Fact-checked by Grok 2 weeks ago

Gravitational redshift

Gravitational redshift is the phenomenon wherein light or other emitted from a of strong experiences a decrease in (increase in ) as it propagates to a of weaker , a direct consequence of the and the of in . This effect, first theoretically derived by in 1911 using the before the full formulation of , arises because clocks run slower in stronger gravitational fields, leading to a relative that manifests as a for outgoing photons. In the weak-field approximation applicable to most terrestrial and solar system contexts, the fractional frequency shift \Delta \nu / \nu is given by \Delta \Phi / c^2, where \Delta \Phi is the difference in gravitational potential between emission and observation points, c is the , and \nu is the ; this formula quantifies how the energy of photons is effectively reduced as they climb out of a gravitational well. For stronger fields, such as near compact objects, the full general relativistic treatment in the yields $1 + z = \sqrt{(1 - 2GM/(c^2 r_e)) / (1 - 2GM/(c^2 r_o))}, where z is the , G is the gravitational constant, M is the mass of the gravitating body, and r_e and r_o are the radial coordinates of the emitter and observer, respectively (r_o > r_e). The effect is distinct from the cosmological redshift due to cosmic expansion and the Doppler redshift from relative motion, though all contribute to observed spectral shifts in astronomy. The gravitational redshift was experimentally confirmed in 1960 by Robert Pound and Glen Rebka at , who measured the frequency shift of gamma rays traveling up and down a 22.6-meter tower using the , achieving agreement with to within 10% accuracy, later improved to 1%. Further validations include observations of white dwarfs, where the redshift correlates with stellar mass and radius as predicted, as seen in spectroscopic data from the combined with , confirming the effect in over 3,000 objects. In practical applications, the effect is crucial for the (GPS), where satellite clocks experience a gravitational redshift causing them to tick faster by about 45 microseconds per day relative to ground clocks, necessitating relativistic corrections to maintain positional accuracy within meters. These tests underscore gravitational redshift's role as a cornerstone verification of , with ongoing observations in extreme environments like black hole vicinities providing probes of gravity's limits.

Theoretical Foundations

Equivalence Principle Prediction

The , as formulated by , states that the effects of a uniform are locally indistinguishable from those of an accelerated reference frame in the absence of gravity. This principle implies that any physical process, including the propagation of light, behaves identically in both scenarios over small regions of where the gravitational field can be approximated as uniform. Consider a involving light emission in such an equivalent setup: imagine an accelerating upward with constant g in flat , or equivalently, at rest in a uniform of strength g. A pulse of \nu is emitted from the floor (lower position) toward the ceiling (higher position), separated by height h. In the accelerating frame, the time interval between successive wave crests at emission is \Delta \tau_e = 1/\nu, measured by a clock at the emitter. However, by the time the reaches the at the ceiling, the elevator has moved upward, causing the receiver to recede from the approaching light during transit. This relative motion induces a Doppler-like shift, lowering the observed frequency \nu_o at the receiver, where the proper time interval is \Delta \tau_o > \Delta \tau_e. The shift arises from the differential s in the accelerated frame, equivalent to in the gravitational case. To derive the approximate formula, note that the transit time for over h is approximately t \approx h/c, during which the receiver accelerates away by a (1/2) g t^2 \approx (1/2) g h^2 / c^2. The fractional velocity change relative to c is v/c \approx g h / c^2, leading to a first-order Doppler shift \Delta \nu / \nu \approx -v/c = -g h / c^2, where the negative sign indicates a (decrease in ). More precisely, the equates clock rates at different heights: a clock at the lower position (deeper in the potential) ticks slower by the factor $1 + g h / c^2 relative to one at the higher position, so the emitted appears reduced when compared to the faster-ticking receiver clock. Physically, this redshift can be interpreted as the losing equivalent to the work done against the over height h, \Delta E = m g h, where the photon's effective gravitational mass is E/c^2, yielding \Delta \nu / \nu = -g h / c^2 since E = h \nu. However, the effect is fundamentally rooted in the geometry of spacetime, where the reveals local time dilation due to the interplay of and . Einstein first predicted this gravitational redshift in 1911 using the , prior to the full development of , estimating a shift of about $2 \times 10^{-6} for light from the Sun's surface observed on . This local approximation is extended in to global curved spacetimes.

General Relativity Derivation

In , gravity manifests as the curvature of induced by the presence of and , with propagating along in this . The of a , as measured by a local observer, is given by the scalar product of the observer's u^\mu and the photon's k^\mu, specifically \nu = -u_\mu k^\mu / h, where h is Planck's constant. For static spacetimes, the timelike Killing vector ensures of the k_t component along the geodesic, enabling the frequency shift to be computed between and events. The ratio of the observed \nu_\mathrm{obs} to the emitted \nu_\mathrm{em} for photons traveling between two static observers is thus \nu_\mathrm{obs} / \nu_\mathrm{em} = (u^\mu k_\mu)_\mathrm{obs} / (u^\mu k_\mu)_\mathrm{em}. In a static, spherically symmetric ds^2 = g_{tt}(r) dt^2 + g_{rr}(r) dr^2 + r^2 d\Omega^2, observers at rest have u^\mu = (1/\sqrt{-g_{tt}}, 0, 0, 0), normalized such that u^\mu u_\mu = -1. The conserved k_t = -E (energy at infinity) leads to the frequency ratio simplifying to \nu_\mathrm{obs} / \nu_\mathrm{em} = \sqrt{ -g_{tt}(r_\mathrm{em}) / -g_{tt}(r_\mathrm{obs}) }, assuming radial propagation and r_\mathrm{obs} > r_\mathrm{em}. For the specific case of the Schwarzschild metric, which describes the vacuum spacetime around a spherically symmetric, non-rotating mass M, the line element is ds^2 = -\left(1 - \frac{2GM}{c^2 r}\right) c^2 dt^2 + \left(1 - \frac{2GM}{c^2 r}\right)^{-1} dr^2 + r^2 d\Omega^2, where G is the gravitational constant and c is the speed of light. Here, g_{tt}(r) = -\left(1 - \frac{r_s}{r}\right), with the Schwarzschild radius r_s = 2GM/c^2. Substituting into the frequency ratio yields \nu_\mathrm{obs} / \nu_\mathrm{em} = \sqrt{ \left(1 - \frac{r_s}{r_\mathrm{em}}\right) / \left(1 - \frac{r_s}{r_\mathrm{obs}}\right) }. The redshift z is defined as z = (\nu_\mathrm{em} - \nu_\mathrm{obs}) / \nu_\mathrm{obs}, so $1 + z = \sqrt{ \left(1 - \frac{r_s}{r_\mathrm{obs}}\right) / \left(1 - \frac{r_s}{r_\mathrm{em}}\right) }. In the weak-field limit, where r_s / r \ll 1, the metric component expands as g_{tt}(r) \approx - (1 + 2\Phi(r)/c^2), with Newtonian \Phi(r) = -GM/r. The redshift then approximates to z \approx (\Phi_\mathrm{em} - \Phi_\mathrm{obs}) / c^2, recovering the form predicted by the as the local limit of this global effect. This holds for typical solar-system scales, where |\Phi| / c^2 \sim 10^{-6}.

Approximations and Limits

Newtonian Limit

The Newtonian limit of gravitational redshift applies in the weak-field, low-velocity regime, where the absolute value of the Newtonian |Φ| ≪ c² and velocities are much less than the c, with Φ = -GM/r for a M at distance r from the center. In this approximation, the general relativistic prediction simplifies to a form that connects directly to classical through potential differences. The derivation begins with the weak-field metric of , where the time-time component is g_{tt} ≈ -(1 + 2Φ/c²). For a emitted at a point with potential Φ_em and observed at Φ_obs, the redshift z is given by z ≈ (Φ_obs - Φ_em)/c² = ΔΦ/c², assuming the light climbs out of a deeper (Φ_em < Φ_obs). This arises from the gravitational time dilation encoded in the metric, where clocks run slower deeper in the potential, stretching the 's wavelength relative to a distant observer. A representative example is the vertical frequency shift near Earth's surface, where for a height difference h ≪ R (Earth's radius), the potential difference ΔΦ ≈ gh with g = GM/R², yielding z ≈ gh/c². This recovers the intuitive result from the equivalence principle but derives it explicitly from the potential, as verified in laboratory tests like the Pound-Rebka experiment measuring gamma-ray shifts over 22.5 meters. Unlike the Doppler shift in special relativity, which results from relative motion, the gravitational redshift stems from position-dependent time dilation in a static gravitational field, with no source-observer velocity involved. This distinction highlights the geometric nature of the effect in general relativity, even in the Newtonian limit. The approximation breaks down in strong fields, such as near black holes where |Φ| approaches c², or at relativistic velocities, requiring the full general relativistic treatment. Historically, while Johann Georg von Soldner calculated light bending in a Newtonian framework in 1801, the redshift effect was first predicted by in 1911 using an early equivalence principle argument, later refined in the weak-field limit of general relativity.

Semi-Classical Photon Approach

In the semi-classical approach to gravitational redshift, photons are treated as relativistic particles with energy E = h \nu, where h is and \nu is the frequency, and momentum p = E / c = h \nu / c, with c the . Although photons possess no rest mass, their interaction with gravity can be heuristically modeled by assigning an effective inertial mass m_\text{eff} = E / c^2, drawing from the equivalence of mass and energy in . This perspective bridges classical particle dynamics with relativistic effects, allowing an intuitive understanding of frequency shifts without invoking the full machinery of general relativity's curved spacetime. Consider a photon emitted at a location with gravitational potential \Phi_e and observed at a higher potential \Phi_o, such that \Delta \Phi = \Phi_o - \Phi_e > 0. In a weak, static gravitational field, the photon "climbs" against the potential gradient, analogous to a massive particle losing kinetic energy to gain potential energy. The effective gravitational force arises from the field gradient, but for energy conservation along the photon's trajectory in this static background, the change in photon energy mirrors the potential difference scaled by the effective mass: \Delta E = - m_\text{eff} \Delta \Phi = - (E / c^2) \Delta \Phi. Since E = h \nu, the frequency shift follows as \Delta \nu / \nu = \Delta E / E = - \Delta \Phi / c^2, yielding a redshift for light escaping a gravitational well. This derivation assumes a Newtonian-like potential \Phi = - G M / r in the weak-field limit, where higher potentials correspond to regions farther from the mass source. This heuristic aligns with the particle description of light in phenomena like the Compton effect, where photons collide with electrons as if carrying definite and related by E = p c, reinforcing the de Broglie-like wave-particle duality for massless . However, it remains non-rigorous, as photons lack rest and do not experience "forces" in the classical sense; the approach serves primarily for intuition rather than precise calculation. The semi-classical picture proves useful for rough estimates in solar system contexts, such as the redshift of sunlight observed from Earth, where the potential difference across the Sun's radius gives |\Delta \Phi| / c^2 \approx 2 \times 10^{-6}, predicting a fractional frequency shift of similar magnitude. Despite its approximations, it highlights how energy conservation in a static field implies the observed frequency adjustment without invoking time dilation directly. Nonetheless, this force-based analogy is limited to weak fields and overlooks the geometric nature of gravity; the true relativistic treatment relies on spacetime curvature affecting null geodesics, not particle trajectories under potential.

Experimental Verification

Laboratory Tests

The first direct laboratory confirmation of gravitational redshift was achieved in the Pound-Rebka experiment conducted in 1959 at , utilizing the to measure frequency shifts in gamma rays emitted from an iron-57 source over a height difference of 22.5 meters in the Jefferson Physical Laboratory tower. The experiment detected a fractional frequency shift of (2.5 ± 0.3) × 10^{-15}, aligning with the general relativity prediction of 2.46 × 10^{-15} derived from the in the weak-field limit, where the shift is approximately gh/c² with g as Earth's , h as height, and c as the . Subsequent refinements improved the precision significantly. In a 1960 follow-up by Pound and Rebka, the setup was optimized to achieve about 10% agreement with by better accounting for effects and source stability. Further enhancement came in 1964 with the Pound-Snider experiment, which used improved and data analysis to measure the shift to within 1% of the predicted value, confirming the effect with reduced systematic errors from and environmental noise. Later laboratory experiments incorporated cryogenic sources to minimize thermal noise and advanced interferometric techniques for higher resolution. For instance, experiments in the 1970s at Harvard employed cooled emitters and multi-channel detectors, reaching sensitivities around 10^{-17} for the fractional shift over similar tower heights. These Mössbauer-based tests demonstrated the redshift's consistency with general relativity while probing the local validity of the equivalence principle. Modern laboratory verifications leverage atom with cold atoms and optical clocks placed at varying elevations, enabling comparisons of atomic transition sensitive to differences. Precursors to the 2017 space mission included ground-based atom setups that confirmed the to uncertainties below 10^{-15}, using laser-cooled or cesium atoms to create matter waves whose phase shifts encode the difference. More recent optical clock experiments, such as those comparing clocks separated by centimeters to meters, have achieved confirmations to parts in 10^{18}, with fractional stabilities exceeding 10^{-18} over short baselines. For example, a 2023 blinded test using a network of five optical clocks measured the gravitational gradient over a 1 cm height, achieving a of ~3 × 10^{-19} in the fractional shift, rejecting deviations at high . These experiments typically involve to compare oscillator frequencies at distinct gravitational potentials, with careful cancellation of first-order Doppler shifts via counter-propagating beams or simultaneous up-down measurements, alongside corrections for second-order relativistic effects and environmental gradients. The results validate in controlled terrestrial settings and the universality of , as the ties directly to the of inertial and gravitational . Recent applications extend to searches for , where anomalies in clock comparisons could signal ultralight scalar fields modulating effective masses and thus mimicking variations. Despite these advances, laboratory tests face inherent limitations: the minuscule signal strength over short heights (typically meters or less) demands sources with fractional stabilities better than 10^{-17}, challenging from vibrations, , and fluctuations.

Space-based Experiments

Space-based experiments on gravitational utilize satellites and orbital platforms to achieve greater height differences above Earth's surface, enabling stronger signals and tests in varying gravitational potentials compared to setups. These experiments typically involve precise clocks or comparisons between space and ground stations, providing direct verification of general 's predictions over baselines of thousands of kilometers. By isolating the gravitational component from kinematic effects like velocity-induced Doppler shifts, such tests confirm the redshift effect with high accuracy, contributing to broader validations of principles and in dynamic environments. The core methodology in these experiments compares clock rates or signal frequencies between ground-based references and orbiting instruments, with data processing to separate the gravitational redshift from special relativistic time dilation and other perturbations. For instance, one-way or two-way ranging signals are analyzed to extract the fractional frequency shift \Delta \nu / \nu, predicted by integrating the gravitational potential along the signal path. This approach leverages the Newtonian approximation for Earth's potential in orbital contexts, where the redshift scales with the difference in gravitational potential between emitter and receiver. One of the earliest space-based tests was Gravity Probe A in 1976, which launched a clock aboard a Scout rocket to an apogee altitude of approximately 10,000 km. The experiment measured the gravitational redshift by comparing the rocket's clock to ground-based masers during ascent and descent, yielding a fractional shift consistent with the general prediction of ~4.5 × 10^{-10} to within 70 parts per million. This result confirmed the effect to better than 0.01% relative accuracy, marking a seminal verification in a controlled orbital . The (GPS) constellation provides ongoing operational validation of , with satellites at about 20,000 km altitude requiring clock corrections of approximately 45 μs/day to account for the effect, as satellite clocks run faster in weaker . Daily comparisons between satellite and ground clocks, integrated into GPS , confirm the to high precision, typically better than 0.1%, ensuring system accuracy for . These routine validations demonstrate the practical incorporation of corrections in a global network. More recent efforts include the GREAT experiment (2019–2023), which exploited the eccentric orbits of Galileo navigation satellites with a semi-major axis of 29,000 km to analyze one-way ranging signals for signatures. By processing tracking data over multiple years, the experiment confirmed general relativity's prediction to 0.5% accuracy, improving upon prior tests by a factor of several through enhanced and signal modeling. The results, published in 2020, highlight the utility of navigation satellites for precision gravitation tests. The Atomic Clock Ensemble in Space (ACES), deployed on the in April 2025, employs cold-atom cesium and clocks aiming for $10^{-16} fractional frequency stability to measure and effects. ACES facilitates space-to-ground comparisons via microwave and optical links, targeting an absolute measurement with uncertainty below $2 \times 10^{-6} after short integration periods, while enabling long-term tests of fundamental physics. This mission extends space-based clock technology for verification in microgravity. Recent advances include 2025 simulations exploring the integration of space clocks with the European Laser Timing (ELT) experiment on missions like ACES, using two-way laser time transfer to enhance tests by 3–4 orders of magnitude in precision. These simulations also link measurements to checks of Lorentz invariance, probing potential violations in through clock comparisons in varying potentials. Such developments underscore the growing role of ground-space systems in gravitational tests. Overall, space-based experiments offer larger baselines and exposure to Earth's inhomogeneous field, achieving percent-level confirmations that complement other verification methods and support applications in precise timing networks.

Astrophysical Observations

One of the earliest astrophysical confirmations of gravitational came from observations of the Sirius B, the companion to Sirius A. In 1925, Walter Adams measured a in the of Sirius B, reporting a shift corresponding to approximately 23 km/s, which was interpreted as partial evidence for , though the value was underestimated due to incomplete of the star's and the presence of Doppler effects from orbital motion. Modern analysis attributes the intrinsic gravitational component to the deep of the compact star, with z \approx \frac{GM}{Rc^2} \approx 2.7 \times 10^{-4} based on its of about 1 and of roughly 0.0084 solar radii. This was precisely confirmed in 2004 using spectroscopy of Balmer lines, yielding a gravitational velocity of 80.42 ± 4.83 km/s after isolating it from Doppler contributions via orbital modeling. Direct measurements of the Sun's also provided key evidence in the 1960s through ranging experiments involving planetary echoes. Signals bounced off and Mercury during superior conjunctions, when the paths grazed the solar limb, exhibited frequency shifts consistent with the integrated along the trajectory, measuring z \approx 2 \times 10^{-6} for light emerging from the solar surface, in agreement with the prediction z = \frac{GM_\odot}{c^2 R_\odot} \approx 2.12 \times 10^{-6}. These observations, combined with earlier implications from the 1919 light deflection, helped validate weak-field effects over cosmic distances. In strong-field regimes, of has enabled inferences of through pulse profile modeling. For instance, the Interior Composition Explorer (NICER) observed the PSR J0030+0451 in 2019, deriving a of 1.44 +0.15 −0.14 solar masses and of 13.02 +1.24 −1.06 km by fitting thermal emission hotspots, where the observed pulse phases incorporate the redshift factor \sqrt{1 - \frac{2GM}{Rc^2}} \approx 0.82, corresponding to [z](/page/Z) \approx 0.22. Similar analyses for accretion disks, such as those from X-ray binaries, further constrain strong-field predictions, though isolation from remains challenging. Galaxy clusters serve as natural laboratories for measuring subtle gravitational redshifts on large scales, often using stacked spectra to detect shifts from the cluster potential. The Wojtak effect, identified in 2011 through analysis of 7800 Sloan Digital Sky Survey (SDSS) clusters, reveals a net blueshift in central galaxy velocities relative to the outskirts due to photons climbing out of the potential well, with an average shift of about 3-5 km/s (z \approx 10^{-5}) consistent with general relativity mass profiles. In lensed quasars behind clusters, such as those in SDSS data, differential redshifts in multiply imaged light paths probe the potential depth, with 2015 studies confirming patterns matching Wojtak predictions for cluster masses around 10^{14} solar masses. A 2023 study in Astronomy & Astrophysics analyzed velocity distributions in 72 dynamically relaxed clusters from the SDSS, measuring an average gravitational redshift velocity shift of -11.4 ± 3.3 km/s (z ≈ 3.8 × 10^{-5}) after subtracting peculiar velocities, aligning with general relativity expectations from weak lensing-derived mass profiles. These observations rely on methodologies that isolate gravitational shifts from Doppler and cosmological redshifts through multi-wavelength modeling, such as kinematic mapping of velocities or gravitational lensing to reconstruct potentials without assuming . For instance, asymmetries in cluster galaxies are fitted using N-body simulations to disentangle effects, while X-ray data from or complements optical spectra for mass modeling. The significance of these astrophysical tests lies in probing in regimes inaccessible to laboratories, including strong fields near neutron stars and cumulative weak fields in clusters, where 2025 studies derive relativistic corrections to gravitational redshift signals in galaxy clusters to third order in the weak-field approximation, enabling refined tests of models. Such measurements test alternative theories and constrain models through potential-depth consistency. Challenges include contamination from line-of-sight Doppler motions in unresolved systems and cosmological expansion, necessitating high-resolution and deep multi-epoch data to achieve signal-to-noise ratios above 100 for precise isolation.

References

  1. [1]
    [PDF] On the Influence of Gravitation on the Propagation of Light
    * Of course we cannot replace any arbitrary gravitational field by a state of motion of the system without a gravitational field, any more than, by a trans-.
  2. [2]
    General Relativity - HyperPhysics
    In the early 60's physicists Pound, Rebka,and Snyder at the Jefferson Physical Laboratory at Harvard measured the shift to within 1% of the predicted shift. By ...
  3. [3]
    Apparent Weight of Photons | Phys. Rev. Lett.
    Apparent Weight of Photons, RV Pound and GA Rebka, Jr., Lyman Laboratory of Physics, Harvard University, Cambridge, Massachusetts.
  4. [4]
    A Gravitational Redshift Measurement of the White Dwarf Mass ...
    Aug 25, 2020 · Here we observationally measure this relation using the gravitational redshift effect, a prediction of general relativity that depends on the ratio between ...Abstract · Introduction · Data Analysis · Systematic Effects
  5. [5]
    [PDF] Gravity 2.0 Relativity in Your Hand - Physics of the Cosmos
    1959Gravity's redshift​​ One of the central ideas of general relativity is the equivalence principle, which asserts that everything responds to gravity, whether ...
  6. [6]
    [PDF] On the Influence of Gravitation on the Propagation of Light
    On the Influence of Gravitation on the Propagation of Light ... In The Collected Papers of Albert Einstein, Volume 3: The Swiss Years: Writings,. 1909-1911, ...
  7. [7]
    [PDF] General Relativity Fall 2019 Lecture 4: the equivalence principle
    Sep 12, 2019 · Gravitational redshift is an immediate consequence of the undistinguishability of a uniform gravitational field and a uniformly accelerated ...
  8. [8]
    [PDF] Gravitational redshift revisited: inertia, geometry, and charge - arXiv
    In §§2–4, we derive and discuss the gravitational redshift effect in three ways: (i) from the framework of general relativ- ity (GR), (ii) using the equivalence ...
  9. [9]
    [PDF] Physics 161: Black Holes: Lecture 5: 22 Jan 2013
    Jan 22, 2013 · The equivalence principle says that experiments done in this falling frame will give the same results as experiments done floating in free space ...
  10. [10]
    [PDF] Charles W. MISNER Kip S. THORNE John Archibald WHEELER
    Gravitational Redshift Implies Spacetime Is Curved 187. 4. Gravitational Redshift as Evidence for the Principle of Equivalence 189. 5. Local Flatness, Global ...
  11. [11]
    [PDF] the gravitational redshift, deflection of light, and Shapiro delay
    Nov 7, 2012 · The gravitational redshift: a photon in a static spacetime changes its energy in accordance with the gravitational potential, ∆ν/ν = −∆Φ. The ...
  12. [12]
    [PDF] Part 3 General Relativity - DAMTP
    These are lecture notes for the course on General Relativity in Part III of the. Cambridge Mathematical Tripos. There are introductory GR courses in Part II.Missing: primary | Show results with:primary
  13. [13]
    [PDF] arXiv:physics/9907017v2 [physics.ed-ph] 27 Jul 1999
    The first laboratory measurement of the gravitational redshift was performed at Harvard in. 1960 by Robert Pound and Glenn Rebka [4], [5] (with 10% accuracy) ...
  14. [14]
    Johann Georg von Soldner and the gravitational bending of light ...
    In a paper published in 1804, Soldner derived the gravitational bending of light on the classical Newtonian basis and calculated its value around the sun with ...
  15. [15]
    [PDF] 19800011717.pdf - NASA Technical Reports Server (NTRS)
    GRAVITATIONAL REDSHIFT SPACE-PROBE EXPERIMENT. GP-A Project Final Report ... over the I % tests made by Pound became evident, and the redshift clock experiment.<|separator|>
  16. [16]
    A lab-based test of the gravitational redshift with a miniature clock ...
    Aug 12, 2023 · Here we perform a laboratory-based, blinded test of the gravitational redshift using differential clock comparisons within an evenly spaced array of 5 atomic ...
  17. [17]
    The Confrontation between General Relativity and Experiment
    The gravitational redshift could be improved to the 10−10 level using an array of laser cooled atomic clocks on board a spacecraft which would travel to within ...
  18. [18]
    [1907.12320] Gravitational redshift test with the future ACES mission
    Jul 29, 2019 · We investigate the performance of the upcoming ACES (Atomic Clock Ensemble in Space) space mission in terms of its primary scientific objective, ...
  19. [19]
    Relativity in the Global Positioning System - PMC - NIH
    The Global Positioning System (GPS) uses accurate, stable atomic clocks in satellites and on the ground to provide world-wide position and time determination.
  20. [20]
    A new test of gravitational redshift using Galileo satellites
    By analyzing several years of Galileo tracking data, we have been able to improve the Gravity probe A test (1976) of the gravitational redshift by a factor of ...
  21. [21]
    Simulating Gravitational Redshift Test Using the European Laser ...
    Our simulation results indicate that using the two-way laser time transfer (TWLTT) link via the ELT experiment achieves precision levels 3-4 orders of magnitude ...
  22. [22]
    gravitational redshift of Sirius B - Oxford Academic
    Sep 4, 2018 · In the following year Adams published his result of 23 km s−1 (Adams 1925) and extended Eddington's original motivation for the observation to ...ABSTRACT · INTRODUCTION · ANALYSIS · DISCUSSION
  23. [23]
    PSR J0030+0451 Mass and Radius from NICER Data ... - IOP Science
    Dec 12, 2019 · Here we report estimates of the mass and radius of the isolated 205.53 Hz millisecond pulsar PSR J0030+0451 obtained using a Bayesian inference approach.
  24. [24]
    New measurements of gravitational redshift in galaxy clusters
    We analysed the velocity distribution of the cluster member galaxies to make new measurements of the gravitational redshift effect inside galaxy clusters.
  25. [25]
    Gravitational Redshift from Galaxy Clusters -- a Relativistic Approach
    Mar 14, 2025 · Abstract:The light that we receive from clusters of galaxies is redshifted by the presence of the clusters' gravitational potential.