature that says experimental results are independent of the orientation or the boost velocity of the laboratory through space".

A classical field theory is a physical theory that predicts how one or more physical fields interact with matter through field equations, without considering effects of quantization; theories that incorporate quantum mechanics are called quantum field theories. In most contexts, 'classical field theory' is specifically intended to describe electromagnetism and gravitation, two of the fundamental forces of nature.

The Kerr–Newman metric is the most general asymptotically flat, stationary solution of the Einstein–Maxwell equations in general relativity that describes the spacetime geometry in the region surrounding an electrically charged, rotating mass. It generalizes the Kerr metric by taking into account the field energy of an electromagnetic field, in addition to describing rotation. It is one of a large number of various different electrovacuum solutions, that is, of solutions to the Einstein–Maxwell equations which account for the field energy of an electromagnetic field. Such solutions do not include any electric charges other than that associated with the gravitational field, and are thus termed vacuum solutions.

When studying and formulating Albert Einstein's theory of general relativity, various mathematical structures and techniques are utilized. The main tools used in this geometrical theory of gravitation are tensor fields defined on a Lorentzian manifold representing spacetime. This article is a general description of the mathematics of general relativity.

In theoretical physics, Nordström's theory of gravitation was a predecessor of general relativity. Strictly speaking, there were actually two distinct theories proposed by the Finnish theoretical physicist Gunnar Nordström, in 1912 and 1913 respectively. The first was quickly dismissed, but the second became the first known example of a metric theory of gravitation, in which the effects of gravitation are treated entirely in terms of the geometry of a curved spacetime.

In physics, precisely in the study of the theory of general relativity and many alternatives to it, the post-Newtonian formalism is a calculational tool that expresses Einstein's (nonlinear) equations of gravity in terms of the lowest-order deviations from Newton's law of universal gravitation. This allows approximations to Einstein's equations to be made in the case of weak fields. Higher-order terms can be added to increase accuracy, but for strong fields, it may be preferable to solve the complete equations numerically. Some of these post-Newtonian approximations are expansions in a small parameter, which is the ratio of the velocity of the matter forming the gravitational field to the speed of light, which in this case is better called the speed of gravity. In the limit, when the fundamental speed of gravity becomes infinite, the post-Newtonian expansion reduces to Newton's law of gravity.

Scalar theories of gravitation are field theories of gravitation in which the gravitational field is described using a scalar field, which is required to satisfy some field equation.

In theoretical physics, a scalar–tensor theory is a field theory that includes both a scalar field and a tensor field to represent a certain interaction. For example, the Brans–Dicke theory of gravitation uses both a scalar field and a tensor field to mediate the gravitational interaction.

In general relativity, the Newtonian gauge is a perturbed form of the Friedmann–Lemaître–Robertson–Walker line element. The gauge freedom of general relativity is used to eliminate two scalar degrees of freedom of the metric, so that it can be written as:

In cosmological perturbation theory, the scalar–vector–tensor decomposition is a decomposition of the most general linearized perturbations of the Friedmann–Lemaître–Robertson–Walker metric into components according to their transformations under spatial rotations. It was first discovered by E. M. Lifshitz in 1946. It follows from Helmholtz's Theorem The general metric perturbation has ten degrees of freedom. The decomposition states that the evolution equations for the most general linearized perturbations of the Friedmann–Lemaître–Robertson–Walker metric can be decomposed into four scalars, two divergence-free spatial vector fields, and a traceless, symmetric spatial tensor field with vanishing doubly and singly longitudinal components. The vector and tensor fields each have two independent components, so this decomposition encodes all ten degrees of freedom in the general metric perturbation. Using gauge invariance four of these components may be set to zero.

<span class="mw-page-title-main">Post-Newtonian expansion</span> Method of approximation in general relativity

In general relativity, post-Newtonian expansions are used for finding an approximate solution of Einstein field equations for the metric tensor. The approximations are expanded in small parameters that express orders of deviations from Newton's law of universal gravitation. This allows approximations to Einstein's equations to be made in the case of weak fields. Higher-order terms can be added to increase accuracy, but for strong fields sometimes it is preferable to solve the complete equations numerically. This method is a common mark of effective field theories. In the limit, when the small parameters are equal to 0, the post-Newtonian expansion reduces to Newton's law of gravity.

In theoretical physics, scalar field theory can refer to a relativistically invariant classical or quantum theory of scalar fields. A scalar field is invariant under any Lorentz transformation.

The Newman–Penrose (NP) formalism is a set of notation developed by Ezra T. Newman and Roger Penrose for general relativity (GR). Their notation is an effort to treat general relativity in terms of spinor notation, which introduces complex forms of the usual variables used in GR. The NP formalism is itself a special case of the tetrad formalism, where the tensors of the theory are projected onto a complete vector basis at each point in spacetime. Usually this vector basis is chosen to reflect some symmetry of the spacetime, leading to simplified expressions for physical observables. In the case of the NP formalism, the vector basis chosen is a null tetrad: a set of four null vectors—two real, and a complex-conjugate pair. The two real members often asymptotically point radially inward and radially outward, and the formalism is well adapted to treatment of the propagation of radiation in curved spacetime. The Weyl scalars, derived from the Weyl tensor, are often used. In particular, it can be shown that one of these scalars— in the appropriate frame—encodes the outgoing gravitational radiation of an asymptotically flat system.

Alternatives to general relativity are physical theories that attempt to describe the phenomenon of gravitation in competition with Einstein's theory of general relativity. There have been many different attempts at constructing an ideal theory of gravity.

Newton–Cartan theory is a geometrical re-formulation, as well as a generalization, of Newtonian gravity first introduced by Élie Cartan and Kurt Friedrichs and later developed by Dautcourt, Dixon, Dombrowski and Horneffer, Ehlers, Havas, Künzle, Lottermoser, Trautman, and others. In this re-formulation, the structural similarities between Newton's theory and Albert Einstein's general theory of relativity are readily seen, and it has been used by Cartan and Friedrichs to give a rigorous formulation of the way in which Newtonian gravity can be seen as a specific limit of general relativity, and by Jürgen Ehlers to extend this correspondence to specific solutions of general relativity.

f(R) is a type of modified gravity theory which generalizes Einstein's general relativity. f(R) gravity is actually a family of theories, each one defined by a different function, f, of the Ricci scalar, R. The simplest case is just the function being equal to the scalar; this is general relativity. As a consequence of introducing an arbitrary function, there may be freedom to explain the accelerated expansion and structure formation of the Universe without adding unknown forms of dark energy or dark matter. Some functional forms may be inspired by corrections arising from a quantum theory of gravity. f(R) gravity was first proposed in 1970 by Hans Adolph Buchdahl. It has become an active field of research following work by Starobinsky on cosmic inflation. A wide range of phenomena can be produced from this theory by adopting different functions; however, many functional forms can now be ruled out on observational grounds, or because of pathological theoretical problems.

<span class="mw-page-title-main">Gauge theory</span> Physical theory with fields invariant under the action of local "gauge" Lie groups

In physics, a gauge theory is a type of field theory in which the Lagrangian, and hence the dynamics of the system itself, do not change under local transformations according to certain smooth families of operations. Formally, the Lagrangian is invariant.

Lagrangian field theory is a formalism in classical field theory. It is the field-theoretic analogue of Lagrangian mechanics. Lagrangian mechanics is used to analyze the motion of a system of discrete particles each with a finite number of degrees of freedom. Lagrangian field theory applies to continua and fields, which have an infinite number of degrees of freedom.

References

  1. Fry, J. N. (April 1984). "The Galaxy correlation hierarchy in perturbation theory". The Astrophysical Journal. 279: 499. Bibcode:1984ApJ...279..499F. doi: 10.1086/161913 .
  2. Bharadwaj, Somnath (June 1994). "Perturbative growth of cosmological clustering. I: Formalism". The Astrophysical Journal. 428: 419. Bibcode:1994ApJ...428..419B. doi:10.1086/174254. ISSN   0004-637X.
  3. Bharadwaj, Somnath (March 1996). "Perturbative Growth of Cosmological Clustering. II. The Two-Point Correlation". The Astrophysical Journal. 460: 28–50. arXiv: astro-ph/9511085 . Bibcode:1996ApJ...460...28B. doi:10.1086/176950. S2CID   17179734.
  4. Bharadwaj, Somnath (20 November 1996). "The Evolution of Correlation Functions in the Zeldovich Approximation and Its Implications for the Validity of Perturbation Theory". The Astrophysical Journal. 472 (1): 1–13. arXiv: astro-ph/9606121 . Bibcode:1996ApJ...472....1B. doi: 10.1086/178036 .
  5. Dodelson, Scott; Schmidt, Fabian (2020). Modern Cosmology (2 ed.). Academic Press. Bibcode:2020moco.book.....D. doi:10.1016/C2017-0-01943-2. ISBN   978-0-12-815948-4. S2CID   241570171.
  6. Baumann, Daniel (2022). Cosmology. Cambridge University Press. doi:10.1017/9781108937092. ISBN   9781108838078.
  7. Bardeen, James M. (1980-10-15). "Gauge-invariant cosmological perturbations". Physical Review D. American Physical Society (APS). 22 (8): 1882–1905. Bibcode:1980PhRvD..22.1882B. doi:10.1103/physrevd.22.1882. ISSN   0556-2821.
  8. Kodama, Hideo; Sasaki, Misao (1984). "Cosmological Perturbation Theory". Progress of Theoretical Physics Supplement. Oxford University Press (OUP). 78: 1–166. Bibcode:1984PThPS..78....1K. doi: 10.1143/ptps.78.1 . ISSN   0375-9687.
  9. Lifshitz E M (1946) J. Phys. (USSR), 10, 116
  10. Mukhanov, V (1992). "Theory of cosmological perturbations". Physics Reports. Elsevier BV. 215 (5–6): 203–333. Bibcode:1992PhR...215..203M. doi:10.1016/0370-1573(92)90044-z. ISSN   0370-1573.
  11. Hu W, Sugiyama N (1995). "Toward Understanding CMB Anisotropies and Their Implications". Physical Review D. 51 (6): 2599–2630. arXiv: astro-ph/9411008 . Bibcode:1995PhRvD..51.2599H. doi:10.1103/PhysRevD.51.2599. PMID   10018735. S2CID   12811112.
  12. Ehlers J (1971) General Relativity and Cosmology (Varenna), R K Sachs (Academic Press NY)
  13. Ellis G F R, (1971) General Relativity and Cosmology(Varenna), R K Sachs (Academic Press NY)
  14. Hawking S W (1966) ApJ. 145, 44
  15. Ellis, G. F. R.; Bruni, M. (1989-09-15). "Covariant and gauge-invariant approach to cosmological density fluctuations". Physical Review D. American Physical Society (APS). 40 (6): 1804–1818. Bibcode:1989PhRvD..40.1804E. doi:10.1103/physrevd.40.1804. ISSN   0556-2821. PMID   10012011.
  16. Tsagas, C. G.; Challinor, A; Maartens, R (2008). "Relativistic cosmology and large-scale structure". Physics Reports. 465 (2–3): 61–147. arXiv: 0705.4397 . Bibcode:2008PhR...465...61T. doi:10.1016/j.physrep.2008.03.003. ISSN   0370-1573. S2CID   119121482.
  17. Maartens R, Gebbie T, Ellis GF (1999). "Cosmic microwave background anisotropies: Nonlinear dynamics". Physical Review D. 59 (8): 083506. arXiv: astro-ph/9808163 . Bibcode:1999PhRvD..59h3506M. doi:10.1103/PhysRevD.59.083506. S2CID   119444449.
  18. Thorne, Kip S. (1980-04-01). "Multipole expansions of gravitational radiation" (PDF). Reviews of Modern Physics. American Physical Society (APS). 52 (2): 299–339. Bibcode:1980RvMP...52..299T. doi:10.1103/revmodphys.52.299. ISSN   0034-6861.
  19. Ellis, G.F.R; Treciokas, R; Matravers, D.R (1983). "Anisotropic solutions of the Einstein-Boltzmann equations. II. Some exact properties of the equations". Annals of Physics. Elsevier BV. 150 (2): 487–503. Bibcode:1983AnPhy.150..487E. doi:10.1016/0003-4916(83)90024-6. ISSN   0003-4916.

Bibliography

See physical cosmology textbooks.