Fact-checked by Grok 2 weeks ago

Speed of light

The speed of light in vacuum, commonly denoted as c, is a fundamental physical constant that represents the propagation speed of electromagnetic radiation, including visible light, in empty space, fixed exactly at 299,792,458 meters per second since its definition in 1983 to establish the meter as the distance light travels in vacuum during 1/299,792,458 of a second. This value is invariant, independent of the motion of the source or observer, and serves as the universal speed limit for matter, energy, and information transfer in the cosmos, as no particle or signal can exceed it according to established physics. First estimated in 1676 by Danish astronomer Ole Rømer through observations of Jupiter's moon Io, which revealed delays in eclipses attributable to light's finite travel time across Earth's orbit, the speed of light was later refined via terrestrial experiments, such as Hippolyte Fizeau's 1849 toothed-wheel method yielding approximately 313,000 km/s and Léon Foucault's 1862 rotating-mirror apparatus measuring about 298,000 km/s in air. By the late 19th century, electromagnetic theory by James Clerk Maxwell predicted c as the ratio of electromagnetic constants, aligning with measurements and confirming light as an electromagnetic wave. In the 20th century, laser interferometry and cavity resonator techniques achieved precision to within parts per billion, culminating in the 1983 redefinition that eliminated measurement uncertainty for c itself. Central to Albert Einstein's 1905 theory of special relativity, c underpins the equivalence of mass and energy via E = mc² and the relativity of simultaneity, time dilation, and length contraction for objects approaching this speed, while general relativity extends its role in spacetime curvature and gravitational effects on light paths. In modern physics, c appears in quantum field theory, cosmology (e.g., horizon distances), and engineering applications like GPS, where relativistic corrections for satellite clocks are essential, and it remains a cornerstone constant in the International System of Units (SI).

Definition and Numerical Value

Exact Value and Notation

The speed of light in vacuum, commonly denoted by the symbol c, is a fundamental physical constant with an exact value of 299792458 meters per second. This precise numerical value was established by the 17th Conférence Générale des Poids et Mesures (CGPM) in 1983, which redefined the metre as the distance traveled by light in vacuum during a time interval of 1/299792458 of a second, thereby fixing c exactly as part of the International System of Units (SI). Prior to this redefinition, c was determined through experimental measurements that carried uncertainties, but the 1983 convention assigned zero uncertainty to its value, making it a defining constant of the SI. The notation c specifically refers to the speed of light in vacuum, an invariant universal constant, and is distinguished from other wave propagation speeds such as phase velocity (v_p) or group velocity (v_g), which can differ in contexts involving dispersive media or materials. In classical electromagnetism, c arises directly from Maxwell's equations as the propagation speed of electromagnetic waves in vacuum. The relevant vacuum equations (in SI units, with no charges or currents) are Faraday's law, \nabla \times \mathbf{E} = -\frac{\partial \mathbf{B}}{\partial t}, and Ampère's law with Maxwell's displacement current correction, \nabla \times \mathbf{B} = \mu_0 \epsilon_0 \frac{\partial \mathbf{E}}{\partial t}, where \mathbf{E} is the electric field, \mathbf{B} is the magnetic field, \epsilon_0 is the vacuum permittivity, and \mu_0 is the vacuum permeability. To derive the wave nature, take the curl of Faraday's law: \nabla \times (\nabla \times \mathbf{E}) = -\frac{\partial}{\partial t} (\nabla \times \mathbf{B}) = -\mu_0 \epsilon_0 \frac{\partial^2 \mathbf{E}}{\partial t^2}. Using the vector identity \nabla \times (\nabla \times \mathbf{E}) = \nabla (\nabla \cdot \mathbf{E}) - \nabla^2 \mathbf{E} and noting that \nabla \cdot \mathbf{E} = 0 in vacuum (from Gauss's law), this simplifies to the wave equation \nabla^2 \mathbf{E} = \mu_0 \epsilon_0 \frac{\partial^2 \mathbf{E}}{\partial t^2}. A similar equation holds for \mathbf{B}. The general solution describes waves propagating at speed v = \frac{1}{\sqrt{\mu_0 \epsilon_0}}, which matches the measured speed of light, identifying electromagnetic waves as light and yielding c = \frac{1}{\sqrt{\mu_0 \epsilon_0}}.

Units and Constants

In the International System of Units (SI), the speed of light in vacuum, denoted c, plays a defining role in establishing the metre as the base unit of length. The metre is defined as the length of the path travelled by light in vacuum during a time interval of $1/299\,792\,458 of a second. This definition, adopted by the 17th General Conference on Weights and Measures in 1983 and retained in the 2019 revision of the SI, fixes the numerical value of c at exactly 299,792,458 metres per second. As a result, the metre derives its realization from the second—the SI base unit of time, defined by the frequency of the caesium-133 hyperfine transition—and c, ensuring the metric system's internal consistency and universality without reliance on physical artefacts. In natural unit systems prevalent in theoretical and particle physics, c is conventionally set to 1, simplifying equations by equating units of length and time. This choice, often combined with the reduced Planck's constant \hbar = 1, expresses physical quantities like energy and momentum in units of inverse length (e.g., electronvolts or GeV), facilitating calculations in quantum field theory and relativity. Planck units extend this approach by incorporating the gravitational constant G, setting c = 1, \hbar = 1, and G = 1 to define scales where quantum mechanics, gravity, and relativity intersect, such as the Planck length (\approx 1.616 \times 10^{-35} m). These systems highlight c's role as a conversion factor between spatial and temporal dimensions, underscoring the spacetime unity in modern physics. The speed of light is integral to the fine-structure constant \alpha, a dimensionless fundamental constant that measures the electromagnetic force's strength. Expressed as \alpha = \frac{e^2}{4\pi \epsilon_0 \hbar c}, where e is the elementary charge, \epsilon_0 is the vacuum permittivity, and \hbar is the reduced Planck's constant, \alpha emerges as a pure number because the dimensions of its components cancel: e^2 / (4\pi \epsilon_0) has units of action times velocity, which \hbar c matches. With a value of approximately $1/137.035999177 (relative uncertainty $1.6 \times 10^{-10}), \alpha governs atomic spectra fine structure and quantum electrodynamics processes, with c's inclusion reflecting the theory's relativistic foundations. For practical applications outside SI, c is often expressed in context-specific units. In engineering, particularly for electromagnetic signals, c \approx 0.98357 feet per nanosecond, commonly approximated as 1 foot per nanosecond to estimate propagation delays in circuits and cables. Astronomically, light traverses approximately 173 AU per day, derived from the 499-second travel time across one astronomical unit (the mean Earth-Sun distance), emphasizing c's scale in solar system dynamics. These representations aid intuitive understanding without altering c's fundamental value.

Fundamental Role in Physics

Invariance in Relativity

One of the foundational postulates of special relativity, as formulated by Albert Einstein, states that the speed of light in vacuum, denoted c, is constant and the same for all inertial observers, irrespective of the motion of the light source or the observer. This invariance implies that classical notions of absolute time and space must be abandoned, leading to the relativity of simultaneity and the unification of space and time into spacetime. To reconcile this postulate with the principle of relativity—which asserts that the laws of physics are identical in all inertial frames—Einstein derived the Lorentz transformations, which replace the Galilean transformations of classical mechanics. Consider two inertial frames S and S', where S' moves at velocity v along the x-axis relative to S. The transformations for coordinates and time are: x' = \gamma (x - vt), \quad y' = y, \quad z' = z, \quad t' = \gamma \left(t - \frac{vx}{c^2}\right), where \gamma = \frac{1}{\sqrt{1 - \frac{v^2}{c^2}}} is the Lorentz factor. These equations ensure that if a light pulse emitted at the origin of S at t=0 satisfies x = ct (and similarly for other directions), then in S' it satisfies x' = ct', preserving the invariance of c. The invariance of c directly implies time dilation and length contraction. For time dilation, imagine a light clock in S' where two mirrors separated by proper distance L_0 reflect a light pulse back and forth; the proper time \Delta \tau for one tick is \Delta \tau = \frac{2L_0}{c}. In S, the light path elongates due to the frame's motion, yielding dilated time \Delta t = \gamma \Delta \tau. Similarly, length contraction follows from the requirement that speeds measured in S match those in S'; a rod of proper length L_0 parallel to the motion appears contracted to L = \frac{L_0}{\gamma} in S. These effects arise solely from enforcing c's constancy across frames. Hermann Minkowski reformulated special relativity in 1908 using four-dimensional spacetime, where the invariant interval is ds^2 = c^2 dt^2 - dx^2 - dy^2 - dz^2. This metric remains unchanged under Lorentz transformations, highlighting c's role as a conversion factor between time and space units. Events connected by null intervals (ds = 0) lie on light cones, which define the causal structure: future-directed light cones bound the region where signals can propagate at or below c, ensuring no faster-than-light causation. Worldlines of massive particles remain within these cones, while light rays trace the cone surfaces. Experimental support for c's invariance includes the 1887 Michelson-Morley experiment, which sought to detect Earth's motion through a hypothetical luminiferous ether by measuring light speed differences in perpendicular directions but yielded a null result, with the observed shift less than one-fortieth of the expected ether drift. In the context of special relativity, this isotropy confirms that light speed is independent of the observer's velocity relative to any medium, providing indirect evidence for the postulate. Subsequent tests, such as Kennedy-Thorndike experiments, further validated the Lorentz transformations by ruling out alternative explanations involving variable c.

Causal Limit on Speeds

In special relativity, the speed of light in vacuum, denoted as c, serves as the ultimate speed limit for any object with nonzero rest mass. Accelerating a massive particle to approach c requires progressively greater energy input, as described by the relativistic total energy formula E = \gamma m c^2, where m is the rest mass and \gamma = \frac{1}{\sqrt{1 - \frac{v^2}{c^2}}} is the Lorentz factor. As the particle's speed v nears c, \gamma diverges to infinity, implying that an infinite amount of energy would be needed to reach or exceed c. This prohibition arises directly from the Lorentz transformations and the invariance of c, ensuring that no massive object can attain superluminal speeds. Hypothetical particles known as tachyons, which would always travel faster than c with imaginary rest mass (m^2 < 0), have been proposed but face significant theoretical challenges. Introduced in the context of quantum field theory, tachyons would require a spacelike four-momentum, leading to issues with Lorentz invariance and particle creation processes. Moreover, their presence in a quantum field theory typically signals an unstable vacuum state, where the field's potential has a maximum rather than a minimum, causing spontaneous decay and rendering the theory unphysical for stable systems. No experimental evidence for tachyons exists, and their incorporation into consistent frameworks remains problematic. The causal limit imposed by c is essential for preserving the principle of causality, which dictates that cause precedes effect in all inertial frames. Superluminal signaling would allow information to propagate along spacelike paths, potentially enabling paradoxes such as sending messages backward in time relative to some observers, including the formation of closed timelike curves in curved spacetime extensions of relativity. By restricting information transfer to speeds at or below c, special relativity ensures a consistent ordering of events without such violations, aligning with the observed unidirectional flow of cause and effect. This limit manifests in the relativistic velocity addition formula, which prevents speeds from combining to exceed c. For two objects moving at speeds u and v' relative to an observer, the combined speed v in the observer's frame is given by v = \frac{u + v'}{1 + \frac{u v'}{c^2}}. Even if both u and v' approach c, v asymptotes to c but never surpasses it, reinforcing the universal prohibition on superluminal motion for causal propagation.

Propagation and Variations

In Vacuum

In vacuum, light propagates as an electromagnetic wave, consisting of mutually perpendicular oscillating electric and magnetic fields that are transverse to the direction of travel. The speed of light c in this medium relates the wave's frequency f and vacuum wavelength \lambda through the fundamental relation c = f \lambda, which holds for all electromagnetic radiation regardless of frequency. This propagation speed is isotropic, meaning c remains constant and identical in all directions for any inertial observer, independent of the source's or observer's motion—a postulate central to special relativity that resolves apparent asymmetries in classical electrodynamics. James Clerk Maxwell's equations in vacuum predict electromagnetic wave propagation via the derived wave equation for the electric field \mathbf{E}: \nabla^2 \mathbf{E} - \frac{1}{c^2} \frac{\partial^2 \mathbf{E}}{\partial t^2} = 0, where c = \frac{1}{\sqrt{\epsilon_0 \mu_0}}, with \epsilon_0 as the electric permittivity and \mu_0 as the magnetic permeability of vacuum; this derivation unifies electricity, magnetism, and light as manifestations of the same phenomenon. Physically, vacuum denotes the absence of matter, enabling unimpeded wave travel at c, though quantum field theory describes it as permeated by fleeting virtual particle-antiparticle pairs from zero-point fluctuations; these do not alter light's speed but manifest in effects like the Casimir force, an attraction between nearby uncharged conductors arising from restricted fluctuation modes between them.

In Media and Effective Speeds

When light propagates through a material medium, its speed v is reduced relative to the vacuum speed c, quantified by the refractive index n = \frac{c}{v}. This reduction arises from the interaction of the electromagnetic wave with the charged particles in the medium, which become polarized and radiate secondary waves that interfere constructively with the incident wave but with a phase delay, resulting in a net propagation speed lower than c. The refractive index typically exceeds 1 for common media like water (n \approx 1.33) or glass (n \approx 1.5), meaning light travels at about 75% or 67% of c, respectively. The refractive index governs refraction at interfaces between media, as described by Snell's law: n_1 \sin \theta_1 = n_2 \sin \theta_2, where \theta_1 and \theta_2 are the angles of incidence and refraction. This law explains phenomena such as the bending of light rays when entering a denser medium, leading to effects like mirages or the apparent depth of objects in water. In non-vacuum environments, the phase velocity v_p = \frac{\omega}{k} = \frac{c}{n} represents the speed of constant-phase surfaces, but it does not necessarily carry information. For signal propagation, the relevant quantity is the group velocity v_g = \frac{d\omega}{dk}, which is the velocity of the wave packet's envelope and determines how energy or information travels. In dispersive media, where n varies with frequency, v_g differs from v_p and remains subluminal (v_g \leq c), ensuring no violation of causality. Dispersion, the wavelength dependence of n, causes shorter wavelengths (e.g., blue light) to experience higher n and slower speeds than longer ones (e.g., red light) in materials like glass. This dispersive effect separates white light into a spectrum when passing through a prism, producing rainbows by differential refraction and total internal reflection in water droplets. In optical fibers, dispersion leads to pulse broadening, where different wavelengths in a signal spread out over distance, limiting high-speed data transmission unless compensated by dispersion-shifted fibers. Cherenkov radiation occurs when a charged particle traverses a medium at a speed exceeding the local phase velocity (v > \frac{c}{n}) but below c, analogous to a sonic boom for light. The emitted coherent shock wave of blue light forms a cone with angle \cos \theta = \frac{c}{n v}, observable in nuclear reactors or particle detectors using water or aerogel. This phenomenon was first observed experimentally in 1934 by Pavel Cherenkov and theoretically explained in 1937 by Igor Tamm and Ilya Frank, providing a method to measure particle velocities without exceeding the universal speed limit.

Apparent Faster-Than-Light Phenomena

Observational Examples

One prominent astronomical observation of apparent faster-than-light motion occurs in the relativistic jets emanating from quasars, where radio-emitting components appear to traverse angular distances on the sky at speeds exceeding c. For instance, in the quasar 3C 279, very long baseline interferometry (VLBI) observations have revealed knot-like features moving with apparent transverse velocities up to approximately 4.3 times the speed of light, as measured over multiple epochs between 1979 and 1981. This illusion arises from relativistic beaming, where the jet is oriented close to our line of sight; the apparent speed is given by v_{\rm app} = \frac{v \sin \theta}{1 - \frac{v \cos \theta}{c}}, with v the actual jet speed, \theta the angle to the line of sight, and c the speed of light, amplifying the projected motion without any material exceeding c. Similar superluminal ejections, with v_{\rm app} reaching 10c or more, have been documented in other quasars like 3C 345 and BL Lac objects, confirming the effect in over 100 sources through coordinated global VLBI campaigns. In binary pulsar systems, timing delays due to light travel time across the orbit provide another example where the effect can lead to misinterpretations of superluminal motion if not properly modeled. For the relativistic binary pulsar PSR B1913+16, pulsar timing observations spanning decades show periodic delays in pulse arrival times due to the varying light path length as the pulsar orbits its companion, with delays up to tens of seconds corresponding to the 7.75-hour orbital period. Without accounting for this geometric light travel time effect—known as the Roemer delay—the observed timing variations could erroneously suggest orbital velocities exceeding c, as the signal's propagation delay mimics an impossibly rapid positional shift. This effect has been precisely quantified in PSR B1913+16, where the orbital inclination and eccentricity confirm the delays align with special relativistic predictions, avoiding any true superluminal inference. Terrestrial optical illusions further illustrate apparent superluminal phenomena without violating causal limits. A classic demonstration involves rapidly closing a pair of long scissors, where the intersection point of the blades sweeps along a distant wall or screen at speeds far exceeding c, as the local motion of each blade remains subluminal. Similarly, projecting a shadow from a moving object, such as sweeping a laser pointer across the Moon, can produce a spot that traverses the lunar surface at apparent speeds orders of magnitude greater than c, since no information or energy travels faster than light—the shadow is merely an absence of illumination propagating geometrically. These effects highlight how phase velocities or image projections can exceed c locally while preserving relativity's prohibition on signal transmission. The expansion of the universe provides a cosmological example of apparent recession speeds surpassing c, observed in the redshift of distant galaxies. According to Hubble's law, galaxies beyond about 14 billion light-years recede at velocities greater than c due to the metric expansion of spacetime itself, rather than local motion through space; for instance, galaxies at redshift z > 1.5 exhibit recession speeds up to several times c. This superluminal recession does not allow causal influence beyond the cosmic horizon, as the expansion stretches the light paths without enabling faster-than-light communication. Observations from the Hubble Space Telescope confirm this in surveys of thousands of supernovae and galaxies, showing the effect increases with distance, consistent with the Friedmann-Lemaître-Robertson-Walker metric.

Experimental Results

Laboratory experiments exploring evanescent waves in quantum tunneling have reported apparent superluminal group velocities, where the peak of a microwave or optical pulse traverses a barrier in less time than expected for light in vacuum. For instance, experiments using frustrated total internal reflection or undersized waveguides demonstrated pulse peaks emerging before the time light would take to travel the same distance, suggesting group velocities exceeding c. However, detailed analysis reveals that these effects arise from the non-local nature of evanescent fields and do not permit superluminal information transfer, as the signal's leading edge propagates at or below c, and reshaping of the pulse ensures causality is preserved. In waveguide structures, phase velocities greater than c are routinely observed for electromagnetic waves propagating below the cutoff frequency, where the wave's phase advances faster than light to compensate for the evanescent field in the transverse direction. This superluminal phase velocity v_p = \frac{\omega}{\beta}, with \beta < \frac{\omega}{c}, has been confirmed in microwave and optical experiments, but the associated group velocity v_g = \frac{d\omega}{d\beta}, which determines the speed of energy and information transport, remains subluminal (v_g < c). Such configurations highlight that while phase propagation can exceed c, no violation of relativity occurs for observable signals. During the early 2000s, several experiments investigated light pulse propagation in dispersive media with anomalous absorption or gain, reporting apparent faster-than-light travel. A prominent example is the 2000 study by Wang et al., where a laser pulse in a cesium vapor cell with coherent population oscillation emerged with its peak advanced relative to c, implying a group velocity over 300 times c. Subsequent analyses, including direct pulse shape measurements, resolved these observations by demonstrating pulse distortion and reformation: the superluminal advance is an artifact of the medium's response, with no part of the pulse or information carrier exceeding c, thus upholding the causal limit. Quantum entanglement experiments, such as those involving Bell state measurements on photon pairs, exhibit correlations that appear instantaneous across large separations, raising questions about faster-than-light influences. The no-communication theorem rigorously demonstrates that such entanglement cannot enable superluminal signaling, as the local density operator for one party's subsystem remains unchanged by the distant measurement, preventing any controllable information transfer without classical communication. This theorem, foundational to quantum information theory, has been verified in numerous entanglement distribution and measurement protocols.

Historical Development

Ancient and Early Concepts

In ancient Greece during the 5th century BCE, Empedocles was the first philosopher to propose that light travels at a finite speed, suggesting that the interval between a stimulus and its perception by the eye indicated a non-instantaneous propagation, though he did not quantify it. This view contrasted sharply with that of Aristotle in the 4th century BCE, who argued that light moves instantaneously, likening it to a presence rather than a motion that requires time, a position that dominated philosophical thought for centuries. During the medieval Islamic Golden Age, the polymath Ibn al-Haytham (known as Alhazen in the Latin West), writing in the early 11th century, advanced the understanding by proposing light propagates in straight lines at a large but finite velocity, which varies depending on the medium's density, laying foundational ideas for later optics without providing a numerical value. In his seminal Book of Optics, he described these concepts based on optical observations. In the 17th century, Galileo Galilei attempted the first experimental measurement of light's speed in 1638, using lanterns positioned on distant hilltops where one observer would uncover a light and time the moment another observer, miles away, responded by uncovering theirs; however, he concluded that the speed was too great to detect with human reaction times and terrestrial distances. This effort marked a shift toward empirical approaches, though unsuccessful in yielding a value. Shortly after, in 1676, Danish astronomer Ole Rømer provided the first quantitative estimate by observing discrepancies in the predicted eclipses of Jupiter's moon Io, attributing them to the finite time light takes to cross varying Earth-Jupiter distances; his calculation yielded approximately 220,000 km/s, remarkably close to the modern value despite assumptions about orbital sizes. In 1728, English astronomer James Bradley refined astronomical measurements using stellar aberration, estimating the speed at approximately 301,000 km/s.

19th-Century Advances

In the early 19th century, Thomas Young conducted the double-slit experiment in 1801, which provided key evidence for the wave nature of light through observed interference patterns. By passing sunlight through two closely spaced slits onto a screen, Young demonstrated that light produces alternating bright and dark fringes, consistent with wave superposition rather than particle behavior. Building on Christiaan Huygens' 17th-century wave theory, Augustin-Jean Fresnel refined the model in the 1810s and 1820s, incorporating the Huygens-Fresnel principle to explain diffraction and polarization. These advancements mathematically predicted a finite speed for light waves propagating through space, aligning with emerging experimental evidence against instantaneous transmission. The first terrestrial measurement of light's speed came in 1849 from Hippolyte Fizeau, who used a toothed-wheel apparatus over an 8.6 km path in air. Light from a source passed through gaps in a rapidly rotating wheel, reflected off a distant mirror, and returned; by adjusting the wheel's 720-tooth rotation to 12.6 turns per second for the first minimum visibility, Fizeau calculated a speed of approximately 313,000 km/s. James Clerk Maxwell's 1865 electromagnetic theory unified electricity, magnetism, and optics, deriving light as an electromagnetic wave with speed given by c = \frac{1}{\sqrt{\epsilon_0 \mu_0}} where \epsilon_0 and \mu_0 are the permittivity and permeability of free space. Using contemporary values, Maxwell obtained approximately 310,000 km/s, closely matching astronomical and Fizeau's results, thus confirming light's electromagnetic character. In the 1880s, Albert A. Michelson improved the rotating-mirror method, achieving higher precision over longer baselines. His 1879–1880 experiments at the U.S. Naval Academy yielded a value of about 299,850 km/s in air, refined further in subsequent work to within 0.01% accuracy, setting a benchmark for vacuum measurements.

Modern Measurements

Terrestrial Techniques

Terrestrial techniques for measuring the speed of light have evolved from mechanical and optical methods to highly precise laser-based approaches, enabling laboratory determinations with exceptional accuracy. One foundational modern variant of the time-of-flight method involves sending a light pulse over a known distance and timing its return, with significant improvements in the mid-20th century. In 1941, W. C. Anderson employed a Kerr cell electro-optic modulator as a high-speed shutter to precisely capture the transit time of light pulses, achieving a value of c = 299776 \pm 14 km/s, corresponding to an accuracy better than 0.01%. Contemporary laser time-of-flight experiments build on this principle using picosecond or femtosecond pulsed lasers over short baselines, such as 20 cm paths with beam splitters and fast photodetectors, yielding results consistent with the defined value of c to within parts per million. Cavity resonance techniques represent a cornerstone of modern terrestrial measurements, leveraging the relationship c = f \lambda, where frequency f is determined from microwave or optical cavity resonances and wavelength \lambda from interferometric calibration. In the 1970s, researchers at the National Institute of Standards and Technology (NIST) pioneered stabilized laser systems in Fabry-Pérot cavities to measure optical frequencies directly against cesium time standards. A seminal 1973 experiment by Evenson et al. used a methane-stabilized helium-neon laser, measuring its frequency and wavelength to obtain c = 299792456.2 \pm 1.1 m/s, reducing uncertainty by two orders of magnitude compared to prior values and establishing a benchmark for subsequent refinements. These methods, refined with frequency combs in later decades, have achieved relative uncertainties below $10^{-11}, tying the speed of light to atomic clocks without relying on physical length standards. Interferometric approaches, particularly variants of the Michelson interferometer, have been instrumental in terrestrial measurements by precisely determining light wavelengths, which, when multiplied by measured frequencies, yield c. Albert A. Michelson adapted his interferometer in the early 20th century for baseline determinations in time-of-flight setups, but later applications focused on wavelength metrology in vacuum paths. For instance, interferometers with evacuated arms and stabilized light sources measure fringe shifts to calibrate \lambda, enabling c computations with accuracies approaching $10^{-8} in pre-laser eras. In the context of length standards, these techniques historically linked the meter to light wavelengths (e.g., the krypton-86 line), indirectly affirming c through the relation c = f \lambda, though post-1983 definitions fix c exactly and derive length from it. Electromagnetic methods derive c from independent measurements of the vacuum permittivity \epsilon_0 and permeability \mu_0 via the relation c = 1 / \sqrt{\epsilon_0 \mu_0}, using capacitance and inductance standards in controlled laboratory conditions. A key early 20th-century implementation by E. B. Rosa and Louis B. Dorsey in 1907 at the Bureau of Standards involved electrostatic and electromagnetic unit comparisons with precision capacitors and coils, yielding c = 299710 \pm 22 km/s and validating Maxwell's theoretical prediction to within 0.03%. Modern codifications, as in CODATA recommendations, incorporate such measurements into adjusted values; prior to the 1983 SI redefinition, these contributed to c = 299792458 m/s with uncertainties around $10^{-8}, now exact by convention while \epsilon_0 is derived accordingly.

Astronomical Methods

The explosion of Supernova 1987A in the Large Magellanic Cloud provided a modern astronomical confirmation that light and neutrinos travel at effectively the same speed. Neutrino detectors, including Kamiokande-II and the Irvine-Michigan-Brookhaven (IMB) experiment, recorded a burst of about 20 electron antineutrinos on February 23, 1987, UTC, while the first visible light photons arrived roughly three hours later on February 24. Over the supernova's distance of approximately 168,000 light-years, any significant speed difference would have produced a arrival time gap of years or more; the observed short delay instead reflects the physics of the explosion, where neutrinos escape the collapsing core almost immediately, whereas photons are trapped and diffuse slowly through the dense stellar envelope before emerging. This near-simultaneous arrival, with neutrinos traveling at more than 99.999% of light speed (consistent with their tiny mass), validates that both massless photons and near-massless particles propagate at the universal speed limit c in vacuum. Gravitational lensing time delays offer a contemporary method to probe the speed of light in the curved spacetime predicted by general relativity. In strong lensing systems, such as quasars multiply imaged by foreground galaxies, the differing geometric paths cause light rays to arrive at Earth with measurable delays, typically days to weeks, between images. These delays arise from the Shapiro time delay effect, where light slows in the gravitational potential well near the lens, combined with path length differences; the formula incorporates c explicitly, assuming it remains constant locally even in curved regions. Observations of systems like the Einstein Cross or COSMOGRAIL-monitored lenses have yielded time delays precise to hours, enabling tests of general relativity by comparing predicted delays (using known distances and lens models) to measurements, thereby confirming that light travels at c without variation in strong gravitational fields. For instance, analyses of lensed supernovae or quasars constrain potential deviations in c to less than 1 part in 10^5, supporting the theory's invariance of light speed across cosmic scales. The detection of gravitational waves (GWs) from binary neutron star mergers, such as GW170817 observed by LIGO/Virgo in 2017, provides a precise astronomical test of c's invariance through multi-messenger astronomy. The GW signal arrived at Earth 1.7 seconds before the gamma-ray burst counterpart detected by Fermi and INTEGRAL, over a distance of about 40 megaparsecs. This near-coincidence, combined with subsequent kilonova observations, constrains the speed of GWs to equal c within 10^{-15} (at 1σ confidence), confirming that massless gravitons (if they exist) propagate at the same speed as photons in vacuum and ruling out significant deviations predicted by some quantum gravity models.

Practical Implications

Terrestrial Effects

The finite speed of light manifests in various terrestrial applications, where propagation delays and related effects influence technology and natural observations on human and planetary scales. In the Global Positioning System (GPS), electromagnetic signals from satellites orbiting at approximately 20,200 km altitude take about 67 milliseconds to reach ground-based receivers, as the one-way distance corresponds to this travel time at the speed of light. This propagation delay is fundamental to pseudorange calculations for positioning, but GPS accuracy also demands corrections for relativistic effects, including special relativistic time dilation from satellite velocities (reducing clock rates by about 7 μs/day) and general relativistic gravitational redshift (increasing rates by about 45 μs/day), ensuring positional errors remain below 10 meters. A common natural demonstration of light's finite speed occurs during thunderstorms, where the flash of lightning is visible almost instantaneously due to the near-instantaneous propagation of light over distances of several kilometers, while the accompanying thunder—sound waves traveling at roughly 343 m/s in air—arrives seconds later. For instance, a 5-second delay between seeing the flash and hearing the thunder indicates the strike was approximately 1.7 km away, calculated as the time for sound to cover that distance (using the rule of thumb: 5 seconds per mile or 3 seconds per kilometer). This disparity highlights light's vastly superior speed compared to sound, allowing observers to gauge storm proximity without specialized equipment. On microscopic scales within electronic devices, light's travel time becomes relevant in high-speed computing. In modern computer chips, where silicon dies measure on the order of 1–2 cm across, electromagnetic signals propagate at speeds close to that of light in vacuum, resulting in delays of approximately 30–70 picoseconds for signals to traverse the chip. Similarly, in cathode-ray tube (CRT) displays, electron beams accelerate to velocities around 10^6 m/s (about 0.003c), taking microseconds to sweep across a 50 cm screen, whereas light from the phosphor would traverse the same distance in mere nanoseconds, underscoring why beam speed limits refresh rates in older televisions. In particle accelerators, the approach to light speed introduces synchrotron radiation as a key terrestrial effect, where relativistic electrons (accelerated to 0.999c or higher) emit intense electromagnetic radiation when forced into curved paths by magnetic fields. This energy loss imposes practical limits on accelerator design; for example, in electron synchrotrons like those at CERN, restoring the radiated energy requires powerful radio-frequency systems, constraining maximum energies to tens of GeV without prohibitively large or costly facilities, as the radiation power scales with the fourth power of the Lorentz factor γ.

Astrophysical and Space Applications

In astrophysics, the finite speed of light, denoted as c \approx 3 \times 10^8 m/s, serves as a fundamental scale for measuring vast cosmic distances, most notably through the unit of the light-year. A light-year is defined as the distance light travels in one Julian year (365.25 days) in vacuum, equivalent to approximately $9.46 \times 10^{12} km. This unit is essential for expressing scales in astronomy, such as the distance to the nearest star beyond the Sun, Proxima Centauri, which lies about 4.24 light-years away, allowing astronomers to conceptualize interstellar and intergalactic separations without cumbersome numerical values. The speed of light imposes significant constraints on spaceflight operations, particularly communication delays due to signal propagation times. For missions to Mars, the round-trip light time varies with planetary positions but typically ranges from 6 to 40 minutes, with an average of about 20 minutes during closer approaches, precluding real-time control of rovers or spacecraft. This lag necessitates autonomous systems for immediate decision-making, as commands from Earth arrive long after events unfold, impacting mission design and safety protocols. In observational astronomy, the finite c enables insights into the universe's history via redshift and look-back time. Cosmological redshift occurs as light from distant galaxies stretches to longer wavelengths due to the expansion of space, quantified by the redshift parameter z, which correlates with recession velocity and distance. Look-back time represents the duration light has traveled to reach Earth, meaning observations of remote objects reveal their past states; for instance, light from galaxies 10 billion light-years away shows the universe as it was 10 billion years ago, facilitating studies of cosmic evolution from the Big Bang onward. The boundary of black holes, known as the event horizon, is defined in general relativity as the surface where the escape velocity equals c, preventing any matter or radiation from escaping once crossed. For a non-rotating black hole, this horizon radius, the Schwarzschild radius r_s = \frac{2GM}{c^2}, marks the point of no return, with G as the gravitational constant and M the mass, underpinning phenomena like gravitational lensing and the information paradox in theoretical physics.

References

  1. [1]
    speed of light in vacuum - CODATA Value
    speed of light in vacuum $c$ ; Numerical value, 299 792 458 m s ; Standard uncertainty, (exact).
  2. [2]
    Physics of Light and Color - Molecular Expressions Microscopy Primer
    Finally in 1983, more than 300 years after the first serious measurement attempt, the speed of light was defined as being 299,792.458 kilometers per second by ...
  3. [3]
    DOE Explains...Relativity - Department of Energy
    The speed of light is incredibly high. Because the speed of light is squared in Einstein's equation, tiny amounts of mass contain huge amounts of energy.
  4. [4]
    Probing Question: Can anything travel faster than the speed of light?
    Mar 19, 2006 · According to Einstein's theory of special relativity, published in 1905, nothing can exceed the speed of light.
  5. [5]
    Ole Roemer Profile: First to Measure the Speed of Light | AMNH
    In 1676, the Danish astronomer Ole Roemer (1644–1710) became the first person to measure the speed of light.
  6. [6]
    July 1849: Fizeau Publishes Results of Speed of Light Experiment
    Jul 1, 2010 · A French scientist named Armand-Hippolyte-Louis Fizeau figured out how to measure the speed of light by means of a terrestrial experiment.
  7. [7]
    The Speed of Light - Las Cumbres Observatory
    In the 1850s, French physicist Léon Foucault measured the speed of light in a laboratory using a light source, a rapidly rotating mirror and a stationary mirror ...
  8. [8]
    [PDF] History of the Speed of Light ( c )
    The speed of light is a very important fundamental constant known with great precision today due to the contribution of many scientists. Up until the late ...
  9. [9]
    [PDF] Speed of Light From Direct Frequency and Wavelength Measurements
    In fact, a 1958 measurement of the speed of light by Froome [3] was done by determining the frequency and wavelength of a microwave source at 72GHz. The ...
  10. [10]
    Relativity - University of Oregon
    The key premise to special relativity is that the speed of light (called c = 186,000 miles per sec) is constant in all frames of reference, regardless of their ...
  11. [11]
    Why Constant Speed of Light | Physics Van | Illinois
    Many experiments have shown that the speed of light is a constant in all directions. This is in agreement with the theory of Special Relativity. As to the ...
  12. [12]
    Physics: Relativity - The Learning Commons | TCTC
    Mar 25, 2025 · The speed of light in a vacuum, c, is a fundamental constant of nature. It occurs in many scientific equations, such as Einstein's famous E ...
  13. [13]
    The International System of Units (SI): Defining constants - BIPM
    The SI is defined in terms of a set of seven defining constants. The complete system of units can be derived from the fixed values of these defining constants.
  14. [14]
    Resolution 1 of the 17th CGPM (1983) - BIPM
    that a new definition of the metre has been envisaged in various forms all of which have the effect of giving the speed of light an exact value, equal to the ...
  15. [15]
    speed of light in a vacuum (S05854) - IUPAC Gold Book
    The speed of light in vacuum, c , is one of seven fixed constants defining the International System of Units, the SI, with c = 299 792 458 m s − 1 .
  16. [16]
    18 The Maxwell Equations - Feynman Lectures - Caltech
    18–5The speed of light ... and ∇×B=jϵ0c2. If we take any arbitrary definition of a unit of charge, we can determine experimentally the constant ϵ0 required in Eq.
  17. [17]
    Maxwell's equations and light
    c = 1/(e0m0)1/2 = 2.998 X 108m/s. Light is an electromagnetic wave: this was realized by Maxwell circa 1864, as soon as the equation c = 1/(e0m0)1/2 = 2.998 X ...
  18. [18]
    [PDF] Mise en pratique - metre - Appendix 2 - SI Brochure - BIPM
    May 20, 2019 · This definition implies that 'The metre is the length of the path travelled by light in vacuum during a time interval of 1/299 792 458 of a ...
  19. [19]
    [PDF] A concise summary of the International System of Units, SI - BIPM
    ... unit Hz, which is equal to s. −1 . length. The metre, symbol m, is the SI unit of length. It is defined by taking the fixed numerical value of the speed of ...<|separator|>
  20. [20]
    [PDF] Natural Units - UF Physics Department
    It is common in the realm of the elementary particle physics to redefine units so that speed of light and Plank's constant become equal to one: c=1 and = =1. ...
  21. [21]
    natural units in nLab
    Jun 16, 2023 · 1. Idea. In physics, by natural units one means physical units of measurements formed from combinations of fundamental physical constants, such ...
  22. [22]
    Ask Ethan: What Is The Fine Structure Constant And Why Does It ...
    Jun 1, 2019 · In terms of the other constants known at the time, α = e²/(4πε_0)ħc, where: e is the electron's charge,; ε_0 is the electromagnetic constant for ...
  23. [23]
    Current advances: The fine-structure constant
    The usual fine-structure constant α is defined as the square of the completely screened charge, that is, the value observed at infinite distance or in the limit ...
  24. [24]
    speed of light [vacuum] foot/nanosecond - Convert Units
    Quick conversion chart of speed of light [vacuum] to foot/nanosecond. 1 speed of light [vacuum] to foot/nanosecond = 0.98357 foot/nanosecond.
  25. [25]
    Astronomical Unit - an overview | ScienceDirect Topics
    At that distance, light from the Sun takes 499.004 782 s to travel to the Earth. By way of comparison, light from the Sun's nearest stellar neighbour, Proxima ...<|control11|><|separator|>
  26. [26]
    [PDF] ON THE ELECTRODYNAMICS OF MOVING BODIES
    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 ...
  27. [27]
    [PDF] Space and Time - UCSD Math
    It was Hermann Minkowski (Einstein's mathematics professor) who announced the new four- dimensional (spacetime) view of the world in 1908, which he deduced from ...
  28. [28]
    [PDF] On the Relative Motion of the Earth and the Luminiferous Ether (with ...
    For on the emission theory the velocity of light must be greater in the water telescope, and therefore the angle of aberration should be less; hence, in order ...
  29. [29]
    Zur Elektrodynamik bewegter Körper - Einstein - Wiley Online Library
    First published: 1905. https://doi.org/10.1002/andp.19053221004. Citations ... Learn more. Copy URL. Share a link. Share on. Email · Facebook · x · LinkedIn ...
  30. [30]
    Possibility of Faster-Than-Light Particles | Phys. Rev.
    A quantum field theory of noninteracting, spinless, faster-than-light particles is described. The field theory is Lorentz-invariant, but must be quantized with ...Missing: original URL
  31. [31]
    Tachyon Paradoxes | American Journal of Physics - AIP Publishing
    May 1, 1971 · It is sometimes inferred from tachyon paradoxes that there is no consistent motion of a system permitting faster than light signals.Missing: source | Show results with:source
  32. [32]
    Causality and Tachyons - Nature
    The special theory of relativity which “still purport to show that causality arguments forbid the existence of faster-than-light particles”Missing: source | Show results with:source
  33. [33]
    15.2 The Behavior of Electromagnetic Radiation - Physics | OpenStax
    Mar 26, 2020 · In a vacuum, all electromagnetic radiation travels at the same incredible speed of 3.00 × 108 m/s, which is equal to 671 million miles per hour.
  34. [34]
    VIII. A dynamical theory of the electromagnetic field
    Oct 27, 2025 · A Dynamical Theory of the Electromagnetic Field. By J. Clerk Maxwell, F.B.S. ... PROFESSOR CLERK MAXWELL ON THE ELECTROMAGNETIC FIELD.
  35. [35]
    [PDF] On the attraction between two perfectly conducting plates
    -. On the attraction between two perfectly conducting plates. By H. B. G. CASIMIR. (Communicated at the meeting of May 29. 1948.) In a recent paper by POLDER ...
  36. [36]
    Refractive Index - StatPearls - NCBI Bookshelf
    May 20, 2023 · The index of refraction is given by the equation: n = c / v. c = speed of light (~300,000 km/s). v = speed of light in a material. Determining ...Definition/Introduction · Issues of Concern · Clinical Significance
  37. [37]
    Science, Optics, and You: Light and Color - Refraction of Light
    Nov 13, 2015 · The index of refraction, n, of other transparent materials is defined through the equation: n = c/v. where c is the speed of light, and v is ...<|control11|><|separator|>
  38. [38]
    Refraction, Snell's law, and total internal reflection
    Mar 20, 1998 · When light crosses an interface into a medium with a higher index of refraction, the light bends towards the normal. Conversely, light traveling ...
  39. [39]
    Law of Refraction - Richard Fitzpatrick
    The law of refraction, which is generally known as Snell's law, governs the behaviour of light-rays as they propagate across a sharp interface between two ...
  40. [40]
    Phase and group velocities
    phase velocity: watch the motion of one position in the "average" wave · group velocity: watch the motion of one position in the "envelope" wave.
  41. [41]
    [PDF] Lecture 11: Wavepackets and dispersion
    . • The pulse travels with the group velocity vg. = dω dk k=kc evaluated at the carrier wavenumber/frequency. Note that because σk ≪ kc (σν ≪ νc), the ...Missing: v_p = v_g =
  42. [42]
  43. [43]
    [PDF] Mathematical description of EM waves
    Here, phase velocity = group velocity (the medium is non-dispersive). 8 ... Refractive index of blue light > red light. 10. Dispersion: phase/group ...
  44. [44]
    [PDF] Superluminal and Subluminal Lasers for Precision Metrology
    Because the index in glass is wavelength-dependent, a prism separates white light into a rainbow, where red has the longest visible wavelength and violet has ...
  45. [45]
    [PDF] Optical Wavelength Measurement Based on Microwave-Photonics ...
    Material dispersion is caused by the wavelength dependence of refractive index of the fiber core material. Waveguide dispersion arises because the light ...
  46. [46]
    An Introduction to Cherenkov Radiation - Stanford University
    Mar 15, 2014 · When a charged particle is moving faster than light speed inside a medium, a faint radiation would appear which is called Cherenkov radiation.
  47. [47]
    [PDF] 34. Passage of Particles Through Matter
    Jun 1, 2020 · 34.7.2 Coherent radio Cherenkov radiation. Coherent Cherenkov radiation is produced by many charged particles with a non-zero net charge.
  48. [48]
    [PDF] c:r v.vz OTUJUVJO , Ur-L - OSTI
    The Cherenkov effect describes the radiation of photons by charged particles traveling through an infinite dielectric medium with a nearly uniform speed that ...
  49. [49]
  50. [50]
    Relativistic Beaming
    Superluminal motion is observed between this stationary point in the nozzle and moving shock fronts or other inhomogeneities in the relativistic outflow. In the ...
  51. [51]
    Timing Measurements of the Relativistic Binary Pulsar PSR B1913+16
    Nov 2, 2010 · We present results of more than three decades of timing measurements of the first known binary pulsar, PSR B1913+16.
  52. [52]
    Is Faster-Than-Light Travel or Communication Possible?
    If you project the shadow of your finger using a nearby lamp onto a distant wall and then wag your finger, the shadow will move much faster than your finger. If ...
  53. [53]
    [PDF] Tunneling time, the Hartman effect, and superluminality: A proposed ...
    We conduct a critical analysis of the key experiments that report superluminal tunneling velocities in order to determine what it is that they actually measure.
  54. [54]
    [1111.2402] Superluminal phase and group velocities - arXiv
    Nov 10, 2011 · Superluminal phase and group velocities: A tutorial on Sommerfeld's phase, group, and front velocities for wave motion in a medium, with ...
  55. [55]
    Superluminal signals: an engineer's perspective - ScienceDirect.com
    Abstract. Recent experiments with evanescent waveguide modes demonstrated the existence of superluminal group velocities.
  56. [56]
    Clear message for causality - Physics World
    Dec 3, 2003 · In recent years, the superluminal propagation of light pulses through certain media has led to renewed controversy. In 1995, for example ...
  57. [57]
    Light through the ages: Ancient Greece to Maxwell - MacTutor
    Empedocles, in the fifth century BC, postulated that everything was composed of four elements; fire, air, earth and water. He believed that Aphrodite made the ...
  58. [58]
    [PDF] 650 Years of Optics: From Alhazen to Fermat and Rømer
    The legacy of Ibn al-Haytham in the Book of Optics about the different finite speeds of light spanned six centuries culminating with “Fermat's principle” in ...
  59. [59]
    [PDF] Galileo, measurement of the velocity of light, and the reaction times
    According to the commonly accepted view, Galileo Galilei promoted in 1638 an experiment that seemed able to show that the speed of light is finite. An analysis ...
  60. [60]
    Thomas Young and the Nature of Light - American Physical Society
    ... double-slit experiment to demonstrate the interference of light waves. The demonstration would provide solid evidence that light was a wave, not a particle.
  61. [61]
    July 1816: Fresnel's Evidence for the Wave Theory of Light
    But by the early 19th century, the wave theory was making a comeback, thanks in part to the work of a French civil engineer named Augustin-Jean Fresnel.Missing: finite speed
  62. [62]
    VIII. A dynamical theory of the electromagnetic field - Journals
    Mathematical theories of statical electricity, of magnetism, of the mechanical action between conductors carrying currents, and of the induction of currents ...
  63. [63]
    '…a paper …I hold to be great guns': a commentary on Maxwell ...
    Apr 13, 2015 · For the case of air, for which μ=1, the speed of propagation of light was measured by Foucault to be 298 000 000 m s−1 compared with the value ...
  64. [64]
    The Site Where Albert Michelson Measured the Speed of Light 145 ...
    May 15, 2024 · On a seawall at the U.S. Naval Academy, now an APS Historic Site, Michelson measured the speed of light more accurately than anyone had before.Missing: 1880s 299850
  65. [65]
    Speed of Light - Physclips.
    The speed of light is measured using a time-of-flight method, over a distance of 20 cm, using a high speed pulsed laser, a beam splitter, mirrors and fast ...Missing: 1941 | Show results with:1941
  66. [66]
    [PDF] 6. Laser Frequency Measurements, the Speed of Light, and the Meter"
    Three different techniques have been used to measure the speed of light: 1) time of flight techniques; 2) ratio of electrostatic to electro- magnetic units; ...
  67. [67]
    Measuring The Speed Of Light In 1927 - Hackaday
    May 8, 2020 · Michelson used an 8-sided mirror device that rotates while bouncing light to the reflector and also collecting return light. The speed of the ...
  68. [68]
    Michelson Interferometer
    A device to measure distances to an accuracy of fractions of the wavelength of light, it also provided the critical experiment for the non-existence of the ...
  69. [69]
    CODATA Values of the Fundamental Constants
    speed of light in vacuum · vacuum electric permittivity · vacuum magnetic permeability ... constant.
  70. [70]
    Roemer's Speed of Light
    She concludes that Roemer's speed of light was about 215,000 km/sec., or about 134,000 miles per second, not very far from Kopal's value.5 Fontenelle and ...
  71. [71]
    SUPERNOVA 1987A - Annual Reviews
    Dec 1, 2024 · On February 23.316 UT, 1987, light and neutrinos from the brightest supernova in 383 years arrived at Earth, shocking astrophysicists into a.Missing: peer- | Show results with:peer-
  72. [72]
    Gravitational lens time-delay as a probe of a possible time variation ...
    22 jun 2021 · Strong gravitational lensing systems (SGL) can be used to investigate gravitational and cosmological theories and fundamental physics.
  73. [73]
    The Global Positioning System - NOAA's National Ocean Service
    Aug 12, 2024 · To find the distance from a satellite to a receiver, use the following equation: (186,000 mi/sec) x (signal travel time in seconds) = Distance ...
  74. [74]
    [PDF] Lecture 4: Interconnect RC
    Nov 4, 1997 · Speed of light delay is proportional to the length of the wire, while. RC delay increases with the square of the wire length. Thus, for long ...<|control11|><|separator|>
  75. [75]
    Does it contradict special relativity that an electron beam in a ...
    Dec 28, 2010 · The electron beam in a typical television moves with a speed no higher than 106m/s, which is much less than the speed of light. I wonder what ...
  76. [76]
    Particle accelerator - Synchrotrons, Particles, Physics | Britannica
    The practical limit on the energy of an electron synchrotron is set by the cost of the radio-frequency system needed to restore the energy the electrons lose by ...
  77. [77]
    [PDF] SYNCHROTRON RADIATION
    Synchrotron radiation is of major importance in the design of electron synchrotrons and storage rings. The energy loss strongly affects the size of accelerator ...
  78. [78]
    Glossary term: Light Year - IAU Office of Astronomy for Education
    It is defined as the distance that light travels in vacuum in one year: 9.46 trillion kilometers (km), that is 9.5x1012 km.
  79. [79]
    Key Differences in Operating a Rover on the Moon vs. Mars
    Jun 26, 2017 · For an asset on Mars, the round trip light time varies, from roughly fourteen minutes to up to forty minutes. For a Lunar asset the round-trip ...
  80. [80]
    Hubble Cosmological Redshift - NASA Science
    Sep 17, 2024 · Light from galaxies stretches to increasingly longer, redder wavelengths as it travels through an ever-expanding universe.Missing: than | Show results with:than
  81. [81]
    Look Back in Time to the Early Universe! - NASA Science
    May 30, 2023 · As the universe expands, it stretches out the waves of light coming from distant galaxies. Now you can measure this stretching, or “redshift” ...Missing: explanation authoritative
  82. [82]
    Black Hole Anatomy - NASA Science
    We can think of the event horizon as the black hole's surface. Inside this boundary, the velocity needed to escape the black hole exceeds the speed of light, ...