Fact-checked by Grok 2 weeks ago

Coordinate time

Coordinate time is the temporal component of a coordinate system used to label events in spacetime, particularly within the frameworks of special and general relativity, where it provides a conventional measure for ordering occurrences rather than directly representing physical duration experienced by observers. Unlike proper time, which quantifies the invariant interval along an object's worldline as measured by its own clock, coordinate time depends on the chosen reference frame and can vary due to relative motion or gravitational fields. In mathematical terms, the relationship between infinitesimal proper time d\tau and coordinate time dt is given by the spacetime metric, such as d\tau^2 = dt^2 - dx^2 - dy^2 - dz^2 (in units where c=1) for flat Minkowski space in special relativity. In , coordinate time corresponds to the synchronized time across an inertial frame, where is relative, and it equals only for observers at rest in that frame. extends this concept to curved , emphasizing : all coordinate systems, including their time labels, are equally valid for describing physics, with no absolute frame privileged, as per the . For instance, in noninertial frames like those near , coordinate time incorporates adjustments for and , differing from inertial coordinate time through metric components that account for . Practically, coordinate time underpins astronomical time scales, such as (TDB), which serves as the independent variable for ephemerides in solar system dynamics within general relativity's framework. This realization, approximating (TCB) with periodic corrections, ensures precise synchronization for space missions, light propagation, and calculations.

Fundamental Concepts

Coordinate Time vs. Proper Time

In , coordinate time serves as a global parameter that labels events within a chosen , typically measured by a network of synchronized clocks at rest relative to an inertial observer, and remains independent of the motion of individual clocks or observers. This time coordinate, often denoted as t, provides a frame-dependent measure of temporal progression across space, allowing events separated in position to be assigned simultaneous timestamps through conventions. In contrast, represents the invariant interval along the worldline of a specific observer or particle, measured directly by a single clock traveling along that path, and is independent of the choice of . Denoted as \tau, it quantifies the duration experienced locally by the clock, unaffected by relative motion between frames, and serves as the fundamental measure of aging or physical processes for that observer. The distinction arises because is an intrinsic property tied to the geometry of , while coordinate time is an extrinsic label imposed by the observer's frame. Albert Einstein introduced the foundational concepts of this distinction in his 1905 paper on , where he demonstrated through that the time measured by a moving clock differs from the synchronized time in a stationary frame, a difference that becomes particularly pronounced when considering in non-inertial frames. Hermann Minkowski formalized as an invariant in his 1908 reformulation of into a four-dimensional framework, emphasizing its role as the proper duration along timelike paths. A classic illustration of this difference is the , where one twin remains on while the other travels at relativistic speeds and returns; the traveling twin accumulates less due to their accelerated worldline, aging slower than the Earth-bound twin, whose coincides with the coordinate time in the inertial frame. In , this relationship is captured by the for : d\tau = \sqrt{dt^2 - \frac{dx^2 + dy^2 + dz^2}{c^2}} where dt is the infinitesimal coordinate time interval, dx, dy, dz are spatial displacements, and c is the , showing that proper time is always shorter than or equal to coordinate time, with equality only for observers.

Clock Synchronization in Relativity

In , clock establishes a coordinate time across an inertial reference frame, but the procedure reveals fundamental limitations due to the theory's postulates. first discussed the synchronization of moving clocks in his 1905 memoir, where he introduced the concept of "local time" derived from signal exchanges between observers, highlighting the conventional nature of in relative motion. elaborated on this in his seminal 1905 paper, defining for spatially separated events through the assumption that propagates at constant speed c in all directions within the frame. The Einstein synchronization convention provides the standard procedure for inertial frames: two clocks at positions A and B are synchronized if a light signal emitted from A at coordinate time t_A arrives at B at t_B = t_A + \frac{d}{c}, and the return signal from B arrives back at A at t_A' = t_B + \frac{d}{c}, ensuring the round-trip travel time is symmetric and light speed appears isotropic. This method assumes the one-way speed of light is c, a convention that aligns with the principle of relativity but is not empirically verifiable without prior synchronization assumptions. However, special relativity introduces the relativity of simultaneity, where clocks synchronized in one inertial frame appear desynchronized in another frame moving relative to the first, as the transformation of time coordinates depends on spatial separation and relative velocity. This frame-dependence means no absolute simultaneity exists, rendering synchronization a local convention rather than a universal truth. Extending to accelerated frames, where inertial methods fail due to non-constant , requires alternative protocols. The method uses a single clock to emit and receive or radio signals from a distant point, assigning coordinate time to events based on the midpoint of emission and reception times, scaled by the assumed round-trip speed; this defines an approximate for the observer's instantaneous . Complementarily, slow clock transport involves moving a clock at low from one location to another, where the accumulated during transport approximates the coordinate time difference, yielding equivalent to the Einstein convention in the low-speed limit and avoiding direct reliance on signals. These approaches provide practical consistency in non-inertial settings but inherit the when comparing across frames. A key practical application arises in the (GPS), where clocks must synchronize with Earth-based coordinate time despite orbital velocities causing special relativistic effects. GPS employs atomic clocks pre-adjusted for relativistic effects, including both velocity-induced and gravitational contributions, and uses two-way signal ranging to maintain synchronization, ensuring positional accuracy within meters by aligning time readings to the scale on .

Mathematical Foundations

Formulation in Special Relativity

In , the of flat Minkowski geometry provides the foundational framework for defining coordinate time. The in Minkowski is given by the ds^2 = -c^2 dt^2 + dx^2 + dy^2 + dz^2, where dt represents the coordinate time interval in a given inertial , c is the , and dx, dy, dz are spatial differentials. This , introduced by , treats time as a coordinate on equal footing with space, forming a four-dimensional continuum where coordinate time t parameterizes events along worldlines. Coordinate time transforms between inertial frames via the , which accounts for the . For two frames where the primed frame moves at v along the x- relative to the unprimed frame, the time component is t' = \gamma \left( t - \frac{v x}{c^2} \right), with \gamma = \frac{1}{\sqrt{1 - v^2/c^2}}. This transformation, derived from the constancy of the and the principle of , demonstrates that coordinate time is frame-dependent: events simultaneous in one frame (dt' = 0) are not in another unless x = 0. The relationship between coordinate time and \tau—the time measured by a clock along its worldline—arises from the for timelike paths (ds^2 < 0). For a clock moving at velocity \mathbf{v}(t) in the coordinate frame, the proper time interval is d\tau = dt \sqrt{1 - v^2/c^2}, leading to the integrated form \tau = \int \sqrt{1 - \frac{v^2(t)}{c^2}} \, dt along the worldline. This integral shows that proper time accumulates more slowly than coordinate time for moving observers, a direct consequence of the invariance of the spacetime interval. For constant velocity v, the time dilation formula simplifies to \Delta \tau = \Delta t / \gamma = \Delta t \sqrt{1 - v^2/c^2}, where \Delta t is the coordinate time elapsed in the rest frame of the coordinates. Consider an example: a clock moving at constant speed v = 0.8c relative to an inertial frame, with coordinate time \Delta t = 10 years elapsing in that frame. Here, \gamma = 1 / \sqrt{1 - (0.8)^2} = 1.6667, so the proper time is \Delta \tau = 10 / 1.6667 \approx 6 years. This calculation illustrates how coordinate time overestimates the clock's elapsed proper time by the factor \gamma > 1.

Formulation in

In , is described by a equipped with a g_{\mu\nu}, which defines the as ds^2 = g_{\mu\nu} \, dx^\mu \, dx^\nu, where the coordinates x^\mu include the coordinate time t (typically as x^0 = ct) alongside spatial coordinates, and the metric components encode the curvature induced by mass and energy. This formulation generalizes the flat Minkowski metric of to curved , allowing coordinate time to serve as a global parameter for events in asymptotically flat regions. The proper time \tau experienced by an observer along a timelike worldline is obtained by integrating the infinitesimal interval d\tau = \sqrt{ -\frac{g_{\mu\nu} \, dx^\mu \, dx^\nu}{c^2} }, which extends the special-relativistic definition to account for gravitational effects through the metric's dependence on position. For observers at rest in the coordinate system (i.e., dx^i = 0 for spatial indices i), this reduces to a relation between proper time and coordinate time influenced solely by the temporal component g_{00}. A prominent example of gravitational time dilation arises in the , which describes the exterior around a spherically symmetric, non-rotating M: 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. For stationary observers at fixed radial coordinate r > 2GM/c^2, the elapses as d\tau = dt \sqrt{1 - \frac{2GM}{c^2 r}}, demonstrating that clocks deeper in the run slower relative to coordinate time measured at . In static metrics like Schwarzschild, the coordinate time t is associated with a timelike \xi = \partial_t, which generates time translations and remains orthogonal to spatial hypersurfaces, ensuring the metric's time-independence. This vector is timelike in the asymptotically flat region (r \to \infty), where g_{00} \to -1, allowing coordinate time to approximate for distant observers and facilitating the matching of local physics to global coordinates. To illustrate the distinction, consider radial light signals (null geodesics, ds^2 = 0) in the Schwarzschild solution. For an outgoing from radius r_e to , the coordinate time interval is \Delta t = \int_{r_e}^\infty \frac{dr}{c \left(1 - \frac{2GM}{c^2 r}\right)}, which diverges logarithmically as r_e approaches the event horizon at r_s = 2GM/c^2, whereas the for a stationary emitter at r_e is finite but redshifted. Conversely, an infalling observer measures a finite to reach the horizon, while coordinate time for the signal remains undefined inside r_s. This highlights how coordinate time captures asymptotic behavior without directly corresponding to local clock readings.

Implementation and Measurement

Observational and Experimental Methods

Atomic clocks serve as the primary standards for realizing coordinate time in relativistic frameworks, with cesium-beam and clocks providing the necessary precision for both terrestrial and space-based applications. Cesium atomic clocks, based on the hyperfine transition frequency of the cesium-133 atom at approximately 9.192 GHz, achieve long-term and accuracy essential for defining scales like (), where they contribute to the weighted ensemble average through primary frequency standards such as , with an accuracy of about 10^{-16}. clocks, utilizing the 21 cm hyperfine transition of neutral , offer superior short-term on the of 10^{-15} over integration times of 100 seconds, making them ideal for high-precision timing in () and satellite missions, where they maintain phase coherence for coordinate time synchronization. These clocks incorporate relativistic corrections for and velocity effects to ensure alignment with coordinate time definitions in . Interferometric techniques and pulsar timing arrays provide astronomical-scale verification of coordinate time predictions, leveraging the stability of pulsars as celestial clocks to test relativistic effects over vast distances. pulsars, rotating hundreds of times per second with pulse arrival times predictable to within nanoseconds over years, enable measurements of delay and that confirm general relativistic and periastron advance, as demonstrated in analyses of binary systems like PSR B1913+16. Pulsar timing residuals, after accounting for interstellar dispersion and solar system , yield tests of the strong with precision approaching 10^{-5} of the predicted relativistic effects, using arrays such as the North American Nanohertz Observatory for (NANOGrav). These methods validate coordinate time scales by comparing observed pulse phases against models incorporating (TCB). The Hafele-Keating experiment in 1971 provided early empirical confirmation of coordinate time variations due to both special and general relativistic effects, using four cesium-beam atomic clocks flown on commercial airliners around the world. Clocks on eastward flights lost 59 ± 10 ns relative to ground stations, while those on westward flights gained 273 ± 7 ns, combining kinematic and gravitational contributions, with results agreeing with relativistic predictions to within experimental uncertainties of about 10%. Complementing this, the Gravity Probe A rocket experiment in 1976 launched a clock to 10,000 km altitude, measuring a of 4.5 × 10^{-10} over two hours, consistent with to 0.01% precision, as the clock rate increased by the predicted factor (1 + gh/c²) where h is altitude. These terrestrial and suborbital tests established the feasibility of coordinate time realization under dynamic conditions. Satellite-based astrometric missions like and further verify coordinate time scales through precise measurements of stellar positions and s, incorporating relativistic light propagation delays in their data reduction. The mission (1989–1993) achieved parallax accuracies of 1 milliarcsecond by modeling observations in the barycentric celestial reference system, applying corrections for annual aberration and gravitational deflection using , which ensured positional consistency at the microarcsecond level. Similarly, the mission, operational since , processes billions of stellar observations with a relativistic astrometric model that includes coordinate time-dependent effects like the gravitomagnetic field, yielding precisions of 0.1 milliarcsecond per year and validating time scales through consistency checks against pulsar-based ephemerides. Modern experimental setups achieve relativistic corrections in coordinate time realization at precisions of parts in 10^{15}, enabling tests of fundamental physics beyond classical limits. For instance, comparisons of optical clocks, such as single-ion and types separated by millimeter to centimeter heights, demonstrate with fractional shifts on the order of 10^{-17} per meter, as in early measurements of (7.8 ± 1.3) × 10^{-17} and more recent clock tests approaching 10^{-18}. Error analyses in cesium fountain clocks reveal systematic uncertainties from and relativistic Doppler shifts below 10^{-16}, with hydrogen masers in VLBI networks maintaining timing residuals under 10 femtoseconds after corrections, thus supporting coordinate time scales with uncertainties far below astronomical signal levels.

Practical Applications

In global navigation satellite systems (GNSS) such as GPS, coordinate time forms the basis for precise positioning and timing, requiring relativistic corrections to account for gravitational and velocity effects on satellite clocks. These corrections include a net adjustment of approximately 38 microseconds per day to compensate for the faster rate of satellite clocks relative to ground clocks, arising from weaker gravitational fields and orbital velocities. Additionally, the due to and periodic variations from , with amplitudes around 46 nanoseconds, are incorporated into the coordinate time calculations to ensure synchronization accuracy within nanoseconds. GPS time itself serves as a continuous coordinate time scale in the frame, enabling reliable broadcast of time signals for user receivers. For deep space missions, coordinate time is critical for signal timing and navigation, particularly through the use of (TCB) to model propagation delays over vast distances. In the , which has operated since 1977, TCB-based ephemerides compute one-way light times exceeding 20 hours for , ensuring accurate command sequencing and data return via the Deep Space Network. Similarly, the mission to and beyond relies on for ranging and Doppler measurements, achieving position accuracies better than 0.3 AU from the solar system barycenter during interstellar phases. In satellite telecommunications networks, coordinate time synchronization prevents errors in data transmission, particularly in time-division multiple access (TDMA) systems where precise timing aligns bursts across geostationary or low-Earth orbit constellations. GNSS-derived coordinate time, such as GPS time, provides the reference for clock alignment, mitigating delays from signal propagation and relativistic effects to maintain bit error rates below 10^{-10}. This ensures seamless global coverage for services like mobile backhaul and , where even offsets could disrupt packet synchronization. Astronomy employs coordinate time in ephemerides to determine planetary positions with , as seen in the Laboratory's DE430 model, which integrates orbits using (TDB), a scaled version of . DE430 spans from 1550 AD to 2650 AD and achieves positional accuracies of 1 meter for inner planets over decades, facilitating precise predictions for observations and mission planning. This coordinate time framework accounts for relativistic light-time corrections, enabling accurate modeling of celestial events like planetary conjunctions. Future applications of coordinate time leverage quantum technologies for enhanced precision in relativistic environments, including optical lattice clocks that achieve stabilities below 10^{-18}, surpassing current atomic standards. These clocks, using trapped neutral atoms in lattices, promise to refine coordinate time in distributed networks for deep space communication and gravitational mapping. In quantum-enhanced relativistic networks, entanglement-shared optical clocks could probe curvature on Earth-scale baselines, supporting with sensitivities to height differences of centimeters.

Established Coordinate Time Scales

Geocentric and Terrestrial Scales

Geocentric and terrestrial time scales provide standardized coordinate times referenced to Earth's or surface, incorporating relativistic corrections to account for gravitational and rotational effects in the geocentric frame. These scales form the foundation for both scientific computations, such as ephemerides, and civil timekeeping, ensuring uniformity across applications like astronomy and global navigation. International Atomic Time (TAI) is a continuous, high-precision realized by the International Bureau of Weights and Measures (BIPM) through the weighted average of readings from over 450 atomic clocks maintained by institutions worldwide. It serves as the primary realization of the second on the rotating and acts as the foundational coordinate time scale for deriving other geocentric times, with its stability ensured by primary frequency standards and its accuracy by a subset of the most precise clocks. TAI began continuous computation in , retroactively defined from January 13, 1958, at 0h UT1, and is disseminated via BIPM Circular T publications. Geocentric Coordinate Time (TCG) is the relativistic coordinate time defined in the Geocentric Celestial Reference System (GCRS), with its origin at the geocenter (Earth's center of mass); it represents the proper time for a hypothetical clock at infinite distance from Earth, unaffected by local gravitational fields. TCG was established by the International Astronomical Union (IAU) in 1991 as part of resolutions on relativistic reference systems. Its rate is adjusted relative to TAI through a linear scaling that accounts for Earth's gravitational potential, given by the defining constant L_G = 6.969290134 \times 10^{-10}, such that the difference accumulates as \Delta t = L_G \times ( \text{JD}_{TT} - 2443144.5 ) seconds, where JD is the Julian Date in Terrestrial Time (TT) and 2443144.5 corresponds to the epoch 1977 January 1, 0h TT. This adjustment ensures TCG runs faster than terrestrial realizations by approximately 0.7 microseconds per day due to the absence of gravitational redshift in its definition. Terrestrial Time (TT) is a scaled version of TCG specifically designed for astronomical observations and calculations, providing a uniform time coordinate in the geocentric system that aligns closely with atomic time for practical use. Defined by IAU Resolution A4 in 1991 and refined in 2000, TT differs from TCG by a prescribed constant rate to maintain continuity with earlier times, expressed as TT = TCG - L_G \times (JD_{TT} - 2443144.5) in seconds, ensuring the scale unit matches the SI second as realized by plus a fixed of 32.184 seconds (TT = + 32.184 s) at the 1977 . This scaling eliminates secular drift, making TT suitable for modeling Earth's motion without ongoing relativistic offsets. TT is realized as TT(BIPM), computed annually by BIPM from international data for enhanced long-term stability. Coordinated Universal Time (UTC) is the civil time standard derived from by the insertion or deletion of leap seconds, as determined by the International Earth Rotation and Reference Systems Service (IERS), to maintain alignment with (UT1) within 0.9 seconds. This adjustment compensates for irregularities in , with leap seconds announced in advance and typically added at the end of or ; as of 2025, 37 leap seconds have been introduced since 1972. UTC is disseminated globally via radio signals and network time protocols, serving as the basis for international time zones and everyday applications while preserving the uniform second of . Relativistic adjustments to these geocentric scales account for Earth's oblateness and , which influence coordinate times through variations in and velocity-dependent effects. Earth's oblateness, characterized by the dynamical ellipticity J_2 \approx 1.0826 \times 10^{-3}, perturbs the , leading to a correction in the metric that affects by up to several nanoseconds per day at the compared to poles; this is incorporated into the definition of L_G as the average potential at the geocenter and further refined for proper times via the post-Newtonian parameter \Phi / c^2, where \Phi includes oblateness terms. Rotational effects introduce a velocity term -v^2 / (2c^2) in the proper time formula for clocks on the surface, causing a slowing of about 100 per day due to equatorial speeds of 465 m/s, while the in rotating frames adds path-dependent delays for signal propagation; these are modeled in the IERS Conventions for transforming to terrestrial reference times. Such corrections ensure consistency between coordinate times and observed clock rates in applications like .

Barycentric and Cosmological Scales

(TCB) serves as the standard time coordinate within the Barycentric Celestial Reference System (BCRS), which is centered at the solar system's barycenter and aligns with the International Celestial Reference System (ICRS) at infinite distance. is the coordinate time in the Barycentric Celestial Reference System (BCRS), with its unit defined as the proper time interval of an ideal clock at rest at infinity in the BCRS, unaffected by solar system gravitational fields, and it functions as the independent argument for the relativistic of solar system bodies. This timescale is independent of Earth's motion, providing a uniform temporal framework for describing the dynamics across the solar system. The relativistic foundation of TCB relies on the post-Newtonian formalism, which approximates for weak gravitational fields and slow velocities relevant to the solar system. In this framework, TCB incorporates the gravitational potentials of and planets, ensuring that the metric in the BCRS includes post-Newtonian corrections to the coordinate time, such as those from the isotropic parameterized post-Newtonian gauge. The IAU 2000 resolutions establish TCB such that its rate is adjusted relative to atomic time scales, with the linear scaling factor chosen so that TCB and Geocentric Coordinate Time (TCG) coincide at the Earth's geocenter on the 1977 January 1 epoch. Barycentric Dynamical Time (TDB) is derived from through a periodic adjustment to align with Earth's , making it suitable for use in ephemerides and dynamical theories referenced to the solar system barycenter. Specifically, TDB differs from by small, periodic variations arising from relativistic effects, approximated as TDB = − ΔTDB, where ΔTDB is a with on the order of 1.5 milliseconds, ensuring TDB remains close to (TT) over long periods. This adjustment facilitates practical astronomical computations by avoiding secular drifts while preserving the relativistic accuracy of . In applications such as detection via methods, TDB provides the precise timing reference needed for Barycentric Julian Dates (BJD_TDB), enabling accurate determination of orbital parameters. For instance, timing variations (TTVs) in systems like those observed by Kepler require BJD_TDB corrections to achieve sub-minute precision, revealing additional through deviations in predicted epochs. This use of TDB ensures that light travel time across the solar system and relativistic effects are accounted for, yielding reliable ephemerides for exoplanetary orbits. On galactic and cosmological scales, coordinate time is often defined in the rest frame of the (), which serves as a universal reference for isotropic expansion and large-scale structure studies. In this frame, corresponds to the measured by comoving observers, providing a synchronized temporal coordinate across the that underpins models of from CMB anisotropies. This CMB-based coordinate time facilitates analyses of galaxy distributions and the universe's age, estimated at 13.80 billion years, by aligning observations with the homogeneous, isotropic Friedmann-Lemaître-Robertson-Walker metric.

References

  1. [1]
    [PDF] Coordinates and Proper Time
    Coordinate Transformations. General relativity drops the assumption that the spatial coordinates are Cartesian and the time coordinate measures physical time ...
  2. [2]
    Special Relativity: Proper Time, Coordinate Systems, and Lorentz ...
    'Time' is also called coordinate time or real time, to distinguish it from 'proper time'. Proper time is also called clock time, or process time, and it is a ...Proper Time · The STR Relationship... · Cartesian Coordinates for Space
  3. [3]
    that all coordinate systems are created equal - Einstein-Online
    In general relativity, all possible coordinate systems are equally valid. This is quite a generalization from special relativity, where only inertial observers ...
  4. [4]
    [PDF] Noninertial Coordinate Time
    Noninertial coordinate time is the time given by a fixed remote clock in a noninertial reference frame synchronized to a fixed master clock in that frame, ...
  5. [5]
    [PDF] SECTION 2 TIME SCALES AND TIME DIFFERENCES - DESCANSO
    Ephemeris time (ET) means coordinate time, which is the time coordinate of general relativity. It is either coordinate time of the Solar-System barycentric.
  6. [6]
    1.2: The Nature of Time
    ### Definitions and Distinction
  7. [7]
    [PDF] Minkowski's Proper Time and the Status of the Clock Hypothesis
    In the limit, one recovers Minkowski's formula for the proper time, (2). Or one can begin with the differential form for proper time (1), and simply.
  8. [8]
    [PDF] ON THE ELECTRODYNAMICS OF MOVING BODIES - Fourmilab
    It is known that Maxwell's electrodynamics—as usually understood at the present time—when applied to moving bodies, leads to asymmetries which do.
  9. [9]
    [PDF] Poincaré on clocks in motion - HAL-SHS
    Nov 26, 2015 · In the month of May, 1905, Henri Poincaré (1905) was arguably the first sci- entist to express the principle of relativity in terms of the form ...
  10. [10]
    [PDF] ON THE ELECTRODYNAMICS OF MOVING BODIES - Fourmilab
    This edition of Einstein's On the Electrodynamics of Moving Bodies is based on the English translation of his original 1905 German-language paper. (published as ...Missing: primary | Show results with:primary
  11. [11]
    [PDF] A look at Einstein's clocks synchronization - arXiv
    This has the advantage of improving the transparence of the symmetry that exists between space and time in special relativity. We can see this clearly in ...
  12. [12]
    Discovering the Relativity of Simultaneity - University of Pittsburgh
    It is routinely assumed that Einstein discovered the relativity of simultaneity by thinking about how clocks can be synchronized by light signals.Missing: primary source
  13. [13]
    On the radar method in general-relativistic spacetimes - ar5iv - arXiv
    If a clock, mathematically modeled by a parametrized timelike curve in a general-relativistic spacetime, is given, the radar method assigns a time and a ...
  14. [14]
    I. Simultaneity and shock synchronization. - ADS
    Synchronization by the Einstein procedure and by slow clock transport turn out to be equivalent if and only if the time dilatation factor is given by the ...
  15. [15]
    Relativity in the Global Positioning System - PMC - NIH
    GPS time is steered so that, apart from the leap second differences, it stays within 100 ns UTC(USNO). Usually, this steering is so successful that the ...
  16. [16]
    [PDF] Space and Time - UCSD Math
    Minkowski's realization that the relativity principle implies many times and spaces, which in turn implies that the world is four-dimensional, naturally ...
  17. [17]
    [PDF] Chapter 6 Curved spacetime and General Relativity
    For a curved spacetime, the proper time dτ is defined to be dτ2 = −. 1 c2 ds2 = −. 1 c2 gαβdxαdxβ . (6.69). Page 12. CHAPTER 6. CURVED SPACETIME AND GR. 76.
  18. [18]
    On the gravitational field of a mass point according to Einstein's theory
    May 12, 1999 · Plain TeX, 7 pages, English translation of the original paper by K. Schwarzschild. Subjects: History and Philosophy of Physics (physics.hist ...
  19. [19]
    [PDF] 5 Schwarzschild metric
    motion so c2dτ2 = c2dt2 ie dτ = dt ie proper time is coordinate time as ... So we can instantly do gravitational redshift for stationary observers as these.
  20. [20]
    7. The Schwarzschild Solution and Black Holes
    The procedure will be to first present some non-rigorous arguments that any spherically symmetric metric (whether or not it solves Einstein's equations) must ...<|separator|>
  21. [21]
    [PDF] General Relativity Fall 2018 Lecture 19: Symmetries, spherically ...
    Killing vector field should be timelike in the asymptotically flat region of spacetime. ... Change to a new time coordinate t(˜t, r). The tr component of ...
  22. [22]
  23. [23]
    [PDF] Timekeeping and time dissemination in a distributed space-based ...
    The net effect of time dilation and gravitational redshift is that the satellite clock appears to run fast by approximately 38 µs per day when compared to a ...
  24. [24]
    [PDF] Relativity in Global Satellite Navigation Systems
    Clocks at rest on geoid beat at equal rates, defining. International Atomic Time. They are synchronized in the underlying inertial frame. Centripetal potential,.
  25. [25]
    [PDF] The Evolution of Deep Space Navigation: 1962-1989 - DESCANSO
    This paper covers the time interval from the launch of NASA's. Mariner 2 mission to Venus in 1962 to the Voyager flyby of Neptune in 1989. NAVIGATIONAL ...
  26. [26]
    [PDF] Telecommunications Synchronization Overview - Reeve
    Synchronization in telecommunications networks is the process of aligning the time scales of transmission and switching equipment so equipment operations occur ...
  27. [27]
    New ITU clock concept for more resilient synchronization networks
    Mar 13, 2024 · GNSS-based synchronization keeps telecoms, data centres, power grids, transport, and public safety systems working on the same time clock.
  28. [28]
    [PDF] The Planetary and Lunar Ephemerides DE430 and DE431 - NASA
    Feb 15, 2014 · The coordinate system and units are similar to those used for DE421 and DE405, but include changes to the definition of coordinate time and the ...
  29. [29]
    Optical atomic clocks | Rev. Mod. Phys.
    Jun 26, 2015 · In this article a detailed review on the development of optical atomic clocks that are based on trapped single ions and many neutral atoms is provided.
  30. [30]
    Quantum networks can probe general relativity on Earth - IQUIST
    Jul 22, 2025 · A new study published in the journal PRX Quantum demonstrates how a network of quantum computers employing optical clocks can probe ...
  31. [31]
    BIPM technical services: Time Metrology
    TT(BIPM) is a realization of Terrestrial Time as defined by the International Astronomical Union (IAU). It is computed in deferred time, each January, based on ...
  32. [32]
  33. [33]
    THE LEAP SECOND - Paris Observatory IERS Centers
    The Coordinated Universal Time (UTC, replacing GMT) is the reference time scale derived from The Temps Atomique International (TAI) calculated by the Bureau ...
  34. [34]
    [PDF] 10 General relativistic models for space-time coordinates and ...
    Before 1991, previous IAU definitions of the time coordinates in the barycen- tric and geocentric frames required that only periodic differences exist between.
  35. [35]
    Lunar Time in General Relativity - arXiv
    Sep 23, 2024 · TCB is defined as the independent argument of the relativistic equations of motion of solar system bodies referred to the Barycentric Celestial ...
  36. [36]
    [PDF] TDB or TCB: does it make a difference?
    • IAU 1976: TDB is a time scale for the use for dynamical modelling of the Solar system motion which differs from TT only by periodic terms. • This definition ...
  37. [37]
    TRANSIT TIMING OBSERVATIONS FROM KEPLER. IX. CATALOG ...
    We started by folding the PDC-MAP Kepler long-cadence data, with the BJDTDB timings, using the ephemeris of NASA Exoplanet Archive in order to obtain a good ...
  38. [38]
    [PDF] 22. Big-Bang Cosmology - Particle Data Group
    Dec 1, 2023 · using the microwave background and large-scale structure: t0 = (13.80 ± 0.02)Gyr [24], where the extra accuracy comes at the price of ...
  39. [39]
    Monopole fluctuation of the CMB and its gauge invariance
    where the coordinate (time) lapse δ η o represents the difference of the ... CMB rest frame provides an absolute frame for all observers in the Universe.