Fact-checked by Grok 2 weeks ago

Hyperfine structure

Hyperfine structure is the smallest observable splitting in the energy levels and spectral lines of atoms and molecules, arising from the interaction between the nuclear magnetic dipole moment (and higher-order multipoles like the electric quadrupole) and the generated by the orbiting and spinning electrons. This effect produces energy shifts on the order of 10^{-6} or less, much smaller than the splitting caused by relativistic effects and spin-orbit coupling, which is typically around 10^{-4} . The F, combining the electron's total angular momentum J and the nuclear spin I (where F = |I + J| to |I - J|), governs the multiplicity of these split levels. In the , hyperfine structure is particularly prominent in the (n=1, L=0, S=1/2, J=1/2), where the and proton spins interact via their magnetic moments, splitting the level into two: the lower-energy (F=0) and the higher-energy (F=1), separated by an energy difference corresponding to a of approximately 1,420 MHz (the famous 21 cm radio line). The measured value is precisely 5.88 × 10^{-6} . The 21 cm line has been crucial in for mapping neutral in galaxies, revealing the Way's spiral structure. Beyond , hyperfine structure manifests in heavier atoms through similar interactions, often complicated by quadrupole moments, and is observable in alkali metals like cesium and via laser spectroscopy. Its precise measurement enables applications in atomic clocks, where the hyperfine in ^{133}Cs (frequency: 9,192,631,770 Hz) defines the international second, achieving timekeeping accuracies better than 1 part in 10^{15}. Hyperfine effects also play a role in with trapped ions and in precision tests of fundamental symmetries, such as parity violation in nuclei.

Fundamentals

Definition and Physical Origin

Hyperfine structure refers to the finest level of splitting observed in the spectral lines of atoms and molecules, arising from the between the magnetic and electric moments of the and the surrounding electrons or molecular fields. This splitting occurs in otherwise of the atomic or molecular ground and excited states, where the total angular momentum \mathbf{F} is the vector sum of the spin angular momentum \mathbf{I} and the electronic angular momentum \mathbf{J}, such that \mathbf{F} = \mathbf{I} + \mathbf{J}. For atoms with nuclear spin I > 0, this coupling lifts the degeneracy, producing multiple hyperfine levels labeled by the F, which range from |I - J| to I + J. The physical origin of hyperfine structure lies in two primary interactions. The magnetic dipole interaction stems from the coupling with the generated by the electrons, which includes contributions from the electron's orbital motion and ; relativistic effects on the electron orbits, such as those described in the , produce an effective at the nucleus that interacts with the . Additionally, the electric quadrupole interaction arises from the non-spherical distribution of the charge, creating an electric quadrupole moment that couples with the produced by the asymmetric electron cloud around the nucleus. These effects reveal properties that are otherwise invisible in the gross spectra dominated by transitions. In terms of energy scale, hyperfine splittings are typically $10^{-6} to $10^{-3} times smaller than splittings, which themselves arise from coarser electron spin-orbit couplings. A prominent example is the hyperfine in the of neutral , known as the 21 cm line, corresponding to a of 1420 MHz and an splitting of about 5.9 \mueV between the F=1 and F=0 levels. This , driven by the interaction between the proton and electron spins, is crucial for in mapping interstellar . Hyperfine structure was first resolved in the optical spectra of metals like sodium in the 1930s, marking the experimental confirmation of these subtle nuclear-electronic couplings.

Relation to Other Spectral Splittings

Hyperfine structure represents the smallest scale of splitting in and molecular spectra, arising from interactions between the nuclear spin and the electronic angular momentum. It fits into a broader of spectral features that refine the basic energy levels predicted by the non-relativistic . The gross structure originates from the dominant Coulomb interactions and orbital angular momentum quantization, producing energy differences on the order of $10^{15} Hz for typical optical transitions in light atoms like . The , due to spin-orbit coupling and relativistic corrections, introduces smaller splittings on the scale of $10^{9} to $10^{11} Hz (GHz to hundreds of GHz), depending on the Z, as the splitting scales roughly as Z^4 \alpha^2 times the gross energy, where \alpha is the . Hyperfine structure follows at even lower energies, typically $10^6 to $10^9 Hz (MHz to GHz), while the —a quantum electrodynamic correction—provides an intermediate scale of around 1 GHz in , resolving degeneracies within the . A key distinction of hyperfine structure is its dependence on nuclear properties, particularly a non-zero nuclear spin I > 0, which is absent in fine structure phenomena that involve only electronic degrees of freedom. For atoms with I = 0, such as ^{12}C or ^{16}O, no hyperfine splitting occurs. In contrast, fine structure splits levels based on total electronic angular momentum j = l \pm s, independent of the nucleus. A classic example is the hydrogen ground state (n=1, l=0), where fine structure leaves the $1s level unsplit (as l=0), but hyperfine interaction couples the electron spin s = 1/2 with the proton spin I = 1/2, yielding total angular momentum F = 0 or F = 1 levels separated by 1420 MHz. This splitting reveals nuclear magnetic properties, such as the proton's magnetic moment, whereas fine structure probes electronic relativistic effects. Additionally, electric quadrupole hyperfine interactions (for I \geq 1) expose nuclear charge distributions, a feature unrelated to fine or gross structure. The energy scales highlight hyperfine structure's position as the finest resolution in this hierarchy, enabling precise probes of structure. The following table summarizes typical frequencies for , illustrating the orders-of-magnitude differences:
Splitting TypePhysical OriginTypical Frequency (Hydrogen)Example Transition
Gross Structure + orbital ~$10^{15} HzLyman-α (1s–2p): 2.47 × 10^{15} Hz
Fine StructureSpin-orbit + relativistic corrections~10 GHz2p_{3/2}–2p_{1/2}: 10.9 GHz
QED vacuum fluctuations~1 GHz2s–2p_{1/2}: 1058 MHz
Hyperfine Structure spin–electron coupling~1 GHz ()1s F=1–F=0: 1420 MHz
These values scale with atomic number and quantum numbers; for heavier atoms, fine structure can reach ~10^4 GHz due to Z^4 enhancement. Hyperfine patterns exhibit strong isotope dependence, as they rely on the nuclear spin I and magnetic moment \mu_I, which vary across isotopes. For instance, in , the common isotope ^1H (protium, I = 1/2, \mu_I \approx 2.79 \mu_N) shows a 1420 MHz ground-state splitting, while ^2H (, I = 1, \mu_I \approx 0.86 \mu_N) has a much smaller splitting of 327 MHz due to the quadrupled moment of inertia and reduced magnetic moment per spin unit. This variation allows isotopic identification in spectra and underscores hyperfine structure's sensitivity to nuclear composition, unlike , which is isotope-independent to first order.

Historical Development

Early Observations

The hyperfine structure in atomic spectra was first observed in the late through high-resolution of lines. In 1928, Hermann Schüler resolved the hyperfine components of the sodium D lines, revealing each line as a closely spaced doublet with separations on the order of 0.01 cm⁻¹, which was initially attributed to the presence of isotopes rather than nuclear interactions. Independent observations by A. N. Terenin and L. N. Dobretsov in the same year confirmed this splitting in sodium vapor, marking the initial empirical detection of these fine details beyond the resolution limit. During the 1930s, further key experiments expanded these findings to other elements, particularly alkali metals. Ernst Back and investigated hyperfine multiplets in the spectra of and , identifying complex patterns in multiple lines that varied systematically with , using grating spectrographs to achieve the necessary resolution. Their work on lines in 1928 demonstrated multiplet structures with up to four components, highlighting the prevalence of hyperfine effects in heavy elements. These studies relied on advancements in , such as the Fabry-Pérot interferometer, which allowed precise measurement of splittings as small as 0.01 cm⁻¹, and ruled diffraction gratings that improved for detailed line profiles. Early observations also revealed isotopic variations in hyperfine patterns, providing clues to properties. For , differences in the hyperfine structure between the isotopes ⁶Li and ⁷Li were noted in the mid-1930s, with the patterns enabling the assignment of nuclear spins I = 1 for ⁶Li and I = 3/2 for ⁷Li through analysis of level splittings in optical spectra. A notable event was the prediction of the 21 cm line by Hendrik van de Hulst, arising from the hyperfine transition in neutral , although its experimental detection came later; this line's anticipated radio emission stemmed from early spectral insights into hyperfine effects in light atoms.

Theoretical Advancements

The theoretical foundations of hyperfine structure emerged in the mid-1920s when Wolfgang Pauli proposed that the small splittings observed in atomic spectral lines arose from an angular momentum associated with the atomic nucleus, which he termed nuclear spin. This concept marked a departure from purely electronic models of atomic spectra, attributing the phenomenon to interactions between the nuclear spin and the electron's magnetic moment. Building on this, Pauli further formalized the role of the nuclear spin quantum number I in 1926, recognizing it as the primary cause of hyperfine splitting through coupling with the total electronic angular momentum J. A significant advancement came in 1930 with Enrico Fermi's derivation of the contact interaction term, which specifically described the magnetic hyperfine coupling for s-electrons where the electron probability density at the nucleus is non-zero. This term, proportional to the product of the nuclear spin and the electron spin density at the nucleus, provided a quantitative framework for calculating splitting magnitudes in alkali atoms. During the 1930s, Hendrik Casimir and others extended these ideas by deriving the general forms of the and interaction terms in the , accounting for both point-like and distributed nuclear charge effects. These developments were integrated with the Dirac relativistic theory of the electron, enabling more accurate predictions for fine and hyperfine splittings in atoms where relativistic corrections to electron wavefunctions became relevant. Key milestones in the post-1930s era included the 1931 Breit-Rabi formula, which precisely described the hyperfine energy levels in hydrogen-like atoms under external magnetic fields, resolving the intermediate-field regime between weak and strong Zeeman effects. This formula, essential for atomic beam experiments, allowed for the separation of hyperfine and Zeeman contributions to energy shifts. Following , the advent of the by and J. Hans D. Jensen in 1949 provided a microscopic understanding of nuclear structure, enabling improved predictions of electric moments from hyperfine data and explaining variations in quadrupole hyperfine splittings across isotopic chains. The evolution of hyperfine theory represented a fundamental shift from models considering only electronic degrees of freedom to those incorporating nuclear quantum properties, such as spin I. This inclusion facilitated the determination of nuclear spins from experimental spectra; for instance, the hyperfine splitting in the ground state of hydrogen confirmed the proton's spin as I = 1/2. In the 1950s, theoretical advancements addressed discrepancies in hyperfine splittings for heavy atoms, where simple point-nucleus approximations failed. The Bohr-Weisskopf effect, introduced in 1950, accounted for the finite distribution of nuclear magnetization, explaining isotopic anomalies in hyperfine constants. Concurrently, core polarization models, as developed by Abragam and colleagues in 1955, incorporated the distortion of the electronic core by the nuclear moment, enhancing the effective magnetic field at the nucleus. Relativistic effects, including corrections to electron wavefunctions near the nucleus, were further refined during this period to resolve anomalous splittings in elements like and mercury.

Atomic Hyperfine Interactions

Magnetic Dipole Mechanism

The magnetic dipole mechanism arises from the interaction between the \vec{\mu}_I = g_I \mu_N \vec{I}—where g_I is the g-factor, \mu_N is the , and \vec{I} is the nuclear spin angular momentum—and the generated by the electrons' orbital motion and spin. This interaction splits the degenerate fine-structure energy levels into hyperfine components, with the strength determined by the near the and the properties. In atoms with spherical nuclear charge distributions, such as light elements, this mechanism dominates the hyperfine structure, providing key insights into nuclear g-factors through precise measurements. The effective Hamiltonian for this interaction is H_{hf} = A \vec{I} \cdot \vec{J}, where \vec{J} is the total electron and A is the hyperfine that encapsulates the magnetic interaction strength. The constant A derives from three contributions: the Fermi contact term for s-electrons, the orbital term, and the -dipolar term. The Fermi contact interaction, originating from the polarization at the nucleus, is expressed as A_s = \frac{8\pi}{3} g_e g_I \mu_B \mu_N |\psi(0)|^2, where g_e is the electron g-factor, \mu_B is the Bohr magneton, and |\psi(0)|^2 is the electron probability density at the nucleus; this term vanishes for orbitals with l > 0. For l > 0, the orbital contribution involves \langle L \cdot I / r^3 \rangle, coupling the nuclear spin to the electron's orbital angular momentum, while the dipolar term accounts for the classical dipole-dipole coupling between \vec{I} and the \vec{S}. In the of (^1H), where I = 1/2 and J = 1/2, the hyperfine levels are labeled by the total angular momentum F = I + J = 1 and F = |I - J| = 0, with the energy splitting \Delta E = A between the F=1 and F=0 states corresponding to the famous 1420 MHz (21 cm) transition. This splitting exemplifies the mechanism in atoms, where the unpaired s-electron enhances the contact term, leading to observable hyperfine structure in their spectra; for instance, the ground-state splitting in directly probes the proton's . In intermediate magnetic s, the Breit-Rabi describes the nonlinear Zeeman shifts of these levels: E(F, m_F) \approx (\Delta E / 2) \left(1 + x^2 / 2 \pm x \sqrt{1 + x^2 / 4}\right), with x = (g_J - g_I) \mu_B B / \Delta E, where B is the external , g_J the electron g-factor, and m_F the of F; this , derived for systems like , enables precise of from field-dependent splittings.

Electric Quadrupole Mechanism

The electric quadrupole mechanism in atomic hyperfine structure arises from the between the nuclear electric moment and the electric field gradient () produced by the asymmetric distribution of surrounding electrons. This coupling becomes relevant for nuclei with I \geq 1, where the nucleus possesses a non-spherical charge distribution, leading to a tensorial that further splits the hyperfine levels beyond the magnetic dipole effect. The quadrupole moment Q quantifies the deviation from spherical in the nuclear charge density, positive for prolate (elongated) shapes and negative for (flattened) shapes, providing insights into nuclear deformation. The interaction is described by the quadrupole Hamiltonian: H_Q = \frac{eQ}{2I(2I-1)} \vec{I} \cdot \nabla E \cdot \vec{I}, where e is the , \vec{I} is the nuclear spin operator, and \nabla E is the EFG tensor at the , with components derived from the second derivatives of the electrostatic potential V due to the electrons. In atomic systems, the EFG originates primarily from valence electrons in non-s-state orbitals, such as p or d orbitals, where the electron density lacks spherical symmetry; the axial component is given by e_q = \partial^2 V / \partial z^2 evaluated at the , assuming a principal along the quantization . For , the energy shifts depend on the nuclear |m_I|, resulting in distinct splittings for different total F = J + I states. This mechanism vanishes for nuclei with I = 1/2, as no quadrupole moment exists, distinguishing it from the isotropic interaction. In atoms like (^{35}\mathrm{Cl}, I = 3/2), the ^2P_{3/2} exhibits hyperfine splitting into F=2 and F=1 levels, with the quadrupole coupling constant B measured via atomic beam magnetic resonance, enabling determination of the nuclear moment Q \approx -0.079 . Similarly, for oxygen (^{17}\mathrm{O}, I = 5/2), atomic hyperfine studies yield Q \approx -0.0256 , revealing its shape through the negative sign, independent of contributions alone. These measurements highlight the quadrupole effect's role in probing nuclear structure via .

Molecular Hyperfine Interactions

Nuclear Spin-Spin Coupling

Nuclear spin-spin coupling in molecules arises from the indirect interaction between two nuclear spins, \vec{I_1} and \vec{I_2}, mediated by the bonding electrons that transmit magnetic fields through the molecular framework, distinguishing it from the direct electron-nuclear interactions in hyperfine structure. This through-bond primarily involves second-order effects from the hyperfine interactions between each and the surrounding electrons, resulting in an effective scalar or tensorial coupling that splits spectral lines in polyatomic systems. Unlike cases where hyperfine splitting stems from direct moments, the molecular variant relies on and delocalization along sigma bonds. The interaction is described by the coupling Hamiltonian H_{SS} = 2\pi J \vec{I_1} \cdot \vec{I_2}, where J is the coupling constant in hertz, representing the strength of the interaction; for isotropic cases common in solution, J originates mainly from the Fermi contact mechanism, which depends on the s-electron density at the nuclei. This Hamiltonian captures the scalar coupling that leads to observable multiplet patterns in spectra, with J being either isotropic (dominant in fluids) or anisotropic in oriented systems. Theoretically, J emerges from second-order applied to the electron-nuclear hyperfine , where the indirect mechanism dominates over direct nuclear-nuclear dipolar coupling in most molecules. The direct dipolar contribution, given approximately by J \approx \frac{\mu_0 \gamma_1 \gamma_2 \hbar^2}{4\pi r^3} for nuclei separated by distance r, is typically small and anisotropic, averaging to zero in isotropic environments, whereas the indirect term—arising from virtual excitations of spins—provides the primary isotropic J via mechanisms like Fermi contact in covalent bonds. This electron-mediated nature makes J sensitive to the electronic structure and bond type. In diatomic molecules like , the proton-deuteron is J_{\mathrm{HD}} \approx 43 Hz, manifesting as a splitting in NMR spectra due to the heteronuclear interaction. Similarly, in organic molecules, the one-bond ^1\mathrm{H}-{}^{13}\mathrm{C}\) coupling (^1J_{\mathrm{CH}}$) typically ranges from 120 to 200 Hz, producing characteristic multiplets that reflect the sp³-hybridized carbon environment./05%3A_Structure_Determination_Part_II_-_Nuclear_Magnetic_Resonance_Spectroscopy/5.06%3A_Spin-Spin_Coupling) These couplings provide insights into molecular geometry and bonding characteristics, with larger J values observed in multiple bonds—such as ^1J_{\mathrm{CH}} \approx 150{-}250 Hz in alkynes due to increased s-character—allowing inference of hybridization and torsion angles. The effect is isotope-specific, vanishing for nuclei with zero spin (I=0), like {}^{12}\mathrm{C} or ${}^{16}\mathrm{O}), which do not contribute to or experience the coupling.

Spin-Rotation and Other Effects

In molecular hyperfine structure, the spin-rotation represents a key mechanism beyond direct nuclear spin-spin coupling, arising from the magnetic coupling between a nucleus's angular momentum \vec{I} and the molecule's rotational angular momentum \vec{N}. This effect originates from the internal magnetic field generated during molecular rotation, primarily due to the motion of electrons and the nuclear magnetic moments themselves, which interact with the rotating charge . In diatomic and linear molecules, the is often isotropic, but in asymmetric tops, it takes a tensorial form to account for the of the molecular frame. The spin-rotation Hamiltonian is expressed as H_{SR} = \sum_k \vec{I}_k \cdot \boldsymbol{\epsilon}_k \cdot \vec{N}, where the sum runs over nuclei with nonzero spin k, and \boldsymbol{\epsilon}_k is the spin-rotation tensor for nucleus k, with components typically on the order of MHz. This tensor arises from second-order perturbation theory involving the nuclear magnetic moment interacting with the rotational magnetic field; for diatomic molecules, its parallel component scales as \epsilon_\parallel \approx \gamma_I \frac{\mu_0}{4\pi} \frac{8\pi}{3} g_e \mu_B \mu_N / h, where \gamma_I is the nuclear gyromagnetic ratio, g_e is the electron g-factor, \mu_B and \mu_N are the Bohr and nuclear magnetons, respectively, reflecting the dominant electronic contribution to the field at the nucleus. Off-diagonal elements of \boldsymbol{\epsilon} become significant in asymmetric rotors, leading to more complex splittings. A prominent example occurs in isotopologues, where the ^{12}C has zero spin (I=0), yielding no hyperfine splitting in ^{12}CO, but the I=1/2 ^{13}C in ^{13}CO introduces observable spin-rotation effects in the J=1 \leftarrow 0 rotational transition, with hyperfine splittings on the order of 40 kHz due to the tensor components. Similarly, in (NH_3), a symmetric top, the inversion doubling of rotational levels is modulated by hyperfine interactions from the and nuclear spins, where spin-rotation coupling contributes to the complex multiplet structure observed in the ground-state inversion-rotation transitions around 24 GHz, enhancing the resolution of the tunneling-split levels. Unique manifestations of spin-rotation effects include hyperfine-induced modifications to tunneling splittings in symmetric top molecules, where the interaction lifts degeneracies in the rotational-nuclear spin basis, producing "superfine" cluster splittings smaller than the primary hyperfine but resolvable in high-precision spectra, as seen in trigonal and tetrahedral rotors like phosphine derivatives. In chiral molecules, recent advancements have exploited spin-rotation hyperfine structure for parity mixing, enabling distinction between enantiomers through field-induced level repulsions sensitive to parity-violating interactions; precision spectroscopy in the 2020s has demonstrated sensitivities to new physics beyond the Standard Model, such as P- and T-violating forces, with resolutions down to mHz in molecules like chiral alcohols. Additionally, in small free radicals such as the OH radical, spin-rotation hyperfine splittings (e.g., \sim 50-100 MHz in the ground state) are crucial for astrophysical applications, allowing hyperfine-resolved lines to trace interstellar cloud densities and temperatures via collisional excitation models.

Measurement Techniques

Spectroscopic Methods

Optical spectroscopy plays a central role in resolving hyperfine structure in atomic spectra, particularly through high-resolution laser absorption and emission techniques. In alkali atoms, Fabry-Pérot interferometers are employed to calibrate frequencies and achieve sub-MHz precision in measuring hyperfine splittings of optical lines, such as the D lines in and cesium. These setups often involve stabilizing lasers to the interferometer's fringes, enabling the identification of hyperfine components separated by tens of GHz. For finer , Doppler-free spectroscopy eliminates first-order , allowing MHz-level discrimination of hyperfine transitions in alkali vapors; counterpropagating pump and probe beams create dips at exact resonance, as demonstrated in precise measurements of the 5S_{1/2} to 5P_{3/2} transition in . Radiofrequency and targets direct hyperfine transitions, often using to modulate frequencies for detection. In hydrogen, the 21 cm hyperfine line (1420 MHz) is observed via radiofrequency techniques in laboratory settings and with large radio telescopes for astrophysical contexts, revealing the ground-state splitting with resolutions down to Hz in controlled environments. The magnetic resonance (ABMR) method, pioneered in , provides precise determination of hyperfine constants (A) by detecting resonant flips in collimated under applied magnetic fields; it has been applied to isotopes like potassium-39, with relative uncertainties on the order of 1%. For molecular systems, molecular electric resonance isolates electric interactions by applying to orient rotational states, enabling measurement of quadrupole coupling constants (e_qQ) in diatomic molecules like LiF and TlCl. Advanced variants like further enhance resolution for hyperfine features in emission or absorption spectra, particularly useful for resolving closely spaced components in transition metals such as . The exemplifies high-precision radiofrequency measurement, oscillating at the hyperfine of 1420.405751768(0.00002) MHz through in a storage bulb, serving as a with stability better than 10^{-13} over seconds. To disentangle hyperfine structure from shifts, especially for low-abundance isotopes, enriched samples are utilized in optical and beam , as seen in measurements of isotopes where enrichment enabled the first determination of the ^{29} hyperfine structure and improved the accuracy of the 28Si-30Si shift by approximately two orders of magnitude. These techniques collectively interpret spectra based on underlying and electric interactions, providing quantitative access to properties.

Advanced Detection Approaches

Advanced detection approaches for hyperfine structure extend beyond conventional spectroscopic methods by leveraging quantum control, high-resolution magnetic resonance, and ultrafast techniques to probe hyperfine interactions with unprecedented precision and in challenging environments. These methods enable the coherent manipulation and readout of hyperfine states in isolated systems, resolving subtle couplings that are inaccessible to ensemble-averaged techniques. In ion traps, quantum logic spectroscopy allows for the coherent manipulation of hyperfine states in single ions, such as the ^43Ca^+ clock transition between the 4^2S_{1/2} F=4 and F=3 levels, achieving gate fidelities exceeding 99.9% through microwave-driven operations. This approach uses sympathetic cooling and state-dependent for non-destructive readout, enabling hyperfine frequency measurements with uncertainties below 10^{-14}. For instance, experiments with ^43Ca^+ ions demonstrate robust initialization and high-fidelity two-qubit gates based on hyperfine qubits, facilitating precise determination of hyperfine splittings in processing contexts. Nuclear magnetic resonance (NMR) and electron paramagnetic resonance (EPR) techniques resolve hyperfine interactions in solid-state and solution environments, particularly through electron-nuclear double resonance (ENDOR) for paramagnetic centers. ENDOR enhances resolution of small hyperfine couplings (e.g., <1 MHz) by applying radiofrequency pulses to excite nuclear transitions while monitoring EPR signals, as applied to iron-sulfur proteins where hyperfine tensors reveal ligand environments. In modern spintronics, pulsed EPR variants like ELDOR-detected NMR bridge EPR and NMR for hyperfine spectroscopy in molecular magnets, enabling detection of nuclear spins coupled to electron spins in devices for spin-based information processing. These methods address limitations in nuclear hyperfine resolution by combining high sensitivity with site-specific information, such as in ENDOR studies of Mn(II) complexes where hyperfine parameters inform electronic structure. Laser cooling confines ultracold atoms to the Lamb-Dicke regime, where the recoil energy is much smaller than the trap frequency (η << 1), allowing resolved hyperfine sidebands in for precise state preparation. In this regime, atoms like ^87Rb experience minimal motional heating, enabling hyperfine coherence times up to seconds. Atom interferometers exploit these cooled atoms to measure hyperfine-induced phase shifts, as in where hyperfine interactions between F=1 and F=2 components cause differential phase accumulation during free fall, with sensitivities reaching 10^{-12} rad/√Hz. Such setups, using light-pulse diffraction, quantify hyperfine phase shifts from atomic interactions, enhancing inertial sensing applications. Developments in the 2020s have pushed optical lattice clocks to resolve hyperfine structure with fractional frequency precision of 10^{-18}, using neutral atoms like ^87Sr in magic-wavelength lattices to minimize differential light shifts. These clocks leverage hyperfine transitions in bosonic species, such as M1/E2 clocks between hyperfine-split states, achieving stabilities below the standard quantum limit through spin squeezing. For example, ^171Yb lattice clocks demonstrate hyperfine-resolved interrogations with systematic uncertainties under 10^{-18}, enabling tests of quantum gravity and time dilation. Quantum sensing with nitrogen-vacancy (NV) centers in diamond probes nuclear hyperfine via optically detected magnetic resonance, resolving couplings to ^13C and ^15N nuclei with linewidths ~1 kHz at room temperature. NV centers enable nanoscale mapping of hyperfine fields, as in studies of single nuclear spins where dipolar interactions yield coherence times up to milliseconds for quantum registers. Femtosecond pump-probe spectroscopy captures transient hyperfine dynamics in excited states, revealing coherent evolution of hyperfine coherences on picosecond timescales following ultrafast excitation. Post-2015 advancements, such as two-dimensional coherent spectroscopy on cold atoms, resolve hyperfine splittings in Rydberg states by tracking population transfers and dephasing in the time domain. In molecular systems, these techniques probe hyperfine-modulated excited-state relaxation, as in fullerenes where nonadiabatic dynamics couple electronic and nuclear spins, providing insights into transient magnetic interactions not observable in steady-state methods.

Applications

Metrology and Fundamental Constants

The definition of the SI second relies on the unperturbed ground-state hyperfine transition frequency of the cesium-133 atom, fixed at exactly 9,192,631,770 Hz, a standard established in 1967 and reaffirmed in the 2019 revision of the (SI). This microwave transition between the two hyperfine levels of the 6s ^2S_{1/2} ground state provides the basis for primary frequency standards, enabling atomic clocks with long-term stability suitable for global timekeeping. As an alternative realization, the hydrogen maser utilizes the hyperfine transition in neutral hydrogen atoms at approximately 1,420 MHz, offering superior short-term stability (on the order of 10^{-15} τ^{-1/2}, where τ is averaging time) compared to cesium fountains, though it requires corrections for cavity pulling and wall shifts to achieve comparable accuracy. The meter, defined since 1983 as the distance traveled by light in vacuum in 1/299,792,458 of a second, is realized indirectly through this time standard combined with the fixed speed of light, but practical length metrology often employs for high-precision interferometry. For instance, the stabilized to a hyperfine component of the (specifically the a_{17} component at 633 nm) serves as a recommended secondary standard, with its frequency measured against the to derive the wavelength, achieving uncertainties below 10^{-11}. Such stabilization leverages to lock the laser to the narrow iodine hyperfine lines, enabling traceable realizations of the meter in dimensional metrology. Hyperfine structure also facilitates stringent tests of fundamental constants and quantum electrodynamics (QED). The hyperfine anomaly, defined as the deviation in the ratio of hyperfine splittings between isotopes from the ratio of their nuclear magnetic moments μ_I / I, arises from nuclear structure effects like the and probes finite nuclear size and magnetization distributions, with measurements in systems such as cadmium and indium yielding anomalies up to 0.1% that inform nuclear models. Variations in μ_I / I across atomic and muonic systems test QED predictions; for example, the ground-state hyperfine splitting in ordinary hydrogen agrees with QED to 10 parts per billion, but muonic hydrogen measurements reveal discrepancies of about 0.2% in the splitting, partly attributed to enhanced proton structure effects and contributing to resolutions of the through refined Zemach radius extractions. By 2022, updated CODATA recommendations incorporated refined electronic hydrogen spectroscopy data, converging the proton charge radius to 0.84075(64) fm, largely resolving the puzzle and affirming QED calculations of hyperfine splitting to parts-per-billion precision. Proposals for redefining the second using optical transitions in atoms like ^{87}Sr and ^{171}Yb leverage hyperfine-resolved states for enhanced precision, with lattice clocks achieving systematic uncertainties of 10^{-18} by 2022 and stability approaching 10^{-19} in comparisons by 2024, surpassing cesium standards and enabling future SI revisions. These clocks operate on electric-octupole-forbidden transitions between hyperfine Zeeman sublevels (e.g., ^{1}S_{0}(F=9/2) to ^{3}P_{0}(F=9/2) in ^{87}Sr), where hyperfine structure ensures magnetic insensitivity, and post-2019 advancements in blackbody radiation shift evaluations have solidified their candidacy for redefinition by the late 2020s.

Astrophysics and Nuclear Physics

In astrophysics, the hyperfine structure of neutral hydrogen plays a pivotal role in mapping the distribution of atomic gas in galaxies through the 21 cm emission line, arising from the spin-flip transition between the parallel and antiparallel states of the proton and electron spins. This line enables observations of HI regions, revealing spiral arm structures, rotation curves, and the extent of galactic disks via radio telescopes, providing insights into galaxy formation and evolution without significant dust obscuration. Similarly, the hyperfine splitting in the CN radical's rotational transitions allows probing of interstellar magnetic fields through the Zeeman effect, where line polarization splits further in the presence of magnetic fields, yielding field strengths on the order of microgauss in molecular clouds. Recent Atacama Large Millimeter/submillimeter Array (ALMA) observations in the 2020s have resolved hyperfine components in formaldehyde (H₂CO) isotopologues, such as ¹³CH₂O, aiding in the characterization of molecular cloud chemistry and serving as foreground contaminants in cosmic microwave background (CMB) studies by distinguishing kinematic components. The hyperfine transition of deuterium at 327 MHz (92 cm wavelength) provides a direct measure of the deuterium-to-hydrogen (D/H) abundance ratio in the interstellar medium, offering constraints on Big Bang nucleosynthesis (BBN) models that predict primordial D/H values around 2.5 × 10⁻⁵, as subsequent stellar processing reduces this ratio. Observations of this line in low-metallicity regions help isolate the primordial signature, tightening BBN bounds on the baryon density parameter Ω_b h² ≈ 0.022 and testing for new physics beyond the standard model. In nuclear physics, hyperfine splitting in atomic spectra serves as a sensitive probe of nuclear properties, including magnetic dipole moments, electric quadrupole moments, and charge radii, particularly for exotic isotopes produced in radioactive ion beams. At facilities like , collinear laser spectroscopy resolves hyperfine structures in fast ion beams, enabling isotope-shift measurements that reveal changes in nuclear radii across chains, such as in neutron-rich sodium isotopes where radii increase nonlinearly with neutron number. These techniques have mapped moments for over 100 short-lived nuclides, informing shell-model calculations and nuclear deformation trends far from stability. Hyperfine interactions also inform nucleosynthesis processes; in the r-process occurring during core-collapse supernovae or neutron star mergers, isotopic abundance patterns in metal-poor stars are deduced from hyperfine splitting in spectral lines of elements like barium and europium, where odd-even isotope effects broaden or split lines, constraining neutron-capture yields and site conditions. In Mössbauer spectroscopy, recoilless nuclear gamma emission reveals hyperfine fields influenced by nuclear moments and lattice vibrations, providing nuclear-level insights into isotope-specific dynamics in solids, such as quadrupole splitting in ⁵⁷Fe that quantifies electric field gradients at the nucleus.

Quantum Technologies and Precision Tests

Hyperfine structure plays a pivotal role in precision tests of quantum electrodynamics (QED), particularly through discrepancies observed in the hyperfine splitting between muonic and electronic hydrogen, which contributed to the proton radius puzzle. Measurements of the 2P_{3/2}-2S_{1/2} hyperfine transition in muonic hydrogen yielded a Zemach radius of the proton that conflicted with values from electronic hydrogen spectroscopy, highlighting potential inconsistencies in proton structure effects on QED predictions. By 2022, updated CODATA recommendations incorporated refined electronic hydrogen spectroscopy data, converging the proton charge radius to 0.84075(64) fm, largely resolving the puzzle and affirming QED calculations of hyperfine splitting to parts-per-billion precision. Additionally, hyperfine splitting in muonium (a muon-electron bound state) provides a sensitive probe for the muon's anomalous magnetic moment (g-2), where QED contributions from virtual particles tie directly to the observed 4.2σ discrepancy between experiment and Standard Model predictions, motivating searches for new physics. In quantum technologies, hyperfine states enable robust qubit implementations due to their long coherence times and insensitivity to magnetic field fluctuations. In trapped-ion quantum computing, the ^43Ca^+ ion utilizes hyperfine levels in the electronic ground state (F=4, m_F=0 and F=3, m_F=0) as a clock qubit, facilitating two-qubit entangling gates via the Mølmer-Sørensen protocol with fidelities exceeding 99.9%. Recent advancements by groups at NIST and Quantinuum have achieved two-qubit gate fidelities of 99.914(3)% using hyperfine-encoded ions, enabling scalable error-corrected operations in ion-trap arrays. Similarly, neutral atom platforms employ hyperfine clock states (e.g., |F=1, m_F=0⟩ and |F=2, m_F=0⟩ in rubidium or cesium) in optical tweezer arrays for qubit encoding, supporting high-fidelity Rydberg-mediated entangling gates and parallel operations across hundreds of atoms. Hyperfine interactions also underpin precision tests of fundamental symmetries, such as atomic parity violation (APV) in cesium, where weak neutral currents induce mixing between hyperfine levels of the 6S_{1/2} ground state, measurable to 0.3% precision and constraining extensions to the Standard Model. In diamond nitrogen-vacancy (NV) centers, hyperfine coupling between the electron spin and ^{14}N nuclear spin enables room-temperature quantum sensing of magnetic fields with nanoscale resolution, while 2024 developments in scalable NV arrays advance hybrid quantum computing by integrating these states for error-corrected qubits. Furthermore, in semiconductor quantum dots, hyperfine-mediated nuclear spin registers in silicon or diamond provide topological protection and fault-tolerant quantum memories, with recent demonstrations achieving high-fidelity initialization (F > 99.999%) for multi-qubit error correction.

References

  1. [1]
    [PDF] Hyperfine Structure in Atoms
    The higher order multipole moments produce small corrections to the atomic structure that are known generally as hyperfine effects.
  2. [2]
    [PDF] Quantum Mechanics and Atomic Physics
    Nov 23, 2022 · ▫ This is why it's called “fine” structure! Page 18. Hyperfine Structure. ▫ Proton and neutrons are also spin 1/2 particles. ▫ Proton and ...Missing: definition | Show results with:definition
  3. [3]
    12 The Hyperfine Splitting in Hydrogen - Feynman Lectures
    These spins are responsible for the “hyperfine structure” in the energy levels, which splits all the energy levels into several nearly equal levels. The ...
  4. [4]
    [PDF] Hyperfine Spectrum of Rubidium: laser spectroscopy experiments
    these two terms there is hyperfine structure, a very, very. small energy perturbation compared with the energy gap. between the two terms...and we want to see ...
  5. [5]
    Hyperfine Structure Constants on the Relativistic Coupled Cluster ...
    The hyperfine structure (HFS) constants parametrize the interaction between the electronic and the nuclear electromagnetic moments. The HFS consequently ...
  6. [6]
    [PDF] Hyperfine Structure in Hydrogen and Alkali Atoms
    Hyperfine structure occurs on a smaller energy scale than the fine structure, hence the name. The physical origin of hyperfine effects, however, is quite ...
  7. [7]
    [PDF] Hyperfine splitting - bingweb
    In atoms, hyperfine structure occurs due to the energy of the nuclear magnetic dipole moment in the magnetic field generated by the electrons, and the energy ...
  8. [8]
    Precision measurement of the 1S ground-state Lamb shift in atomic ...
    Oct 1, 1995 · Our result of 8172.874(60) MHz for the 1S Lamb shift in hydrogen is in agreement with the theoretical value of 8172.802(40) MHz.
  9. [9]
    In-beam measurement of the hydrogen hyperfine splitting and ...
    Jun 12, 2017 · The measured value of νHF=1,420,405,748.4(3.4) (1.6) Hz with a relative precision of 2.7 × 10−9 constitutes the most precise determination of ...
  10. [10]
    [PDF] arXiv:2007.06798v2 [physics.atom-ph] 16 Jul 2020
    Jul 16, 2020 · Hyperfine structure (HFS) of atomic energy levels arises due to interactions of atomic electrons with a hierarchy of nuclear multipole moments, ...
  11. [11]
    The Nuclear Spin and Magnetic Moment of Sodium from Hyperfine ...
    The hyperfine structure of the sodium D lines from a liquid-air cooled Schüler tube was observed with a Fabry-Perot interferometer.Missing: Sawyer Barker 1923
  12. [12]
    [PDF] Wolfgang Pauli - Nobel Lecture
    1945 W.PAULI was discovered, I proposed to use the assumption of a nuclear spin to inter- pret the hyperfine-structure of spectral lines15. This proposal ...
  13. [13]
    Pauli's 1924 note on hyperfine structure
    In other words, Pauli's nuclear “spin” did not include the spin of the nucleons, as we would say today, for in 1924 he did not think for a moment that the ...
  14. [14]
    Electron Spin and Its History - Annual Reviews
    May 15, 2012 · Also in 1927, Goudsmit and Back, unaware of Pauli's 1924 paper on hyperfine structure, proposed the idea of nuclear spin to account for.
  15. [15]
    Über die magnetischen Momente der Atomkerne - ADS
    Über die magnetischen Momente der Atomkerne. Fermi, E. Abstract. Publication: Zeitschrift fur Physik. Pub Date: May 1930; DOI: 10.1007/BF01339933.Missing: Moment des
  16. [16]
    Enrico Fermi – Biographical - NobelPrize.org
    “Uber die magnetischen Momente der AtomKerne”, Z. Phys., 1930, is a quantitative theory of the hyperfine structures of spectrum lines. The magnetic moments ...Missing: den Atomkernes
  17. [17]
    On the Interaction between Atomic Nuclei and Electrons - SpringerLink
    Cite this chapter. Casimir, H.B.G. (1936). On the Interaction between Atomic Nuclei and Electrons. In: On the Interaction between Atomic Nuclei and ...
  18. [18]
    Theory of Atomic Hyperfine Structure - ScienceDirect
    This chapter discusses the principal experimental techniques in the study of hyperfine interactions, such as the Hamiltonian for hyperfine interactions; ...
  19. [19]
    [PDF] Atomic Beam Frequency
    Breit-Rabi formula, Eq. (22). This energy is a function of the magnitude of the field intensity H. The fields of the deflection magnets are conserva- tive ...
  20. [20]
    The Influence of Nuclear Structure on the Hyperfine Structure of ...
    A study of such anomalies may give some information regarding the structure of nuclear moments, in particular, regarding the nuclear -factor. References (9). F.
  21. [21]
    [PDF] Core Polarization in Chrmium-53
    In 1955 Abragam, Horowitz, and Pryce showed that the magnetic field at the nucleus tnat was responsible for hyperfine structure in the ground states of ...Missing: heavy | Show results with:heavy
  22. [22]
    On the “contact” hyperfine interaction | American Journal of Physics
    Jun 1, 2024 · In 1936, Casimir analyzed the magnetic interaction of proton and neutron spins in the ground state of the deuteron using the Schrödinger ...
  23. [23]
    Magnetic hyperfine interaction made easier | AIP Publishing
    May 1, 2024 · We present two derivations of the hyperfine interaction in the ground state of hydrogen using classical electrodynamics.
  24. [24]
    Measurement of Nuclear Spin | Phys. Rev.
    Measurement of Nuclear Spin. G. Breit and II Rabi New York University, Columbia University. PDF Share. Phys. Rev. 38, 2082 – Published 1 December, 1931.Missing: title | Show results with:title
  25. [25]
    Year-2017 nuclear quadrupole moments - Taylor & Francis Online
    Spectroscopic quadrupole splittings act as a gauge of the electron distribution, and studies of molecular dynamics require Q in systems where nuclear quadrupole ...
  26. [26]
    Hyperfine structure and electric quadrupole transitions in the ...
    Jan 11, 2021 · We report here the results of the calculations of the hyperfine structure of the laser-induced electric quadrupole transitions between a large set of ...
  27. [27]
    [PDF] Atomic hyperfine structure studies using temperature/current tuning ...
    Because of the orbital motion of the electron in the electric field of the nucleus, it experiences an apparent magnetic field. Bl proportional to the orbital ...
  28. [28]
    The Hyperfine Structure and Nuclear Moments of the Stable ...
    The hyperfine structure of the 2 P 3 2 ground states of C l 3 5 and C l 3 7 is investigated by the atomic beam magnetic resonance method.
  29. [29]
    Hyperfine Structure and Nuclear Electric Quadrupole Moment of 17 O
    Hyperfine Structure and Nuclear Electric Quadrupole Moment of 17O. R A Kamper, K R Lea and C D Lustig. Published under licence by IOP Publishing Ltd
  30. [30]
    [PDF] TABLE OF NUCLEAR ELECTRIC QUADRUPOLE MOMENTS
    This Table is a compilation of experimental measurements of static electric quadrupole moments of ground states and excited states of atomic nuclei ...
  31. [31]
    Nuclear Spin-Spin Coupling - an overview | ScienceDirect Topics
    Nuclear spin-spin coupling refers to the indirect interaction between magnetic nuclei within a molecule, mediated by the spins of bonding electrons, ...
  32. [32]
    Theory and calculation of nuclear spin–spin coupling constants
    A full and successful theory of the indirect nuclear spin–spin coupling based on the complete Hamiltonian for electron–nuclear interactions was first outlined ...
  33. [33]
    Perspective on Norman Ramsey's theories of NMR chemical shifts ...
    The theories connecting the observed NMR chemical shifts and nuclear spin—spin coupling constants to electronic wave functions were published by Norman Ramsey ...
  34. [34]
    Spin–spin coupling in the HD molecule determined from 1H and 2H ...
    The spin–spin coupling constant of the HD molecule is equal to 43.136(7) Hz ... The indirect spin–spin coupling of hydrogen deuteride, J(D, H), was ...
  35. [35]
    Hyperfine interactions and internal rotation in methanol
    We present a rigorous derivation of the nuclear spin-rotation and spin-torsion coupling terms in the hyperfine Hamiltonian for molecules with internal rotation.
  36. [36]
    Hyperfine structure in the rotational spectra of trans-formic acid
    The spin-rotation Hamiltonian,. HSR = K. IK · CK · J,. (2) with K running over all nuclei with a non-zero nuclear spin IK. (hydrogens in the present case) ...
  37. [37]
    [astro-ph/0403059] Hyperfine Structure in H^{13}CO^{+} and ^{13}CO
    Mar 2, 2004 · We report discovery of hyperfine structure in the H^{13}CO^{+}(1-0) transition. The doublet splitting is observed to be 38.5 +/- 5.2 kHz, and is ...
  38. [38]
    The hyperfine structure of the inversion-rotation transition JK = of ...
    The complex hyperfine structure (hfs) of the inversion-rotation transitions of ammonia is due to the interactions of the electric and magnetic multipole moments ...
  39. [39]
    Constraining P and T violating forces with chiral molecules
    Nov 7, 2024 · In this work, we show that precision spectroscopy of the hyperfine structure of chiral molecules is sensitive to new physics sources of parity ...Missing: 2020s | Show results with:2020s
  40. [40]
    Hyperfine excitation of CH and OH radicals by He
    The present paper focuses on the calculation of inelastic rate coefficients among the lowest OH/CH hyperfine levels in their ground vibrational state induced ...
  41. [41]
    Hyperfine structure in the and states of and | Phys. Rev. A
    Jun 19, 2015 · These Doppler-free spectra exhibit peaks in the transmission of the 780.03-nm beam passing through the cell—peaks that occur when Laser 2 scans ...
  42. [42]
    Determination of the hyperfine structure constants of the - 8 7 R b
    Nov 10, 2014 · The hyperfine anomaly between different isotopes of the same element, arising from the isotope differences in the distribution of nuclear charge ...
  43. [43]
    [PDF] High-precision measurement of hyperfine structure in the D lines of ...
    Aug 19, 2010 · Abstract. We have measured hyperfine structure in the first-excited P state (D lines) of all the naturally-occurring alkali atoms.<|control11|><|separator|>
  44. [44]
    [PDF] Hyperfine Transition Frequency
    SERIES of measurements involving hydrogen masers, cesium beam frequency standards, and the NBS clock system was made at the National. Bureau of Standards (NBS) ...Missing: splitting | Show results with:splitting<|control11|><|separator|>
  45. [45]
    [2401.09095] Signal in the Hyperfine Structure Line of the Ground ...
    Jan 17, 2024 · We analyze the formation of the signal in the 21 cm hydrogen line of the Dark Ages (30\le z\le300) in different cosmological models and discuss the possibility ...
  46. [46]
    Hyperfine Structure of in the State | Phys. Rev.
    The atomic-beam magnetic resonance method has been used to investigate the hyperfine structure of the $4P$ state of ${\mathrm{K}}^{39}$.
  47. [47]
    L i 7 - F 1 9 - by the Molecular Beam Electric Resonance Method
    The electric resonance method of molecular beam spectroscopy was used to observe the L i 7 F 1 9 spectra arising from transitions of the type ( J , m J ) ...
  48. [48]
    The Molecular Beam Electric Resonance Method Study of Thallium ...
    The hyperfine structure of the first rotational state of Tl C l 3 5 and Tl C l 3 7 was studied by the molecular beam electric resonance method.
  49. [49]
    Hyperfine Structure Constants of Sc i and Sc ii with Fourier ...
    Feb 28, 2022 · Abstract. We report new experimental hyperfine structure (HFS) constants of neutral and singly ionized scandium (Sc I and. Sc II).
  50. [50]
    [PDF] Measurement of the Hyperfine Structure and Isotope Shifts of the 3s
    The present work provides a measurement of the hyperfine structure of 29Si and improves the accuracy of the 28Si-30Si isotope shift by almost two orders of ...
  51. [51]
    Nuclear Magnetic Moment, Hyperfine Structure, and Hyperfine ...
    The atomic-beam magnetic-resonance technique was used to measure the free-atom ground-state hyperfine structure separation ...
  52. [52]
    A primer in pulse EPR-based hyperfine spectroscopy for NMR ...
    This primer is an attempt to bridge some of the gaps between the fields, and at the same time it highlights some applications of pulse EPR.
  53. [53]
    Advanced paramagnetic resonance spectroscopies of iron–sulfur ...
    This review describes the ENDOR and ESEEM techniques and then provides a series of case studies on their application to a wide variety of FeS proteins.
  54. [54]
    Electron spin resonance in emerging spin-driven applications
    Oct 13, 2023 · The electron-nuclear double resonance (ENDOR) technique can resolve the unresolved hyperfine splitting in conventional CW ESR spectra.
  55. [55]
    Ultracold atom interferometry in space | Nature Communications
    Feb 26, 2021 · Here we explore matter-wave fringes of multiple spinor components of a BEC released in free fall employing light-pulses to drive Bragg processes ...
  56. [56]
    Nanoscale quantum sensing with Nitrogen-Vacancy centers in ...
    In this review, we introduce the field of ODMR spectroscopy of NV-nanodiamonds and how it can be used to sense different quantities.
  57. [57]
    Ultrafast Nonadiabatic Electron Relaxation Dynamics in ...
    Feb 20, 2025 · The nonradiative electron-relaxation dynamics in a C 60 molecule is investigated after chosen initial photoexcitations.Missing: post- | Show results with:post-
  58. [58]
    second - BIPM
    It is defined by taking the fixed numerical value of the caesium frequency ΔνCs, the unperturbed ground-state hyperfine transition frequency of the caesium-133 ...Missing: 9192631770 source
  59. [59]
    [PDF] Frequency Standards Based on Atomic - Hydrogen
    Hydrogen masers have the best short-term stability due to long storage times and large atom numbers. Long-term stability is as good as primary standards, and ...
  60. [60]
    [PDF] Mise en pratique - metre - Appendix 2 - SI Brochure - BIPM
    May 20, 2019 · This definition implies that 'The metre is the length of the path travelled by light in vacuum during a time interval of 1/299 792 458 of a ...
  61. [61]
    [PDF] MEP 2003 IODINE (λ ≈ 633 nm) - BIPM
    Ne Laser Stabilized to an Iodine Hyperfine Component, Appl. Opt ... Stabilized Lasers -. Radiations Recommended for the Definition of the Meter, IEEE Trans.
  62. [62]
    [PDF] QED tests, nuclear structure and fundamental constants - arXiv
    The gross structure (∆n 6= 0) is well explained by the Bohr theory (so-called 'old quantum theory') and Schrödinger theory which also predicts the hyperfine ...
  63. [63]
    Proton-structure corrections to hyperfine splitting in muonic hydrogen
    Additionally, hyperfine splittings have an impact on the muonic hydrogen Lamb shift and proton radius measurement, however the correction we advocate has a ...
  64. [64]
    [PDF] Improved Frequency Ratio Measurement with 87Sr and 171Yb ...
    These achievements are driving the discussion of the redefinition of the second using optical clocks. By using two optical clocks with different atomic species,.Missing: structure | Show results with:structure
  65. [65]
    171Yb+ optical clock with systematic uncertainty and absolute ...
    Jun 17, 2024 · 171Yb+ optical clock with $2.2\times 10^{-18}$ 2.2 × 10 − 18 systematic uncertainty and absolute frequency measurements.
  66. [66]
    2024-02-20-redefinition-second - BIPM
    Feb 20, 2024 · The redefinition of the second, based on optical transitions, is being considered by the Consultative Committee for Time and Frequency (CCTF).Missing: lattice hyperfine 18 2025
  67. [67]
    Towards the optical second: verifying optical clocks at the SI limit
    To ensure continuity with the current definition, based on the microwave hyperfine transition in Cs 133 , it is necessary to measure the absolute frequency of ...<|control11|><|separator|>
  68. [68]
    [PDF] II Neutral Atomic Hydrogen (HI) Regions
    This section reviews the physics of the HI hyperfine transition and of HI 21-cm line formation in both absorption and emission. Hyperfine Splitting of HI: A ...
  69. [69]
    [PDF] NRAO ONLINE 54 HI (Hydrogen 21 cm Line)
    (The Observatory article had reported: “Van de Hulst has suggested the hyperfine-structure transition of the ground state of hydrogen, located near. 1410 MHz ...
  70. [70]
    [PDF] A molecular absorption line survey toward the AGN of Hydra-A
    molecular gas, produced using our ALMA observations of CO(2-1) emission. ... • H2CO (formaldehyde) is highly prevalent toward H II ... observed hyperfine structure ...
  71. [71]
    Deuterium Abundance in the Interstellar Gas of the Galactic ...
    We report a measurement of the deuterium-to-hydrogen (D/H) ratio from a 6 σ detection of the hyperfine ground-state transition of deuterium at 327 MHz (92 cm)Missing: BBN | Show results with:BBN
  72. [72]
    [PDF] Deuterium and the baryonic density of the universe
    The goal is to measure the primordial abundance ratios of the light nuclei made in BBN. We normally measure the ratios of the abundances of two nuclei in the ...
  73. [73]
    Collinear laser spectroscopy at ISOLDE: new methods and highlights
    Apr 19, 2017 · Nuclear spins are directly obtained from the relative hyperfine structure intervals provided the electron angular momentum is greater than 1.
  74. [74]
    [PDF] Measurement of nuclear moments and radii by collinear fast-beam ...
    Aug 5, 1991 · At ISOLDE-2 various laser spectroscopy programmes have yielded a wealth of infor. » 1 Introduction - Laser Spectroscopy at ISOLDE-2 ical ...
  75. [75]
    [PDF] Mössbauer Spectrometry
    The method is often useful because nuclear energy levels are altered by hyperfine interactions between the nucleus and its nearby electrons. With some ...
  76. [76]
    Measurement of the proton Zemach radius from the hyperfine ...
    The hyperfine splitting of muonic hydrogen derives the proton Zemach radius, which is defined as a convolution of the charge distribution with the magnetic ...
  77. [77]
    Bottom-up fabrication of scalable room-temperature diamond ...
    Jul 24, 2025 · An integrated diamond quantum device harnesses the collective properties of multiple NV centres, enabling room-temperature quantum computing and ...