Fact-checked by Grok 2 weeks ago

Distance measure

In cosmology, distance measures quantify the separation between astronomical objects, accounting for the , which complicates direct application of . Unlike static spaces, the universe's —described by the Friedmann–Lemaître–Robertson–Walker (FLRW) —means that light from distant objects travels through an evolving geometry, leading to multiple specialized distance definitions based on observables like , angular size, and . These measures arose historically from efforts to interpret observations of galaxies and supernovae. Early work by in the 1920s established the linear velocity-distance relation (), but as cosmology developed in the mid-20th century with and the model, theorists like Weinberg (1972) and (1993) formalized distances incorporating curvature and . Today, they are crucial for mapping cosmic structure, testing models like ΛCDM, and measuring parameters such as the Hubble constant. Key line-of-sight measures include the comoving distance (invariant for expanding objects) and proper distance (at a given ), while transverse measures encompass (relating physical to observed angular size) and (relating intrinsic to observed brightness). These are interconnected via the scale factor and z, where $1 + z = 1/a(t) with a(t) the scale factor, and obey Etherington's reciprocity relation in smooth spacetimes. Later sections detail these in the FLRW framework, corrections for peculiar velocities, and their observational applications.

Introduction

Definition and context

In cosmology, the conventional Euclidean distance, which assumes a static, flat geometry, fails to accurately describe separations between objects in a curved and expanding spacetime governed by general relativity, as the distances between comoving objects continuously evolve due to cosmic expansion. This necessitates multiple specialized distance measures, each derived from distinct observables such as the observed flux of radiation, the angular diameter of celestial objects, and time delays in propagating signals, to infer physical separations in an evolving universe. A distance measure is formally defined as a function that connects the z—a key indicator of cosmic expansion—to the effective physical distance between sources in Friedmann-Lemaître-Robertson-Walker (FLRW) models, which parameterize the 's large-scale geometry. These models assume the universe is homogeneous, meaning any measurable remains uniform when averaged over sufficiently large scales, and isotropic, appearing the same in all directions from any vantage point, collectively known as the . Underpinning this framework is , which provides the gravitational dynamics for the expansion. Redshift z acts as a primary for distance, encoding the cumulative expansion along the line of sight. For example, in a flat, non-expanding , all distance measures reduce to a single invariant quantity, but in cosmology, expansion causes them to diverge, reflecting the dynamic interplay of geometry and evolution.

Historical development

The concept of distance measures in cosmology emerged in the early alongside the recognition of the 's expansion. In 1929, published observations demonstrating a linear relationship between the radial velocities of galaxies and their distances, establishing the foundation for initial distance estimates based on recession velocities. This discovery, derived from spectroscopic data of 24 extra-galactic nebulae, implied an expanding and provided the first empirical framework for scaling cosmic distances beyond the . During the 1930s, the theoretical underpinnings advanced with the independent work of Howard Robertson and Arthur Walker, who formalized the Friedmann-Lemaître-Robertson-Walker (FLRW) metric, introducing comoving coordinates to describe distances in an expanding, homogeneous, and isotropic universe. These coordinates, which remain fixed relative to the while physical distances scale with the universe's , enabled consistent distance calculations across different epochs. Concurrently, concepts like were formalized in early relativistic cosmological models, initially explored in kinematic frameworks of the era, and later adapted to the paradigm following its acceptance in the 1940s and 1950s. In the 1940s, and Rudolf Minkowski advanced practical distance measurements by classifying supernovae into Type I and II based on spectral observations, leveraging their as standard candles to introduce estimates for distant galaxies. In 1998, observations of Type supernovae by the High-Z Supernova Search Team and Supernova Cosmology Project revealed that the universe's is accelerating, using to provide evidence for and refining cosmological parameters. The late 20th century saw distance measures integrated with () observations, providing geometric constraints on . The Cosmic Background Explorer (COBE) mission in 1992 detected temperature anisotropies, providing initial constraints on cosmic geometry. Subsequent ground-based and satellite experiments in the late 1990s and 2000s offered the first precise measurements of the to the last scattering surface at z ≈ 1100, which refined models of cosmic expansion. Subsequent Planck satellite data from 2013 to 2018 further tightened these constraints, yielding a Hubble constant of 67.4 ± 0.5 km/s/Mpc and highlighting the Hubble tension—a discrepancy between CMB-inferred values and local measurements around 73 km/s/Mpc. By the 2020s, (JWST) observations of high-redshift (high-z) galaxies and supernovae have refined distance measures at z > 10, revealing unexpectedly luminous early galaxies that challenge pre-2020 dark energy models and suggest possible modifications to the or . These data, including detections up to z ≈ 14, have addressed gaps in high-z distance ladders, enhancing precision in luminosity and angular diameter distances while probing dynamics.

Cosmological Foundations

FLRW metric and expansion

The Friedmann–Lemaître–Robertson–Walker (FLRW) metric provides the mathematical framework for describing a homogeneous and isotropic universe in general relativity, serving as the foundation for modern cosmological distance measures. This metric arises from the , which posits that the universe is homogeneous ( on large scales) and isotropic (appearing the same in from any point), implying a spacetime symmetry that constrains the possible forms of the . To derive it, consider a coordinate system adapted to comoving observers—those at rest relative to the expanding cosmic background—where the spatial coordinates remain fixed while the proper distances between them evolve with time. The line element in such a spacetime takes the general form ds^2 = -c^2 dt^2 + g_{ij} dx^i dx^j, with the spatial part g_{ij} reflecting isotropy and homogeneity. For isotropy around a point, the spatial must be that of a three-dimensional hypersurface of constant curvature, leading to the form dl^2 = \frac{dr^2}{1 - k r^2} + r^2 (d\theta^2 + \sin^2\theta d\phi^2), where r is a comoving radial coordinate, d\Omega^2 = d\theta^2 + \sin^2\theta d\phi^2 is the on the unit sphere, and k is the curvature parameter (with k = +1, 0, -1 corresponding to closed, flat, and open geometries, respectively, up to a normalization of the curvature radius). Allowing for time-dependent , the full becomes ds^2 = -c^2 dt^2 + a(t)^2 \left[ \frac{dr^2}{1 - k r^2} + r^2 d\Omega^2 \right], where a(t) is the dimensionless scale factor that encodes the universe's history, normalized such that a(t_0) = 1 today. The scale factor a(t) drives the , quantified by the H(t) = \frac{\dot{a}(t)}{a(t)}, which measures the fractional rate of change of distances at time t. This parameter relates the dynamics of the universe to its energy content through the , derived by applying Einstein's field equations to the FLRW . The first is H^2 = \left( \frac{\dot{a}}{a} \right)^2 = \frac{8\pi G}{3} \rho - \frac{k c^2}{a^2} + \frac{\Lambda c^2}{3}, where \rho is the total energy density (including matter, radiation, and dark energy), G is Newton's constant, \Lambda is the cosmological constant, and the curvature term -\frac{k c^2}{a^2} reflects the geometry's influence on expansion. The second Friedmann equation, \frac{\ddot{a}}{a} = -\frac{4\pi G}{3} \left( \rho + \frac{3p}{c^2} \right) + \frac{\Lambda c^2}{3}, governs acceleration, with p as pressure; for matter-dominated eras (p = 0), expansion decelerates, while dark energy (p = -\rho c^2) drives late-time acceleration. These equations link H(t) directly to the densities of components: matter \rho_m \propto a^{-3}, radiation \rho_r \propto a^{-4}, and dark energy \rho_\Lambda constant in the standard model. The parameter k plays a crucial role in distance calculations by determining the global geometry: positive k = +1 implies a closed, finite ; k = 0 a flat, infinite one; and k = -1 an open, . In the \LambdaCDM model, observations favor a spatially flat (k = 0), where the parameter \Omega_k = -\frac{k c^2}{H_0^2 a^2} \approx 0, as evidenced by cosmic microwave background anisotropies and large-scale structure data. This flatness simplifies distance measures, assuming Euclidean geometry on large scales, though non-zero k would introduce corrections scaling with redshift. For light propagation, which underlies astronomical distance measures, photons follow null geodesics (ds = 0) in the FLRW . Setting ds = 0 and assuming radial motion (d\Omega = 0), the geodesic equation simplifies to c dt = \pm a(t) \frac{dr}{\sqrt{1 - k r^2}}. Integrating along the past from emission at t_e to observation at t_0, this yields the comoving distance in terms of conformal time \eta, defined as d\eta = \frac{c dt}{a(t)} or \eta = \int_{t_e}^{t_0} \frac{c dt}{a(t)}, which rescales the metric to a conformally static form ds^2 = a(\eta)^2 (-c^2 d\eta^2 + d\chi^2 + \sinh^2(\sqrt{|k|} \chi) d\Omega^2 ) (adjusted for curvature), facilitating solutions for photon paths.

Redshift and scale factor

In cosmology, the redshift z quantifies the stretching of photon wavelengths due to the universe's expansion, defined as $1 + z = \frac{\lambda_\mathrm{obs}}{\lambda_\mathrm{emit}}, where \lambda_\mathrm{obs} is the observed wavelength and \lambda_\mathrm{emit} is the emitted wavelength. This cosmological relates directly to the cosmic scale factor a(t), normalized such that a(t_0) = 1 today, via $1 + z = \frac{a(t_0)}{a(t_e)} = \frac{1}{a(t_e)}, where t_e is the emission time. This relation arises because the expansion stretches the wavelength proportionally to the increase in the scale factor between emission and observation. Redshift phenomena include three main types: Doppler redshift from relative peculiar motions of objects, gravitational redshift from differences in , and cosmological redshift from the overall of . In the context of large-scale , gravitational effects are typically negligible compared to expansion, and peculiar velocities contribute only small corrections at low , making cosmological redshift the primary mechanism for observed shifts in distant galaxies and quasars. The Doppler component, while important for local motions, diminishes in significance over cosmological distances where expansion dominates. The connection between and recession is approximated at low z (z \ll [1](/page/1)) by v \approx c z, forming the basis of , which links to distance via the Hubble parameter. For higher redshifts, this non-relativistic approximation fails as recession speeds can exceed c without violating , since the is a coordinate effect in the Friedmann-Lemaître-Robertson-Walker (FLRW) framework; the full relation requires integrating the scale factor's over the history. This integral form accounts for the cumulative stretching along the photon's path, providing a more accurate for distant objects. Cosmological redshifts are measured primarily through , identifying shifts in or lines such as the line at rest wavelength 121.6 , which appears at observed wavelengths \lambda_\mathrm{obs} = 121.6 (1 + [z](/page/Z)) . High-resolution spectra from quasars and galaxies allow precise z , mapping of the large-scale structure. For the (), the blackbody temperature scales inversely with the scale factor as T(z) = T_0 (1 + [z](/page/Z)), where T_0 \approx 2.725 K today, providing an independent redshift probe at z \approx 1100. The observed z is essential for measures, as it fixes the lookback time—the elapsed time since emission—through over the expansion rate H(z), and encodes the cumulative expansion history needed to compute in the FLRW model. Without knowledge of z, the scale factor at emission cannot be inferred, rendering subsequent calculations impossible.

Line-of-Sight Distance Measures

Comoving distance

The comoving distance, often denoted as \chi(z), represents the line-of-sight separation between two events in comoving coordinates, which remains under the cosmic expansion. It is defined as the along the photon's path in an expanding universe, given by \chi(z) = \int_0^z \frac{c \, dz'}{H(z')}, where c is the and H(z) is the Hubble parameter at z'. This distance corresponds to the proper distance that two comoving objects (those moving solely with the Hubble flow) would have today if extrapolated backward without accounting for expansion effects. In a flat \LambdaCDM universe, the Hubble parameter takes the form H(z) = H_0 \sqrt{\Omega_m (1+z)^3 + \Omega_\Lambda}, where H_0 is the present-day Hubble constant, \Omega_m is the present-day matter density parameter, and \Omega_\Lambda is the present-day density parameter (with \Omega_m + \Omega_\Lambda = 1 for flatness). Substituting this into the integral yields the specific expression for \chi(z) in this model. Galaxies or other structures at fixed comoving distance \chi maintain constant separation in these coordinates over , providing a fixed "" for mapping the large-scale structure of the . For universes with nonzero spatial curvature (k \neq 0), the comoving distance \chi(z) is still computed via the same integral form, but H(z) now includes a term: H(z) = H_0 \sqrt{\Omega_m (1+z)^3 + \Omega_k (1+z)^2 + \Omega_\Lambda}, where \Omega_k = -k c^2 / (H_0^2 a_0^2) measures the (k > 0 for closed, k < 0 for open, and a_0 = 1 today). In the general Friedmann-Lemaître-Robertson-Walker (FLRW) metric, the line-of-sight proper distance at any epoch is a(t) \chi, but the comoving \chi itself absorbs the expansion history. While the radial \chi remains the integral, the full spatial separation in curved geometries involves -dependent functions applied to \chi, such as the sk(\chi) term in the metric (sinh for open and sin for closed universes). This measure serves as the foundation for computing comoving volume elements in cosmological surveys, where the differential volume is dV_c = \chi^2 \, d\chi \, d\Omega in flat space, enabling the estimation of galaxy number densities and clustering statistics. It is particularly crucial in baryon acoustic oscillation (BAO) analyses, where the comoving sound horizon scale at recombination (\sim 150 \, \mathrm{Mpc}) acts as a standard ruler; by measuring the observed angular or redshift separation of this feature, surveys infer \chi(z) and constrain cosmological parameters like H_0 and \Omega_m.

Proper distance

The proper distance, denoted d_p(z), is the physical distance between an observer and an object at redshift z, measured at a fixed cosmic time—specifically, the present epoch—along a radial line in the Friedmann-Lemaître-Robertson-Walker (FLRW) metric. It corresponds to the length that would be measured by a set of rigidly held rods at the current scale factor a(t_0) = 1. In a flat universe, the formula is d_p(z) = \chi(z), where \chi(z) is the comoving distance given by the integral \chi(z) = \int_0^z \frac{c \, dz'}{H(z')}, with H(z) = H_0 \sqrt{\Omega_m (1+z')^3 + \Omega_\Lambda} for a flat \LambdaCDM model (assuming \Omega_k = 0). This measure distinguishes itself from the comoving distance by incorporating the expansion up to the present day, yielding the current physical separation rather than a time-independent coordinate. At the object's emission time t_e, the proper distance scales as d_p(t_e) = a(t_e) \chi(z) = \frac{\chi(z)}{1+z}, highlighting how physical rulers were shorter in the past due to the smaller scale factor. The comoving distance \chi(z) forms the foundational integral, equivalent to \int_{t_e}^{t_0} \frac{c \, dt}{a(t)}, which proper distance at present directly adopts since a(t_0) = 1. Proper distance finds key applications in galaxy clustering analyses, where it quantifies the present-day physical scales of overdensities and voids, enabling tests of structure growth models against observations like the two-point correlation function in proper coordinates. For nearby objects (z \ll 1), it supports proper motion distance estimates, linking observed angular displacements to transverse peculiar velocities, as in kinematic studies of extragalactic jets.

Light-travel distance

The light-travel distance d_{LT}(z) to a source at redshift z is defined as d_{LT}(z) = \int_0^z \frac{c \, dz'}{H(z') (1 + z')}, where c is the speed of light and H(z) is the Hubble parameter as a function of redshift. This distance measure represents the path length traversed by the light ray from the source to the observer, accounting for the cosmic expansion during transit; it is equal to c times the lookback time, the coordinate time interval between emission and observation. Unlike the proper distance d_p(z), which is the physical separation at the current epoch and given by d_p(z) = \int_0^z \frac{c \, dz'}{H(z')}, the light-travel distance incorporates the varying scale factor along the light's path, resulting in d_{LT}(z) < d_p(z) for z > 0 in an expanding . The light-travel distance finds applications in estimating of the , obtained as the limit d_{LT}(\infty)/c, which integrates the cosmic history from the to the present. It also informs the timing of light curves, where the observed duration is stretched by a factor of (1 + z) due to , allowing models to align emission events with the lookback time derived from d_{LT}(z).

Transverse Distance Measures

Transverse comoving distance

The transverse comoving distance, denoted d_M(z), measures the proper distance between two points at redshift z that are separated by a small angular displacement perpendicular to the line of sight, in a coordinate system that expands with the universe's Hubble flow. It generalizes the line-of-sight comoving distance to transverse directions, remaining fixed over cosmic time for objects comoving with the expansion. In a flat universe, d_M(z) equals the line-of-sight comoving \chi(z) = \int_0^z \frac{c \, dz'}{H(z')}, where c is the and H(z) is the Hubble parameter at z'. For curved universes, modifies this relation to account for the of . The general expression is d_M(z) = \frac{c}{H_0 \sqrt{|\Omega_k|}} \sinh\left( \sqrt{|\Omega_k|} \int_0^z \frac{H_0 \, dz'}{H(z')} \right) for an open universe (\Omega_k > 0), where H_0 is the present-day Hubble and \Omega_k is the density parameter; the hyperbolic sine is replaced by the sine function for a closed universe (\Omega_k < 0). In the flat-space limit (\Omega_k = [0](/page/0)), this reduces to \chi(z). This distance plays a crucial role in interpreting observations of angular scales in the large-scale structure of the universe, such as the angular size of (BAO) in galaxy surveys, where it relates observed angular separations to comoving scales. It is also essential for gravitational lensing calculations, providing the transverse separation needed to compute deflection angles and magnification effects between lens and source. At high redshifts, the effects of spatial become pronounced, altering the transverse separation: in (open) geometries, distances grow more rapidly than in flat space, while in spherical (closed) geometries, they are compressed due to the positive . These deviations allow constraints on \Omega_k from transverse measurements, complementing radial probes.

Angular diameter distance

The , denoted d_A(z), relates the physical transverse size l of an object at z to its observed size \theta (in radians) via the relation \theta = l / d_A(z) for small angles. This measure is essential for interpreting the apparent sizes of cosmological objects as affected by the universe's . In the Friedmann–Lemaître–Robertson–Walker (FLRW) metric, the line element for transverse separations is ds^2 = a^2(t_e) \left[ d\chi^2 + f_K(\chi)^2 d\Omega^2 \right], where a(t_e) = 1/(1+[z](/page/Z)) is the scale factor at emission time t_e, \chi is the radial comoving coordinate, and f_K(\chi) accounts for spatial (with f_K(\chi) = \chi in a flat universe). For a small separation d\theta, the proper transverse size at emission is l = a(t_e) f_K(\chi) d\theta, leading to d_A(z) = l / d\theta = a(t_e) f_K(\chi). Since the transverse comoving distance is d_M(z) = f_K(\chi), the follows as d_A(z) = \frac{d_M(z)}{1 + [z](/page/Z)}. The factor of (1+[z](/page/Z))^{-1} arises from the scale factor at emission, which diminishes the proper size relative to the comoving frame due to . Physically, d_A(z) incorporates the effects of cosmic expansion on angular appearances, causing distant objects to subtend larger angles than in a static universe beyond a certain redshift. In decelerating models like Einstein–de Sitter, d_A(z) increases with z up to a maximum around z \approx 1.25, then decreases at higher redshifts; in accelerating \LambdaCDM universes dominated by dark energy, the maximum shifts to around z \approx 1.6, with a slower decrease thereafter. This shift in the maximum provides a geometric probe of acceleration. Applications of d_A(z) include measuring galaxy physical sizes from their angular extents, such as in surveys of disk galaxies to test expansion history. In the cosmic microwave background (CMB), d_A(z_\star) to the last scattering surface at z_\star \approx 1100 determines the angular scale of acoustic peaks in the power spectrum, with the first peak at \ell \approx 220 constraining and . Gravitational lensing arcs also rely on d_A(z) ratios between lens and source redshifts to infer mass distributions and distances.

Luminosity distance

The distance d_L(z), denoted as a function of z, quantifies the effective to a source based on its observed brightness in an expanding . It is defined such that the bolometric F received from a source with intrinsic bolometric L satisfies the relation F = \frac{L}{4\pi d_L^2(z)}, where the flux is integrated over all frequencies. In the standard Friedmann-Lemaître-Robertson-Walker (FLRW) cosmology, for a flat universe, d_L(z) = (1 + z) d_M(z), with d_M(z) being the transverse comoving distance. This definition incorporates the effects of cosmic expansion on propagation. The additional factor of (1 + [z](/page/Z)) relative to d_M(z) stems from two redshift-dependent phenomena: the energy loss of each , which decreases its observed by a factor of $1/(1 + [z](/page/Z)) due to the stretching of wavelengths, and the of the source's emission, which dilutes the arrival rate by another factor of $1/(1 + [z](/page/Z)). Together, these yield an overall dimming effect of (1 + [z](/page/Z))^2 compared to a static-space , making d_L(z) larger than the by precisely that factor, as per Etherington's distance duality . Type Ia supernovae (SNe Ia) are widely used as standard candles for luminosity distance measurements because their peak absolute magnitudes are empirically uniform after corrections for light-curve shape and host-galaxy properties, enabling reliable inference of d_L(z) from observed peak fluxes. These observations have mapped the universe's expansion history, revealing deviations from a decelerating model. In particular, analyses of high-redshift SNe Ia in 1998 by the Supernova Cosmology Project (led by Perlmutter et al.) and the High-Z Supernova Search Team (led by Riess et al.) demonstrated that luminosity distances at z \approx 0.5 were larger than expected in a matter-dominated universe, providing the first observational evidence for an accelerating expansion driven by dark energy. Contemporary applications of SNe Ia luminosity distances contribute to the Hubble tension, a discrepancy in the Hubble constant H_0. Local measurements, calibrated via Cepheid variables and extending to SNe Ia at low redshifts (z < 0.1), yield H_0 \approx 73 km/s/Mpc, while (CMB) inferences from high-redshift physics give H_0 \approx 67 km/s/Mpc, highlighting potential tensions in the standard cosmological model.

Relations and Corrections

Etherington's distance duality

Etherington's distance duality, also known as the reciprocity theorem, establishes a between the d_L(z) and the d_A(z) at z, given by d_L(z) = d_A(z) (1 + z)^2. This equation arises from the conservation of number and the of propagation in an expanding , as originally derived by Etherington in 1933 within the framework of applied to . The links the observed flux from a , which defines d_L, to the apparent of an object, which defines d_A, through the factor accounting for cosmological expansion. The proof of this duality follows from the consistency of surface areas and fluxes in the . In a Friedmann-Lemaître-Robertson-Walker (FLRW) , the infinitesimal subtended by a source corresponds to a physical area at the source, while the bolometric flux received is diluted by the expansion-induced in energy and . Equating these via the for both area and flux yields the (1 + z)^2 factor, assuming geodesics for paths and no or . This holds more generally in any theory of gravity where propagate without , meaning the is the same for all polarizations, independent of the specific dynamics like the . Observational tests of the duality have utilized multiple probes to verify consistency. Type Ia supernovae provide d_L(z) measurements up to z \approx 2, while angular diameter distances from radio quasars or galaxy cluster Sunyaev-Zel'dovich effects extend to similar ranges; combined analyses confirm the relation to within 1-2% precision. Cosmic microwave background (CMB) anisotropies, which fix d_A(z) through acoustic peak positions at recombination (z \approx 1100), align with supernova-derived d_L when the duality is imposed, supporting standard \LambdaCDM cosmology. Any violation, parametrized as \eta(z) = d_L(z) / [d_A(z) (1 + z)^2] \neq 1, would signal new physics, such as theories with varying speed of light or non-metric gravity effects, though current data show no significant deviations at the percent level. The duality unifies transverse distance measures by enabling the interconversion of d_L and d_A, facilitating robust cosmological parameter estimation across datasets. It assumes a homogeneous and isotropic on large scales; in inhomogeneous models, such as those with significant local voids or structures, the relation can break down, leading to direction-dependent distances that require generalized frameworks beyond the FLRW .

Peculiar velocity

In cosmology, peculiar velocity refers to the component of an object's that deviates from the smooth Hubble flow, representing local deviations due to gravitational interactions within the large-scale structure of the . These velocities are typically measured relative to the (CMB) rest frame, which serves as the reference for the overall expansion. At low redshifts (z \ll 1), peculiar velocities induce a Doppler shift in the observed redshift, approximated as z_D \approx v_{\rm pec}/c, where v_{\rm pec} is the line-of-sight peculiar velocity and c is the speed of light. Peculiar velocities perturb standard distance measures by altering the inferred , leading to errors in quantities like the luminosity distance d_L or d_A. The relative perturbation is on the order of \delta d / d \sim v_{\rm pec} / [c (1+z)], which becomes significant for nearby objects where the cosmological is small compared to the Doppler contribution. On larger scales, bulk flows—coherent peculiar motions of galaxy clusters and groups—amplify these effects, with velocities reaching hundreds of km/s in regions like the Local Supercluster or towards the . To mitigate these perturbations, corrections are applied using models of , such as the effect, which describes the linear enhancement of structure along the line of sight in galaxy surveys due to coherent inflows. The dipole provides a direct measure of our own peculiar relative to the rest frame, estimated at approximately 370 km/s towards galactic coordinates (l, b) \approx (264^\circ, 48^\circ). In the nearby (z < 0.1), where peculiar motions dominate over expansion, field reconstructions from surveys like the Redshift Survey and the ongoing Cosmicflows program enable precise mapping and correction of these effects up to distances of several hundred Mpc. These reconstructions, incorporating data from fundamental plane and Tully-Fisher relations, reveal bulk flows converging towards major structures and support tests of on local scales. Redshift distortions from peculiar velocities, including nonlinear effects, are further analyzed in the context of cosmological foundations to refine distance inferences.

References

  1. [1]
    [PDF] Metric Spaces - UC Davis Math
    A metric space is a set X that has a notion of the distance d(x, y) between every pair of points x, y ∈ X. The purpose of this chapter is to introduce metric ...
  2. [2]
    [PDF] Comprehensive Survey on Distance/Similarity Measures between ...
    From the scientific and mathematical point of view, distance is defined as a quantitative degree of how far apart. two objects are. Synonyms for distance ...
  3. [3]
    [astro-ph/9905116] Distance measures in cosmology - arXiv
    May 11, 1999 · Title:Distance measures in cosmology. Authors:David W. Hogg (IAS). View a PDF of the paper titled Distance measures in cosmology, by David W.
  4. [4]
    Cosmology Tutorial - Part 2
    Jun 12, 2009 · To say the Universe is homogeneous means that any measurable property of the Universe is the same everywhere.
  5. [5]
    Planck 2018 results - VI. Cosmological parameters
    We present cosmological parameter results from the final full-mission Planck measurements of the cosmic microwave background (CMB).
  6. [6]
    Exploring the dark energy equation of state with JWST
    Aug 9, 2024 · Observations from the James Webb Space Telescope (JWST) have unveiled several galaxies with stellar masses at redshifts .
  7. [7]
    Über die Krümmung des Raumes | Zeitschrift für Physik A Hadrons ...
    Über die Krümmung des Raumes. Published: December 1922. Volume 10, pages 377–386, (1922); Cite this article. Download PDF · Zeitschrift für Physik. Über die ...
  8. [8]
    [PDF] A homogeneous universe of constant mass and increasing radius ...
    Jun 13, 2013 · Original paper: Georges Lemaître, Un univers homogène de masse constante et de rayon croissant, rendant compte de la vitesse radiale des ...
  9. [9]
    [PDF] Testing the mapping between redshift and cosmic scale factor - arXiv
    Mar 17, 2016 · The canonical redshift−scale factor relation, 1/a = 1+ z, is a key element in the standard ΛCDM model of the big bang cosmology.
  10. [10]
    [PDF] Computing the Expansion History of the Universe - Nasa Lambda
    By measuring the redshift, z, of an astronomical source as a shift in wavelength, we learn the value of the scale factor at the time the light was emitted.
  11. [11]
    Gravity Probe B - Special & General Relativity Questions and Answers
    There are three known types: Doppler shifts ( due to motion through space away from the observer); gravitational redshifts ( due to light leaving a strong ...
  12. [12]
    [PDF] GCN - arXiv
    It is well known that there are three kinds of redshifts in nature, namely, the cosmological redshift, the gravitational redshift and the Doppler redshift ...
  13. [13]
    [PDF] Testing ΛCDM at the lowest redshifts with SN Ia and galaxy velocities
    At low redshifts z 1, the Hubble law applies, and we have cz ≈ H0d, where H0 is the Hubble constant and d is the proper distance. In this limit, there is ...
  14. [14]
    [PDF] The kinematic component of the cosmological redshift - arXiv
    Mar 8, 2011 · In the specially-relativistic formula for a Doppler shift,. zD = p(1 + v/c)/(1 − v/c), the recession velocity v is an inertial velocity of a ...
  15. [15]
    [astro-ph/9810142] Cosmology with the Lyman-alpha Forest - arXiv
    Oct 8, 1998 · Abstract: We outline the physical picture of the high-redshift Ly-alpha forest that has emerged from cosmological simulations, describe ...
  16. [16]
    [PDF] measuring the redshift dependence of the cosmic microwave ...
    1999). The adiabatic expansion of the universe and photon number conservation imply that the CMB temperature evolves with redshift as T (z) = T0(1 + z).
  17. [17]
    [astro-ph/0501171] Detection of the Baryon Acoustic Peak in ... - arXiv
    Jan 10, 2005 · Detection of the Baryon Acoustic Peak in the Large-Scale Correlation Function of SDSS Luminous Red Galaxies. Authors:D. J. Eisenstein, I. Zehavi ...
  18. [18]
    [PDF] Distances in Cosmology - MPA Garching
    The reason lies in our definition of proper distance, as the distance between two events measured in a frame of reference where those two events happen at ...
  19. [19]
    [PDF] Distance measures in cosmology - arXiv
    The transverse comoving distance DM is not a proper distance—it is a proper distance divided by a ratio of scale factors. 5.
  20. [20]
    Acoustic Peaks and the Cosmological Parameters
    The distance to the surface of last scattering, corresponding to its angular size, is given by what is called the angular diameter distance. It has a simple ...Missing: applications | Show results with:applications
  21. [21]
    [PDF] The need for accurate redshifts in supernova cosmology - arXiv
    The CMB redshift is used to define a range of integration in equation (2.1), and hence a small change in this value can significantly change the luminosity ...
  22. [22]
    [astro-ph/9805201] Observational Evidence from Supernovae for an ...
    May 15, 1998 · Observational Evidence from Supernovae for an Accelerating Universe and a Cosmological Constant. Authors:Adam G. Riess, Alexei V. Filippenko, ...Missing: discovery | Show results with:discovery
  23. [23]
    Hubble Constant and Tension - NASA Science
    Aug 4, 2025 · But measurements derived from studying the CMB provide a Hubble Constant of about 67-68 kilometers per second per megaparsec. Scientists call ...
  24. [24]
    On the perturbation of the luminosity distance by peculiar motions
    We consider some aspects of the perturbation to the luminosity distance d(z) that are of relevance for SN1a cosmology and for future peculiar velocity surveys ...
  25. [25]
    Reconstructed density and velocity fields from the 2MASS Redshift ...
    Abstract. We present the reconstructed real-space density and the predicted velocity fields from the Two-Micron All-Sky Redshift Survey (2MRS). The 2MRS is.
  26. [26]
    Density and velocity fields using CosmicFlows-4
    This article publicly releases 3D reconstructions of the local Universe gravitational field below z = 0.8 that were computed using the CosmicFlows-4 (CF4) ...
  27. [27]
    On the perturbation of the luminosity distance by peculiar motions
    Nov 24, 2014 · We consider some aspects of the perturbation to the luminosity distance d(z) that are of relevance for SN1a cosmology and for future peculiar velocity surveys.