Fact-checked by Grok 2 weeks ago

Chronometry

Chronometry is the and of measuring time with extreme accuracy, encompassing the study of timekeeping devices and methods to establish standard units of time. It enables precise across various fields, from to scientific research, by tracking intervals and periods through mechanical, electrical, and atomic means. The history of chronometry traces back to ancient civilizations, where early humans relied on natural phenomena like the sun's position for basic time division, leading to the invention of sundials around 1500 BCE in and . Water clocks (clepsydrae) emerged shortly after, around 1400 BCE, allowing time measurement independent of , while incense clocks and clocks provided similar functions in by the 6th century . The mechanical escapement, pivotal to modern chronometry, was developed in monasteries around 1270–1300 , enabling the first weight-driven clocks that marked hours with bells. Significant advancements accelerated in the 17th century with ' invention of the in 1656, which improved accuracy to within 15 seconds per day by regulating oscillation periods. In the 18th century, 's H4 (1761) revolutionized by maintaining accuracy on long sea voyages, solving the problem and earning a British government prize. Early 20th-century innovations included crystal oscillators developed in the 1920s, providing vibrations precise enough for electric clocks, while later developments in the century introduced atomic clocks based on cesium-133 oscillations, defining since 1967 with accuracies better than 1 second in 300 million years. Today, chronometry extends to global positioning systems (GPS), , and scientific experiments, where atomic time standards like (UTC) ensure worldwide . Recent optical clocks offer even greater , surpassing traditional atomic standards. Subfields include horology, focused on the design and craftsmanship of timepieces, and applications in astronomy, physics, and for measuring short intervals or .

Fundamentals

Etymology

The term chronometry derives from the khronos (χρόνος), meaning "time," and metron (μέτρον), meaning "measure" or "meter." This combination reflects the discipline's focus on the scientific of time intervals and durations. The suffix -metry itself originates from the same Greek metron, commonly used in English to denote systems of , as seen in terms like and . The word chronometry first appeared in English in the early , with its earliest recorded use in 1833 by the Sir in his treatise Astronomy. Herschel employed the term to describe the precise fixing of temporal moments in astronomical observations, stating that "chronometry … enables us to fix the moments in which phenomena occur, with the last degree of precision." This introduction aligned with growing advancements in timekeeping instruments during the era, such as marine chronometers, which demanded rigorous scientific terminology for accuracy in and studies. Astronomers like Johann Heinrich von Mädler also contributed to early applications through chronometrical observations in 1833 for the Russian government, underscoring the term's relevance in 19th-century astronomical practice. In contrast to horology, which derives from Greek hōra (ὥρα, "hour" or "season") and logos (λόγος, "study" or "account") and primarily refers to the art and craft of constructing mechanical timepieces like clocks and watches, chronometry encompasses a broader scientific scope. Horology emphasizes the design, manufacture, and maintenance of horological devices, whereas chronometry extends to all methods of precise time measurement, including non-mechanical techniques and theoretical principles across disciplines. This distinction highlights chronometry's role as a foundational metrological science rather than a specialized craft.

Core Concepts

Chronometry is the quantitative science of measuring time intervals and epochs with , encompassing the and application of methods to quantify durations and moments in physical processes. This distinguishes it from qualitative aspects of , such as subjective experiences of duration influenced by psychological factors, which do not rely on standardized . The foundational unit in chronometry is the second (s), the base unit of time in the (SI). It is defined as the duration of exactly 9,192,631,770 periods of the corresponding to the transition between the two hyperfine levels of the of the cesium-133 atom, at rest and at a of 0 . This atomic definition ensures a stable and reproducible standard for time measurement, independent of astronomical or mechanical variations. A core principle of chronometry in is the uniformity of time, positing that time progresses at a constant rate everywhere and is unaffected by spatial location or motion. However, introduces effects on time measurement, including , where the passage of time for an observer differs based on relative velocity, such that a moving clock appears to run slower from the perspective of a stationary observer. Chronometry further differentiates between , which represents the temporal coordinate in a specific reference frame, and , the invariant interval measured directly by a clock along its path through .

Subfields

Biological Chronometry

Biological chronometry, also known as biochronometry, is the scientific study of endogenous biological clocks that govern timing processes in living organisms, particularly focusing on rhythmic phenomena such as circadian cycles. These internal timekeepers enable organisms to anticipate and adapt to environmental changes, synchronizing physiological functions like , , and hormone release with external cues. In mammals, the (SCN), located in the , serves as the primary master clock, coordinating approximately 24-hour rhythms through interconnected neuronal networks that generate self-sustaining oscillations via transcriptional-translational loops involving clock genes like CLOCK and PER. Key techniques in biological chronometry include , a non-invasive method that uses wearable accelerometers to monitor movement patterns and infer sleep-wake cycles over extended periods, providing objective data on rest-activity rhythms without the need for laboratory confinement. For assessing cellular aging, telomere length measurement—often via quantitative (qPCR) or —quantifies the progressive shortening of protective chromosomal caps, which correlates with replicative as described by the , where human fibroblasts typically undergo around 50 population doublings before halting division. In humans, the intrinsic circadian period averages approximately 24.2 hours under constant conditions, slightly longer than the solar day, which requires daily by zeitgebers like to maintain with the 24-hour environment. Disruptions to these rhythms, such as those induced by transmeridian travel or irregular work schedules, desynchronize the SCN from external cues, leading to conditions like disorder—characterized by , fatigue, and —or , which increases risks for metabolic and cardiovascular issues due to chronic misalignment.

Psychological Chronometry

Psychological chronometry, also known as , is the scientific study of the duration of mental processes through the measurement of reaction times in response to stimuli. This field emerged in the as a method to quantify cognitive operations by analyzing the time elapsed between a sensory input and a behavioral output, providing insights into the sequential stages of in the . The foundational work was conducted by Dutch physiologist Franciscus Donders in 1868, who introduced the subtractive method to isolate the time required for specific mental stages by comparing reaction times across tasks of varying complexity. In simple reaction time (SRT) tasks, where participants respond to a single, predictable stimulus, average human response latencies are approximately 200 milliseconds, encompassing basic sensory and motor execution. Choice reaction time (CRT) tasks, involving discrimination among multiple stimuli and selection of an appropriate response, typically take around 400 milliseconds, reflecting additional cognitive demands. Donders' subtractive approach decomposes these latencies into discrete stages: stimulus identification (the time to perceive and categorize the input), response selection (choosing the appropriate action), and response execution (initiating the motor output). By subtracting SRT from more complex tasks like go/no-go or choice reactions, he estimated the duration of identification and selection processes, establishing that mental operations occur in measurable, additive intervals rather than instantaneously. This method has been widely adopted and refined, revealing that identification and selection each add roughly 100-200 milliseconds to baseline motor times. In contemporary research, psychological chronometry integrates with neuroimaging techniques such as (fMRI) to map the neural correlates of these temporal stages. Latency-resolved fMRI allows researchers to track the sequence of activations during reaction tasks, correlating hemodynamic responses with behavioral timings to identify regions involved in stimulus (e.g., ) versus (e.g., prefrontal areas). For instance, studies have shown that response selection engages the with latencies aligning to the 150-250 range observed in subtractive paradigms. These applications extend mental chronometry beyond behavioral measures, enhancing understanding of cognitive timing in clinical contexts like attention deficits. Variations in , influenced by circadian , can modulate these reaction times by up to 20-30 milliseconds across the day.

Geological Chronometry

Geological chronometry, also known as geochronometry, encompasses methods to determine the absolute ages of rocks, minerals, and geological events on timescales ranging from thousands to billions of years, primarily through techniques that exploit the predictable decay of radioactive isotopes. These methods provide quantitative timelines for Earth's history, enabling the construction of the and understanding of processes like and mountain building. The foundational concept in is the of unstable isotopes into stable daughter isotopes, governed by the of . This is expressed as: N = N_0 e^{-\lambda t} where N is the number of atoms remaining at time t, N_0 is the number of atoms, \lambda is the constant (specific to each ), and t is the elapsed time. The , the time for half the atoms to , is related to \lambda by t_{1/2} = \ln(2)/\lambda, providing a constant rate independent of environmental conditions, which allows age calculation by measuring the -daughter ratio in a sample. One of the most precise methods is uranium-lead (U-Pb) dating, which measures the decay of to lead-206 (half-life 4.468 billion years) and to lead-207 (half-life 704 million years) in accessory minerals like crystals, which resist alteration and trap isotopes effectively during . This technique has dated the oldest terrestrial materials, such as crystals from at approximately 4.4 billion years, and contributed to establishing Earth's age at 4.54 billion years through analysis of meteorites and lead isotope ratios. For younger geological events, particularly in volcanic contexts, potassium-argon (K-Ar) dating is widely applied to measure the decay of to argon-40 (half-life 1.25 billion years) in potassium-bearing minerals like and within igneous rocks. Upon cooling below the argon closure temperature (around 300–500°C), argon gas is trapped, allowing the accumulation of radiogenic to be dated, with applications to volcanic rocks from millions of years ago, such as those in the Yellowstone region. Radiocarbon dating, suitable for more recent timescales, relies on the decay of (half-life 5,730 years) in organic materials, formed in the atmosphere and incorporated into living organisms until death, after which it decays without replenishment. This method is effective up to about 50,000 years, as beyond this, the remaining levels become too low for accurate measurement, and it is calibrated using tree rings and lake varves for precision. Complementing radiometric methods, stratigraphic correlation integrates by matching rock layers (strata) across regions based on shared , fossils, and sedimentary features, providing a to assign ages from dated points and resolve gaps in the continuous record of Earth's history.

Physical Chronometry

Physical chronometry examines the fundamental nature of time within the framework of physical laws, spanning scales from the quantum realm to the vast expanse of the . At the and subatomic levels, time is quantized in ways that challenge classical notions, while on cosmic scales, it serves as a intertwined with . This subfield distinguishes itself by focusing on universal principles that govern time's behavior, independent of biological or geological contexts, emphasizing theoretical foundations over practical measurement devices. One key concept is the Planck time, defined as the smallest theoretically meaningful interval of time in current physical theories, approximately $5.39 \times 10^{-44} seconds. This unit arises from combining fundamental constants—the c, the G, and the reduced Planck constant \hbar—as t_p = \sqrt{\frac{\hbar G}{c^5}}, marking the scale where and must unify to describe phenomena like the Big Bang's earliest moments. Below this duration, itself may lose its classical structure, rendering traditional time measurements inapplicable. In the , time is not absolute but relative, exhibiting dilation for observers in motion relative to one another. The , \gamma = \frac{1}{\sqrt{1 - \frac{v^2}{c^2}}}, quantifies this effect, where v is the and c is the ; for speeds approaching c, \gamma increases, slowing time for the moving observer as measured by a stationary one. This principle, derived from the invariance of the , underscores how time propagates through , with c \approx 299792458 m/s serving as the universal constant that links spatial and temporal dimensions across all inertial frames. On cosmic scales, physical chronometry employs observables like the (CMB) radiation to estimate the universe's age at approximately 13.8 billion years. The CMB, the relic glow from the epoch of recombination about 380,000 years after the , provides a snapshot of the early universe's uniformity and expansion history, allowing precise modeling of time's evolution since the . This age determination relies on parameters such as the Hubble constant and matter density, refined through satellite missions analyzing CMB fluctuations.

Time Metrology

Measurement Standards

International Atomic Time (TAI) serves as a primary international time standard, providing a continuous, uniform scale realized through the weighted average of readings from over 450 atomic clocks contributed by more than 80 national metrology institutes worldwide. The BIPM computes TAI monthly in deferred time, ensuring its stability and accuracy by incorporating data from these clocks, with primary frequency standards used for calibration to maintain traceability to the SI second. Established by the 13th General Conference on Weights and Measures (CGPM) in 1967, TAI realizes Terrestrial Time (TT) as defined by the International Astronomical Union, with a fixed offset of TT - TAI = 32.184 seconds. Coordinated Universal Time (UTC), the global civil time standard, maintains synchronization with while adopting the uniform rate of , differing by an integer number of seconds to account for Earth's irregular . UTC is computed by the BIPM using as its base, with adjustments introduced as leap seconds when the difference between UTC and UT1 (a measure of ) approaches 0.9 seconds. As of November 2025, 27 leap seconds have been inserted since 1972, the most recent on December 31, 2016, with none added since then due to stable Earth trends. In , the 27th CGPM resolved to discontinue the practice of inserting leap seconds into UTC after 2035 to ensure long-term stability of time standards. The evolution of time measurement standards reflects a shift from solar-based systems to atomic precision, beginning with (ET) in 1952, which derived uniform time from Earth's orbital motion around the Sun to address irregularities in rotational time. This was supplanted by atomic time following the 1967 CGPM resolution, which defined the SI second—the core unit underlying and UTC—as the duration of 9,192,631,770 periods of the radiation corresponding to the transition between the two hyperfine levels of the of the cesium-133 atom at rest at 0 K. Current discussions within the Consultative Committee for Time and Frequency (CCTF) explore redefining the second using optical transitions in atoms like or , potentially improving accuracy by orders of magnitude; as of the September 2025 CCTF meeting, evaluations of optical frequency standards continue, with a draft proposal targeted for the 2026 CGPM. The Bureau International des Poids et Mesures (BIPM) holds primary responsibility for maintaining through international coordination of clock data and dissemination via publications like Circular T. The International Earth Rotation and Reference Systems Service (IERS) monitors and announces leap seconds six months in advance, ensuring UTC's alignment with astronomical observations while preserving the continuity of atomic time scales.

Precision Techniques

Precision techniques in chronometry encompass advanced instrumentation and methodologies that achieve fractional frequency accuracies on the order of 10^{-15} or better, enabling unprecedented in time for scientific and technological applications. These methods primarily rely on and optical phenomena to define and maintain time standards, surpassing earlier and quartz-based systems by orders of magnitude. Key developments include beam and clocks operating in the domain, as well as more recent optical and clocks that exploit higher-frequency transitions for superior . Cesium beam atomic clocks represent a foundational , utilizing a beam of cesium-133 atoms excited by microwave radiation to measure the hyperfine transition frequency of 9,192,631,770 Hz, which defines in the (). In these clocks, atoms pass through a where they interact with microwaves in a resonant cavity, and the frequency is adjusted to maximize the detection of atoms in the upper energy state, achieving fractional accuracies around 10^{-15}. For instance, the NIST-7 cesium beam clock, operational since 1993, maintains time such that it neither gains nor loses a second in approximately 6 million years. Hydrogen maser clocks provide exceptional short-term stability for precision timing, employing a maser (microwave amplification by stimulated emission of radiation) with hydrogen atoms in a storage bulb to oscillate at the 1,420 MHz hyperfine transition. Unlike cesium beam clocks, hydrogen masers prioritize stability over absolute accuracy, with fractional frequency instabilities as low as 10^{-16} over seconds to hours, making them ideal for applications requiring continuous, low-noise signals, though long-term accuracy is limited to about 10^{-13} due to cavity phase shifts. They are commonly used in ensembles to support time scales like UTC(NIST), where multiple masers contribute to averaging out individual drifts. Optical atomic clocks mark a significant advancement, leveraging electronic transitions in the optical domain (hundreds of terahertz) rather than microwave frequencies, which allows for narrower linewidths and longer interrogation times, yielding accuracies up to 10^{-18} or better. Strontium-87 lattice clocks trap thousands of neutral atoms in an optical lattice formed by interfering laser beams, where the clock transition at 429 THz is probed, achieving systematic uncertainties of 8.1 × 10^{-19} in recent implementations by mitigating environmental perturbations like blackbody radiation shifts. For example, a strontium-87 lattice clock at the National Time Service Center (China) achieved a total uncertainty of 2 × 10^{-18} as of June 2025, further demonstrating advancements in mitigating environmental shifts. Similarly, ytterbium-171 optical lattice clocks at NIST have demonstrated total uncertainties of 3.3 × 10^{-18} as of 2023, with the clock laser locked to the atomic resonance for stable operation. These clocks outperform microwave standards by factors of 100 in accuracy and stability, paving the way for redefining the SI second. A critical technique enhancing cesium-based precision is the atomic fountain method, which employs to slow cesium atoms to microkelvin temperatures using six counter-propagating beams in an optical molasses configuration, reducing thermal motion and enabling interrogation times up to one second. In a fountain clock, the cooled atomic cloud is launched upward by laser momentum, passes through a twice during (Rabi interrogation), and is detected after , achieving fractional accuracies of 2–5 × 10^{-16}, as realized in NIST-F2. This approach minimizes Doppler shifts and second-order Zeeman effects, making fountain clocks the current basis for primary standards. Frequency combs serve as essential tools for linking optical clock frequencies to the domain, generating a spectrum of evenly spaced modes (mode-locked lasers) that act as a "ruler" in space, with the comb's repetition rate (gigahertz) and carrier-envelope offset enabling direct division of optical signals to outputs. Developed in the late , these combs have enabled the first optical clocks to surpass cesium standards in accuracy by providing phase-coherent transfer, as demonstrated in NIST comparisons where optical-to- synthesis achieved noise levels below 10^{-15}. Recent advances as of 2025 include clocks, which use sympathetic cooling and readout of "clock" ions (e.g., aluminum-27) via entangled "logic" ions (e.g., magnesium-25) to detect forbidden optical transitions with minimal perturbation, attaining systematic uncertainties below 10^{-18}, such as NIST's Al+ clock at 9.4 × 10^{-19}. This quantum-entanglement approach enhances readout fidelity for ions with weak , improving overall precision for fundamental physics tests. Additionally, space-based experiments like the Ensemble in Space (ACES) mission, launched to the in April 2025, integrate a cold-atom cesium clock () with hydrogen masers to verify through measurements and , achieving comparisons with ground clocks at 10^{-16} over intercontinental links.

Historical Development

Ancient and Medieval Periods

The earliest known devices for measuring time emerged in ancient civilizations, relying on natural phenomena such as the sun's shadow or the flow of water. In around 1500 BCE, sundials appeared as simple shadow clocks, often portable L-shaped instruments that divided the sunlit day into ten parts using a to cast shadows on a marked surface. These devices marked a foundational step in chronometry by providing a visual means to track daytime hours based on solar position. Concurrently, water clocks, or clepsydrae, were developed in , including , by approximately 1600 BCE, utilizing the steady outflow of water from a container to measure intervals, particularly useful for nighttime or cloudy conditions when sundials failed. This innovation allowed for more consistent timekeeping in administrative and ritual contexts, as the regulated drip or level drop indicated elapsed time. Advancements in ancient chronometry built upon these basics, integrating astronomical observations for greater precision. The Greek astronomer , working between 147 and 127 BCE, proposed dividing the full day into 24 equinoctial hours of equal length, a system derived from his studies of solar and stellar movements to standardize time beyond varying seasonal daylight. This conceptual shift from unequal "temporal hours" to fixed divisions influenced subsequent calendars and influenced Roman adaptations. In the Roman era, sundials proliferated and diversified; by 30 BCE, the architect documented 13 distinct styles across , Asia Minor, and , including conical and planar designs calibrated for local latitudes to enhance accuracy in public forums, temples, and private estates. These instruments, often inscribed with seasonal adjustments, underscored chronometry's role in civic life, from scheduling to astronomical alignments. During the medieval period, chronometry transitioned toward mechanical solutions, driven by monastic needs for prayer timings and urban synchronization. In , the first mechanical clocks emerged in the 13th century, employing verge-and-foliot escapements powered by weights to regulate motion, marking a departure from natural flows. The , installed in 1386, represents one of the oldest surviving examples, featuring a striking mechanism that chimed hours via bells, though it lacked a visible dial and focused on auditory signals for communal use. Parallel innovations in the advanced water-based automata; the Isma'il al-Jazari described the in his 1206 treatise The Book of Knowledge of Ingenious Mechanical Devices, a monumental clepsydra resembling an elephant with internal gears, floats, and figures that released a bird every half hour to mark time, blending engineering precision with symbolic artistry. These developments laid groundwork for escapement technology that would evolve in later centuries, emphasizing reliability in diverse cultural contexts.

Modern and Contemporary Advances

The marked a pivotal era in chronometry, beginning with ' invention of the in 1656, which dramatically improved timekeeping accuracy to about 15 seconds per day by leveraging the pendulum as a to regulate the mechanism. This innovation addressed the limitations of earlier spring-driven clocks, which deviated by up to 15 minutes daily, enabling more reliable astronomical observations and . Building on this foundation, in the 18th century, developed the marine H4, completed in 1761, which solved the problem at sea by maintaining accuracy within seconds per day despite environmental challenges like temperature variations and motion. Harrison's design, resembling a large , used a fusion of materials and innovative compensation techniques to achieve this precision, revolutionizing maritime and earning him a substantial reward from the British . The 20th century brought electronic advancements, starting with the quartz clock invented by Warren Marrison at Bell Laboratories in 1927, which utilized the piezoelectric vibrations of a quartz crystal at around 50,000 Hz to drive a synchronous motor, achieving stabilities orders of magnitude better than mechanical clocks. This technology became the basis for standard timekeeping in observatories and later consumer devices, with errors reduced to fractions of a second per month. The era's crowning achievement was the development of atomic clocks; the first practical cesium-beam atomic clock was built in 1955 at the UK's National Physical Laboratory by Louis Essen and Jack Parry, using the hyperfine transition frequency of cesium-133 atoms to define time with an accuracy of one second in 300 years. This device laid the groundwork for the 1967 redefinition of the second in the International System of Units (SI). In 1978, the launch of the first NAVSTAR GPS satellite integrated atomic clock timekeeping into global navigation, enabling precise positioning by synchronizing satellite signals with ground-based clocks to within nanoseconds. Post-2000 developments have pushed chronometry into the optical and quantum realms. Optical atomic clocks, emerging in the early 2000s, employ laser-cooled atoms such as or probed at optical frequencies—about 100,000 times higher than frequencies—yielding fractional uncertainties below 10^{-18}, far surpassing cesium standards. These clocks, exemplified by NIST's clock, have demonstrated that would neither gain nor lose a second over of the . As of 2025, efforts to redefine the SI second based on an optical transition continue under the Consultative for Time and , with a targeting by 2030 pending on the reference atom and international validation. In July 2025, researchers conducted the world's largest intercontinental comparison of optical atomic clocks, involving ten clocks across multiple countries, demonstrating at levels that support the anticipated redefinition of the second by 2030. Complementing this, has enabled advanced time transfer techniques; by entangling ensembles of atoms across optical clocks, researchers have achieved precision beyond the standard , reducing phase noise in distributed synchronization for applications like GPS enhancements and fundamental physics tests. For instance, entangled atoms in clocks have improved by factors of up to 10, facilitating ultra-precise comparisons over long distances.

Applications

Scientific Uses

In astronomy, chronometry plays a pivotal role in pulsar timing arrays, which detect gravitational waves through precise measurements of pulsar pulse arrival times. The Hulse-Taylor binary pulsar, discovered in 1974, provided the first indirect evidence of gravitational wave emission via the observed decay in its orbital period, consistent with general relativity predictions; this work earned Russell A. Hulse and Joseph H. Taylor the 1993 Nobel Prize in Physics. In 2023, the North American Nanohertz Observatory for Gravitational Waves (NANOGrav) collaboration, along with international pulsar timing arrays, reported evidence for a low-frequency stochastic gravitational-wave background using 15 years of precise pulsar timing data, marking a significant milestone in gravitational-wave astrophysics. Similarly, transit timing variations (TTVs) enable the detection of exoplanets by monitoring deviations in the timing of planetary transits across a star's disk, revealing the gravitational influence of unseen companions in multi-planet systems. In physics, chronometric techniques test fundamental theories like using atomic clocks to measure effects. The 1971 Hafele-Keating experiment flew cesium-beam atomic clocks on commercial airliners eastward and westward around the world, observing time gains and losses of approximately 59 nanoseconds (eastward) and 273 nanoseconds (westbound) relative to stationary clocks, aligning with relativistic predictions for velocity and gravitational potentials. In and , chronometry synchronizes experimental protocols and calibrates methods with time standards for accurate temporal analysis. In studies, precise timekeeping ensures consistent entrainment and measurement of ~24-hour oscillations in constant conditions, allowing researchers to quantify phase shifts and period lengths critical for understanding clock and health implications. In geological applications, techniques function as natural clocks by measuring in rocks, with decay constants calibrated against atomic time to establish absolute ages on the , enabling correlation of events like volcanic eruptions or sediment deposition over millions of years.

Technological and Everyday Uses

Chronometry plays a pivotal role in modern navigation systems, where precise timekeeping is essential for accurate positioning. The (GPS) relies on atomic clocks aboard satellites to achieve frequency accuracies on the order of 2 × 10^{-12}, enabling positioning errors as low as a few meters by measuring signal travel times at the . This level of chronometric precision ensures that even small timing discrepancies, on the order of nanoseconds, do not compromise location accuracy for applications ranging from to autonomous vehicles. As a terrestrial backup to mitigate GPS vulnerabilities such as or spoofing, enhanced Long Range Navigation (eLoran) provides robust timing and positioning signals with no common failure modes to satellite-based systems, supporting continuity in during disruptions. In computing and telecommunications, chronometry underpins network synchronization through protocols like the Network Time Protocol (NTP), which enables distributed systems to maintain clock coherence over the internet with accuracies typically better than 1 millisecond in local networks. NTP version 4, as standardized by the IETF, uses hierarchical stratum servers synchronized to UTC via atomic clocks or GPS, facilitating reliable timestamping in applications such as data logging and distributed computing. Blockchain technologies further leverage precise chronometry for transaction timestamping, where synchronized clocks prevent double-spending and ensure chronological integrity; for instance, Ethereum nodes often derive timestamps from NTP-synced sources to validate block orders with sub-second resolution. Everyday applications integrate chronometry seamlessly into consumer devices and high-stakes operations. Smartphones synchronize their internal clocks using cellular network signals via the Network Identity and Time Zone (NITZ) protocol, which broadcasts UTC-aligned time and timezone data from base stations, achieving accuracies sufficient for calendar functions and alarms without internet dependency. In financial trading, atomic time standards are critical for high-frequency operations, where regulations like the EU's MiFID II mandate clock synchronization to within 100 microseconds of UTC to timestamp trades accurately and enable regulatory surveillance, far surpassing the 50-millisecond tolerance for general automated systems in the US.

Institutions and Resources

Museums and Libraries

Several prominent museums and libraries around the world house significant collections of chronometry artifacts, ranging from mechanical timekeepers to modern precision instruments, preserving the evolution of time measurement for public education and research. In Europe, the in maintains a notable collection of marine chronometers, including examples featuring John Harrison's maintaining power mechanism, which were pivotal in 18th-century advancements in determination at sea. These items, such as pocket chronometers with spring escapements, illustrate early horological innovations and are displayed alongside related historical timepieces. The Musée International d'Horlogerie in , , stands as one of the largest dedicated clock museums globally, featuring an extensive array of timekeeping devices from antiquity to the present, including atomic clocks that demonstrate post-World War II precision timing breakthroughs. Its collection encompasses over 4,500 objects, with the atomic section highlighting instruments that contributed to the 1967 redefinition of based on cesium atom vibrations, underscoring 's role in horological research. In , the National Watch and Clock Museum in , operated by the National Association of Watch & Clock Collectors, boasts the continent's largest assembly of timepieces, with galleries tracing chronometry from sundials and early mechanical clocks to -based technologies that revolutionized consumer timekeeping in the late 20th century. Exhibits on quartz history include pivotal devices like the first quartz wristwatches from the 1960s and 1970s, showcasing the shift from mechanical to electronic regulation and its impact on accuracy and affordability. Complementing this, the in , preserves extensive archival collections on timekeeping, including rare books, manuscripts, and maps that document the scientific and of chronometry, such as treatises on astronomical clocks and calculations. Beyond these regions, the Beijing Ancient Observatory in exemplifies preservation efforts in , with its 15th-century platform housing original bronze instruments like armillary spheres, celestial globes, and sundials used for celestial observations and time reckoning. Built in 1442, the site features eight large astronomical tools that integrated time measurement with calendar-making, offering insights into pre-modern East Asian chronometry.

Professional Organizations

The Bureau International des Poids et Mesures (BIPM), established in 1875 under the Metre Convention, serves as the intergovernmental organization responsible for coordinating international time metrology, including the calculation and dissemination of Coordinated Universal Time (UTC) through integration of atomic clock data from global laboratories. The BIPM Time Department maintains reference time scales such as UTC and Terrestrial Time (TT), ensuring worldwide synchronization for scientific, navigational, and telecommunication applications. The International Union of Pure and Applied Physics (IUPAP) advances chronometry through its Commission on Atomic, Molecular, and Optical Physics (C15), established in 1996, which promotes research in precision frequency standards and optical clocks essential for high-accuracy timekeeping. This commission fosters international conferences and collaborations that drive innovations in relevant to chronometric standards, complementing efforts in symbols, units, and fundamental constants via Commission C2. Nationally, the (USNO) leads in atomic timekeeping as the official source of time for the U.S. Department of Defense, maintaining the Master Clock ensemble of over 100 atomic clocks that realizes UTC(USNO) with femtosecond precision. The USNO's Precise Time Department provides timing signals via GPS and other systems, supporting military navigation and global positioning. The National Institute of Standards and Technology (NIST) Time and Frequency Division, part of the U.S. Department of Commerce, develops and maintains national standards for time and frequency, including primary frequency standards like cesium fountains and optical clocks that contribute to () and UTC(NIST). As of 2025, NIST leads advancements in quantum timekeeping and frequency dissemination for applications in , , and fundamental physics. In the , the National Physical Laboratory (NPL) functions as the national institute, operating the UTC(NPL) time scale and primary frequency standards, including caesium fountains and optical clocks that contribute to (). NPL's Time and Frequency Group develops transportable optical clocks and frequency dissemination techniques, enhancing precision for applications in and fundamental physics. In 2025, global collaborations under initiatives like the international optical clock network have enabled simultaneous comparisons of ten next-generation clocks across six countries, using links and connections to share optical clock signals with unprecedented stability, paving the way for potential redefinition of the second. These efforts, involving institutions such as NIST, PTB, and NPL, focus on quantum-enhanced time transfer and entanglement for distributed chronometry networks.

Key Terms

Definitions of Essential Concepts

In chronometry, an refers to a specific, fixed instant in time designated as a reference point for establishing time scales, calendars, celestial coordinate systems, star catalogs, and orbital parameters. This standardization ensures consistency in measurements across astronomical and navigational applications. For instance, the J2000.0 epoch is defined as noon on January 1, 2000 (Julian Date 2451545.0), serving as the standard reference for modern and ephemerides. Sidereal time measures the Earth's rotation relative to distant rather than , providing a basis for tracking celestial positions independent of solar motion. It is calculated as the of the vernal , with mean sidereal time using the mean equinox and apparent sidereal time using the true equinox to account for . A complete sidereal day, representing one full relative to the , lasts approximately 23 hours, 56 minutes, and 4 seconds of mean . Dynamical time denotes a class of relativistic time scales that superseded in , incorporating to model the uniform motion of celestial bodies. Defined by the (IAU), it includes variants such as (TT) for geocentric ephemerides and (TDB) for barycentric coordinates, ensuring predictions align with observed solar system dynamics. These scales maintain a constant rate based on the second as realized by atomic clocks, adjusted for relativistic effects near or the solar system barycenter.

Specialized Terminology

In chronometry, the serves as a key metric for quantifying the of precision oscillators and clocks, particularly in timekeeping systems. Developed to analyze the statistical properties of s in standards, it provides a time-dependent measure that distinguishes between different noise types, such as white , , and random walk noise, which affect long-term . The Allan deviation, the square root of the variance, is defined as \sigma_y(\tau) = \sqrt{\frac{1}{2} \langle (y_{n+1} - y_n)^2 \rangle}, where \tau is the averaging time, y_n represents the fractional between consecutive measurements, and the angle brackets denote an ensemble average; this formulation allows evaluation of over varying intervals, revealing optimal windows for clocks like cesium fountains or optical lattices. Coherence time in atomic clocks refers to the duration over which the quantum superposition of atomic states—essential for precise interrogation of hyperfine or optical transitions—persists without significant decoherence from environmental perturbations such as fluctuations, , or collisions. In optical lattice clocks, for instance, achieving coherence times exceeding several seconds enables fractional frequency uncertainties below $10^{-18}, as longer coherence allows more Ramsey interrogation cycles and reduces contributions; techniques like dynamical decoupling or lattice engineering can extend this time by mitigating scattering losses and thermal effects. This parameter directly impacts the clock's short-term stability and accuracy, with state-of-the-art systems demonstrating coherence times up to minutes in neutral atom ensembles. The is a discrete adjustment inserted into (UTC) to account for the irregular slowing of relative to time scales, ensuring that UTC remains within 0.9 seconds of UT1, the solar-based timescale. Administered by the International Earth Rotation and Reference Systems Service (IERS), the procedure involves continuous monitoring of the UT1-UTC difference using and ; when predictions indicate the difference will approach 0.6 seconds, the IERS announces a leap second approximately six months in advance, typically adding one second (positive leap) at the end of or by extending the final minute to 61 seconds. Since 1972, 27 such insertions have occurred, all positive, with no negative leaps to date, reflecting the net deceleration of due to friction. In November 2022, the International Bureau of Weights and Measures (BIPM) and the (ITU) agreed to phase out leap seconds starting in 2035, allowing UTC and UT1 to diverge gradually by up to one second until at least 2135.

References

  1. [1]
    CHRONOMETRY definition in American English - Collins Dictionary
    chronometry in British English​​ (krəˈnɒmɪtrɪ ) noun. the science or technique of measuring time with extreme accuracy.
  2. [2]
    CHRONOMETRY Definition & Meaning - Merriam-Webster
    The meaning of CHRONOMETRY is the measuring of time ... : the science of measuring time especially by periods or intervals. Love ...
  3. [3]
    A Walk Through Time - The Evolution of Time Measurement through ...
    This brief essay on the history of timekeeping was conceived and written by Kent Higgins and illustrated by Darwin Miner, of the Program Information Office of ...
  4. [4]
    A Walk Through Time - A Revolution in Timekeeping | NIST
    Aug 12, 2009 · In 1656, Christiaan Huygens, a Dutch scientist, made the first pendulum clock, regulated by a mechanism with a "natural" period of oscillation.
  5. [5]
    A Brief History of Atomic Time | NIST
    Aug 20, 2024 · Since the first societies, humans have needed to keep track of time. Ancient farmers needed to know when to plant and harvest crops.
  6. [6]
    Chronometry - an overview | ScienceDirect Topics
    Chronometry is defined as the measurement of time intervals, which can be applied in various studies such as time and motion analysis. AI generated ...
  7. [7]
  8. [8]
    Timescales and the Factors Influencing Time Perception
    Aug 10, 2025 · This article explores timescales within absolute and psychological times, and identifies the many factors that affect our perception of time passing.
  9. [9]
    second - BIPM
    The second, symbol s, is the SI unit of time. It is defined by taking the fixed numerical value of the caesium frequency Δν Cs, the unperturbed ground-state ...
  10. [10]
    1.2: The Nature of Time - Physics LibreTexts
    Jan 11, 2023 · The feature that best distinguishes proper time from coordinate time is the fact that a coordinate system is not needed to measure proper time.
  11. [11]
    Neuroanatomy, Nucleus Suprachiasmatic - StatPearls - NCBI - NIH
    It is the central pacemaker of the circadian timing system and regulates most circadian rhythms in the body.[1] Multiple afferent neuronal tracts project to the ...
  12. [12]
    Generation of circadian rhythms in the suprachiasmatic nucleus
    Jun 22, 2018 · The suprachiasmatic nucleus (SCN) is the principal circadian clock of the brain. Although individual cells of the SCN contain a self-sustaining ...
  13. [13]
    Actigraphy-Based Assessment of Sleep Parameters - PMC - NIH
    Feb 13, 2020 · Actigraphy, a method for inferring sleep/wake patterns based on movement data gathered using actigraphs, is increasingly used in population-based epidemiologic ...
  14. [14]
    Hayflick Limit - an overview | ScienceDirect Topics
    Hayflick limit defines the number of possible cell divisions and depends on the length of chromosomal telomeres, which decreases in standard cells with every ...
  15. [15]
    What Is Circadian Rhythm? - Sleep Foundation
    Jul 23, 2025 · “An adult's natural internal clock is on average 24.2 hours. We use external stimuli to help entrain this rhythm daily to 24 hours. Changing ...
  16. [16]
    Jet lag disorder - Symptoms and causes - Mayo Clinic
    Nov 19, 2022 · Jet lag, also called jet lag disorder, is a temporary sleep problem that can affect anyone who quickly travels across several time zones.Overview · Symptoms · Causes
  17. [17]
    Shift Work and Shift Work Sleep Disorder - PubMed Central - NIH
    A subset of shift workers develops shift work disorder (SWD), a condition that is triggered by circadian misalignment and which results in insomnia and/or ...
  18. [18]
    Factors influencing the latency of simple reaction time - PMC
    SRT latencies increased from 217.8 ms (200 ms when latencies were corrected ... Stimulus detection times averaged 131 ms and did not change with age ...
  19. [19]
    Reaction time - Human Homo sapiens - BNID 110799
    Laming (Reference) concluded that simple reaction times averaged 220 msec but recognition reaction times averaged 384 msec. This is in line with many studies ...
  20. [20]
    [PDF] MENTAL CHRONOMETRY: Beyond Reaction Time
    MENTAL CHRONOMETRY: ... These findings bear on all models of hu- man cognitive performance that have been developed within the mental chronometry tradition.
  21. [21]
    Mental chronometry using latency-resolved functional MRI - PNAS
    The ability to correlate psychophysical parameters such as reaction time with latency-resolved fMRI allows the determination of which neural substrates are ...
  22. [22]
    Tracking cognitive processes with functional MRI mental chronometry
    Functional magnetic resonance imaging (fMRI) is used widely to determine the spatial layout of brain activation associated with specific cognitive tasks.
  23. [23]
    Timing the Brain: Mental Chronometry as a Tool in Neuroscience
    Feb 15, 2005 · Over the last ten years, fMRI has improved in spatial and temporal resolution and can now provide evidence of quite specific brain areas, in the ...
  24. [24]
    Geologic Time: Radiometric Time Scale
    Jun 13, 2001 · Currently Accepted Half-Life Values ... 5,730 years. The radiocarbon clock has become an extremely useful and efficient tool in dating ...Missing: range | Show results with:range
  25. [25]
    Radiometric Age Dating - Geology (U.S. National Park Service)
    Oct 3, 2018 · The effective dating range of the carbon-14 method is between 100 and 50,000 years. ... Half-life (years). Dated materials. Uranium-238. Lead-206.
  26. [26]
    [PDF] The application of uranium-thorium - USGS Publications Warehouse
    Integration of equation (1) yields the exponential law for the decay of a radioactive species. o where N^ is the number of atoms present at t = 0. Applying ...
  27. [27]
    Geologic Time: Age of the Earth - USGS Publications Warehouse
    Jul 9, 2007 · The age of 4.54 billion years found for the Solar System and Earth is consistent with current calculations of 11 to 13 billion years for the ...
  28. [28]
    [PDF] Radiometric Dating, Geologic Time, And The Age Of The Earth
    Feb 19, 1982 · In a recent reevaluation Tera (1980) concluded that the age of the Earth was about 4.54 billion years. Tera also summarized several other lead ...
  29. [29]
    Going, going, argon! Determining volcanic eruption ages with argon ...
    Jun 10, 2024 · Argon dating is an advancement of the long-used potassium-argon (K-Ar) dating method. Both techniques use the decay of unstable 40K to stable 40 ...
  30. [30]
    A beginner's guide to dating (rocks) | U.S. Geological Survey
    Apr 8, 2024 · Ar/Ar geochronology is based on the decay of the isotope 40K (potassium 40) to 40Ar (argon 40) and is used to determine when volcanic rocks ...
  31. [31]
    Carbon-14 dating, explained - UChicago News
    Carbon-14 has a half-life of about 5,730 years. That means half the atoms in a sample will change into other atoms, a process known as “decay,” in that amount ...
  32. [32]
    Geologic Time - Tulane University
    Oct 17, 2017 · Over the past 150 years detailed studies of rocks throughout the world based on stratigraphic correlation have allowed geologists to correlate ...
  33. [33]
    Planck time - Einstein-Online
    It amounts to about 5·10 -44 seconds [see exponential notation] and is the time it takes light to traverse one Planck length's worth of distance.
  34. [34]
    [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 ...
  35. [35]
    Cosmic History - NASA Science
    Oct 22, 2024 · Around 13.8 billion years ago, the universe expanded faster than the speed of light for a fraction of a second, a period called cosmic inflation.Cosmic Inflation · Big Bang and Nucleosynthesis · Recombination
  36. [36]
    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 ...
  37. [37]
    2021-12-21-record-tai - BIPM
    Dec 21, 2021 · International Atomic Time (TAI) achieves its stability from more than 450 atomic clocks world-wide and its accuracy from a small number of ...
  38. [38]
    Recommendation 2 (1970) - BIPM
    International Atomic Time (TAI) is the time reference coordinate established by the Bureau International de l'Heure on the basis of the readings of atomic ...
  39. [39]
    [PDF] Recommendation on the definition of time-scales The Consultative ...
    Feb 21, 2017 · - International Atomic Time (TAI) is a realization of Terrestrial Time (TT) as defined by the IAU Resolution B1.9 (2000) with TT-TAI = 32.184 s ...
  40. [40]
    2024-02-23-circularT-updates - BIPM
    Feb 23, 2024 · UTC - Coordinated Universal Time - is a time scale produced by the BIPM with the same rate as International Atomic Time (TAI). It differs from ...
  41. [41]
    bulletinc-068.html - IERS Datacenter
    Jul 4, 2024 · NO leap second will be introduced at the end of December 2024. The difference between Coordinated Universal Time UTC and the. International ...
  42. [42]
    Leap Second Announcement - EO | USNO
    NO leap second will be introduced at the end of December 2025. The last leap second was positive and WAS introduced in UTC at the end of December 2016.
  43. [43]
    [PDF] TIME SCALES - BIPM
    Ephemeris Time, ET, is a dynamic time, derived from the theory of the Earth rotation around the Sun: it is provided through an expression for the mean ...
  44. [44]
    committees/cg/cgpm/13-1967 - BIPM
    Resolutions/Recommendations. CGPM Resolution 1 (1967). SI unit of time (second). DOI : 10.59161/CGPM1967RES1E. CGPM Resolution 2 (1967). Unité SI de temps ( ...
  45. [45]
    2024-02-20-redefinition-second - BIPM
    Feb 20, 2024 · * TAI - Temps atomique international / International Atomic Time - is a continuous time scale produced by the BIPM based on the best ...
  46. [46]
    Latest Updates from 2025 CCTF Meeting - BIPM
    Oct 7, 2025 · Work is progressing on the evaluation of optical frequency standards for a future redefinition of the second. Several international and ...
  47. [47]
    FAQ-redefinition-second - BIPM
    Mar 21, 2025 · After the redefinition, caesium standards will have an additional uncertainty (of the order of 1 x 10−16), but this will only be significant in ...
  48. [48]
    Circular T - BIPM
    Circular T provides the values of the differences [UTC – UTC(k)] every five days, for about 80 institutes regularly contributing clock and clock comparison data ...Missing: atomic | Show results with:atomic
  49. [49]
    Clocks Galore | NIST - National Institute of Standards and Technology
    Aug 22, 2024 · A hydrogen maser uses atoms of hydrogen, the lightest chemical element, to keep highly precise time.
  50. [50]
    This New Atomic Clock Is So Precise It Won't Lose a Second for 140 ...
    May 2, 2025 · 1993 – NIST-7 Cesium Beam Clock​​ NIST introduced NIST-7, achieving an accuracy where it wouldn't gain or lose a second in 6 million years.
  51. [51]
    [PDF] Cesium Beam Atomic Time and Frequency Standards
    Observations over a few years are necessary to attain precision comparable to the other factors which limit accuracy. Observations over centuries are necessary.
  52. [52]
    How UTC(NIST) Works
    Nov 18, 2020 · About 2/3 of the atomic clocks that support the UTC(NIST) time scale are hydrogen masers, and about 1/3 are cesium beam clocks.
  53. [53]
    [PDF] Preliminary Evaluation of Time Scales Based on Hydrogen Masers
    H-masers with cavity autotuning can never be a primary reference for producing accurate frequency, due to limits in evaluating the wall shift of 5 10-13 at best.
  54. [54]
    Optical Clocks: The Future of Time | NIST
    Aug 22, 2024 · The best of these clocks are now 100 times more accurate and stable than cesium fountain clocks. And they are still improving, as physicists ...
  55. [55]
    Clock with Systematic Uncertainty | Phys. Rev. Lett.
    Through precise atomic and environmental control, we have realized a strontium optical lattice clock with a total systematic accuracy of 8.1 × 10 - 19 as ...Missing: ytterbium | Show results with:ytterbium
  56. [56]
    A clock with 8×10⁻¹⁹ systematic uncertainty - arXiv
    Mar 15, 2024 · This represents greater than a factor of 2 2 2 2 improvement in systematic accuracy over the previously most accurate strontium optical lattice ...
  57. [57]
    NIST's Cesium Fountain Atomic Clocks
    This frequency is the natural resonance frequency of the cesium atom (9,192,631,770 Hz), or the frequency used to define the SI second. The combination of laser ...
  58. [58]
    [PDF] Lasers, Cold Atoms and Atomic Clocks: Realizing the Second Today
    With respect to the previous Cs atomic beam clocks, the fountain advantages are several, and the accuracy is improved from 5×10-14 to 2-5×10-16. An atomic ...
  59. [59]
    20 years of developments in optical frequency comb technology and ...
    Dec 6, 2019 · Optical frequency combs were developed nearly two decades ago to support the world's most precise atomic clocks. Acting as precision optical ...
  60. [60]
    Optical frequency combs: Coherently uniting the electromagnetic ...
    Jul 17, 2020 · The frequency comb connects the microwave gear in a phase-coherent way, i.e., without slippage, to the smaller fast-spinning gears, each of ...<|control11|><|separator|>
  61. [61]
    Quantum-Logic Clock with a Systematic Uncertainty below
    Here, we report the systematic uncertainty evaluation of an optical atomic clock based on quantum-logic spectroscopy of Al + 27 with a fractional frequency ...Abstract · Erratum · See Also · Article Text
  62. [62]
    NIST's Quantum Logic Clock Returns to Top Performance
    Jul 15, 2019 · Back in 2010, NIST's quantum logic clock had the best performance of any experimental atomic clock. The clock also attracted attention for ...
  63. [63]
    ESA - ACES: Atomic Clock Ensemble in Space
    The European instrument launched on 21 April 2025 on Space X Commercial Resupply Mission 32 and docked with the International Space Station the next day. On ...
  64. [64]
    A Walk Through Time - Early Clocks | NIST
    Aug 12, 2009 · Another Egyptian shadow clock or sundial, possibly the first portable timepiece, came into use around 1500 BCE. This device divided a sunlit ...
  65. [65]
    [PDF] 1 Calendar, Clock, Tower John Durham Peters What is time ... - MIT
    Clepsydrae (water-clocks) were in use in Egypt and Babylon by 1600 BCE. A chief kind of clepsydra in Greece was a container with holes that, when filled with ...
  66. [66]
    [PDF] Scientific American: Why is a minute divided into 60 seconds, an ...
    Mar 5, 2007 · Hipparchus, whose work primarily took place between 147 and 127 B.C., proposed dividing the day into 24 equinoctial hours, based on the 12 hours ...
  67. [67]
    [PDF] recalibration - Time and Frequency Division
    Aug 1, 2011 · The clepsydra, or water clock, could also measure time dur- ing the night and was perhaps the most accurate clock of the ancient world. A ...Missing: BCE | Show results with:BCE
  68. [68]
    [PDF] Islamic Automata in the Absence of Wonder - DSpace@MIT
    May 20, 2010 · (Fig 3.4) The two automata, both from al-Jazari's Compendium act in a large scale pop-ornamental capacity. One is al-Jazari's elephant clock, ...
  69. [69]
    June 16, 1657: Christiaan Huygens Patents the First Pendulum Clock
    Jun 16, 2017 · His designs proved far more accurate at keeping time than the basic spring-driven table clocks of the era, with a drift of only fifteen seconds ...
  70. [70]
    Longitude found - the story of Harrison's Clocks
    In order to solve the problem of Longitude, Harrison aimed to devise a portable clock which kept time to within three seconds a day. This would make it far more ...
  71. [71]
    Standard Quartz Clock | Smithsonian Institution
    In this kind of clock, first built at Bell Telephone Laboratories in 1927, a small crystal of quartz takes the place of a pendulum or balance wheel.
  72. [72]
  73. [73]
    Louis Essen - About us - NPL
    Louis Essen built the world's first working caesium clock and paved the way for a more accurate definition of the second. Louis Essen re-measured the velocity ...Missing: 1949 | Show results with:1949
  74. [74]
    Global Positioning System History - NASA
    Oct 27, 2012 · DoD then followed through and launched its first Navigation System with Timing and Ranging (NAVSTAR) satellite in 1978. The 24 satellite system ...
  75. [75]
    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.
  76. [76]
    Future Directions: Atomic Clocks Meet Quantum Entanglement | NIST
    Apr 4, 2025 · Entangled clocks. As much as quantum entanglement may benefit from atomic clocks, atomic clocks could benefit just as much from entanglement.
  77. [77]
    Entanglement-enhanced optical atomic clocks - AIP Publishing
    Nov 21, 2022 · In this Perspective article, we describe the commonly used experimental methods to create many-body entangled states to operate quantum sensors beyond the SQL.
  78. [78]
    Press release: The 1993 Nobel Prize in Physics - NobelPrize.org
    Gravity investigated with a binary pulsar. The discovery rewarded with this year's Nobel Prize in Physics was made in 1974 by Russell A. Hulse and Joseph H. ...
  79. [79]
    Timing Variations | The Planetary Society
    As of February 2020, the NASA Exoplanet Archive listed 21 planets discovered by the transit timing variation method (also known as transit photometry), 16 ...Missing: chronometry | Show results with:chronometry
  80. [80]
    Around-the-World Atomic Clocks: Predicted Relativistic Time Gains
    During October 1971, four cesium beam atomic clocks were flown on regularly scheduled commercial jet flights around the world twice, once eastward and once ...
  81. [81]
    Keeping an eye on circadian time in clinical research and medicine
    Dec 25, 2022 · The goal of this manuscript is to provide guidance on best practices in measuring metrics of endogenous circadian rhythms in humans and promote ...
  82. [82]
    John Harrison's Horological Legacy - Google Arts & Culture
    John Harrison (1693–1776) was an English clockmaker, celebrated for developing numerous mechanisms which improved the technology of timekeeping devices.
  83. [83]
    chronometer watch | British Museum
    Object Type: chronometer watch ; Museum number: 1958,1201.1576 ; Description: POCKET-CHRONOMETER MOVEMENT WITH ARNOLD'S SPRING DETENT ESCAPEMENT. Pocket- ...
  84. [84]
    Home | Musée International d'Horlogerie | La Chaux-de-Fonds
    Discover the history of watchmaking at the Musée International d'Horlogerie. Explore unique exhibitions and watchmaking treasures in La Chaux-de-Fonds.
  85. [85]
    Man and Time - Musée International d'Horlogerie
    This fascinating collection, which is in the heart of the watchmaking town of La Chaux-de-Fonds, a UNESCO World Heritage Site, reveals the mysteries of time.
  86. [86]
    National Watch and Clock Museum
    514 Poplar Street, Columbia, PA, USA ; Time for Everyone. Public clocks from around the world ; Time in Lancaster. Explore the history of Lancaster watchmaking in ...
  87. [87]
    National watch & clock museum - FHH Certification
    Chronologically, the exhibits take you on a tour through the entire history of timekeeping technology from early non-mechanical devices to today's quartz ...
  88. [88]
    Ancient Beijing Observatory and Ancient Chinese Astronomical ...
    Built in the 7th year of the Zengtong reign (1442) of the Ming Dynasty (1368-1644), the platform is 14 meters high with eight large astronomical instruments ...
  89. [89]
    Honouring the Past and Building the Future through Museums and
    Jul 22, 2025 · It is designed for cultural institutions to promote intergenerational transmission on the liberation movements, and it contains best practices, ...
  90. [90]
    Time - BIPM
    To calculate, disseminate, and improve the world reference time scale UTC through the integration of data from atomic clocks and time transfer measurements at ...
  91. [91]
    C15: Atomic, Molecular, and Optical Physics - IUPAP
    Jan 30, 2023 · The Commission on Atomic, Molecular, and Optical Physics (C15) was established by the International Union of Pure and Applied Physics in 1966.
  92. [92]
    C2: Commission on Symbols, Units, Nomenclature, Atomic Masses ...
    Apr 12, 2021 · The Commission on Symbols, Units, Nomenclature, Atomic Masses and Fundamental Constants (C2) was established by the International Union of Pure and Applied ...
  93. [93]
    U.S. Naval Observatory - CNMOC
    The US Naval Observatory (USNO) provides a wide range of astronomical data and products, and serves as the official source of time for the US Department of ...The USNO Master Clock
  94. [94]
    USNO's Precise Time services - CNMOC
    USNO's Precise Time Department is charged with maintaining the DoD reference for Precise Time and Time Interval (PTTI). That reference is UTC(USNO).
  95. [95]
    Time and frequency - NPL - National Physical Laboratory
    NPL operates the national time scale UTC(NPL) and the UK primary frequency standards, and uses these to contribute to global timekeeping.
  96. [96]
    Time scales - NPL - National Physical Laboratory
    We maintain the national time scale and the primary standards for frequency, contribute to global timekeeping, and disseminate accurate time and frequency ...
  97. [97]
    Unprecedented optical clock network lays groundwork for redefining ...
    Jun 12, 2025 · Researchers carried out the most extensive coordinated comparison of optical clocks to date by operating clocks and the links connecting them simultaneously ...
  98. [98]
    Coordinated international comparisons between optical clocks ...
    May 10, 2025 · The European network of phase-stabilized fiber links connects the optical clocks at NPL in the UK, SYRTE in France, PTB in Germany and INRIM in ...<|control11|><|separator|>
  99. [99]
    Terrestrial Time (TT) - Astronomical Applications Department
    That is, TT is used for the prediction or recording of the positions of celestial bodies as measured by an observer on Earth.
  100. [100]
    Computing Approximate Sidereal Time
    Sidereal time is a system of timekeeping based on the rotation of the Earth with respect to the fixed stars in the sky.
  101. [101]
    Sidereal Time - Astronomical Applications Department
    Sidereal time is the hour angle of the equinox . If the mean equinox is used, the result is mean sidereal time; if the true equinox is used, the result is ...
  102. [102]
    Glossary - Astronomical Applications Department
    sidereal month: see month, sidereal. sidereal time: the hour angle of the equinox. If the mean equinox is used, the result is mean sidereal time; if the ...
  103. [103]
    History of The Astronomical Almanac
    Ephemeris Time (ET) was replaced with Barycentric Dynamic Time (TDB) and Terrestrial Dynamical Time (TDT). The basis for the ephemerides was changed to Jet ...
  104. [104]
    [PDF] The IAU Resolutions on Astronomical Reference Systems, Time ...
    Oct 20, 2005 · ... Time (UT1). The CIO is analogous to the equinox, the reference point on the celestial sphere for sidereal time. Unlike the equinox, however ...
  105. [105]
    [PDF] Handbook of Frequency Stability Analysis
    Feb 5, 2018 · In the context of frequency stability analysis, the standard variance is used primarily in the calculation of the B1 ratio for noise recognition ...
  106. [106]
    [PDF] Role of the IERS in the leap second - ITU
    • Note that to do this, all clocks are stopped by 1 second. • UTC is adjusted by leap seconds to ensure that |UT1-UTC| < 0.9s. Page 9. Coordinated Universal ...
  107. [107]
    [PDF] Role of the IERS in the leap second - BIPM
    Leap seconds are added or subtracted from the last second of a UTC month. • First preference given to end of June or end of December. • Second preference given ...