Fact-checked by Grok 2 weeks ago

Fine-structure constant

The fine-structure constant, denoted by the Greek letter α, is a fundamental dimensionless physical constant that quantifies the strength of the electromagnetic interaction between elementary charged particles, such as electrons and photons. It is defined as α = e² / (4πε₀ ħ c), where e is the elementary charge, ε₀ is the vacuum permittivity, ħ is the reduced Planck constant, and c is the speed of light in vacuum. The current recommended value, based on the 2022 CODATA adjustment, is α = 7.2973525643(11) × 10⁻³, or equivalently, its inverse 1/α ≈ 137.035999177(21), with a relative uncertainty of about 1.5 × 10⁻¹⁰. Introduced by Arnold Sommerfeld in 1916, the constant arose from his extension of the Bohr model of the hydrogen atom, incorporating relativistic effects to account for the observed fine splitting in atomic spectral lines. This "fine structure" refers to the small deviations from the energy levels predicted by the non-relativistic Bohr theory, which Sommerfeld explained by allowing elliptical orbits and considering the relativistic increase in electron mass. Over time, the fine-structure constant has become central to quantum electrodynamics (QED), the quantum field theory describing electromagnetic interactions, where it determines the magnitude of radiative corrections and governs phenomena like the Lamb shift and anomalous magnetic moments of particles. As one of the key coupling constants in the of , α highlights the fundamental scale of relative to other forces, such as and weak forces. Its dimensionless nature makes it a pure number without units, intriguing physicists because it appears ubiquitously in , molecular, and , influencing everything from the stability of to the of in materials. Although α is not predicted by theory and must be measured experimentally, its value runs with energy scale due to quantum effects like , increasing slightly at higher energies—for instance, α ≈ 1/128 near the Z boson mass. Precise measurements, often using techniques like the or g-2 experiments, continue to refine its value and test the consistency of predictions.

Definition and Fundamentals

Definition

The fine-structure constant, denoted by the symbol \alpha, is a fundamental dimensionless quantity in physics that quantifies the strength of the electromagnetic interaction between elementary charged particles. In SI units, it is precisely defined as \alpha = \frac{e^2}{4\pi \epsilon_0 \hbar c}, where e is the elementary electric charge, \epsilon_0 is the vacuum permittivity, \hbar is the reduced Planck's constant, and c is the speed of light in vacuum. This expression arises from combining the fundamental constants that govern electric charge, quantum mechanics, and relativity, providing a measure of how strongly charged particles couple to the electromagnetic field. The originates from the observation of in the spectral lines of atoms, particularly the closely spaced splittings in the atom's emission , which could not be explained by the non-relativistic alone. In 1916, introduced \alpha while extending the Bohr atomic model to include relativistic corrections, showing that the splitting in hydrogen's spectral lines is proportional to \alpha^2. These splittings represent small deviations in levels due to the electron's relativistic motion and spin-orbit coupling. Being dimensionless, \alpha is independent of any unit system, making it a universal constant that permeates diverse areas of physics without reliance on arbitrary scales. This property underscores its role as a pure number reflecting the intrinsic scale of electromagnetic interactions in nature. In the context of Paul Dirac's 1928 relativistic quantum mechanical treatment of the , the yields energy levels where the splitting factor is exactly \alpha^2 / n^3 (with n the principal quantum number), confirming Sommerfeld's earlier relativistic approximation and elevating \alpha to a of .

Numerical Value and Dimensionless Nature

The fine-structure constant, denoted by α, has the recommended of 7.2973525643(11) × 10^{-3}, or equivalently, its 1/α = 137.035999177(21), as determined by the 2022 CODATA adjustment of fundamental physical constants. This carries a relative standard uncertainty of approximately 1.5 × 10^{-10}, reflecting the high precision achieved through contemporary measurements. The constant is dimensionless, meaning it is a pure numerical independent of the choice of units, which ensures its invariance across different physical systems and scales. In the Gaussian cgs unit system, α is explicitly given by the expression \alpha = \frac{e^2}{\hbar c}, where e is the (in statcoulombs), ℏ is the reduced Planck's constant, and c is the . This formulation highlights its role as the coupling strength of , scaled by fundamental constants. The absence of any theoretical framework predicting the exact numerical value of α remains one of the unresolved mysteries in fundamental physics, with no derivation available from the or beyond. The dimensionless nature of α has profound implications for . It remains unchanged under unitary transformations or rescalings of physical quantities, facilitating comparisons across disparate energy scales. In , where ℏ = c = 1, the expression simplifies such that the square of the is e² = α, underscoring its direct interpretation as the strength of electromagnetic interactions without additional dimensional factors. This property positions α as a fundamental parameter that permeates diverse areas of physics, from spectra to particle interactions, always manifesting as a fixed numerical factor.

Measurement and Determination

Historical Measurements

The fine-structure constant was first estimated in 1916 by through his analysis of the observed in the lines of atoms, particularly the splitting in the . Using a relativistic modification to the Bohr atomic model, Sommerfeld derived the constant as a dimensionless parameter governing the strength of electromagnetic interactions at atomic scales, yielding an approximate value of \alpha \approx 1/137. This estimate was based on precise measurements of separations, such as the H\alpha line, which aligned closely with theoretical predictions from the relativistic fine-structure formula. A significant contribution to the numerical determination of \alpha came from Robert A. Millikan's oil-drop experiment in 1917, which provided one of the earliest accurate measurements of the elementary charge e = (4.774 \pm 0.002) \times 10^{-10} esu. Since \alpha = e^2 / (\hbar c) in Gaussian units, Millikan's value of e, combined with contemporaneous determinations of the speed of light c (from interferometric methods) and the reduced Planck's constant \hbar (derived from Planck's constant h via blackbody radiation or photoelectric effect data), enabled the first quantitative computations of \alpha from fundamental constants. These early \hbar and c values, however, carried uncertainties of several percent, limiting the precision of \alpha to about four significant figures. By the mid-20th century, particularly in the , refinements in —such as high-resolution studies of and fine-structure splittings—and precision for c improved the accuracy of \alpha. These efforts yielded values around \alpha^{-1} \approx 137.04, marking a deviation from the 137 conjectured in some theoretical models and highlighting the constant's empirical nature. Challenges in achieving higher precision persisted due to incomplete knowledge of constants, including variations in h measurements and the nascent understanding of quantum electrodynamic corrections to atomic spectra.

Modern Measurements and Precision

Since the 1980s, the and Josephson junctions have provided foundational high-precision measurements of the e and Planck's constant \hbar, enabling indirect determinations of the fine-structure constant \alpha = e^2 / (4\pi \epsilon_0 \hbar c) through metrological standards. These techniques achieved relative precisions on the order of $10^{-9} for \alpha by the late , establishing a for linking electrical units to fundamental constants before the 2019 SI redefinition fixed e and \hbar. Advancements in atomic recoil experiments using laser-cooled atoms have significantly improved precision in the 21st century. By measuring the recoil velocity of atoms from photon absorption, such as in or cesium, the ratio h/m (where m is the ) is determined, which combines with spectroscopic data to yield \alpha. A notable 2008 experiment at LKB (Laboratoire Kastler Brossel) using Bloch oscillations in an accelerated optical lattice for ^{87}Rb atoms reported \alpha^{-1} = 137.03599945(62), corresponding to a relative of approximately $4.5 \times 10^{-9}. Subsequent refinements, including NIST contributions to , have pushed uncertainties to around $10^{-10}, making recoil methods competitive with other approaches. Additionally, 2024 electron g-factor measurements using Penning traps have refined \alpha via the anomalous magnetic moment a_e = (g-2)/2, with an updated value from Harvard improving agreement with predictions while highlighting minor tensions (around 1-2\sigma) with some higher-order theoretical terms. The current CODATA 2022 recommended value is \alpha = 7.2973525643(11) \times 10^{-3}, or \alpha^{-1} = 137.035999177(21), with a relative of 1.6 \times 10^{-10}, dominated by inputs from g-factor and data. This precision reveals subtle discrepancies with certain theoretical predictions in , such as those involving hadronic contributions, motivating ongoing refinements.

Physical Interpretations

Role in Atomic and Molecular Physics

In , the fine-structure constant α governs the magnitude of relativistic corrections that cause the splitting of spectral lines in hydrogen-like atoms, known as . These corrections arise from the relativistic nature of the electron's motion and spin-orbit coupling, as captured by the . The energy shift ΔE for a given level n,j is ΔE = E_n \frac{\alpha^2}{n^2} \left( \frac{n}{j + \frac{1}{2}} - \frac{3}{4} \right), where E_n is the non-relativistic Bohr energy and j is the . This shift is on the order of α² ≈ 5 × 10^{-5} relative to the gross structure levels, making it a small but observable that refines the Bohr model's predictions. Beyond the Dirac-derived , quantum electrodynamic () effects introduce additional shifts, such as the , which further splits the 2S_{1/2} and 2P_{1/2} levels in by about 1057 MHz. While the originates from and diagrams in , its magnitude is ordered by powers of α, scaling roughly as α³ times the fine-structure scale, emphasizing α's role in quantifying the relative strength of electromagnetic interactions at atomic scales. This correction, though smaller than the Dirac fine structure by a factor of about α, is crucial for matching experimental spectra precisely. The fine-structure constant also influences applications like the , where an external splits atomic levels; in the presence of , this results in the anomalous Zeeman effect, with splitting patterns depending on the that incorporates j and thus the α-induced fine-structure intervals. For instance, in atoms, the modulates Zeeman sublevels, altering selection rules and polarization in spectral lines. Similarly, —arising from electron-nuclear spin interactions—scales with α⁴ times the ratio m_e/m_p, as the interaction energy involves the Fermi contact term proportional to the at the , which itself depends on α³ from the . In , this yields the 21 cm hyperfine transition, with splitting ΔE_hf ∝ (8/3) g_p α⁴ (m_e/m_p) Ry, where Ry is the Rydberg energy. In broader atomic contexts, determines the scale of ionization potentials through the , Ry ≈ (1/2) α² m_e c², which sets the for ic ions and influences multi-electron atoms via screening. For elements like and , relativistic fine-structure effects subtly affect chemical bonding; increasing α would weaken covalent bonds by enhancing spin-orbit coupling, potentially altering dissociation energies in H₂ by up to a few percent for α variations of order 1%. These influences highlight α's foundational role in ordering electromagnetic effects from atomic spectra to molecular in systems.

Significance in Quantum Electrodynamics

In (QED), the fine-structure constant α acts as the fundamental coupling parameter that governs the strength of electromagnetic interactions between charged particles, such as electrons and . Developed through the work of , , Sin-Itiro Tomonaga, and in the late 1940s, perturbative QED relies on an expansion series in powers of α (or more precisely α/π for loop contributions), which converges effectively due to α ≈ 1/137 being much less than 1 at low energies. This expansion is visualized through Feynman diagrams, where each vertex representing a emission or absorption is scaled by the e, related to α by e = √(4πα) in ; higher-order diagrams incorporate additional loops and vertices, allowing precise calculations of scattering amplitudes and decay rates. A key success of this perturbative framework is the prediction of the electron's anomalous , a_e = (g_e - 2)/2, where g_e is the electron's . In the Dirac theory, g_e = 2 exactly, but corrections yield a_e = α/(2π) at leading order, as first computed by Schwinger using proper-time methods in the . Higher-order terms include contributions like α²/(π²) and beyond, with the full prediction now known to over five loops, matching experimental measurements to 10 decimal places and serving as one of the most stringent tests of the theory; for instance, the tenth-order term contributes only about 10^{-13} to a_e. Despite its successes at accessible energies, exhibits nontrivial behavior at high scales due to the positive sign of its beta function, leading to a where α diverges. This theoretical singularity occurs around 10^{37} GeV in the context of the full , far beyond the electroweak scale, signaling that pure lacks and requires embedding in a larger theory for UV completion; the running of α from effects increases its value logarithmically with energy. At the electroweak scale (around the Z boson mass of 91 GeV), α(M_Z) ≈ 1/128, a value crucial for unification discussions, as it approaches the weak and strong couplings in grand unified theories, though discrepancies persist without new physics.

Dependencies and Variations

Variation with Energy Scale

In quantum electrodynamics (QED), the fine-structure constant \alpha exhibits a scale dependence known as "running," arising from vacuum polarization effects where virtual fermion-antiparticle pairs screen the bare electromagnetic charge. This behavior is captured by the renormalization group equation (RGE) at one-loop order: \frac{d\alpha}{d \ln \mu} = \frac{\alpha^2}{3\pi} \sum_f Q_f^2 n_c, where \mu is the renormalization scale, the sum runs over active fermion flavors f with electric charge Q_f (in units of the elementary charge e), and n_c = 1 (3) for leptons (quarks) accounting for color degrees of freedom. The leading-logarithmic solution to this equation, valid away from mass thresholds, takes the approximate form \alpha(\mu) = \frac{\alpha(0)}{1 - \frac{\alpha(0)}{3\pi} \left( \sum_f Q_f^2 n_c \right) \ln\left( \frac{\mu^2}{m^2} \right)}, with \alpha(0) \approx 1/137.036 the low-energy value (at \mu \sim m_e, the electron mass) and m a reference scale below the lightest threshold. The positive beta function coefficient in QED leads to an increase in \alpha(\mu) with \mu, reflecting reduced screening at shorter distances. At the electroweak scale \mu = M_Z \approx 91 GeV, \alpha(M_Z) \approx 1/128.9. Below M_Z, lepton loop contributions dominate the running, with the three charged lepton generations (electron, muon, tau; each Q_f = 1, n_c = 1) providing \sum Q_f^2 n_c = 3, augmented by lighter quark effects through hadronic vacuum polarization. Above M_Z, heavier particles enter, including the top (Q_f = 2/3, n_c = 3) whose large mass enhances its impact near the , and electroweak bosons (, ). At even higher scales, (QCD) influences the evolution via gluon-mediated corrections to loops, slowing the hadronic contribution's growth. These multi-loop and effects are incorporated in higher-order calculations for precision. The predicted running has been experimentally verified through e^+ e^- scattering cross-sections at the Large Electron-Positron Collider (LEP), where the effective \alpha at the Z-pole resonance was extracted from the hadronic and leptonic event rates, yielding \alpha(M_Z)^{-1} = 128.89 \pm 0.07 in agreement with . Complementary confirmation arises from global electroweak precision fits to LEP/SLD , which constrain \alpha(M_Z) to better than 0.1% accuracy and test consistency within the . Lower-energy running, including hadronic effects, has been measured in processes like e^+ e^- \to \mu^+ \mu^- \gamma by experiments such as KLOE and .

Temporal and Spatial Variations

The fine-structure constant has been scrutinized for possible temporal variations over cosmological timescales using absorption spectra, which probe distant intervening gas clouds. Early analyses by Webb et al. in suggested a potential change of Δα/α ≈ (-0.72 ± 0.18) × 10^{-5} at redshifts z ≈ 0.5–1.6, corresponding to roughly 10 billion years ago. Subsequent studies expanded the dataset, with claims of a dipole-like pattern implying variations up to ~10^{-5} over similar epochs. However, comprehensive reanalyses, including a 2024 study incorporating convergence properties of measurements from multiple systems, indicate no statistically significant deviation, with Δα/α consistent with zero at the 10^{-6} level or better. Tighter constraints on past variations come from natural nuclear reactors and primordial . The Oklo natural fission reactor in , operating approximately 2 billion years ago, provides isotopic ratios sensitive to α; analyses yield |Δα/α| < 5.8 × 10^{-8} over that interval. Big Bang (BBN) models, calibrated against observed primordial abundances of light elements like deuterium and helium, impose even stricter limits from the early universe at z ≈ 10^9; recent evaluations constrain |Δα/α| < 10^{-3} relative to the present value, though some specialized BBN studies suggest bounds as tight as <10^{-10} when incorporating correlated variations in other constants. At present epochs, laboratory tests using atomic clocks and lunar laser ranging set the most stringent limits on the rate of temporal change. Comparisons of optical atomic clocks, such as those based on ytterbium and dysprosium transitions, over intervals of years detect no variation exceeding |dα/α/dt| < 10^{-18} yr^{-1}. Lunar laser ranging experiments, monitoring Earth-Moon dynamics, complement these by constraining coupled variations in fundamental constants, yielding no detectable drift beyond |dα/α/dt| < 10^{-17} yr^{-1} up to 2025. These results affirm the constancy of α on human timescales. Spatial variations in α have been proposed based on directional dependencies in quasar data. Webb et al. in 2010 reported evidence for a dipole anisotropy, with Δα/α ≈ 10^{-6} aligned toward the southern sky, using Keck and VLT observations of ~300 absorption systems. This claim suggested possible spatial inhomogeneity at the 4σ level. However, independent analyses, including 2023–2024 studies leveraging large galaxy samples and DESI survey data, refute the dipole, finding no confirmed anisotropy and constraining spatial fluctuations to <10^{-6} across the observable universe.

Historical Development

Discovery and Early Interpretations

The fine-structure constant, denoted as \alpha, was first introduced by in 1916 as part of his extension of the within the framework of . sought to explain the observed fine splitting of spectral lines, such as the doublet in the hydrogen , which could not be accounted for by 's original 1913 model. By incorporating relativistic corrections to the electron's orbital motion, Sommerfeld derived that the splitting arises from the ratio of the electron's velocity v in the innermost to the speed of light c, yielding v/c \approx \alpha \approx 1/137. This dimensionless parameter encapsulated the strength of the electromagnetic interaction relative to relativistic effects, marking \alpha as a fundamental constant in early . In 1928, Paul Dirac advanced this understanding through his relativistic wave equation for the electron, which precisely predicted the fine structure of hydrogen spectral lines without ad hoc assumptions. Dirac's equation incorporated both quantum mechanics and special relativity, naturally yielding the fine-structure splitting as a function of \alpha, where the constant appeared as the coupling parameter governing electromagnetic interactions. This formulation elevated \alpha from an empirical correction in the to a core element of relativistic quantum theory, confirming Sommerfeld's approximate value while providing an exact theoretical basis. Prior to the 1940s, the fine-structure constant was primarily interpreted in the context of the as the characteristic ratio of the electron's orbital velocity to the speed of light in the ground state. This view framed \alpha as a measure of how closely atomic electrons approach relativistic speeds, influencing the quantization of angular momentum and energy levels in hydrogen-like atoms. Early calculations, such as those for the hydrogen fine structure, relied on this velocity ratio to match experimental spectral data, underscoring \alpha's role in bridging classical orbits with quantum discreteness. The numerical value of \alpha \approx 1/137 also sparked early numerological interest among physicists. In the 1930s, Arthur Eddington speculated that the inverse fine-structure constant was exactly 136, deriving this from his fundamental theory of the world, which emphasized epistemological and aesthetic principles over empirical measurement. Eddington's arguments, detailed in works like his 1931 paper, posited that \alpha^{-1} must be an integer tied to the structure of physical laws, reflecting a broader fascination with the constant's seemingly arbitrary yet precise magnitude. These interpretations, while influential, remained speculative and were later refined by more rigorous quantum developments.

Evolution Through Quantum Theory

In the 1940s, the fine-structure constant α began to play a central role in through calculations addressing discrepancies in atomic spectra. Hans Bethe calculated the in hydrogen in 1947, attributing the energy level splitting between the 2S_{1/2} and 2P_{1/2} states to radiative corrections involving virtual electron-positron pairs, with the shift scaling as α (up to logarithms) times the Dirac fine structure. This non-relativistic approximation incorporated α as the fundamental measure of electromagnetic coupling strength, resolving infinities via mass renormalization inspired by earlier work. Enrico Fermi contributed foundational QED techniques in the early 1940s, including self-energy evaluations that Bethe adapted for the Lamb shift computation during a train journey following the experimental announcement. The formalization of QED in the late 1940s elevated α to the status of the theory's primary expansion parameter. Sin-Itiro Tomonaga developed a covariant perturbation theory in 1946, enabling consistent handling of relativistic effects in electron-photon interactions. Julian Schwinger and Richard Feynman independently advanced this framework in 1948, introducing functional integrals and path-integral formulations, respectively, where scattering amplitudes expand in powers of α ≈ 1/137, quantifying loop corrections like vacuum polarization. Their work, unified by Freeman Dyson's equivalence proofs, demonstrated QED's predictive power for phenomena such as the anomalous magnetic moment, earning Tomonaga, Schwinger, and Feynman the 1965 Nobel Prize in Physics; α emerged as the dimensionless coupling dictating the theory's perturbative validity up to high energies. From the 1970s to the 2000s, efforts to embed α within broader unification schemes extended its theoretical significance in the Standard Model. Grand Unified Theories (GUTs), starting with the minimal SU(5) model proposed by Howard Georgi and Sheldon Glashow in 1974, aimed to unify the electromagnetic, weak, and strong couplings at a high-energy scale, but the minimal non-supersymmetric version predicted convergence around 10^{12}-10^{14} GeV with α_GUT ≈ 1/25, which was too low to evade proton decay limits. Supersymmetric extensions in the 1980s and 1990s raised the scale to around 10^{16} GeV, implying α evolves via renormalization group running to a unified value α_GUT ≈ 1/25 at that scale, though distinct from the Planck scale of 10^{19} GeV. Subsequent models like SO(10) refined these predictions, incorporating proton decay and neutrino masses, but discrepancies in low-energy coupling measurements—such as α_s at the Z boson mass—challenged minimal GUTs, prompting further supersymmetric adjustments that aligned the unification scale more closely with observations. These frameworks highlighted α's role in testing unification, with logarithmic running governed by β-functions involving particle content, though no exact derivation of its low-energy value emerged.

Theoretical Explanations

Anthropic Principle

The fine-structure constant, denoted as \alpha \approx 1/137, is often cited in fine-tuning arguments within the anthropic principle, which posits that the universe's physical parameters must permit the existence of observers to make such observations possible. Specifically, values of \alpha significantly larger than its observed magnitude would destabilize atomic structures essential for chemistry; for instance, if \alpha > 1/95, iron atoms become unstable, disrupting the formation of complex molecules necessary for life. Conversely, if \alpha < 1/205, would favor nickel over iron as the endpoint of fusion processes, preventing the production of elements critical for stable stars and planetary systems. These narrow bounds highlight how \alpha's value enables the electromagnetic interactions required for stable atoms and molecular bonds, underscoring the apparent tuning for . John D. Barrow and Frank J. Tipler, in their seminal 1986 work, formalized the anthropic cosmological principle, applying it to constants like \alpha to argue that the universe's structure is constrained by the necessity of supporting intelligent life. They distinguish between the weak anthropic principle (observational selection effects) and the strong version (the universe must evolve observers), using \alpha's role in atomic stability to illustrate how deviations would preclude biological complexity, such as the dipole moments in water molecules vital for life's solvent properties. This framework posits that \alpha's specific value is not coincidental but required for the emergence of observers capable of measuring it. In scenarios, particularly the , \alpha can vary across different vacua, with our universe's value selected anthropically because it allows for . The landscape encompasses approximately $10^{500} possible configurations of , where the vacuum expectation values of scalar fields determine coupling constants like \alpha, enabling regions hospitable to observers while others remain barren. This resolves by invoking a vast ensemble where life-bearing universes are statistically inevitable, though only those permitting complex chemistry and are observed. Critics of the approach to \alpha emphasize its limited testability, as predictions rely on unobservable multiverses, rendering it philosophically intriguing but empirically challenging; however, analogies to the —another finely tuned parameter with explanations in similar landscapes—bolster its plausibility by demonstrating consistent explanatory patterns across constants. Efforts to test it involve assessing whether life-permitting values occupy a significant of the parameter space, but current models suggest \alpha's tuning remains a key example of in cosmic ensembles.

Numerological and Speculative Theories

In the early , proposed a combinatorial derivation for the fine-structure constant \alpha, initially aiming to establish its inverse as exactly 136 based on the number of in fundamental particles and cosmological considerations. This approach, outlined in his work, relied on aesthetic and numerical principles rather than empirical derivation, predicting \alpha = 1/136 through relations involving the Eddington number and properties. However, subsequent measurements indicated a value closer to $1/137, prompting Eddington to adjust his combinatorial scheme to fit the new data, highlighting the speculative nature of the method. Religious and esoteric interpretations have also linked the approximate value of $1/137 to numerological significance, particularly in Kabbalistic traditions where the Hebrew word for "Kabbalah" (קַבָּלָה) yields a gematria value of 137, suggesting a mystical connection to the constant's role in governing electromagnetic interactions. Proponents argue this alignment reflects a deeper, non-physical harmony between ancient mysticism and modern physics, with 137 symbolizing reception or parallel structures in creation. Such views, while intriguing, remain outside scientific discourse and lack testable predictions. Contemporary speculative theories continue this numerological tradition, proposing derivations of \alpha \approx 1/137 through geometric or mathematical constructs. For instance, models incorporating the \phi \approx 1.618 relate it to \alpha via scaling laws and geometries. Similarly, recent frameworks explore ratios involving \pi and the Euler-Mascheroni constant \gamma \approx 0.577 to resolve \alpha within extended contexts, though these remain unverified. Classical electron models, reviving pre-quantum ideas, derive \alpha \approx 1/137 from Coulomb-to-gravitational force ratios or radius considerations, linking the 137 to primordial charge dynamics. These numerological and speculative approaches face significant critiques for their lack of and reliance on post-hoc adjustments to match measurements, often prioritizing over falsifiable mechanisms. Unlike theories, they fail to integrate with or yield novel experimental tests, rendering them philosophically appealing but scientifically marginal.

References

  1. [1]
    Current advances: The fine-structure constant
    ... historical portions given here remain ... Our view of the fine-structure constant has changed markedly since Sommerfeld introduced it over 80 years ago.
  2. [2]
    fine-structure constant - CODATA Value
    Concise form, 7.297 352 5643(11) x 10 ; Click here for correlation coefficient of this constant with other constants ; Source: 2022 CODATA
  3. [3]
    [PDF] 1. Physical Constants - Particle Data Group
    fine-structure constant α = e2/4π 0hc. 7.297 352 5693(11)×10−3 = 1/137 ... aUpdated values reflecting the “2022 CODATA recommended values”, released ...
  4. [4]
    [PDF] emunits.pdf - University of California, Berkeley
    In SI units, the MKS (meter-kilogram-second) system is used, while the. Gaussian units use the cgs (centimeter-gram-second) system. ... Fine structure constant.
  5. [5]
  6. [6]
    From Sommerfeld to Dirac: quantum theory and relativity
    The numerical value of the fine-structure constant is today known with great accuracy: α−1=137.035999177±0.000000021. It is definitely not a whole number ...
  7. [7]
    Introduction to the constants for nonexperts 19001920
    Millikan's final value, reported in 1917, was: (4.774 ± 0.002) x 10-10 esu (esu being the electrostatic unit, one of the units of charge in the centimeter-gram- ...
  8. [8]
    Introduction to the constants for nonexperts 19201940
    The classic pre-World War II measurements of the constant c are associated with Albert A. Michelson, a physicist in the United States.
  9. [9]
    [PDF] A note on the fine structure constant - UNL Digital Commons
    Apr 30, 2020 · By the 1940s, however, experimental values for α−1 deviated sufficiently from 137 to refute Eddington's argument [7]. Richard Feynman, one ...
  10. [10]
    Introduction to the constants for nonexperts 19401960
    The fine structure of atomic hydrogen was first measured with high accuracy by a United States physicist, Willis E. Lamb, Jr., and coworkers and reported in a ...
  11. [11]
    (PDF) The application of the Josephson and quantum Hall effects in ...
    Jan 4, 2016 · ... measurement of the response of Josephson junctions ... Prange, Quantized Hall resistance and the measurement of the fine structure constant,.
  12. [12]
    CODATA recommended values of the fundamental physical constants
    Jun 30, 2021 · Of the adjusted constants, the fine-structure constant has the second smallest relative uncertainty. Currently, the two types of measurements ...
  13. [13]
    [PDF] arXiv:0810.3152v1 [physics.atom-ph] 17 Oct 2008
    Oct 17, 2008 · The deduced value of h/mRb leads to a new determination of the fine structure constant α−1 = 137.03599945(62) with a relative uncertainty of 4.6 ...
  14. [14]
    [PDF] Lattice Interferometer for Laser-Cooled Atoms
    Aug 11, 2009 · other well-known constants constitutes a measurement of the fine structure constant [5,7,12]. The inset of Fig. 2 shows the analytical ...
  15. [15]
    Fine structure constant measurements in quasar absorption systems
    Oct 2, 2023 · This paper uses quasar absorption systems to measure the fine structure constant, finding hints of spatial variations, which could indicate a ...Missing: 2024 2025
  16. [16]
    Nuclear clock technology enables unprecedented investigation of ...
    Oct 27, 2025 · Nuclear clock technology enables unprecedented investigation of fine-structure constant stability · The strength of the electromagnetic force.
  17. [17]
    An improved measurement of the electron magnetic moment
    Apr 11, 2024 · Comparison of the measured g-factor and the predicted g-factor using an independent measurement of the fine structure constant yields the most ...
  18. [18]
    Measurement of the fine-structure constant as a test of the Standard ...
    Apr 13, 2018 · The fine-structure constant, α, is a dimensionless constant that characterizes the strength of the electromagnetic interaction between charged elementary ...<|control11|><|separator|>
  19. [19]
    [PDF] Quantum Physics III Chapter 2: Hydrogen Fine Structure
    Apr 2, 2022 · THE DIRAC EQUATION. 33. This equation is a bit reminiscent of the equation L ×. L. = ihL, for angular momentum. Back in the Pauli Hamiltonian ...<|control11|><|separator|>
  20. [20]
    [PDF] Fine Structure in Hydrogen and Alkali Atoms
    In these notes we consider the fine structure of hydrogen-like and alkali atoms, which concerns the effects of relativity and spin on the dynamics of the ...
  21. [21]
    Fine Structure - Richard Fitzpatrick
    ... Dirac equation--see Chapter 11). Let us now apply perturbation theory to a hydrogen-like atom, using ... is the fine structure constant. Actually, Equations ...
  22. [22]
    [PDF] The Zeeman Effect in Hydrogen and Alkali Atoms
    The fine structure energy shifts at zero field strength are. −(5/128)α2 for the 2s1/2 and 2p1/2 levels, where α is the fine structure constant, and −(1/128)α2 ...
  23. [23]
    Hyperfine Structure - Richard Fitzpatrick
    $$ \alpha=1/137$ the fine structure constant. Note that the hyperfine energy-shift is much smaller, by a factor $ m_e/m_p$ , than a typical fine structure ...
  24. [24]
    The discovery of asymptotic freedom and the emergence of QCD
    Landau and collaborators, in the late 1950s, studied the high-energy behavior of QED. They explored the relation between the physical electric charge and ...
  25. [25]
    [PDF] 10. Electroweak Model and Constraints on New Physics
    May 31, 2024 · The electroweak model (SM) is based on SU(2)xU(1) gauge group, with gauge bosons Wµ and Bµ, and a Higgs doublet for mass generation.
  26. [26]
    [PDF] Running electromagnetic coupling constant: low energy ... - arXiv
    Nov 9, 2000 · . The total finite renormalization between the fine structure constant and the MS- scheme EM coupling at Mτ is given by. ∆(4)(M2 τ )=∆lept ...
  27. [27]
    The running fine structure constant alpha(E) via the Adler function
    Jul 26, 2008 · We present an up-to-date analysis for a precise determination of the effective fine structure constant and discuss the prospects for future improvements.Missing: QED | Show results with:QED
  28. [28]
    Measurement of the running of the fine-structure constant
    Mar 9, 2000 · This article interprets measurements of Bhabha scattering by the L3 experiment at LEP in terms of the evolution of α(Q 2 ), using two complementary ...
  29. [29]
    A Search for Time Variation of the Fine Structure Constant - arXiv
    Dec 16, 1998 · A method offering an order of magnitude sensitivity gain is described for using quasar spectra to investigate possible time or space variation in the fine ...
  30. [30]
    Convergence properties of fine structure constant measurements ...
    Feb 19, 2024 · Our recent work has focused on measurements of the fine structure constant α in quasar absorption systems, parametrized using Δα/α = (αz − α0)/α ...ABSTRACT · INTRODUCTION · ASTRONOMICAL, ATOMIC... · SUMMARY AND...
  31. [31]
    Natural nuclear reactor at Oklo and variation of fundamental constants
    Dec 14, 2006 · This allows us to obtain the most accurate limits on the change of the fine structure constant in the past. The paper is organized as follows.
  32. [32]
    Constraints on the variation of the fine structure constant from big ...
    Jul 13, 1999 · We put bounds on the variation of the value of the fine structure constant $\ensuremath{\alpha}$, at the time of big bang nucleosynthesis.
  33. [33]
    A limit on variations in the fine-structure constant from spectra of ...
    Nov 10, 2022 · The fine structure constant α sets the strength of the electromagnetic force. The Standard Model of particle physics provides no explanation ...
  34. [34]
    New Limits on Variation of the Fine-Structure Constant Using Atomic ...
    Aug 6, 2013 · We report on the spectroscopy of radio-frequency transitions between nearly degenerate, opposite-parity excited states in atomic dysprosium (Dy).
  35. [35]
    Indications of a spatial variation of the fine structure constant - arXiv
    Aug 23, 2010 · View a PDF of the paper titled Indications of a spatial variation of the fine structure constant, by J. K. Webb and 5 other authors. View PDF.Missing: refuted 2023
  36. [36]
    Constraints on the Spacetime Variation of the Fine-structure ...
    2023; Webb et al. ... Our large galaxy sample provides a great advantage that allows us to measure the spatial variation of the fine-structure constant.Missing: refuted | Show results with:refuted
  37. [37]
    Constraints on the spacetime variation of the fine-structure constant ...
    We present strong constraints on the spacetime variation of the fine-structure constant \alpha using the Dark Energy Spectroscopic Instrument (DESI).
  38. [38]
    [PDF] The Fine Structure Constant - Indian Academy of Sciences
    Although it was Arnold Sommerfeld who formally intro- duced the fine structure constant in 1916 (as discussed later in this article), its history can be traced ...
  39. [39]
    [PDF] Sommerfeld fine structure constant α and its physical interpretation
    The historical example is the demonstration of “electromag- netic induction” by Michael Faraday himself before a distinguished gathering of members of the ...
  40. [40]
    The quantum theory of the electron - Journals
    It appears that the simplest Hamiltonian for a point-charge electron satisfying the requirements of both relativity and the general transformation theory leads ...
  41. [41]
    Eddington's Theory of the Constants of Nature - jstor
    that the fine-structure constant e2/hc must be the reciprocal of a whole number, which was determined in a second paper (Feb. 1930) to be 137. At the same.Missing: speculations | Show results with:speculations
  42. [42]
    [PDF] On Arthur Eddington's Theory of Everything - arXiv
    Pauli referred specifically to Eddington's identification of the fine-structure constant α with the number 1/136. A main reason for the generally unsympathetic ...
  43. [43]
    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.
  44. [44]
    [PDF] Hans Bethe, Quantum Mechanics, and the Lamb Shift
    An elementary derivation of the Lamb Shift, as originally given by Hans Bethe, is attempted, with a short historical background on the evolu-.
  45. [45]
    The Radiation Theories of Tomonaga, Schwinger, and Feynman
    A unified development of the subject of quantum electrodynamics is outlined, embodying the main features both of the Tomonaga-Schwinger and of the Feynman ...Missing: Nobel fine structure constant coupling
  46. [46]
    [PDF] 93. Grand Unified Theories - Particle Data Group
    Aug 11, 2022 · Since the. GUT scale is close to the Planck scale, higher-dimension operators involving the GUT-breaking. Higgs may modify the predictions ...Missing: 1970s | Show results with:1970s
  47. [47]
    The Anthropic Cosmological Principle - John D. Barrow, Frank J. Tipler
    Known as the Anthropic Cosmological Principle, it holds that the fundamental structure of the Universe is determined by the existence of intelligent observers: ...
  48. [48]
    [PDF] Physical Mathematics and the Fine-Structure Constant - PhilArchive
    Jul 11, 2018 · Arthur Eddington's work leads to a new calculation of the inverse fine-structure constant giving the same approximate value as ancient.
  49. [49]
    Mysticism and the Fine Structure Constant
    Sep 15, 2020 · The mysticism relating to the fine structure constant, including the surprising connection between the number 137 and Kabbalah by means of ...
  50. [50]
    (PDF) Mysticism and the Fine Structure Constant - ResearchGate
    Sep 15, 2020 · Mysticism and the Fine Structure Constant 489. Eddington, A. · S. (1931). On the value of the cosmical constant. Proceedings of the ; Royal ...
  51. [51]
    [PDF] Golden Ratio Geometry and the Fine-Structure Constant - HAL
    Oct 31, 2019 · Atomic Structure and Spectral Lines, New York, NY: Dutton, 1934. [3] Sherbon, M.A. “Fine-structure Constant from Sommerfeld to Feynman,” Journal ...
  52. [52]
    Insights from the π/γ Ratio and Electron Dynamics[v1]
    Dec 14, 2024 · This study explores a novel approach to understanding the fine-structure constant, α, by examining its connection to the anomalous magnetic ...<|control11|><|separator|>
  53. [53]
    [PDF] The links of Pythagorean prime 137 to the fine structure constant ...
    Sep 3, 2023 · We have provided possible links between 137 to the masses of the quarks, proton, W and Z bosons, Higgs boson, the Planck length, and the Coulomb ...<|separator|>
  54. [54]
    Atiyah and the Fine-Structure Constant - Sean Carroll
    Sep 25, 2018 · In the electroweak theory, we have two coupling constants, g and g' (for “weak isospin” and “weak hypercharge,” if you must know). There is also ...
  55. [55]
    On the vital difference between number theory and numerology in ...
    There has been much regrettable unintended or even intended confusion between the use of numerology and number theory in physics which has been the subject of ...Missing: critiques | Show results with:critiques