Fact-checked by Grok 2 weeks ago

Atomic clock

An atomic clock is a precision timekeeping device that measures time by monitoring the resonant frequency of atoms, typically using the stable oscillations of electrons transitioning between energy levels in atoms like cesium-133. These clocks operate by isolating atoms, exciting them with microwaves or lasers tuned to their —such as 9,192,631,770 cycles for the hyperfine transition in cesium—and counting those cycles to define the second, achieving accuracies where the best models lose or gain less than one second over billions of years. Invented in the late 1940s, atomic clocks form the backbone of (UTC) and underpin global standards for time measurement. The development of atomic clocks began with theoretical proposals in the 1870s by scientists like James Clerk Maxwell and , who envisioned using atomic vibrations for timekeeping, but practical realization came in 1948 when Harold Lyons at the National Bureau of Standards (now NIST) built the first ammonia-based atomic clock. In 1955, Louis Essen's cesium clock at the National Physical Laboratory in the UK marked a breakthrough in practicality, leading to the 1967 redefinition of the second by the () based on cesium-133's microwave frequency. Early microwave atomic clocks, such as NIST's NBS-6 in 1975, achieved stability to within one second over 400,000 years, while modern cesium fountain clocks like NIST-F1 (operational since 1999) reach one second in 20 million years. Advancements have shifted toward optical atomic clocks, which use higher-frequency visible light transitions for even greater precision, enabled by the 1999 invention of the optical frequency comb by John Hall and Theodor Hänsch. NIST's 2006 mercury-ion optical clock was the first to surpass cesium standards, and by 2010, their aluminum-ion attained accuracy equivalent to one second in three billion years; recent records, like NIST's 2025 single-ion optical clock, extend this to one second in nearly 60 billion years. These clocks vary in design, from gas cell and beam types to trapped-ion and optical lattice models, with compact versions enabling portable applications. Atomic clocks are indispensable for global positioning systems (GPS), where onboard cesium and clocks synchronize signals to prevent positional errors accumulating at rates of kilometers per day without them. They also support for synchronizing data networks, scientific research in fundamental physics (e.g., testing ), and exploration missions like NASA's Deep Space Atomic Clock for interplanetary navigation. By maintaining UTC through ensembles of clocks at institutions like NIST, they ensure synchronized global infrastructure, from financial transactions to power grids, demonstrating their profound impact on modern society.

Principles of Operation

Basic Mechanism

Atomic clocks are precision timekeeping devices that generate highly stable electromagnetic frequencies derived from quantum transitions between levels in atoms, offering far greater long-term than clocks, which rely on oscillating masses subject to environmental perturbations, or oscillators, which exhibit gradual drift due to aging and variations. Unlike these conventional mechanisms, atomic clocks exploit the intrinsic, reproducible nature of states, where the transition serves as an invariant "tick" unaffected by external factors when properly isolated. The general operating process begins with the preparation of atoms in a controlled , such as a or , to minimize interactions that could perturb their quantum states. These atoms are then interrogated using —typically in the or optical domain—tuned to the corresponding to a specific energy , causing the atoms to absorb or emit photons only at . Detection of this interaction, often through or population changes, generates an error signal that quantifies any deviation between the radiation and the atomic . This signal feeds into a feedback loop, which adjusts a to lock its output precisely to the atomic , producing a stable electrical signal that can be divided down to generate time intervals. A key principle is the use of coherent to probe the levels, where the radiation's is servo-locked to the via the feedback mechanism, ensuring the output mirrors the atomic transition with minimal . This locking process forms the core of frequency synthesis in atomic clocks, converting the quantum phenomenon into a classical timing reference. Atomic clocks often rely on hyperfine transitions as the basis for these standards due to their narrow linewidths and insensitivity to external fields. The setup can be visualized through a simple block diagram: atoms are prepared and exposed to radiation from a tunable local oscillator in the excitation stage; detectors monitor the atomic response to produce an error signal; and a servo controller applies feedback to stabilize the oscillator frequency, closing the loop for continuous operation. To achieve high stability, atomic clocks interrogate ensembles of many atoms simultaneously, averaging their responses to suppress quantum projection noise—the fundamental limit arising from the statistical nature of quantum measurements—with the relative frequency uncertainty scaling as $1/\sqrt{N}, where N is the number of atoms.

Hyperfine Transitions and Frequency Standards

The hyperfine structure in atoms originates from the interaction between the nuclear spin angular momentum \vec{I} and the total electron angular momentum \vec{J}, which splits the otherwise degenerate fine-structure energy levels into multiple sublevels characterized by the total angular momentum quantum number F = I + J, \dots, |I - J|. This interaction is predominantly magnetic in nature, arising from the coupling of the nuclear magnetic dipole moment with the magnetic field generated by the orbiting and spinning electrons. The energy splitting due to this hyperfine interaction is described by the Hamiltonian H_\text{hfs} = A \vec{I} \cdot \vec{J}, where A is the hyperfine structure constant, leading to energy levels given by E_F = \frac{A}{2} [F(F+1) - I(I+1) - J(J+1)], with the splitting between the extreme F levels for J = 1/2 being \Delta E = A (I + 1/2). This form emerges from the relativistic Dirac equation, where the electron's intrinsic spin magnetic moment interacts with the hyperfine magnetic field at the nucleus produced by the nuclear spin; the constant A incorporates Fermi contact, dipolar, and relativistic corrections to the non-relativistic Schrödinger description. In atomic frequency standards, the transition between these hyperfine levels defines a highly stable oscillation frequency \nu = \Delta E / h, where h is Planck's constant; this frequency serves as an invariant unit of time, reproducible across laboratories because it depends solely on universal physical constants and the atom's internal structure rather than external conditions. The precision of such standards is fundamentally limited by the natural linewidth of the transition, given by \Gamma = 1/(2\pi \tau) in frequency units, where \tau is the lifetime (or coherence time) of the upper hyperfine state, determining the minimum uncertainty in the transition frequency and thus the short-term stability of the clock. Ground-state hyperfine transitions are particularly suitable for frequency standards to their exceptionally long coherence times, typically ranging from seconds to hours, which yield high quality factors Q = \nu / \Delta \nu exceeding $10^{15} in optimized systems.

History

Early Developments

The development of clocks began with foundational techniques for measuring and molecular frequencies in the late 1930s. In 1938, physicist Isidor I. Rabi and his collaborators at introduced the molecular beam magnetic resonance (MBMR) method, which involved passing a beam of atoms or molecules through a series of magnets and subjecting them to radiofrequency fields to detect resonance signals corresponding to magnetic moments. This technique achieved precision in measuring nuclear magnetic moments on the order of hertz, laying the groundwork for frequency standards by demonstrating how transitions could be probed non-destructively. Early efforts focused on hyperfine transitions, where the interaction between an atom's and produces stable, reproducible frequencies suitable for timekeeping. Post-World War II demands for precise timing in navigation systems, such as radar and long-range aids like , exposed the limitations of crystal oscillators, which typically offered of only about 1 part in 10^7 over a day due to aging and environmental factors. This spurred the transition to atomic-based methods for superior reproducibility. In 1948, the U.S. National Bureau of Standards (NBS) demonstrated the first operational atomic clock using an , developed by Harold Lyons and his team, which locked a oscillator to the 23.8 GHz inversion transition of the molecule. Although functional, this device suffered from instability caused by cavity pulling and wall interactions, limiting its suitability as a primary frequency standard. A key challenge in these early beam-based experiments was , where the motion of atoms relative to the interrogation fields shifted resonance frequencies and reduced precision. In 1949, Norman F. Ramsey at addressed this with his separated oscillatory fields method, which exposed atoms to two short radiofrequency pulses separated by a field-free drift region, allowing coherent and yielding narrower lines for higher accuracy. This innovation, for which Ramsey received the in 1989, became essential for practical atomic clocks. Building on these advances, in 1955, Louis Essen and Jack Parry at the UK's National Physical Laboratory (NPL) constructed the first cesium atomic clock using a beam apparatus to interrogate the hyperfine transition in cesium-133 atoms, achieving an accuracy of 1 part in 10^9—over an better than quartz standards. Operational from May 24, 1955, this device marked the realization of a stable atomic timekeeper, directly influencing subsequent metrological developments.

Establishment of the Atomic Second

In 1967, General Conference on Weights and Measures (CGPM) formally redefined in the (), marking a pivotal shift from astronomical to atomic standards of time. The new definition established as the duration of exactly 9,192,631,770 periods of the radiation corresponding to the between the two hyperfine levels of the of the cesium-133 atom, at rest and at a temperature of 0 . This precise specification, recommended by the International Committee for Weights and Measures (CIPM) based on measurements from primary cesium frequency standards, tied the unit of time directly to a reproducible quantum in rather than variable celestial motions. This redefinition replaced the ephemeris second, which had been adopted in 1956 as 1/31,556,925.9747 of the mean tropical year for 1900, derived from Earth's orbital motion around the Sun. The ephemeris second aimed to provide a more uniform measure than the mean solar second based on Earth's rotation, but its realization depended on astronomical observations with inherent uncertainties, limiting reproducibility to about 1 part in 10^8. In contrast, cesium atomic standards achieved fractional accuracies of 1 part in 10^13 or better by the mid-1960s, enabling far superior stability and universality across laboratories worldwide. The resulting improvement in precision revolutionized metrology, as the atomic second could be realized independently of geophysical irregularities. The adoption of the atomic second facilitated the establishment of (TAI) in 1967, computed by the International Bureau of Weights and Measures (BIPM) as a weighted of atomic clock readings from multiple national timing laboratories to ensure global consistency. TAI provided a continuous, uniform scale without adjustments for , with its aligned to 1 1958 for compatibility with prior . To maintain synchronization with solar time for civil and navigational purposes, (UTC) was developed, incorporating s into TAI to keep UTC within 0.9 seconds of (UT1). The first leap second was inserted on 30 June 1972. The BIPM played a central role in coordinating international efforts, collecting data from atomic clocks in laboratories such as the National Physical Laboratory (UK), National Bureau of Standards (US), and others to compute TAI. Early validations of clock comparability relied on direct transport of portable atomic standards; in 1964, the first such international comparisons were conducted using a portable cesium clock to align frequencies across distant sites with uncertainties below 10^{-12}. These efforts confirmed the feasibility of a unified atomic timescale, underpinning the 1967 redefinition. The official cesium frequency is thus fixed at \nu_{\text{Cs}} = 9{,}192{,}631{,}770 \, \text{Hz} exactly, representing the inverse of the defined second and anchoring time measurement to fundamental physical constants rather than astronomical phenomena.

Advancements in Microwave and Optical Clocks

Following the establishment of the atomic second in 1967 based on cesium hyperfine transitions, advancements in atomic clock technology have focused on enhancing stability, accuracy, and portability, transitioning from microwave to higher-frequency optical regimes and enabling compact designs. Starting in the , researchers at NIST and other institutions developed optical atomic clocks using trapped ions and neutral atoms; by the early , these achieved fractional stabilities around 10^{-17}, surpassing the 10^{-15} typical of clocks by leveraging higher frequencies for improved . A pivotal innovation enabling this shift was the optical , developed by and , who shared the 2005 for creating methods to measure optical frequencies with unprecedented accuracy by linking them directly to standards through mode-locked laser pulse trains.

Microwave Atomic Clocks

Cesium Beam Clocks

Cesium beam clocks operate by generating a thermal beam of cesium-133 atoms from an oven maintained at approximately 100°C, which effuses into a high-vacuum chamber. The atoms are state-selected using inhomogeneous magnetic fields from sextupole or hexapole magnets to filter those in the lower ground-state hyperfine level (F=3, m_F=0), directing them along a straight trajectory through the apparatus. This beam then passes through a pair of separated radiofrequency (RF) cavities employing the Ramsey interrogation method, where a π/2 in the first partially excites the atoms toward the upper , followed by and a second π/2 in the downstream to interrogate the phase evolution. The Ramsey scheme minimizes first-order and cavity pulling effects, enabling high-resolution of the hyperfine transition. The microwave frequency is locked to the |F=3, m_F=0⟩ to |F=4, m_F=0⟩ hyperfine transition at precisely 9,192,631,770 Hz via a servo feedback loop that adjusts a quartz-crystal oscillator using the detected atomic signal as a reference. After the second cavity, a second state-selection magnet deflects atoms that remain in |F=3, m_F=0⟩ away, while those that have transitioned to |F=4, m_F=0⟩ proceed to a hot-wire detector for ionization and measurement of ion current, or in some designs, to a fluorescence detector where laser-induced emission is counted. This detection provides the error signal for frequency correction, ensuring the local oscillator tracks the atomic resonance with minimal phase noise. A key advancement in cesium clock technology came in the with the development of cesium fountain clocks, which replace the thermal beam with laser-cooled atoms to achieve longer interrogation times. In these devices, cesium atoms are first cooled to microkelvin temperatures using six-beam optical and then launched upward by a pair of vertical beams, forming a symmetric under . As the atom cloud rises and falls through the Ramsey —typically over a 1-meter —the effective interaction time extends to about 1 second, compared to milliseconds in traditional beam clocks, significantly enhancing resolution. State detection in fountains relies on resonant , where atoms in the upper hyperfine state scatter more photons, providing a high-fidelity readout. The phase accumulation during Ramsey interrogation is given by \phi = 2\pi \nu t, where \nu is the hyperfine transition frequency and t is the free-evolution time between cavities, reaching up to 1 s in fountain designs to yield fringe widths below 1 Hz. This extended coherence time contributes to short-term stability characterized by the Allan deviation \sigma_y(\tau) \approx 10^{-13} \tau^{-1/2}, where \tau is the averaging time in seconds, limited primarily by atomic shot noise. Cesium-133's nuclear spin of I = 7/2 results in hyperfine levels with multiple Zeeman sublevels (7 for F=3 and 9 for F=4), enabling precise compensation through observation of field-sensitive transitions to extrapolate the zero-field hyperfine and minimize perturbations from ambient fields. This feature, combined with careful shielding, ensures the clock's insensitivity to environmental magnetic fluctuations, a critical aspect for primary standards.

Rubidium Vapor Clocks

Rubidium vapor clocks, also known as gas cell atomic clocks, utilize a sealed glass cell containing vapor as the core component of their physics package. The cell, typically on the order of centimeters in scale, confines the rubidium atoms and is integrated with a tuned to the hyperfine ground-state transition frequency of approximately 6.835 GHz. is achieved using either a rubidium discharge lamp or a , which selectively excites the atoms to align their spins and create a between the hyperfine levels, preparing them for microwave interrogation. This design enables compact, low-cost implementation, making rubidium vapor clocks prevalent in commercial applications such as and systems. The operation relies on the double-resonance method, where linearly polarized light from the pump passes through the , and a field at 6.835 GHz is applied perpendicular to the light propagation. In the presence of a small DC magnetic field (C-field) of 50–300 , the microwave induces transitions between the hyperfine levels, leading to a detectable change in light transmission due to magnetically induced . This dichroism arises from the Zeeman splitting and provides a dispersive error signal for locking a oscillator to the atomic , ensuring stability. A buffer gas, such as at around 12 , is added to the cell to mitigate wall collisions by promoting diffusive motion of the atoms, though it introduces frequency shifts proportional to pressure (e.g., +548 Hz/Torr for nitrogen). These shifts are temperature-dependent, with a coefficient of +0.46 Hz/Torr/°C, necessitating precise control. The hyperfine transition in rubidium-87, involving the splitting of the ground state due to nuclear spin interaction, serves as the frequency standard, referenced briefly here for context in the clock's interrogation process. Collisional broadening from the buffer gas results in a resonance linewidth \Delta \nu of 10–100 Hz, allowing for high signal-to-noise ratios despite the inhomogeneous environment. Over a one-day averaging time, these clocks achieve fractional frequency stability of $10^{-11} to $10^{-12}, limited primarily by environmental sensitivities rather than quantum noise. Their cm-scale size, combined with production costs far lower than beam-based alternatives, positions them as the workhorse for most commercial atomic clocks. A key challenge in rubidium vapor clocks is their sensitivity to external perturbations, including temperature fluctuations that can cause shifts of $10^{-10}/^\circC and magnetic fields inducing variations of $10^{-12} to $2 \times 10^{-11}/Gauss along the C-field axis. These effects are mitigated through temperature-controlled ovens maintaining the cell at ~65°C and multi-layer shielding, which provides a shielding factor of ~200,000 to suppress ambient fields. Such measures ensure reliable performance in non-laboratory settings, though they require ongoing calibration to counteract long-term drifts from buffer gas interactions.

Hydrogen Maser Clocks

Hydrogen maser clocks operate on the principle of from atoms in the , specifically the 21 cm transition between the F=1 and F=0 levels. Atomic is produced by dissociating molecular in an RF discharge, and a sextupole magnetic state selector focuses atoms in the upper hyperfine state (F=1, m_F=0) into a storage bulb coated with Teflon to minimize wall relaxation. The bulb is placed within a high-Q tuned to the frequency of approximately 1.420 GHz, where the leads to oscillation, with the cavity amplifying the emitted signal. In operation, the active generates a output at the hyperfine , characterized by exceptionally low due to the quantum-limited process. A servo system continuously tunes the cavity to counteract cavity pulling effects, where the finite cavity Q shifts the away from the ; this stabilization maintains the output locked to the . The maser gain, which determines the and output , is proportional to the product of N, stimulated emission cross-section \sigma, and effective storage length L along the cavity axis, expressed as G \propto N \sigma L. The quality factor Q for the hyperfine reaches approximately $10^{10}, reflecting the narrow linewidth and long time of the stored atoms. Hydrogen maser clocks achieve the best short-term frequency among room-temperature standards, with fractional instability as low as $10^{-15} at an averaging time of 1 second, enabling their use in demanding scientific applications such as (VLBI) for and deep space tracking for . However, they exhibit frequency drift on the order of $10^{-14} per year due to interactions between atoms and the storage walls, which cause shifts in the hyperfine frequency. Cryogenic versions of masers, cooled to around 1 K using superfluid , increase by reducing and minimizing spin-exchange collisions, potentially improving short-term by a factor of 100, though wall shift effects persist at lower levels of approximately $3 \times 10^{-11}.

Optical Atomic Clocks

Trapped-Ion Optical Clocks

Trapped-ion optical clocks employ laser-cooled ions, such as aluminum (Al⁺) or (Yb⁺), confined in electromagnetic traps to realize high-precision optical standards. These ions are typically held in Paul traps, which use oscillating radiofrequency fields for confinement, or Penning traps, which combine static magnetic and electric fields for stability. The design isolates the ions from environmental noise, such as collisions or , enabling extended times essential for precision measurements. Interrogation occurs via narrow-linewidth lasers tuned to or visible wavelengths, with the Al⁺ clock transition specifically at 267 nm corresponding to a of 1.12 PHz. The operational principle relies on spectroscopy for non-destructive readout of the clock transition, paired with electron shelving for state detection. In this approach, the clock —lacking a suitable cycling transition for direct —is co-trapped with an auxiliary , such as magnesium (Mg⁺), which facilitates sympathetic cooling and logical coupling. The auxiliary 's strong signals the clock 's state via shared motional modes, while electron shelving shelves the clock in a to preserve phase information during probing. This technique achieves high-fidelity detection without broadening the narrow clock linewidth. These clocks demonstrate exceptional performance, with NIST's Al⁺ system reaching a fractional frequency accuracy of 8.6 × 10^{-18} in the through meticulous control of systematic effects like variations. In 2025, NIST's Al⁺ single-ion clock achieved a systematic of 5.5 × 10^{-19}, a threefold improvement in over previous generations. The stability is quantified by the Allan deviation, \sigma_y(\tau) = \frac{\Delta \nu}{\nu \sqrt{N \tau}}, where \Delta \nu < 10^{-3} Hz represents the narrow linewidth of the optical transition, \nu is the transition frequency, N is the number of ions (often 1 for single-ion clocks), and \tau is the averaging time; this formula captures the shot-noise limit for Ramsey interrogation. Environmental isolation permits interrogation durations up to thousands of seconds, yielding superior short-term stability, though scalability remains constrained to a few ions due to trap depth and heating limitations.

Optical Lattice Clocks

Optical lattice clocks utilize neutral atoms, such as bosonic or fermionic ytterbium-171, which are first laser-cooled to temperatures on the order of microkelvin to reduce thermal motion. These atoms are then confined in one-dimensional or three-dimensional standing-wave optical formed by interfering laser beams, enabling the simultaneous interrogation of thousands of atoms to achieve high stability through parallel operation. The lattice operates at a "magic wavelength," typically around 813 nm for strontium, where the differential light shifts on the clock states are minimized, ensuring that the trapping potential does not perturb the clock transition frequency. In operation, the clock transition for occurs at approximately 429 THz, corresponding to a of 698 , where the atoms are probed by a narrow-linewidth to measure the hyperfine splitting between the ground and excited states. To surpass the standard (SQL) of precision, techniques such as spin-squeezed states and are employed, as demonstrated in 2024 advances by researchers, which correlate the atomic spins to reduce . The gain from entanglement can be quantified as \Delta \phi_{\text{ent}} = \Delta \phi_{\text{SQL}} / \sqrt{\xi}, where \xi < 1 is the squeezing parameter that reflects the degree of quantum correlations achieved. A notable example is the 2024 clock, which interrogated $10^4 atoms to reach a total systematic uncertainty of $8.1 \times 10^{-19}, the lowest reported for such systems at the time. Key challenges in optical lattice clocks include precise control of the lattice depth to prevent atomic tunneling between sites, which could introduce , and mitigation of shifts caused by ambient thermal photons. These shifts are corrected through rigorous temperature stabilization of the , maintaining environmental temperatures to within millikelvin precision to achieve uncertainties below $10^{-19}. Unlike trapped-ion optical clocks, which rely on single-particle precision, lattice clocks excel in averaged stability from ensemble measurements.

Emerging Atomic Clock Technologies

Chip-Scale and Miniaturized Clocks

Chip-scale atomic clocks represent a significant advancement in miniaturizing timekeeping technology, leveraging microelectromechanical systems () and coherent population trapping (CPT) to achieve compact, low-power operation suitable for portable applications. These clocks typically employ CPT in microfabricated vapor cells containing atoms, such as , interrogated by vertical-cavity surface-emitting lasers (VCSELs) that produce dual-frequency light fields, eliminating the need for traditional radiofrequency cavities used in larger vapor clocks. The CPT mechanism creates a dark-state aligned with the atomic hyperfine , enabling precise frequency locking without physical of the atoms. In operation, the dual-frequency laser light drives the atoms into a coherent superposition of hyperfine ground states, forming a non-absorbing that manifests as a narrow linewidth, often on the order of 2-3 kHz. Recent developments, such as symmetric auto-balanced Ramsey (SABR) introduced in 2025, enhance this by applying pulsed interrogation sequences that produce balanced Ramsey fringes, mitigating light-shift sensitivities and improving long-term stability to levels approaching 10^{-12} over 1 day. In January 2025, Microchip released the SA.65-LN, a low-noise with improved frequency mixing capabilities for battery-powered devices. This pulsed SABR approach uses current modulation for power control, allowing compact integration while maintaining contrast. Building on vapor principles for hyperfine interrogation, these miniaturized systems address power constraints below 100 mW, making them viable for battery-operated devices. The DARPA-funded (CSAC) program, initiated in the early , pioneered this technology with prototypes achieving short-term stability of approximately 10^{-10} at 1 second and power consumption around 100-150 mW in volumes of about 10 cm³. By 2025, microcell variants have further shrunk to volumes under 1 cm³, facilitating integration into (IoT) devices for resilient timing in distributed networks. A key challenge in these compact designs is managing frequency shifts from atomic collisions; in the buffer gas regime, inert gases like or are added to reduce wall interactions, but they induce a linear shift given by \delta \nu = \beta P, where \beta is the buffer gas (typically 10-100 Hz/Pa) and P is the gas pressure. In contrast, collisionless regimes using anti-relaxation coatings on cell walls minimize such shifts by preserving over thousands of wall bounces, though they trade off against higher sensitivities to environmental perturbations. Integration with MEMS technologies enhances vibration resistance in chip-scale clocks, crucial for dynamic environments like drones and satellites, where accelerations can exceed 10 and disrupt . MEMS-fabricated components, such as etched vapor cells and optical alignments, enable robust packaging that withstands mechanical stresses while maintaining stability, supporting applications in autonomous and low-Earth orbit timing without reliance on ground-based .

Nuclear Clocks

Nuclear clocks represent an emerging class of timekeeping devices that utilize radiative transitions between low-lying isomeric states in atomic nuclei, offering potential advantages over conventional atomic clocks based on electronic transitions. Unlike electronic orbitals, which are highly sensitive to the surrounding chemical environment, external electromagnetic fields, and perturbations from relativity such as gravitational redshift, nuclear transitions occur within the compact nuclear volume, providing intrinsic shielding from these effects. The leading candidate is the thorium-229 isotope, featuring a unique isomeric state (^{229m}Th) approximately 8 eV above the ground state, corresponding to a transition frequency of about 2 \times 10^{15} Hz in the vacuum ultraviolet (VUV) range near 149 nm. This low energy enables direct laser excitation, a feat unattainable for typical nuclear transitions in the keV to MeV regime. Significant progress toward realizing a thorium-229 was achieved in 2024 through direct resonant laser excitation of the in solid-state hosts, such as thorium-doped aluminum hexafluoride (LiSrAlF_6) crystals. Researchers at the (UCLA) and collaborators precisely measured the transition wavelength at 148.38219(4) nm, equivalent to an excitation energy of 8.355733(2) eV, resolving long-standing uncertainties in the isomeric state energy. This breakthrough demonstrated the feasibility of tabletop VUV laser systems for , marking the first observation of laser-induced excitation and decay with a measured lifetime of 568(13) s in the solid matrix. In October 2025, further studies explored the sensitivity of the Th-229 transition to variations in the , enhancing prospects for fundamental physics tests. Such advancements build on prior optical clock technologies but shift the focus to resonances for enhanced environmental robustness. The potential of nuclear clocks lies in their projected stability and accuracy, far surpassing current optical atomic clocks, with systematic uncertainties approaching 10^{-19} and short-term stabilities potentially reaching 10^{-21}, enabling tests of fundamental physics such as variations in the or effects at unprecedented levels. A key metric is the nuclear quality factor, defined as Q_{\rm nuc} = \frac{\nu_{\rm nuc}}{\Gamma_{\rm nuc}} \gg 10^{18}, where \nu_{\rm nuc} is the transition frequency and \Gamma_{\rm nuc} is the natural linewidth (FWHM), determined by the isomeric lifetime; for thorium-229, this yields Q values up to 10^{19}, orders of magnitude higher than atomic clocks due to the long nuclear coherence time. However, realizing this potential faces major challenges, including the extremely low transition probability—stemming from a radiative lifetime on the order of hours in vacuum—and the need for narrowband VUV lasers to match the millihertz linewidth. Current prototypes in solid-state systems achieve fractional frequency reproducibilities around 10^{-13}, limited by host material interactions and detection efficiency, with ongoing efforts to suppress non-radiative decay and improve laser coherence to reach 10^{-15} stability.

Performance Characteristics

Accuracy and Stability Metrics

The performance of atomic clocks is evaluated through key metrics that quantify their , accuracy, and , which are essential for applications requiring precise timekeeping. Fractional frequency , denoted as σ_y(τ), measures the short- and long-term fluctuations in the clock's output frequency relative to its nominal value, where y(t) = Δf/f is the fractional and τ is the averaging time. This metric is commonly assessed using the Allan deviation, a statistical tool designed to characterize various noise processes in oscillators, such as , , and . The Allan deviation is defined as \sigma_y(\tau) = \sqrt{\frac{1}{2} \left\langle (y_{k+1} - y_k)^2 \right\rangle}, where the angle brackets denote the ensemble average over adjacent frequency averages separated by τ. In atomic clocks, this deviation typically follows a τ^{-1/2} dependence at short averaging times due to quantum projection noise, transitioning to other behaviors at longer times influenced by environmental factors. Accuracy refers to the systematic between the clock's realized and the true , arising from uncorrected environmental and perturbations. It is expressed as a fractional , often at levels below 10^{-18} in state-of-the-art systems, and is determined through detailed evaluations of budgets. assesses the consistency of measurements across realizations of the same clock type or between different laboratories, reflecting the clock's ability to maintain a without drift from manufacturing or operational variations; for instance, comparisons of NIST optical clocks have demonstrated below 10^{-18}. Quantum projection noise sets a fundamental floor for in clocks, originating from the statistical in measuring atomic state populations during interrogation; this limits σ_y(1 s) to approximately the standard quantum limit of 1/√N, where N is the number of atoms, manifesting as white noise in the servo loop. Major sources of systematic errors include relativistic effects, , and interactions. The relativistic Doppler shift, primarily the second-order term due to from atomic motion, contributes a frequency correction proportional to -v^2/(2^2), where v is the atomic velocity and c is the ; in trapped-ion or clocks, this is minimized by cooling atoms to microkelvin s, yielding uncertainties below 10^{-19}. The (BBR) shift arises from fluctuating of thermal photons perturbing atomic energy levels, with the dominant static contribution scaling as Δν_BBR = -C_4 T^4 / h, where C_4 is a material-specific coefficient, T is , and h is Planck's constant; dynamic corrections add a smaller term, and precise thermometry reduces this uncertainty to 10^{-18} or better in controlled environments. The second-order Zeeman shift, induced by ambient , follows Δν_Z = K B^2, where K is the atomic coefficient and B is the field strength; operating at fields around 0.1 μT with active shielding limits this to 10^{-18}. An additional noise source affecting stability is the Dick effect, which aliases high-frequency noise into the servo due to dead time between interrogations in pulsed schemes. This introduces excess , degrading short-term stability as σ_y(τ) ∝ √(T_c / τ), where T_c is the clock cycle time; mitigation via continuous interrogation schemes, such as interleaved ensembles or zero-dead-time protocols, suppresses this to below the quantum noise floor. As of 2025, leading optical clocks, such as NIST's aluminum-ion , achieve σ_y(1 s) ≈ 3.5 × 10^{-16} with total accuracy of 5.5 × 10^{-19}, approaching the limit set by quantum noise while bounding systematic errors through advanced control techniques.

Comparison Across Clock Types

Atomic clocks vary significantly in performance across microwave-based types (such as cesium fountains, vapor, and hydrogen masers), optical types (such as trapped-ion and clocks), and (such as chip-scale and clocks), with trade-offs in , , consumption, and operational maturity. Microwave clocks generally offer reliable long-term performance suitable for continuous operation in time scales, while optical clocks provide superior short-term due to higher frequencies, enabling accuracies beyond 10^{-18}. clocks prioritize but often sacrifice some for portability. The following table summarizes key benchmarks for representative examples, focusing on fractional frequency stability (Allan deviation at 1 day where available), approximate , and power consumption. Stability values reflect achieved long-term performance under controlled conditions; and power are typical for operational systems.
Clock TypeStability (σ_y at 1 day)Size (approximate volume)Power Consumption
Cesium Fountain~10^{-16}~1 m³~500 W
Vapor (chip-scale)~10^{-12}~17 cm³~100 mW
~10^{-15}~1 m³~60 W
Strontium Lattice (optical)~2.5 × 10^{-19}~0.1 m³ (portable)~85 W
Trapped-Ion (optical)~10^{-18}~0.5 m³~100 W
(experimental)Not yet achieved (projected <10^{-19})Lab-scale (~1 m³)Not specified
Microwave clocks like cesium fountains and hydrogen masers excel in long-term reliability, maintaining over days to months with minimal drift, making them ideal for ensemble averaging in international standards; however, their lower transition frequencies limit short-term compared to optical clocks. Optical clocks, conversely, achieve exceptional short-term through higher-frequency interrogations but require more complex systems, leading to higher power needs and sensitivity to environmental perturbations, though recent portable versions mitigate this. clocks hold potential for ultimate insensitivity to external fields due to nuclear transitions but remain immature, with demonstrations limited to excitation measurements rather than full clock operation as of 2025. As of 2025, the () is defined by an ensemble of approximately 450 atomic clocks worldwide, with around 10-15 primary cesium fountain standards providing the key calibrations for accuracy at the 10^{-16} level; optical clocks already outperform these by a factor of about 100 in systematic uncertainty, demonstrating fractional accuracies below 10^{-18}. A key insight into performance scaling arises from the shot-noise limit, where clock stability improves with the of the number of interrogated atoms and averaging time: \sigma_y(\tau) \propto \frac{1}{\sqrt{N \cdot \rho \cdot V \cdot \tau}} Here, N is the number of atoms, \rho the atomic density, V the interrogation volume, and \tau the averaging time; optical lattice clocks excel by trapping thousands of atoms in high-density arrays, achieving near-quantum-limited stability. In a 2024 comparison, JILA's strontium lattice clock and NIST's aluminum-ion clock were linked via a 3.6 km stabilized fiber , agreeing to within 10^{-18} in fractional frequency, validating their interoperability for future networks and highlighting optical clocks' potential to surpass microwave ensembles.

Timekeeping and Synchronization

International Time Standards

The International Bureau of Weights and Measures (BIPM) is responsible for computing (), a continuous derived from a weighted average of readings from approximately 450 atomic clocks located in laboratories worldwide. The International Earth Rotation and Reference Systems Service (IERS) manages the monitoring of 's orientation parameters, including the difference between (UT1) and (UTC), to ensure UTC remains aligned with within 0.9 seconds. BIPM publishes monthly Circular T reports, which include ratings of contributing clocks based on their and accuracy, along with five-day interval differences between UTC and time scales UTC(k). These data enable the computation of TAI as a weighted average, where weights reflect clock performance; UTC is then formed from TAI by subtracting a fixed offset of 10 seconds (established in 1972) and adjusting for leap seconds to synchronize with . Since the introduction of leap seconds in 1972, 27 such adjustments have been added to UTC, with the last occurring on December 31, 2016, resulting in TAI being 37 seconds ahead of UTC as of 2025. Ongoing debates, influenced by the enhanced precision of optical atomic clocks, center on potentially suppressing future leap seconds to avoid disruptions in global systems, with bodies like the (ITU) considering elimination by 2035. Primary frequency standards, such as cesium fountain clocks, realize the SI second through the fixed value of the cesium-133 hyperfine transition frequency of exactly 9,192,631,770 Hz. Secondary standards, including gas cell clocks and masers, provide practical timekeeping but require periodic against primary cesium standards to maintain to the SI second. From 2018 to 2025, BIPM has conducted pilot comparisons of optical clocks, integrating their calibrations into TAI evaluations to assess their potential for future inclusion in international time scales, leveraging techniques like links for remote frequency comparisons. By 2023, optical clocks such as the UK's NPL-Sr1 clock began contributing as secondary frequency standards to TAI calibrations.

Satellite and Network Synchronization

Atomic clocks are disseminated globally through satellite-based systems, enabling precise time for navigation, scientific, and telecommunication applications. The (GPS) utilizes onboard and cesium atomic clocks to broadcast time signals, providing a reference for receivers worldwide by transmitting carrier and code signals derived from these clocks. These broadcasts allow for one-way time transfer, where ground stations compute their relative to GPS time using pseudorange measurements from multiple s. A key technique in this context is the GPS common-view method, which compares remote clocks by having both stations simultaneously observe the same signal, calculating the time offset as \Delta t = (t_{rx1} - t_{rx2}) - (t_{tx1} - t_{tx2}), with corrections applied for orbital parameters to achieve sub-nanosecond . However, clocks require relativistic corrections to account for gravitational and velocity effects; these cause the clocks to run faster by approximately 38 microseconds per day, a net effect pre-compensated by offsetting the nominal clock frequency at the factory, ensuring accuracy despite the satellites' orbital dynamics. For higher precision comparisons between international time laboratories, two-way satellite time and frequency transfer (TWSTFT) employs geostationary satellites to exchange signals bidirectionally, canceling common propagation delays and achieving frequency transfer uncertainties at the $10^{-15} level over intercontinental distances. This method supports the computation of () by linking primary frequency standards with minimal asymmetry in the transfer path. The European Galileo system enhances this capability using passive masers (PHMs) as its primary clocks, offering superior short-term stability compared to standards, which broadcast signals for global in the Galileo navigation constellation. Complementing satellite methods, fiber-optic networks enable ultra-precise time transfer over terrestrial distances using coherent optical techniques. These systems employ dark fibers—unused telecommunication lines—with Doppler cancellation to mitigate from fiber length fluctuations, achieving fractional frequency stabilities of $10^{-19} over links exceeding 1000 km, as demonstrated in European networks during the . Such transfers support comparisons of optical clocks across continents, with bidirectional round-trip configurations ensuring round-trip instabilities below $10^{-18} at integration times of 1000 seconds. Emerging integrations of atomic clock synchronization with quantum technologies further advance secure network timing. By 2025, quantum key distribution (QKD) links incorporate protocols to align distant nodes, enabling time-accurate distribution of cryptographic keys over fiber and hybrids for resilient, tamper-proof networks.

Applications

Atomic clocks are essential to Global Navigation Satellite Systems (GNSS), such as the (GPS), where they provide the precise timing needed for determining user positions through time-of-flight measurements. In GPS, each satellite carries onboard atomic clocks—typically cesium or types—that achieve short-term stability on the order of $10^{-12} to enable pseudorange accuracy better than 10 meters. These clocks experience drift over time due to environmental factors in space, which is continuously monitored and corrected by ground control stations to maintain overall system accuracy. The core principle underlying GNSS positioning is time-of-flight , where a user's position is calculated from the propagation delay \Delta t of signals from multiple satellites, converted to distance via multiplication by the c, with clock errors often dominating the overall user accuracy. Advancements in GNSS technology continue to integrate more stable atomic clocks for enhanced precision. For instance, China's BeiDou-3 system incorporates atomic clocks on its geostationary satellites. In the U.S. GPS, the M-code signal utilizes chip-scale atomic clocks as resilient backups, ensuring timing continuity during GPS denial scenarios and supporting secure, anti-jam operations for positioning. These compact clocks, with low size, weight, and power consumption, maintain for receivers even when satellite signals are unavailable. Experimental deployments further demonstrate the potential of advanced atomic clocks in positioning. In 2025, the European Space Agency's Atomic Clock Ensemble in Space (ACES) mission tested a space-qualified cold-atom cesium clock on the , achieving unprecedented stability to support high-precision timing for navigation applications, with operations ongoing as of November 2025. Additionally, transportable optical lattice clocks have been developed for mobile use, such as in base stations for , offering fractional uncertainty below $10^{-17} over short averaging times to enable centimeter-level accuracy in dynamic environments. A key application enabled by atomic clocks is (DGPS), which uses precise carrier-phase timing from ground reference stations synchronized via atomic standards to achieve positioning resolutions of 1 cm. This technique corrects common errors in pseudorange measurements, leveraging the high stability of atomic clocks to resolve carrier phase ambiguities and support applications like and .

Scientific and Relativistic Experiments

Atomic clocks serve as precision tools for testing fundamental aspects of (GR) and by measuring minute variations in time due to gravitational fields and relative motion. A key prediction of GR is the , where the fractional frequency shift between two clocks separated by height h is given by \frac{\Delta f}{f} = \frac{gh}{c^2}, with g the local and c the . This effect was confirmed at the $10^{-16} level in a 2010 experiment at NIST using two aluminum-ion optical clocks, one positioned 33 cm higher than the other, demonstrating Einstein's theory on a scale. Advancements in optical lattice clocks have pushed these tests to even smaller scales. In 2022, researchers used clocks to detect within a millimeter-scale ensemble, resolving gradients equivalent to a 1 cm height difference with a precision of $10^{-17}, bridging quantum and relativistic regimes. Building on this, a July 2025 theoretical proposal outlines using quantum networks of entangled clocks to probe Lorentz invariance and quantum- effects by comparing clock rates in superimposed states, potentially achieving sensitivities below $10^{-18}. Continental-scale networks of optical clocks, linked by stabilized fiber optics, enable large-baseline tests of , expressed as \frac{\Delta \nu}{\nu} = \frac{\Delta \Phi}{c^2}, where \Delta \Phi is the in between sites. Comparisons over distances exceeding 1000 km in have verified this effect to parts in $10^{18}, providing robust checks of GR in varying potentials. These networks also constrain models of by monitoring potential time variations in the \alpha, with current limits on its fractional change rate at $10^{-18} per year, far tighter than astronomical observations. Emerging integrations of atomic clocks with platforms further expand these investigations. In 2024, researchers demonstrated hybrid systems using quantum control techniques to enhance the precision of multi-ensemble optical clocks, enabling individual control of atoms for improved timekeeping performance.

Commercial and Industrial Uses

Atomic clocks play a pivotal role in infrastructure, particularly in ensuring precise for networks. Synchronous Ethernet (SyncE) standards demand on the order of 10^{-11} to support alignment across base stations, enabling low-latency fronthaul and backhaul essential for massive and technologies. Rubidium-based atomic clocks are commonly deployed in cell towers and base stations due to their compact size, low power consumption, and ability to maintain this over extended periods, outperforming traditional oscillators by several orders of magnitude. In the financial sector, atomic clocks underpin (HFT) by providing the sub-microsecond synchronization required to timestamp transactions accurately and prevent regulatory violations or arbitrage discrepancies. HFT systems often achieve synchronization within 100 microseconds of (UTC), as mandated by frameworks like MiFID II in , using (NTP) servers backed by derived from atomic clocks to ensure traceability and fairness in market operations. This precision mitigates risks from even minor timing errors, which could otherwise lead to invalid trades or compliance issues. Power grid management relies on atomic clocks for synchronizing phasor measurement units (PMUs), which monitor voltage and current waveforms in to maintain grid stability. Chip-scale atomic clocks serve as reliable backups during GPS outages, delivering holdover performance with accuracy better than 1 μs over wide-area networks, thus enabling rapid detection of faults and dynamic load balancing. Consumer applications benefit from atomic clock-derived time signals broadcast via radio stations like NIST's in the United States and Germany's , which transmit cesium-based UTC time codes to synchronize radio-controlled watches and clocks. These low-frequency signals (60 kHz for and 77.5 kHz for ) allow devices to automatically adjust for accurate timekeeping, including daylight saving changes, reaching millions of everyday users without needing direct atomic clock integration. The integration of atomic clocks extends to emerging commercial domains, such as data centers where they support timestamping for secure transaction validation, a practice increasingly adopted to enhance in distributed ledgers. Driven by demand in ecosystems for precise timing in connected devices, the global atomic clock market is projected to reach approximately $1.2 billion by 2035, reflecting growth from , , and sectors. Miniaturized chip-scale variants further enable portable deployment in these industrial settings.

Future Directions

Redefining the Second

The ongoing redefinition of the SI second is being coordinated by the Consultative for Time and (CCTF) Task Force, established in 2020 under the International for Weights and Measures (CIPM), with a target implementation by the 29th General Conference on Weights and Measures (CGPM) in 2030. This effort focuses on transitioning from the hyperfine transition in cesium-133 atoms to optical or potentially transitions, promising accuracies beyond 10^{-18} while maintaining compatibility with existing time scales. Leading candidates include optical lattice clocks based on neutral atoms of strontium-87, operating at a of approximately 429 THz, and ytterbium-171, at around 518 THz, both of which have demonstrated uncertainties below 10^{-18}. These proposals were outlined in the CCTF's updated roadmap, emphasizing three options: a single optical transition, a weighted of multiple transitions, or a tie to a fundamental constant. To qualify for redefinition, candidate optical frequencies must demonstrate agreement with the current cesium standard to within 10^{-16} over averaging periods of up to 10^8 seconds (about three years), ensuring seamless integration into (). International comparisons among national institutes are essential, achieved through methods such as two-way satellite time and frequency transfer (TWSTFT) for stabilities around 10^{-16} per day and emerging links capable of sub-10^{-18} precision over continental distances. Secondary representations of the second, already recommended for and clocks, fix the optical transition frequency directly while measuring its ratio to the cesium hyperfine frequency, thereby decoupling the new definition from ongoing reliance on cesium primary standards. The 2025 roadmap update incorporates a nuclear clock pathway as a long-term option, spurred by the 2024 experimental breakthrough in directly measuring the low-energy nuclear isomer transition in thorium-229 at approximately 8.36 eV, which could enable even greater insensitivity to environmental perturbations. Additionally, linking the second to the fixed Rydberg constant R_\infty = 10\,973\,731.568160 \, \mathrm{m^{-1}}, established in the 2019 SI revision, provides a computable alternative through hydrogen-like transitions, potentially enhancing universality. However, challenges persist in verifying universality—ensuring consistent realizations across global laboratories without systematic offsets—and achieving routine operational reliability for optical standards. The 28th CGPM in 2026 is expected to validate the roadmap and select the redefinition approach, contingent on optical clocks reaching full maturity with widespread contributions to TAI.

Challenges and Technological Hurdles

Atomic clocks, despite their exceptional , remain highly susceptible to environmental perturbations that can introduce frequency shifts and degrade performance. Temperature variations, for instance, affect the atomic transition frequencies through of clock components and changes in fields, leading to measurable errors even in controlled settings. Electromagnetic fields also pose significant challenges, as they can induce Zeeman shifts in atomic energy levels, particularly in clocks using alkali vapors or ions sensitive to magnetic gradients. These sensitivities necessitate advanced shielding and stabilization techniques, such as insulated enclosures and active feedback systems, to maintain stability at levels below 10^{-15}. Deploying atomic clocks in space applications introduces additional hurdles related to scalability and environmental robustness, particularly . Cosmic radiation, including high-energy protons and heavy ions, can cause single-event upsets in electronic components and degrade diodes or vapor cells through and displacement damage, compromising long-term frequency stability. Efforts to mitigate these effects involve radiation-tolerant materials and redundant architectures, as seen in programs like the Radiation Hardened Electronics for Space Environments (RHESE), but achieving the compactness required for satellites while preserving laboratory-grade performance remains a key engineering challenge. In optical atomic clocks, achieving the necessary stability is a critical technological barrier, with requirements typically demanding relative instabilities below 10^{-15} over times of seconds to hours. Cavity-stabilized lasers must minimize noise and mechanical vibrations to probe narrow atomic transitions, such as those in or ions, where even minor linewidth broadening can limit overall clock accuracy. By 2025, limits have been partially overcome through techniques like spin squeezing, enabling entanglement-enhanced clocks to reach instabilities of 10^{-18} with up to 2 dB of squeezing gain. However, decoherence from environmental interactions and imperfect state preparation persists, reducing the effective time and hindering sustained operation at these extremes. Miniaturization for portable applications, such as chip-scale atomic clocks (CSACs), involves inherent trade-offs in stability due to reduced component sizes and simplified . Laboratory-scale vapor-cell or clocks achieve long-term stabilities on the order of 10^{-15} or better, whereas CSACs typically exhibit fractional instabilities around 10^{-10} to 10^{-11}, representing a loss of several orders of magnitude primarily from increased sensitivity to and shorter times. Emerging nuclear clocks face even steeper challenges in excitation efficiency, where achieving reliable to the metastable state, such as the 229Th at ~8.36 eV, is limited to below 10^{-6} due to inefficiencies in vacuum-ultraviolet light generation and coupling. Relativistic and quantum gravity effects further complicate high-precision clock comparisons, manifesting as frequency discrepancies on the order of 10^{-18} that demand refined theoretical models for interpretation. predicts gravitational redshift in clock rates based on potential differences, as verified in NIST comparisons where residual effects must be controlled to avoid systematic biases at this level. These discrepancies, observed in and clock networks, highlight the need for integrated corrections accounting for both special relativistic from motion and general relativistic curvature, pushing the boundaries of tests. Advancements in hybrid quantum systems offer promising avenues for error correction in atomic clocks, with 2024 demonstrations merging optical clock qubits with spin or Rydberg arrays to enable real-time . For example, tweezer-based hybrid platforms combine long-coherence qubits for with optical transitions for readout, potentially suppressing decoherence through encoded logical qubits. Nevertheless, the complexity remains high, involving precise alignment of disparate modalities, scalable entanglement generation, and robust control , which currently limit beyond small ensembles.

References

  1. [1]
    How Do Atomic Clocks Work? | NIST
    Aug 22, 2024 · A clock is, at its core, a marriage of two components: a mechanism that oscillates, or ticks with a steady beat, and a device that counts those beats and ...What, Exactly, Is A Clock? · The Inner Life Of The Atom · The Atomic Clock Age
  2. [2]
    A Brief History of Atomic Time | NIST
    Aug 20, 2024 · In 2006, NIST researchers used the frequency comb to build the first optical atomic clock to outperform the best cesium clock in accuracy. The ...The Atomic Clock Emerges · A Special Atom · Going Optical
  3. [3]
    Knowing Where We Are | NIST
    Jun 30, 2025 · Atomic clocks are essential to modern GPS. Without atomic clocks, errors would build up so fast that the system would quickly become useless. ...
  4. [4]
    How Atomic Clocks Have Changed Our World | NIST
    Nov 4, 2024 · The clocks that NIST and other metrology labs use to produce official time are essentially fancy devices built to measure the microwave ...Read More: How Do We Know... · Read More: A Brief History... · An Opportunity For Awe
  5. [5]
    [PDF] Optical atomic clocks - Physikalisch-Technische Bundesanstalt
    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 ...
  6. [6]
    Prospects of a thousand-ion Sn2+ Coulomb-crystal clock with sub ...
    Jul 5, 2024 · Optical lattice clocks use simultaneous measurements of thousands of atoms to quickly average down quantum projection noise (QPN), a ...Missing: loop | Show results with:loop<|control11|><|separator|>
  7. [7]
    [PDF] Hyperfine Structure in Atoms
    When this is coupled to the nuclear spin i, where I2 = i(i + 1)¯h2, the resulting total angular momentum is f = i ± 1/2. Thus, there are two hyperfine levels, ...
  8. [8]
    12 The Hyperfine Splitting in Hydrogen - Feynman Lectures
    The hyperfine splitting is due to the interaction of the magnetic moments of the electron and proton, which gives a slightly different magnetic energy for each ...Missing: ΔE = AIJ
  9. [9]
    Nuclear clocks for testing fundamental physics - IOPscience
    The conceptual shift from an atomic to a nuclear clock is straightforward: instead of a transition in the electron shell, a radiative transition inside the ...
  10. [10]
    [PDF] arXiv:physics/0511168v1 [physics.atom-ph] 18 Nov 2005
    Nov 18, 2005 · The damping rate Γ = 1/τ describes the relaxation through spontaneous atomic decay and the additional coherence relaxation rate ΓL is used to ...
  11. [11]
    Subpicosecond rotations of atomic clock states | Phys. Rev. A
    May 18, 2018 · Atomic hyperfine states in the ground state can have a long coherence time measured up to tens of seconds [7] ; because of this, they are used ...
  12. [12]
    Isidor Isaac Rabi – Facts - NobelPrize.org
    In 1938, Isaac Isidor Rabi passed a beam of molecules through a magnetic field. When the beam was exposed to radio waves, the direction of rotation could be ...
  13. [13]
    The Molecular Beam Resonance Method for Measuring Nuclear ...
    A new method of measuring nuclear or other magnetic moment is described. The method, which consists essentially in the measurement of a Larmor frequency in ...
  14. [14]
    A Brief History of Atomic Clocks at NIST - Time and Frequency Division
    1949 -- Using Rabi's technique, NIST (then the National Bureau of Standards) announces the world's first atomic clock using the ammonia molecule as the source ...Missing: maser | Show results with:maser<|separator|>
  15. [15]
    Louis Essen - About us - NPL
    By the time this had been published Essen had improved the clock to an accuracy of one part in a million, million, launching the world into an era of practical ...
  16. [16]
    Resolution 1 of the 13th CGPM (1967) - BIPM
    "The second is the duration of 9 192 631 770 periods of the radiation corresponding to the transition between the two hyperfine levels of the ground state of ...Missing: cesium- | Show results with:cesium-
  17. [17]
    A Historical Review of U.S. Contributions to the Atomic Redefinition ...
    This paper was written to commemorate the 50th anniversary of the atomic redefinition of the second in the International System (SI) in 1967.
  18. [18]
    Leap seconds - PTB.de - Physikalisch-Technische Bundesanstalt
    In the following years, however, the actual duration of the ephemeris second could be stated only with an accuracy of 10-8 s so that the advantage over the UT ...Missing: reproducibility | Show results with:reproducibility
  19. [19]
    BIPM technical services: Time Metrology
    The BIPM Time Department is responsible for the realization and dissemination of the international time scales UTC, UTCr and TT, used for different ...
  20. [20]
    Leap second and UT1-UTC information | NIST
    The first leap second was inserted into the UTC time scale on June 30, 1972. Leap seconds are used to keep the difference between UT1 and UTC to within ±0.9 s.
  21. [21]
    [PDF] Time and Frequency Measurement
    Clocks could be compared more precisely using portable atomic clocks, but since this involved flying a fully operational clock from one laboratory to ...
  22. [22]
    A New Era for Atomic Clocks | NIST
    Feb 4, 2014 · The optical lattice clocks at NIST and JILA are much more stable than NIST-F1. NIST-F1 must be averaged for about 400,000 seconds (about five ...
  23. [23]
    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.
  24. [24]
    [PDF] The Chip-Scale Atomic Clock - Low-Power Physics Package - DTIC
    Dec 7, 2004 · provide a stability of 1×10-11 over 1 hour while consuming only 30 mW. While the stability goals are. * E-mail: RLutwak@Symmetricom.com. † 34 ...
  25. [25]
    [PDF] An optical lattice clock based on strontium atoms achieves ... - JILA
    Jul 24, 2024 · Their clock, presented in a Physical Review. Letters paper, exhibits a total systematic uncertainty of 8.1 x 10-19, which is the lowest ...Missing: 10000 | Show results with:10000
  26. [26]
    MIT physicists improve the precision of atomic clocks
    Oct 8, 2025 · An atomic clock keeps time by relying on the “ticks” of atoms as they naturally oscillate at rock-steady frequencies. Today's atomic clocks ...
  27. [27]
    Microcell atomic clock using laser current-actuated power ... - Nature
    Oct 15, 2025 · We present a coherent-population trapping (CPT) microcell atomic clock using symmetric auto-balanced Ramsey (SABR) spectroscopy.
  28. [28]
    Major Leap for Nuclear Clock Paves Way for Ultraprecise Timekeeping
    Sep 4, 2024 · These clocks could lead to improved timekeeping and navigation, faster internet speeds, and advances in fundamental physics research.Missing: CERN 10^- 19
  29. [29]
    [PDF] Nuclear Clocks for Tests of Fundamental Physics - CERN Indico
    Jan 22, 2025 · The nuclear clock promises: • High accuracy (with laser cooled trapped ions). • High stability (in the solid state as a laser Mössbauer system).
  30. [30]
    [PDF] Atomic Frequency Standards: A Survey
    AI1 existing cesium beam tubes feature the Ramsey [25] cavity because of its advantages in reducing first-order Doppler effects, magnetic field problems, and ...Missing: hardware | Show results with:hardware<|control11|><|separator|>
  31. [31]
    [PDF] Caesium Atomic Clocks: Function, Performance and Applications
    wire detector. The atoms are ionized, and the ion current is processed to yield the control signal ID. When fp is tuned across fr, ID exhibits a resonance ...
  32. [32]
    NIST's Cesium Fountain Atomic Clocks
    NIST-F3 and NIST-F4 are referred to as fountain clocks because they use a fountain-like movement of atoms to calibrate the offset of a microwave frequency from ...
  33. [33]
    [PDF] Primary Atomic Frequency Standards at NIST
    NIST develops atomic frequency standards, which are key for timekeeping. The best relative standard uncertainty achieved is 1.5 x 10^-15.
  34. [34]
    [PDF] NIST-F3, a Cesium Fountain Frequency Reference
    The atomic state is detected at the end of the fountain sequence by collecting fluorescence from the atoms as they cross several resonant laser beams. In a ...Missing: method | Show results with:method
  35. [35]
    [PDF] The history of time and frequency from antiquity to the present day
    The nuclear spin in cesium-133 is 7/2, and the hyper-fine splitting between the parallel state with angular momentum F = 7/2+1/2 = 4 and anti-parallel state ...
  36. [36]
    None
    Below is a merged summary of the "Rubidium Frequency Standard Primer (Passive Rb Gas Cell Clocks)" based on the provided segments. To retain all information in a dense and organized manner, I’ll use a combination of narrative text and a table for key details. The response consolidates overlapping information while preserving unique facts from each segment.
  37. [37]
    [PDF] Long-Term Stability of the NIST Chip-Scale Atomic Clock
    The long-term stability of NIST chip-scale atomic clock physics packages is limited by factors like vapor cell and laser temperature, magnetic field, and local ...
  38. [38]
    [PDF] Frequency Standards Based on Atomic - Hydrogen
    Hydrogen masers already exhibit the best short-term stabrlity of any room-temperature atomic clock and this could be improved by a factor of lcx) with the ...
  39. [39]
    [PDF] HYDROGEN MASER FREQUENCY STANDARD R. F. C. Vessot ...
    The field strength and length of the magnet are designed so as to focus the F = 1, mF = 0, and mF = 1 state atoms into the storage bulb through its collimator.
  40. [40]
    [PDF] Hydrogen Maser Clock (HMC) Experiment
    level of stability of any frequency standard currently available. Its frequency stability is typically on the order of several parts in 1016, corresponding.
  41. [41]
    Review of the development of the hydrogen maser technique and a ...
    Sep 13, 2022 · A typical stability performance of 10−15 to 10−16 per day has been reached for the traditional design and a better than 1 × 10−16 has been ...
  42. [42]
    Trapped Ion Optical Clocks | NIST
    Apr 7, 2023 · The trap is the gold structure with the cross-shaped cutout. The inset shows the aluminum ion (blue), the source of the clock's “ticks,” and the ...
  43. [43]
    [PDF] A Linear Paul Trap for Ytterbium Ions - Universität Hamburg
    Apr 30, 2014 · Basic advantages of this ion trap design are high trap frequencies and a high degree of optical access. Mounting parts are not displayed in ...
  44. [44]
    Quantum logic spectroscopy
    Quantum logic spectroscopy is a tool that makes it possible to read out the electronic state of an ion which does not possess a suitable transition for electron ...
  45. [45]
    [PDF] Optical clocks with trapped ions
    Feb 7, 2023 · the trapped ions. Our ion trap is based on the macroscopic linear Paul trap designs successfully implemented in different experiments in our ...
  46. [46]
    Frequency Comparison of Two High-Accuracy A l + Optical Clocks
    In summary, we have built an Al + ion clock with a fractional frequency inaccuracy of 8.6 × 10 - 18 . Its frequency is compared to that of a previously ...
  47. [47]
    An 27Al+ quantum-logic clock with systematic uncertainty below 10-18
    Jul 15, 2019 · Operating the clock with a lower trap drive frequency has reduced excess micromotion, compared to previous 27Al+ clocks, leading to a reduced ...Missing: 1.12 PHz
  48. [48]
    Rapid exchange cooling with trapped ions | Nature Communications
    Feb 5, 2024 · Sympathetic cooling has been employed to address this problem. Here, each ion used in the computation is co-trapped with another ion of a second ...
  49. [49]
    [PDF] Optical atomic clocks - Time and Frequency Division
    This paper reviews the history and the state of the art in optical-clock research and addresses the implementation of optical clocks in a possible future ...
  50. [50]
    Optical Lattice Clocks - Optics & Photonics News
    Most existing lattice clocks use two stages of laser cooling and trapping: one to accumulate large numbers of atoms rapidly, and the other to cool them below ...
  51. [51]
    World's Most Accurate and Precise Atomic Clock Pushes New ...
    Jul 1, 2024 · The new JILA clock uses a web of light known as an “optical lattice” to trap and measure tens of thousands of individual atoms simultaneously.
  52. [52]
    [PDF] The Strontium Optical Lattice Clock - JILA
    ... lattice laser operating at the magic wavelength will yield AC Stark shifts of the clock transition <1 mHz (1 × 10−18 of the clock transition frequency). Of ...
  53. [53]
    [PDF] A strontium lattice clock with 3 × 10 - and its frequency - arXiv
    Aug 15, 2014 · The absolute frequency of the 5s2 1S0 – 5s5p 3P0 transition in 87Sr is 429 228 004 229 873.13(17) Hz. Our result is in excellent agreement with ...
  54. [54]
    Atomic Clocks Surpass Fundamental Precision Limits Through ...
    Mar 19, 2024 · JILA's breakthrough in optical atomic clocks uses quantum entanglement to surpass fundamental precision limits, setting a new standard in timekeeping.
  55. [55]
    Pushing Atomic Clocks Beyond the Standard Quantum Limit - JILA
    Oct 23, 2025 · They found that the spin-squeezed clock not only beat the SQL but also showed a 3.3 dB improvement over the unentangled clock. This confirms ...Missing: MIT 2024
  56. [56]
    [2403.10664] A clock with $8\times10^{-19}$ systematic uncertainty
    Mar 15, 2024 · We report an optical lattice clock with a total systematic uncertainty of 8.1 \times 10^{-19} in fractional frequency units, representing the lowest ...Missing: JILA 10000
  57. [57]
    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: 10000 | Show results with:10000
  58. [58]
    Researchers simplify design of optical atomic clocks without ...
    Sep 16, 2024 · Researchers have demonstrated a new optical atomic clock that uses a single laser and doesn't require cryogenic temperatures.
  59. [59]
    Microfabricated Atomic Clocks at NIST
    Dec 7, 2004 · ... coherent population trapping (CPT). This technique allows for a simple and compact device containing a VCSEL, optics to shape the laser beam ...
  60. [60]
    Microfabricated cells for chip-scale atomic clock based on coherent ...
    The coherent population trapping signals with a typical linewidth of 2–3 kHz and a signal-to-noise ratio of 1500 in the 1-Hz bandwidth were observed, which ...
  61. [61]
    [PDF] arXiv:2503.01681v1 [physics.atom-ph] 3 Mar 2025
    Mar 3, 2025 · Microcell CPT atomic clock using laser current-actuated power modulation with 10-12 range stability at 1 day. Carlos Manuel Rivera-Aguilar1 ...
  62. [62]
    Success Story: Chip-Scale Atomic Clock | NIST
    Dec 2, 2020 · In 2001, DARPA funded a program to develop a commercial CSAC in support of military needs for secure wireless communication and jam-resistant ...
  63. [63]
    [PDF] THE CHIP-SCALE ATOMIC CLOCK – PROTOTYPE EVALUATION
    Over all 10 units the average short-term stability is σy(τ=1)=1.6x10-10, nearly 4X margin over the project goal of σy(τ=1)<6x10-10. The relative performance of ...
  64. [64]
    [PDF] arXiv:2305.00761v1 [quant-ph] 1 May 2023
    May 1, 2023 · The buffer gas induces a temperature-dependent shift of the CPT resonance frequency, which can be suppressed by using a mixture of two gases ...
  65. [65]
    [PDF] Atomic physics with vapor-cell clocks
    curves show, the frequency shift is very nearly linear with the buffer-gas pressure. The slope depends on the particular buffer gas, and also on the temperature ...<|separator|>
  66. [66]
    [PDF] 3.18 MEMS Atomic Clocks - Time and Frequency Division
    Jan 2, 2025 · They provide unmatched frequency stability over long periods of time (from a few seconds to many years), because their resonance frequency is ...
  67. [67]
    Alkali Vapor MEMS Cells Technology toward High-Vacuum Self ...
    One example is the development of the miniature, so-called chip-scale atomic clocks (CSAC), applying the Coherent Population Trapping effect (CPT) [1]. The CPT ...
  68. [68]
    [PDF] Chip Scale Atomic Clocks (CSACs) for Breakthrough LEO Space ...
    GNSS satellites commonly use cesium and rubidium atomic clocks to provide the ubiquitous commercial time and navigation signals from which we all benefit.
  69. [69]
    The thorium-229 low-energy isomer and the nuclear clock - Nature
    Feb 25, 2021 · The 229Th nucleus is the prime candidate for the realization of a nuclear clock because it possesses a low-energy (8 eV) excited state that is ...
  70. [70]
    Laser excitation of the $^{229}$Th nuclear isomeric transition in a ...
    Apr 18, 2024 · Abstract page for arXiv paper 2404.12311: Laser excitation of the $^{229}$Th nuclear isomeric transition in a solid-state host.
  71. [71]
    The $$^{229}$$ Th isomer: prospects for a nuclear optical clock
    Nov 2, 2020 · The proposal for the development of a nuclear optical clock has triggered a multitude of experimental and theoretical studies.<|control11|><|separator|>
  72. [72]
    Frequency reproducibility of solid-state Th-229 nuclear clocks
    ### Summary of Experimental Frequency Stability/Reproducibility of Th-229 Nuclear Clock Prototypes
  73. [73]
    [PDF] A Historical Perspective on the Development of the Allan Variances ...
    Apr 1, 2016 · Abstract—Over the past 50 years, variances have been devel- oped for characterizing the instabilities of precision clocks and oscillators.
  74. [74]
    NIST Atomic Clocks Now Keep Time Well Enough to Improve ...
    Nov 28, 2018 · Reproducibility: How closely the two clocks tick at the same frequency, shown by 10 comparisons of the clock pair, yielding a frequency ...
  75. [75]
    Systematic evaluation of an atomic clock at 2 × 10−18 total ... - NIH
    Our 87 Sr optical lattice clock now achieves fractional stability of 2.2 × 10 −16 at 1 s. With this improved stability, we perform a new accuracy evaluation of ...
  76. [76]
    [PDF] Ultrastable optical clock with two cold-atom ensembles
    Atomic clocks based on optical transitions are the most stable, and therefore precise, timekeepers available. These clocks operate by alternating intervals ...<|separator|>
  77. [77]
    NIST Ion Clock Sets New Record for Most Accurate Clock in the World
    Jul 14, 2025 · Beyond its world-best accuracy, 41% greater than the previous record, this new clock is also 2.6 times more stable than any other ion clock.Missing: 1.12 PHz 18
  78. [78]
    None
    ### Performance Metrics for High-Stability Single-Ion Clock
  79. [79]
    A zero-dead-time strontium lattice clock with a stability at 10^{-19} level
    Sep 18, 2025 · The estimated long-term stability based on the combined data of these measurements reaches 2.5 \times 10^{-19} at one day. These results ...
  80. [80]
    CSAC-SA65 - Microchip Technology
    The CSAC-SA65 is a Chip Scale Atomic Clock with output frequency between 10-16.4 MHz, 3.3V supply, and a temp range of -40 to +80°C.
  81. [81]
    International Atomic Time (TAI) - Time and Date
    International Atomic Time (TAI) is a time scale that uses the combined output of some 450 highly precise atomic clocks. It provides the exact speed at which ...
  82. [82]
    The Cesium Atomic Fountain Clock Developed by NTSC was ...
    May 21, 2024 · And then, after regular calibration by about 10 cesium atomic fountain clocks located in different countries, EAL becomes a continuous, stable ...
  83. [83]
    Redefining the second: Optical atomic clock achieves record ...
    Jan 17, 2025 · These optical clocks are still being assessed, but already now, some are 100 times more accurate than cesium clocks. They will therefore become ...
  84. [84]
    [PDF] A new record in atomic clock performance - JILA
    4.5 Atomic Clock Stability Comparison . ... clock precision scales as 1/√N, where N is the number of atoms in the clock. Until the work described in this ...<|separator|>
  85. [85]
    2021-12-21-record-tai - BIPM
    Dec 21, 2021 · Indeed after redefinition Cs fountains will become secondary standards with enlarged uncertainty while the uncertainty of optical standards will ...Missing: cesium 2025
  86. [86]
    [PDF] Role of the IERS in the leap second - ITU
    Earth rotation (UT1–UTC). • Observations, predictions, algorithms, and software. • IERS works to provide for users future needs. • Whatever decision the ITU ...
  87. [87]
    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 ...
  88. [88]
    Leap Second - What is it? - Time and Date
    Since then, a total of 27 leap seconds have been added (see table). This means that UTC is currently 37 seconds behind TAI (see UTC and TAI clocks).
  89. [89]
    The World Is Going to Lose The Leap Second. Here's Why
    Nov 21, 2022 · Scientists and government representatives meeting at a conference in France voted on Friday to scrap leap seconds by 2035.
  90. [90]
    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 ...
  91. [91]
    Optical clocks started the calibration of the international atomic time
    Mar 4, 2019 · Reliable calibrations by more optical clocks will allow BIPM to foresee a possible maintenance of UTC based on the new optical definition of the ...Missing: comparisons | Show results with:comparisons
  92. [92]
    NIST Time and Frequency Transfer using Common-View GPS
    This technique allows the direct comparison of two clocks at remote locations. A generic common-view setup is illustrated in the figure.Missing: formula | Show results with:formula
  93. [93]
    Relativistic Clock Correction - Navipedia - GSSC
    Jul 8, 2018 · That is, the clock on satellite appear to run faster ( ≃ 38 μ s / d a y ) than on ground [note: Δ f / f = Δ T / T ]. This effect is corrected ( ...
  94. [94]
  95. [95]
    Analysis of synchronization in quantum key distribution networks - ITU
    Mar 14, 2025 · This supplement covers time synchronization in QKDNs, including its function, performance, implementation, and procedures. Time synchronization ...
  96. [96]
    Short‐term GNSS satellite clock stability - AGU Publications - Wiley
    Jul 24, 2015 · Global Navigation Satellite System (GNSS) clock stability is characterized via the modified Allan deviation using active hydrogen masers as the receiver ...2 Development · 5 Clock Stability Results · 6 Implications
  97. [97]
    [PDF] Atomic Clocks and Timing Systems in Global Navigation Satellite ...
    Accurate and ultra-stable atomic clocks have been recognized as the critical equipment for the precision. Global Navigation Satellite Systems (GNSS).
  98. [98]
    [PDF] Basics of the GPS Technique: Observation Equations§
    The unknown receiver clock error can be estimated by the user along with unknown station coordinates. There are 4 unknowns; hence we need a minimum of 4 ...Missing: flight | Show results with:flight
  99. [99]
    BeiDou Navigation Satellite System in 2024 - GPS World
    Nov 15, 2024 · BeiDou Navigation Satellite System in 2024 is achieving stable and precise operations, advancing technological upgrades.Missing: optical | Show results with:optical
  100. [100]
    Could chip-sized atomic clocks replace GPS? - C4ISRNet
    Aug 28, 2019 · Improving the accuracy of chip-scale atomic clocks is key to providing a reliable backup or replacement of GPS timing data for the Department of ...
  101. [101]
    M-Code Time and Frequency Solutions - Microchip Technology
    With built-in 1 PPS disciplining capability, these atomic clocks deliver accuracy with low size, weight and power for applications where performance and ...
  102. [102]
    ESA - ACES: Atomic Clock Ensemble in Space
    By comparing clocks in space and on Earth, ACES will provide scientists with precise measurements to test Einstein's gravitational time dilation effect, search ...
  103. [103]
    [PDF] DGPS Carrier Phase Networks and Partial Derivative Algorithms
    Centimetre to decimetre level positioning accuracy requires the use of carrier-phase GPS measurements. The system must be operated in a double differential mode ...
  104. [104]
  105. [105]
    Quantum networks of clocks open the door to probe how ... - Phys.org
    Jul 14, 2025 · The researchers showed that superpositions of atomic clocks in quantum networks would pick up different time-flows in superposition, and that ...
  106. [106]
    Measuring the stability of fundamental constants with a network of ...
    May 11, 2022 · In this work, we discuss how a network of atomic and molecular clocks can be used to look for such variations with unprecedented sensitivity over a wide range ...
  107. [107]
    A multi-ensemble atomic clock enhanced using quantum computing ...
    Feb 12, 2024 · Using a new technique for individual quantum control, multiple atomic clocks can be made to tick at different rates, like different hands on a watch.
  108. [108]
    Rubidium Oscillators for 5G Networks
    Sep 5, 2025 · Rubidium Oscillators deliver 1E-11 frequency stability over 24 hours, outperforming quartz-based alternatives by 3 orders of magnitude.
  109. [109]
    TCXO, OCXO and Rubidium - understanding oscillators and the role ...
    Aug 19, 2025 · : In 5G and 4G base stations, TCXOs ensure precise clock synchronization ... Their frequency stability ranges from 10^-11 to 10^-12 ppm per ...
  110. [110]
    [PDF] Rb Atomic Clocks for wireless mobile(5G) and Test & Measurement ...
    low power, cost, size and STS. ▫ Lamp offered good stability, but too big. OCXO. Rb-Lamp. Cost.
  111. [111]
    [PDF] Synchronizing Stock Market Clocks to UTC(NIST) - URSI
    It has made markets harder to regulate and accurate time synchronization more important. ✓ HFT allows trades to execute in intervals measured in microseconds. ✓ ...
  112. [112]
    10 Reasons Why Time is Critical for Trading Systems - Safran
    High-Frequency Trading (HFT). In high-frequency trading, where trades are executed in microseconds or even nanoseconds, time synchronization becomes critical.
  113. [113]
    Keeping Stock Trades Fair | NIST
    Jun 30, 2025 · In Europe, clocks involved in high-volume trading must be within 100 microseconds (millionths of a second) of the time provided by an atomic ...Missing: synchronization | Show results with:synchronization
  114. [114]
    [PDF] Time Synchronization in the Electric Power System - naspi
    Synchrophasor technology uses phasor measurement units (PMUs) to measure voltage and current waveforms and calculate phasors at high speeds in real-time, time- ...
  115. [115]
    Utilization of Chip Scale Atomic Clock for Synchrophasor ...
    Aug 6, 2025 · To address this issue, chip scale atomic clock (CSAC) is proposed to be used as a backup solution for time synchronization in this paper. It is ...Missing: 1μs | Show results with:1μs
  116. [116]
    A microsecond-level alternative clock source for power grid ...
    Jun 15, 2025 · ... synchronization networks [21]. For instance, the deployment of chip-scale atomic clocks (CSACs), which boast a frequency accuracy up to 0.05 ...
  117. [117]
    Help with WWVB Radio Controlled Clocks | NIST
    WWVB broadcasts on a frequency of 60 kHz. Your radio controlled clock actually has a miniature radio receiver inside, which is permanently tuned to receive the ...
  118. [118]
    Atomic Clock vs. Radio Clock - Meinberg
    Radio Clocks​​ Once your radio-controlled clock has decoded the signal from DCF77, it will synchronize its own clock to the time signals received by radio.
  119. [119]
    Atomic Time for a Networked World | NIST
    Jun 30, 2025 · Atomic time serves several functions within networks. It is used to time-stamp packets sent over networks, so that data in the packets is properly ordered.
  120. [120]
    Atomic Clock Market | Global Market Analysis Report - 2035
    Oct 17, 2025 · From 2030 to 2035, the market is forecast to grow from USD 850 million to USD 1,200 million, adding another USD 350 million, which constitutes ...
  121. [121]
    Miniature Atomic Clock (MAC-SA5X) - Microchip Technology
    The MAC-SA5X is designed for applications that require rubidium atomic clock holdover performance but are constrained by size and low power requirements.
  122. [122]
    Roadmap towards the redefinition of the second - IOPscience
    This paper outlines the roadmap towards the redefinition of the second, which was recently updated by the CCTF Task Force created by the CCTF in 2020.
  123. [123]
    Redefinition Second - BIPM
    The time and frequency community is highly engaged in the long journey that will eventually lead to the redefinition of the second, possibly in 2030. We will ...Missing: optical | Show results with:optical
  124. [124]
    [1312.3419] A strontium lattice clock with $3 \times 10 - arXiv
    Dec 12, 2013 · The absolute frequency of the 5s^2 ^1S_0-5s5p ^3P_0 transition in ^{87}Sr is 429,228,004,229,873.13(17) Hz. Our result is in excellent agreement ...
  125. [125]
    [0802.0222] Optical Lattice Induced Light Shifts in an Yb Atomic Clock
    Feb 1, 2008 · We present an experimental study of the lattice induced light shifts on the 1S_0-3P_0 optical clock transition (v_clock~518 THz) in neutral ...
  126. [126]
    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 ...
  127. [127]
    New Atom-Based Thermometer Measures Temperature More ...
    Jan 23, 2025 · “Atomic clocks are exceptionally sensitive to temperature changes, which can cause small errors in their measurements,” said NIST research ...
  128. [128]
    [PDF] Challenges for Electronic Circuits in Space Applications
    Finally, the space radiation environment can have damaging effects on spacecraft electronics. There are large variations in the levels of and types of radiation ...
  129. [129]
    radiation hardening technique: Topics by Science.gov
    The Radiation Hardened Electronics for Space Environments (RHESE) project consists of a series of tasks designed to develop and mature a broad spectrum of ...
  130. [130]
    Compact Rb optical frequency standard with 10−15 stability
    Oct 16, 2017 · We achieved a low-cost and small-sized Rb optical frequency standard based on 85 Rb 5S 1/2 → 6P 3/2 transition with 10 −15 stability.
  131. [131]
  132. [132]
    Prospects and challenges for squeezing-enhanced optical atomic ...
    Nov 24, 2020 · Dead time free laser stabilization, basically eliminating the Dick effect, was constructed by anti-synchronized interrogations of two atomic ...Missing: mitigation | Show results with:mitigation
  133. [133]
    Chip-scale atomic devices | Applied Physics Reviews - AIP Publishing
    Aug 14, 2018 · Chip-scale atomic clocks have a short- and long-term frequency stability ... stability of 10−11 at integration times beyond 10 s. In this design ...
  134. [134]
    [PDF] Thorium-229 nuclear clock using a VUV frequency comb - JILA
    This thesis uses a VUV frequency comb for laser spectroscopy of the 229Th nuclear clock transition, which has a high transition frequency and long excited ...
  135. [135]
    [PDF] Recent atomic clock comparisons at NIST
    Jan 2, 2025 · General relativity predicts that the frequency ν of an atomic clock depends on the gravitational potential according to the following ...
  136. [136]
    Test of Einstein Equivalence Principle by frequency comparisons of ...
    Sep 10, 2021 · The test of WEP is realized by the gravity and dynamics effects on clock comparisons. ... Assuming the accuracy of 1 × 10−18 for clock comparison ...Missing: discrepancies | Show results with:discrepancies
  137. [137]
    Hybrid Atom Tweezer Array of Nuclear Spin and Optical Clock Qubits
    Dec 10, 2024 · We demonstrate a hybrid atom array system of nuclear spin qubits and optical clock qubits, where the former act as data qubits with long coherence times.
  138. [138]
    Dual-species atomic arrays show promise for quantum error correction
    Oct 15, 2024 · A study in Nature Physics has realized a dual-species Rydberg array combining rubidium (Rb) and cesium (Cs) atoms to enhance quantum computing and its ...<|control11|><|separator|>