Fact-checked by Grok 2 weeks ago

Lamb shift

The Lamb shift is a subtle energy splitting between the $2^2S_{1/2} and $2^2P_{1/2} states of the , both characterized by the same n=2 and total j=1/2, but differing in orbital (l=0 for S and l=1 for P). This shift, which deviates from the degeneracy predicted by the for relativistic , measures 1057.830(3) MHz, equivalent to an energy difference of approximately $4.374 \times 10^{-6} . It originates from quantum electrodynamic () effects, where the bound interacts with virtual photons in the quantum vacuum, leading to radiative corrections that slightly raise the of the S state relative to the P state. The shift was experimentally detected in 1947 by Willis E. Lamb Jr. and Robert C. Retherford at using a novel technique. They produced a beam of atoms excited to the long-lived metastable $2^2S_{1/2} state via electron bombardment, then applied a resonant field to induce transitions to the short-lived $2^2P_{1/2} state, observing the required frequency for stimulated absorption in zero magnetic field as 1057 MHz. This measurement contradicted the Dirac theory's expectation of exact degeneracy. Within weeks, at offered the first theoretical interpretation using a non-relativistic approximation in , attributing the shift to the electron's from emitting and reabsorbing virtual photons, with a logarithmic divergence cutoff at the Rydberg energy, yielding a predicted value of about 1040 MHz—remarkably close to experiment. The discovery of the Lamb shift catalyzed the post-World War II revival of , resolving infinities in through renormalization techniques developed by , , and Sin-Itiro Tomonaga, who shared the 1965 partly for related advancements. Lamb's work earned him the 1955 "for his discoveries concerning the spectrum of the ." Today, the Lamb shift serves as a precision testbed for , with theoretical predictions matching experimental values to relative uncertainties below $10^{-5} in frequency units, incorporating higher-order , , and recoil effects; recent measurements have also refined the . It has been measured in exotic atoms like muonic and even engineered .

Historical Background

Early Theoretical Predictions

In 1928, formulated a relativistic for the that successfully incorporated both and , providing a theoretical framework for the . The solutions to the for yield energy levels that depend solely on the principal n and the total angular momentum j = l \pm 1/2, where l is the orbital angular momentum quantum number. Consequently, the Dirac theory predicts exact degeneracy between the $2S_{1/2} (n=2, l=0, j=1/2) and $2P_{1/2} (n=2, l=1, j=1/2) states, with no energy splitting between them. The in the Dirac theory arises from the interplay of relativistic kinematic effects—such as the variation in with velocity—and the spin-orbit coupling, where the electron's spin magnetic moment interacts with the generated by the proton's in the electron's . These corrections lift the degeneracy between states of different j but the same n and l, such as splitting the $2P_{3/2} from the degenerate $2S_{1/2} and $2P_{1/2} pair, while preserving the latter's exact equality in energy. This framework accounted remarkably well for the observed fine structure splittings in atoms and heavier elements but left the n=2 levels' degeneracy intact. Early efforts to incorporate quantum electrodynamic effects beyond the began in the 1930s, focusing on radiative corrections to the potential. In 1935, Uehling calculated the influence of virtual electron-positron pairs on the vacuum, demonstrating that this induces a small deviation from the pure $1/r law at short distances, effectively screening the charge. Uehling's qualitative analysis suggested that such effects could perturb atomic energy levels, potentially lifting degeneracies like that between $2S_{1/2} and $2P_{1/2}, though the magnitude was estimated to be too small for direct observation at the time. By the late 1930s, high-resolution optical spectroscopy of hydrogen's Balmer lines, such as H\alpha, began to reveal subtle inconsistencies with the Dirac predictions. Measurements by Simon Pasternack in 1938 indicated a possible small separation in the components attributable to the $2S_{1/2} and $2P_{1/2} levels, hinting at deviations on the order of the unresolved linewidths. Similar observations by R.C. Williams using discharges further supported these hints, suggesting the $2S_{1/2} state lay slightly above the $2P_{1/2} as per Dirac theory, but with an anomalous shift not accounted for by relativistic corrections alone. These early spectroscopic indications, though not definitive due to experimental limitations, set the stage for precise measurements that would confirm the discrepancy.

Lamb–Retherford Experiment

The Lamb–Retherford experiment, performed in 1947 at shortly after , employed to probe the of the , specifically targeting the energy levels with n=2. The setup utilized wartime-developed microwave technology to excite and detect transitions in a controlled atomic beam, enabling precise measurement of small energy splittings that optical methods could not resolve due to . Molecular gas was thermally dissociated in a oven heated to around 2000 , producing a of atomic primarily in the $1S_{1/2} that emerged through a small . This passed through an excitation region where low-energy electrons (about 20 eV) bombarded the atoms, populating the metastable $2S_{1/2} via collisional while minimizing higher s. A state selector, consisting of slits and additional electron bombardment, further purified the to enrich the fraction of $2S_{1/2} atoms (reaching up to 1-2% of the ). The then entered a interaction region with a uniform (up to 100 gauss) for Zeeman resolution and a resonant tuned to approximately 1000 MHz, where a reflex served as the tunable microwave source to induce transitions to the short-lived $2P_{1/2} . Detection occurred at the end of the using a heated , which registered the arrival of $2S_{1/2} atoms through electron upon collision; successful transitions depleted the $2S_{1/2} population, reducing the current and producing a signal. The procedure involved sweeping the microwave frequency or the strength to map the resonance curve, exploiting the Zeeman splitting to isolate the m_F = \pm 1 sublevels of the F=1 hyperfine states for both $2S_{1/2} and $2P_{1/2}, where the transition frequency was independent of the field at low strengths. Resonances were identified by sharp dips in the detector signal, with linewidths limited to about 1 MHz by power broadening and transit-time effects. This method allowed direct comparison to the Dirac theory's prediction of degeneracy between the $2S_{1/2} and $2P_{1/2} levels. Key results revealed a shift of 1057.8 ± 0.1 MHz for the $2S_{1/2} to $2P_{1/2} , placing the $2S_{1/2} approximately 1000 MHz above the $2P_{1/2} —contradicting the Dirac prediction of zero splitting and confirming an anomalous separation on the order of the fine-structure scale. This ~0.03% deviation from Dirac expectations highlighted limitations in . Significant challenges included minimizing perturbations from stray electric fields, which cause the Stark effect and rapidly quench the $2S_{1/2} lifetime (from 1/8 second to microseconds above ~10 V/cm); this was addressed by shielding and grounding the apparatus to maintain gradients below 5 V/cm, verified through quenching measurements dependent on magnetic field strength. Ensuring beam purity required careful control of excitation conditions to suppress unwanted $2P$ population, while the atomic beam geometry inherently reduced Doppler broadening to negligible levels compared to gas discharge methods. These measures enabled the precision necessary to detect the subtle shift. The findings were published in August 1947 in Physical Review, marking a pivotal empirical advance in and earning Willis E. Lamb the 1955 for his contributions to the understanding of 's energy levels.

Theoretical Explanation

Bethe's Non-Relativistic Approach

Following the presentation of the Lamb–Retherford experiment at the Island Conference on in June 1947, Hans sought to provide an immediate theoretical interpretation of the observed energy shift between the 2S1/2 and 2P1/2 states in . Motivated by the discrepancy with Dirac's relativistic theory, Bethe developed a non-relativistic calculation during his train ride home from the conference to , marking a pivotal moment in the revival of (QED). Bethe's approach focused on the self-energy of the bound arising from its with the quantized radiation , treating the non-relativistically and neglecting spin-orbit effects. He modeled the shift as the difference between the self-energy of the in the atomic state and that of a , following Hendrik Kramers' earlier idea of to handle divergences. The leads to a second-order energy shift, but the integral over photon momenta diverges linearly at high frequencies; Bethe subtracted the free- contribution, reducing the divergence to a milder logarithmic form. To regularize this, he introduced an ad hoc cutoff at the 's , corresponding to photon energies up to the rest mass energy mc², arguing that relativistic effects become important beyond this scale. The resulting formula for the energy shift in the 2S of is \Delta E = \frac{4\alpha}{3} \frac{Z^3 e^2}{\hbar} |\psi(0)|^2 \ln \left( \frac{m c^2}{Ry} \right), where |\psi(0)|^2 = \frac{(Z \alpha m)^3}{\pi n^3} for S states (n=2, Z=1). This yields a numerical value of about 1040 MHz, in close agreement with the experimental measurement of 1058 MHz from the Lamb–Retherford experiment. Bethe interpreted the physical origin as the "shaking" of the by zero-point fluctuations of the , which jitter the 's and thus alter its relative to the Dirac prediction. This simple yet insightful calculation, published shortly thereafter, demonstrated the feasibility of handling QED infinities through mass renormalization and inspired the full relativistic treatments that followed.

Full QED Derivation

The full derivation of the emerged in the late 1940s as a relativistic extension of earlier non-relativistic calculations, spearheaded by , , and Sin-Itiro Tomonaga. Their work, spanning 1948 to 1949, reformulated using covariant and introduced to handle divergent integrals arising from interactions, enabling finite predictions for observable quantities like shifts in . This approach resolved the infinities plaguing pre-war QED formulations by redefining bare parameters such as and charge in terms of measurable quantities. In the Feynman diagram formalism of , the Lamb shift originates from order-α radiative corrections beyond the , primarily through three one-loop diagrams: the electron (where the electron emits and reabsorbs a , altering its propagation), the vertex correction (modifying the electron-photon interaction and contributing to the anomalous ), and the (screening the Coulomb potential due to virtual electron-positron pairs in the ). These diagrams collectively shift the 2S_{1/2} and 2P_{1/2} energy levels, with the providing the dominant logarithmic term and the others adding finite corrections. The shift is computed as the expectation value of the radiative in the non-relativistic bound-state wavefunctions, expanded in powers of the α. The leading-order energy shift, of order α^5 relative to the Rydberg energy, is expressed as \begin{equation} \Delta E = \frac{\alpha^3 (Z\alpha)^4 m}{4 \pi n^3} \left[ \ln \frac{m^2}{\langle E_{n,0} \rangle} + \frac{19}{30} \right], \end{equation} for S states (with analogous expression for P states; n=2, Z=1), where \langle E_{n,0} \rangle is the expectation value of the unperturbed energy. Higher-order contributions, including two-loop effects up to order α^6, refine this by adding terms like α^6 m c^2 / π (numerically ~10 MHz for hydrogen), ensuring the theoretical prediction matches experiment across multiple orders. Renormalization is central to the derivation, proceeding by isolating ultraviolet divergences in the Σ(p) and Π(k^2) integrals, then subtracting them via counterterms that adjust the bare mass m_0 = m (1 + δm) and charge e_0 = e (1 + δe), where δm and δe are infinite but cancel in physical amplitudes. This mass and charge renormalization preserves invariance, as verified by Ward identities linking vertex and self-energy corrections, yielding covariant, finite results independent of the regularization scheme (e.g., dimensional or Pauli-Villars). QED predictions for the Lamb shift in achieve a relative of 10^{-12}, validated against high-accuracy spectroscopic measurements that confirm the theoretical to this level.

Applications in Atomic Spectra

Shift in Hydrogen

In the , the Lamb shift appears as an energy splitting between the $2S_{1/2} and $2P_{1/2} states, which are predicted to be degenerate by the due to their identical principal quantum number n=2 and total j=1/2. This degeneracy is lifted by quantum electrodynamic effects, primarily raising the energy of the $2S_{1/2} state above that of the $2P_{1/2} state. The experimental value for this splitting, measured via , is 1057.845(9) MHz. This shift is observable in the of the 2p state and contributes to the precise determination of 's spectral lines, such as those in the , where it provides small but essential corrections on the order of MHz to the transition frequencies. The Lamb shift in arises mainly from three contributions within the full quantum electrodynamic framework: the , the dominant contribution at approximately 1086 MHz through radiative corrections to the 's mass and position; , contributing approximately -27 MHz via modifications to the potential from virtual electron-positron pairs; and transverse exchange, a effect that adds a smaller correction of roughly 0.3 MHz for the n=2 states. For the $2S_{1/2}-2P_{1/2} difference, higher-order terms refine the total to match experiment. These effects are computed using the Bethe logarithm, a logarithmic average of excitation energies that incorporates reduced-mass corrections to the , ensuring accuracy in the non-relativistic limit for the proton- system. The magnitude of the Lamb shift scales with the principal n as approximately $1/n^3, reflecting its origin in relativistic and radiative corrections proportional to \alpha^5 Z^4 / n^3, where \alpha is the and Z=1 for . This scaling influences higher-n levels, such as those in the (n \geq 3 to n=2), where the shift introduces sub-MHz adjustments to line positions and is crucial for extracting the R_\infty from spectroscopic data with parts-per-billion precision. In practice, omitting the Lamb shift would shift the inferred Rydberg value by several parts in $10^7, highlighting its role in despite being a perturbative effect.

Extensions to Other Atoms

In multi-electron atoms, screening effects from inner electrons reduce the felt by outer electrons, thereby diminishing the magnitude of the Lamb shift relative to hydrogen-like systems. This screening is particularly pronounced in metals, where many-body interactions modify the contributions to the radiative shift. In heavier elements such as mercury (Z = 80), relativistic enhancements arise due to high electron velocities in inner shells, amplifying the Lamb shift; for instance, the K-shell in mercury experiences a shift of approximately 38 Rydbergs. The Lamb shift scales with atomic number Z according to the adapted formula \Delta E \propto Z^4 \alpha^5 m c^2 \ln(Z^2 \alpha^{-2}) for hydrogen-like atoms, highlighting its sensitivity to charge and the logarithmic dependence on the \alpha. Isotopic variations introduce finite mass corrections through recoil effects, altering the and thus the shift by small but measurable amounts across isotopes. In muonic atoms, the muon's mass (207 times that of the ) contracts the orbits, yielding dramatically larger Lamb shifts—up to orders of magnitude greater than in electronic atoms—due to enhanced overlap with the . Specific measurements illustrate these extensions: in H-like helium (He⁺), the 2S–2P Lamb shift is approximately 14 GHz. In lithium, precision spectroscopy of the Li^{6++} ion reveals a shift of about 63 GHz, incorporating finite mass and screening adjustments. Alkali atoms like cesium and rubidium benefit from accurate Lamb shift evaluations in their valence states, enabling sub-hertz precision in optical atomic clocks where QED corrections calibrate transition frequencies to parts in $10^{18}. In highly charged ions, such as H-like or lithium-like tin, Lamb shift measurements probe in strong fields (Z\alpha \approx 0.7), achieving agreements at the 0.1% level and offering sensitivity to potential new through discrepancies in two-loop radiative corrections.

Physical Significance

Role in Quantum Electrodynamics

The discovery of the Lamb shift in 1947 provided a critical test for (), highlighting discrepancies in early Dirac theory predictions and prompting the resolution of infinite divergences through techniques. Hans Bethe's non-relativistic calculation, which introduced an energy cutoff to yield a finite shift, marked a pivotal step toward salvaging , influencing the development of covariant methods by Sin-Itiro Tomonaga, , and . Their work, validated by the Lamb shift, earned the 1965 for foundational contributions to . The Lamb shift serves as a cornerstone benchmark for QED's perturbative expansion and framework, demonstrating the theory's predictive power through agreement between calculated and observed values to high precision, such as the determined to about 6 from the shift. This validation underscores QED's status as the most accurately tested physical theory, with the shift's magnitude arising from higher-order radiative corrections. In broader physical context, the Lamb shift empirically confirms the physical reality of particle-antiparticle pairs and quantum fluctuations, as these phenomena generate the radiative corrections responsible for the energy splitting. This success bolsters the of by affirming as its electromagnetic component, where effects propagate to electroweak interactions. Modern high-precision Lamb shift measurements continue to probe for deviations that could signal physics beyond , such as contributions from axion-like particles that might modify the propagator and alter the shift; current limits from and muonic atom in the 2020s constrain such extensions to below detectable levels in standard QED frameworks. Additionally, the Lamb shift exemplifies QED's triumphs in educational contexts, frequently featured in textbooks as the archetypal demonstration of quantum field theory's predictive success.

Precision Tests and Modern Measurements

Since the 1970s, laser spectroscopy techniques have revolutionized precision measurements of the Lamb shift in , enabling relative precisions on the order of 10^{-6} through methods such as Doppler-free two-photon spectroscopy of the 1S-2S transition. These advancements allow direct optical frequency comparisons that isolate effects from fine-structure splittings, with early implementations achieving uncertainties below 1 kHz for the ground-state shift. Two-photon transitions, in particular, minimize and recoil effects, providing a robust platform for testing predictions at progressively higher accuracy. Key experiments have further refined these measurements using advanced trapping and cooling techniques. For instance, a 2019 direct optical measurement of the 2S-2P transition frequency in hydrogen yielded a Lamb shift value of 1057.8298(32) MHz, corresponding to an absolute uncertainty of 3.2 kHz and enabling a proton charge radius determination with 1.2% precision. Complementary efforts incorporate recoil corrections, often benchmarked using Penning traps to precisely measure nuclear masses and finite-size effects, which contribute at the level of a few kHz to the shift in light atoms like hydrogen. These trapped-ion approaches, while more commonly applied to muonic systems, have informed hydrogen spectroscopy by validating recoil models to uncertainties below 1 kHz. Theoretical discrepancies, particularly from hadronic vacuum polarization (HVP), have been addressed through calculations in the 2020s, reducing uncertainties in the leading-order HVP contribution to the Lamb shift by up to 50%. These updates shift the predicted 1S Lamb shift in by approximately -3.4 kHz, aligning experiment and more closely and resolving prior tensions at the 10 kHz level. Such computations, leveraging four-flavor , provide evaluations of non-perturbative QCD effects essential for sub-kHz precision. As of 2025, QED predictions for the Lamb shift agree with experimental values to a relative precision limited by experiment to about 10^{-6} for the direct 2S-2P shift, with higher precision achievable via 1S-2S transitions incorporating the 1S Lamb shift measured to 10^{-14} relative uncertainty. This concordance underscores 's validity across scales. Recent 2024 calculations of two-loop self-energy have further refined predictions for the 1S Lamb shift in , achieving two-fold improvement in accuracy. In contrast, muonic measurements revealed anomalies in the proton radius puzzle, where the 2010 Lamb shift implied r_p ≈ 0.841 fm, discrepant from values; however, 2020s spectroscopy has narrowed the discrepancy to r_p ≈ 0.833-0.841 fm through refined finite-volume corrections and effects, though some tension remains as of 2025. Looking ahead, next-generation atomic clocks based on optical transitions in hydrogen-like ions promise sub-Hz resolutions for Lamb shift tests, potentially probing beyond current limits. Additionally, experiments in strong fields, such as Lamb shift measurements in high-Z ions like , will extend validations to regimes where α Z ≈ 1, revealing nonlinear effects absent in .

References

  1. [1]
    Fine Structure of the Hydrogen Atom by a Microwave Method
    Fine Structure of the Hydrogen Atom by a Microwave Method. Willis E. Lamb, Jr. and Robert C. Retherford. Columbia Radiation Laboratory, Department of Physics ...
  2. [2]
    [hep-ph/0002158] Theory of Light Hydrogenlike Atoms - arXiv
    Abstract: The present status and recent developments in the theory of light hydrogenic atoms, electronic and muonic, are extensively reviewed.
  3. [3]
    The Electromagnetic Shift of Energy Levels | Phys. Rev.
    The Electromagnetic Shift of Energy Levels, HA Bethe Cornell University, Ithaca, New York, PDF Share, Phys. Rev. 72, 339 – Published 15 August, 1947.
  4. [4]
    Willis E. Lamb – Facts - NobelPrize.org
    Willis Eugene Lamb, Nobel Prize in Physics 1955. Born: 12 July 1913, Los Angeles, CA, USA. Died: 15 May 2008, Tucson, AZ, USA.
  5. [5]
    Invention of the Maser and Laser - Physics Magazine
    Jan 27, 2005 · In 1947 Willis Lamb and Robert Retherford, also of Columbia, used ... They sent a beam of excited ammonia molecules into a resonant cavity.
  6. [6]
    The Lamb Shift - HyperPhysics
    The Lamb shift is extremely small and difficult to measure as a splitting in the optical or UV spectral lines.
  7. [7]
    [PDF] Lamb Shift in Atomic Hydrogen
    Careful scrutiny of this figure raises the following issue: the Lamb shift is defined as the energy separation between the 2S1/2 and 2P1/2 states.
  8. [8]
    <i>Landmarks</i>—Lamb Shift Verifies New Quantum Concept
    Jul 27, 2012 · By plotting the critical magnetic field strength for a variety of microwave frequencies, Lamb and Retherford could determine the energy ...Missing: klystron | Show results with:klystron
  9. [9]
  10. [10]
    [PDF] Hans Bethe, Quantum Mechanics, and the Lamb Shift
    Bethe estimated this value as 1040 MHz in amazingly good accord with the Lamb-Retherford value [1]. Discussion. For a historical perspective on how Bethe zoomed ...Missing: ride | Show results with:ride
  11. [11]
  12. [12]
    Schwinger, Tomonaga, Feynman, and Dyson: the triumph of ...
    This chapter looks at Schwinger's method of canonical transformations, his covariant approach to QED, Sin-itiro Tomonaga's covariant formulation of quantum ...
  13. [13]
    Sin-Itiro Tomonaga – Nobel Lecture - NobelPrize.org
    In particular, little by little news arrived that the so-called Lamb-shift was discovered as a manifestation of the electromagnetic field reaction and that ...
  14. [14]
    Lamb Shift - an overview | ScienceDirect Topics
    The nonrelativistic Lamb shift for atomic hydrogen was first calculated by Bethe in 1947, following the experiments of Lamb and Retherford. Whereas Dirac's one- ...
  15. [15]
    [PDF] Derivation of the Lamb Shift using an Effective Field Theory - arXiv
    In this paper, we rederive the complete O(α5) corrections in the non-recoil limit1. (me/mP = 0) using NRQED, an effective field theory developed by Caswell and ...
  16. [16]
    A measurement of the atomic hydrogen Lamb shift and the proton ...
    Sep 6, 2019 · W. E. Lamb, R. C. Retherford, Fine structure of the hydrogen atom by a microwave method. Phys. Rev. 72, 241–243 (1947). GO TO REFERENCE.Missing: klystron | Show results with:klystron<|control11|><|separator|>
  17. [17]
    [PDF] Renormalization theory of quantum electrodynamics
    The radiative corrections to the energy of an electron in a. Coulomb field will produce a shift in the energy levels of hy- drogen-like atoms and modify the ...
  18. [18]
    Separated Oscillatory Field Measurement of the Lamb Shift in H, n=2*
    The final value for the Lamb shift is. S(H,n=2) = 1057.845(9) MHz. Figure 7 shows a plot of all the reported direct measurements of the Lamb shift in the n=2 ...
  19. [19]
    The Rydberg constant and proton size from atomic hydrogen - Science
    The remaining main systematic effects in our experiment are the first-order Doppler shift and apparent line shifts caused by quantum interference of neighboring ...Missing: series | Show results with:series
  20. [20]
    [PDF] Theory of the Lamb shift in hydrogen and light hydrogen-like ions
    The contribution due to the reduced mass in all formulas is summed up and tabulated separately as the relativistic reduced mass (RRM) correction. The ...<|control11|><|separator|>
  21. [21]
    Two-Loop Bethe-Logarithm Correction in Hydrogenlike Atoms
    The two-loop Bethe logarithm has been the largest unknown correction to the hydrogen Lamb shift to date. Together with the ongoing measurement of the proton ...
  22. [22]
    Nobel Prize in Physics 1965 - Presentation Speech - NobelPrize.org
    ... experimentally. This occurred, however, in 1947 when Lamb and his collaborator Retherford discovered that some energy levels of hydrogen which should ...
  23. [23]
    Physics - Quantum Milestones, 1947: Lamb Shift Verifies New ...
    Feb 24, 2025 · In 1947 Willis Lamb and Robert Retherford of Columbia University discovered an unexpected detail in the hydrogen spectrum, later called the Lamb ...
  24. [24]
    Shift to quantum electrodynamics | Nature Physics
    a theory that, in the hands of Feynman, Schwinger and others, would become the most robust and well tested in all of ...
  25. [25]
    History and Some Aspects of the Lamb Shift - MDPI
    Although measurements of the shift were attempted in the 1930s, it was not measured accurately until 1947 when Lamb and Retherford employed rf spectroscopy and ...
  26. [26]
    Axion-modified photon propagator, Coulomb potential, and Lamb shift
    Dec 5, 2018 · In connection, the plausible distortion of the Lamb-shift in hydrogenlike atoms is established and the scopes for searching axionlike particles ...
  27. [27]
    Precision Measurement of the Lamb Shift in Muonium | Phys. Rev. Lett.
    Using the scaling factor of 1 / n 3 as in Ref. [36] , we get a contribution of 5% for 4 S states. This is validated by keeping the magnitude of the 4 S ...
  28. [28]
    Precision measurement at the 1S Lamb shift in atomic hydrogen
    Of all the atomic transitions that can be studied with laser spectroscopy the 1S—2S transition in atomic hydrogen is one of the most fundamental.Missing: review | Show results with:review
  29. [29]
    Precision measurement of the hydrogen and deuterium 1 S ... - OSTI
    Jan 16, 1994 · We report on new precision measurements of the 1[ital S] ground state Lamb shift in hydrogen and deuterium, based on a direct frequency ...
  30. [30]
    (PDF) Precision Spectroscopy of Atomic Hydrogen - ResearchGate
    Aug 10, 2025 · We review advances in optical precision spectroscopy of atomic hydrogen achieved at Garching and Paris since the first symposium on the ...
  31. [31]
    Nuclear recoil corrections to the Lamb shift of hydrogen and light ...
    Jul 1, 2016 · Accurate calculations of the nuclear recoil effect to the Lamb shift of hydrogen-like atoms are presented. Numerical results are reported.Missing: Penning trap
  32. [32]
    Nuclear recoil corrections to the Lamb shift of hydrogen and light ...
    Jun 30, 2016 · In the present work, we will identify the higher-order fns recoil correction that is beyond the reduced-mass part (28) and parametrize it in ...
  33. [33]
    Hadronic vacuum polarization correction to atomic energy levels
    Oct 10, 2022 · The shift of atomic energy levels due to hadronic vacuum polarization is evaluated in a semiempirical way for hydrogenlike ions and for ...
  34. [34]
    Hadronic vacuum-polarization contribution to various QED ...
    Feb 8, 2021 · The result [19] of 1.89(6)\times 10^{-12} is in perfect agreement ... In Table 10, we present for the Lamb shift in hydrogen and muonic ...
  35. [35]
    Theory of the Lamb shift in hydrogen and light hydrogen-like ions
    Sep 3, 2018 · We review the present status of theoretical calculations of the Lamb shift in hydrogen and light hydrogen-like ions with the nuclear charge number up to Z = 5.Missing: NIST | Show results with:NIST
  36. [36]
    New Measurement Fits Another Piece in the Proton Radius Puzzle
    Nuclear physicists may have now solved the proton radius puzzle with a unique new measurement of the charge radius using a novel electron scattering technique.Missing: 2020s | Show results with:2020s
  37. [37]
    Measurement of the transition frequency from 2S 1/2 , F = 0 ... - Nature
    Nov 25, 2022 · The Lamb shift is defined as the difference between the centroid values of the 2S1/2 and 2P1/2 levels.
  38. [38]
    Stringent test of QED with hydrogen-like tin - PMC - PubMed Central
    So far, bound-state QED in heavy highly charged ions has been mostly tested by Lamb-shift measurements; the highest precisions were achieved in lithium-like ...