Fact-checked by Grok 2 weeks ago

Charge radius

The charge radius is a key parameter in nuclear and particle physics that quantifies the spatial distribution of within subatomic particles, such as protons, neutrons, and pions, or within atomic nuclei. It is conventionally defined as the root-mean-square () charge radius, \sqrt{\langle r^2 \rangle}, where \langle r^2 \rangle represents the mean-square charge radius extracted from the electromagnetic F(q^2) via the relation \langle r^2 \rangle = -6 \frac{dF}{dq^2} \big|_{q^2=0}, with q^2 being the squared transfer. This definition arises from the of the distribution, providing a model-independent measure of the particle's or nucleus's effective size as probed by electromagnetic interactions. Measurements of the charge radius are obtained through high-precision experiments, including elastic electron scattering off targets, which reveals the form factor's slope at low momentum transfer, and spectroscopic techniques such as those involving muonic atoms or laser spectroscopy on hydrogen-like ions, where shifts in energy levels reflect the finite nuclear size. For the proton, a case, the CODATA-recommended value as of 2022 is $0.84075 \pm 0.00064 fm, consistent with recent electron-scattering results around $0.831 to $0.841 fm and resolving prior discrepancies known as the "proton radius puzzle" through refined . In nuclei, charge radii exhibit systematic trends, such as the semi-empirical formula R \approx r_0 A^{1/3} (with r_0 \approx 1.2 fm and A the ), but deviate due to shell effects and dependence, offering insights into nuclear structure and the liquid-drop model. The charge radius holds significant theoretical importance, serving as a testing ground for (QCD) in describing quark-gluon dynamics within hadrons and for nuclear models like the or calculations. Discrepancies in measurements, such as those for exotic nuclei or the 's negative mean-square charge radius (\langle r_n^2 \rangle = -0.1155 \pm 0.0017 fm² (PDG 2025), indicating a non-zero charge distribution despite overall neutrality), highlight ongoing challenges in precision electroweak physics and beyond-standard-model searches. Recent advances, including isotope-shift laser spectroscopy, have extended measurements to short-lived isotopes, revealing "kinks" in radius trends across isotopic chains that probe neutron skins and equations of state.

Conceptual Foundations

Definition

The charge radius is a key observable in nuclear and particle physics that quantifies the spatial extent of the density distribution in subatomic particles or atomic nuclei. It is defined as the of the mean-square charge radius, \sqrt{\langle r^2 \rangle_{\rm ch}}, providing a measure of the effective size of the charge distribution beyond the point-like approximation used in basic models. The mean-square charge radius is formally given by \langle r^2 \rangle_{\rm ch} = \frac{\int \rho_{\rm ch}(\mathbf{r}) \, r^2 \, d^3\mathbf{r}}{\int \rho_{\rm ch}(\mathbf{r}) \, d^3\mathbf{r}}, where \rho_{\rm ch}(\mathbf{r}) is the charge density function, normalized such that the denominator equals the total charge (e.g., Ze for a nucleus with atomic number Z), and the integrals extend over all space. This expression captures the second radial moment of the charge distribution, emphasizing the quadratic weighting of distances from the center. In contrast to the matter radius, which describes the overall distribution of nucleons including s, the charge radius specifically probes the proton charge distribution through electromagnetic interactions, making it sensitive to differences in proton versus neutron densities. Values are typically reported in femtometers (, where $1 \, \rm fm = 10^{-15} \, \rm m), on the scale of nuclear dimensions; for instance, the proton's RMS charge radius is approximately 0.84 . Physically, the charge radius accounts for finite-size corrections to point-particle treatments, manifesting in processes like and atomic energy level shifts where the extended charge alters interaction potentials.

Theoretical Framework

In non-relativistic , the root-mean-square () charge radius arises as a key parameter in the description of a particle's charge distribution through the charge , which is the of the normalized charge \rho(\mathbf{r}): F(\mathbf{q}) = \int \rho(\mathbf{r}) e^{i \mathbf{q} \cdot \mathbf{r}} \, d^3\mathbf{r}, where \int \rho(\mathbf{r}) \, d^3\mathbf{r} = 1. For small momentum transfers q = |\mathbf{q}|, the exponential can be Taylor expanded as e^{i \mathbf{q} \cdot \mathbf{r}} \approx 1 + i \mathbf{q} \cdot \mathbf{r} - \frac{1}{2} (\mathbf{q} \cdot \mathbf{r})^2 + \cdots. The linear term vanishes by symmetry for a centered distribution, while the quadratic term yields F(q^2) \approx 1 - \frac{1}{6} \langle r^2 \rangle_\mathrm{ch} q^2, with \langle r^2 \rangle_\mathrm{ch} = \int r^2 \rho(\mathbf{r}) \, d^3\mathbf{r} being the mean-square charge radius (and the RMS charge radius \sqrt{\langle r^2 \rangle_\mathrm{ch}}). This expansion connects the low-energy electromagnetic response to the spatial extent of the charge. In (QED), the finite charge radius introduces corrections to atomic energy levels beyond the point-like nucleus approximation. These finite-size effects modify the potential near the , altering self-energy and contributions. For the —the splitting between $2S_{1/2} and $2P_{1/2} states in hydrogen-like atoms—the leading correction is \Delta E_\mathrm{FS} \propto \langle r^2 \rangle_\mathrm{ch} \delta_{l0}, scaling as \alpha^4 m^3 \langle r^2 \rangle_\mathrm{ch} / n^3 (where \alpha is the , m the , and n the principal ), and is most pronounced for s-states due to their into the nuclear region. Similar finite-size corrections apply to hyperfine splitting, arising from the modified magnetic interaction between the and , with contributions proportional to moments of the charge and magnetization densities convolved via the Zemach radius. These effects are crucial for precision tests of QED, as discrepancies in extracted radii (e.g., from muonic vs. electronic ) highlight nuclear structure influences. Charge density models provide explicit forms for \rho(\mathbf{r}) to compute \langle r^2 \rangle_\mathrm{ch} and F(q^2). The uniform sphere model assumes constant density \rho(r) = 3/(4\pi R^3) for r \leq R and zero otherwise, yielding a form factor F(q^2) = [3 j_1(qR)/(qR)] (with j_1 the spherical Bessel function) and \langle r^2 \rangle_\mathrm{ch} = (3/5) R^2; this simple model approximates heavier nuclei but underestimates surface diffuseness. The exponential model, \rho(r) \propto e^{-r/a}, is more suitable for lighter systems or mesons like the pion, giving F(q^2) = [1 + (q a)^2/12]^{-2} and \langle r^2 \rangle_\mathrm{ch} = 12 a^2, capturing a more diffuse distribution. These models influence \langle r^2 \rangle_\mathrm{ch} calculations by parameterizing the density's radial falloff, with fits to form factors revealing deviations from uniformity in nucleons. Relativistic effects complicate the charge radius definition, as the charge distribution is frame-dependent due to Lorentz boosts. In the Breit frame—the for virtual photon-nucleon scattering—the Sachs electric G_E(Q^2) directly Fourier-transforms a static, three-dimensional \rho(\mathbf{r}) with no initial/final motion, minimizing boost artifacts. However, experimental data are typically in the laboratory frame, where the target is at rest, leading to differences: the observed density involves relativistic quasi-distributions distorted by factors like \sqrt{1 + Q^2/(4M^2)} (with M the mass), and the effective radius can differ by up to 10% for Q^2 \sim 1 GeV² due to contraction along the boost direction. These frame distinctions are reconciled via light-front or infinite-momentum formulations for consistent extractions. The charge radius connects to nucleon structure functions in deep inelastic scattering (DIS) through the charge distribution's role in low-momentum-transfer limits. Elastic form factors relate to moments of generalized parton distributions (GPDs) via G_E(Q^2) = \int_{-1}^1 dx \, H(x, 0, -Q^2), where H(x, \xi, t) encodes charge densities; at small t = -Q^2, this probes the transverse of quarks contributing to the charge, complementing DIS structure functions F_2(x, Q^2) that integrate over higher twists but inform the overall charge profile.

Historical Evolution

Early Discoveries

The Geiger-Marsden experiments, conducted between 1906 and 1913 under Ernest Rutherford's supervision at the , provided the first experimental evidence for a finite nuclear size through α-particle scattering off thin gold foil. In their 1909 report, Geiger and Marsden observed that while most α particles passed through the foil undeflected, a small fraction scattered at large angles, up to 150 degrees, contradicting the prevailing of the atom which predicted only small deflections. This large-angle scattering implied the existence of a dense, positively charged core within the atom, with Rutherford later calculating an upper limit for the gold nucleus radius of approximately 34 fm based on the closest approach distances derived from the scattering data. In , Rutherford formalized these findings in his seminal paper, proposing the model of the where the positive charge is concentrated in a tiny central . He derived the Rutherford formula for the differential cross-section, \sigma(\theta) \propto \frac{1}{\sin^4(\theta/2)}, which quantitatively explained the observed angular distribution of scattered α particles assuming repulsion from a point-like central charge. This formula allowed estimates of size from scattering cross-sections, confirming the as much smaller than the and setting the stage for understanding finite dimensions, though initial focus remained on qualitative evidence against point-like atomic structures. The and saw further developments in nuclear composition that informed early concepts of charge distribution. Rutherford himself hypothesized a in to explain isotopic masses without additional charge, paving the way for nuclear models beyond pure protons. This culminated in James Chadwick's discovery of the through bombardment experiments with α particles on , revealing a of mass similar to the proton that could bind with protons to form stable nuclei. The proton-neutron model thus introduced the idea that nuclear charge arises solely from protons, distributed within a finite volume alongside neutral neutrons, shifting theoretical attention toward the spatial arrangement of charges in composite nuclei. By the mid-1930s, empirical relations for nuclear radii emerged from analyses of α-decay and data, assuming constant nuclear density akin to a liquid drop. In 1935, proposed the formula R \approx r_0 A^{1/3} with r_0 \approx 1.5 in his , derived to fit energies and rates for various nuclei. Early measurements predominantly targeted heavy elements like (A=197, R \sim 8.7 ), where signals were stronger, rather than light hadrons such as the proton, whose precise charge radii awaited later high-energy probes.

Mid-20th Century Developments

Following , the development of high-energy particle accelerators enabled precise measurements of nuclear charge radii through experiments. In the 1950s, and his collaborators at pioneered high-energy on nuclei, using the university's linear electron accelerator to probe the internal structure of protons and other light nuclei. These experiments revealed deviations from point-like scattering predictions, allowing the extraction of charge form factors and the first precise determination of the proton charge radius, approximately 0.8 , via analysis of the cross-section. A key theoretical advancement was the derivation of the Rosenbluth formula for elastic electron-proton scattering, which separated the contributions of charge and magnetic form factors to the differential cross-section. The formula is given by \frac{d\sigma}{d\Omega} = \left( \frac{d\sigma}{d\Omega} \right)_{\rm point} \left[ F_{\rm ch}^2(q^2) + \frac{q^2}{4M^2} F_{\rm mag}^2(q^2) \right], where \left( \frac{d\sigma}{d\Omega} \right)_{\rm point} is the point-like Mott cross-section, F_{\rm ch}(q^2) is the charge form factor, F_{\rm mag}(q^2) is the magnetic form factor, q^2 is the four-momentum transfer squared, and M is the proton mass. This enabled isolation of the charge form factor at low q^2, providing direct access to the charge radius from its slope. A pivotal experimental confirmation of the proton's finite size came in 1955, when Hofstadter and McAllister observed deviations from the point-like Mott cross-section in electron-proton scattering data at electron energies of 100, 188, and 236 MeV, establishing the proton as a composite object with a measurable size. In the , measurements extended to the deuteron and light nuclei, refining charge radii with higher precision using improved detectors and early synchrotrons such as those at Stanford and . These experiments, often employing Cerenkov counters and magnetic spectrometers alongside emerging techniques for particle identification, yielded deuteron charge radii around 2.1 and provided insights into distributions within composite systems. Theoretical progress complemented these efforts, particularly in understanding the neutron's charge properties. In 1958, Leonard Foldy analyzed the electromagnetic interaction between neutrons and electrons, deriving a relativistic correction term—now known as the Foldy term—that contributes negatively to the neutron's mean-square charge radius, \langle r^2 \rangle_n \approx -0.071 \, \rm fm^2, arising from the neutron's anomalous and substructure effects rather than a positive charge distribution. This work highlighted the neutron's effective charge radius as a subtle balance of relativistic and internal dynamics.

Experimental Methods

Scattering Techniques

Scattering techniques for determining charge radii rely on elastic electron scattering from protons or atomic nuclei, where the differential cross-section encodes the Fourier transform of the charge distribution in the form of the electric (charge) form factor F_{\mathrm{ch}}(q^2). At low momentum transfers q^2, the form factor deviates from unity due to the finite spatial extent of the charge distribution, allowing extraction of the mean squared charge radius \langle r^2 \rangle_{\mathrm{ch}} from its expansion. The radius is defined as the slope at q^2 = 0: \langle r^2 \rangle_{\mathrm{ch}} = -6 \left. \frac{d F_{\mathrm{ch}}}{d q^2} \right|_{q^2 = 0}, where q is the transfer. This approach applies to both protons and nuclei, with the F_{\mathrm{ch}}(q^2) obtained by comparing measured cross-sections to the point-like Mott cross-section, corrected for and inelastic contributions. Pioneering measurements by Hofstadter and collaborators in the at Stanford, using beams up to several hundred MeV on and light nuclei targets, first revealed the non-point-like nature of charge distributions through deviations in the cross-section at forward angles. High-precision implementations of elastic electron scattering employ dedicated facilities like Jefferson Laboratory's Hall A, which features two superconducting quadrupole-toroid magnetic spectrometers (High Resolution Spectrometers, HRS) optimized for detecting scattered electrons at low q^2 (down to ~0.01 fm^{-2}) with momentum resolutions better than 10^{-4}. These spectrometers, combined with cryogenic hydrogen or deuterium targets and large solid-angle scintillators, enable precise kinematic reconstruction and background suppression, crucial for form factor determinations at the percent level. Data from such setups are fitted to parametrized form factors (e.g., dipole or polynomial expansions) to extrapolate the slope at q^2 = 0, with uncertainties dominated by statistical precision and model dependence at low q. Radiative corrections are essential in these experiments due to QED effects like virtual photon exchange and real bremsstrahlung, which introduce infrared divergences that must be handled carefully to avoid biasing the low-q^2 slope. The Peierls-Yennie method, developed in the early , provides a systematic for resumming these soft-photon contributions by separating infrared-safe hard radiative processes from infrared-divergent soft ones, ensuring consistent treatment across Born and higher-order diagrams. This approach, often implemented numerically in modern codes like ESEn or RADCOR, reduces systematic uncertainties from radiative effects to below 0.1% for charge radius extractions. Polarized enhances sensitivity to individual form factors, particularly for the , whose electromagnetic charge is zero but possesses a non-zero charge distribution from its content. Experiments at Jefferson Lab's Hall A use polarized ^3He gas targets, where the ^3He acts as an effective polarized target (due to ~90% alignment in the S-wave), measuring the asymmetry in quasi-elastic ^3He(\vec{e}, e' \vec{n})p reactions to extract the electric form factor G_E^n(q^2). The charge radius follows from the low-q^2 of G_E^n, yielding \langle r^2 \rangle_{\mathrm{ch}}^n \approx -0.11 fm^2. Parity-violating methods leverage the interference between electromagnetic and weak neutral currents in polarized , providing complementary access to neutron distributions without uncertainties. The SAMPLE experiment at MIT-Bates (1998–2001) measured the parity-violating asymmetry in of longitudinally polarized electrons from unpolarized targets, sensitive to the 's weak form factors and yielding constraints on the 's axial charge radius via the anapole moment. This technique has been extended to nuclei, where the parity-violating asymmetry A_{PV} scales with the weak charge, dominated by neutrons, allowing isolation of neutron radial distributions. Recent advancements emphasize calorimeter-based detection to minimize acceptance uncertainties at ultra-low q^2. The PRad experiment at Jefferson Lab (data collected 2016, analysis through 2020) used a setup with a high-intensity unpolarized beam (1.1–2.2 GeV) on a target, detecting scattered electrons and positrons in a large-acceptance calorimeter at forward angles (\theta < 6^\circ). This magnetic-spectrometer-free approach achieved ~1% precision on the proton charge radius by directly measuring the e p \to e p cross-section ratio to hydrogen-like scattering, with radiative corrections via the Peierls-Yennie framework ensuring low systematic errors. The result, proton charge radius r_p = 0.831 \pm 0.007 (stat.) \pm 0.012 (syst.) fm (corresponding to \langle r^2 \rangle_{\mathrm{ch}}^p \approx 0.691 \pm 0.023 fm^2, total relative uncertainty \approx 3.3\%), demonstrates the technique's potential for sub-percent accuracy in future iterations.

Spectroscopic Approaches

Spectroscopic approaches to determining charge radii rely on measuring energy level shifts in atomic and exotic atom spectra, where quantum electrodynamics (QED) predictions incorporate finite nuclear size effects as corrections to otherwise point-like nucleus calculations. These methods achieve high precision by isolating the charge radius-dependent terms in transitions like the , which arises from the difference between s- and p-state energies, or hyperfine splittings influenced by nuclear structure. The proton serves as the primary target for such spectroscopic studies due to its fundamental role in atomic physics. In ordinary hydrogen, the finite nuclear size introduces a correction to the Lamb shift proportional to the mean-square charge radius ⟨r²⟩_ch divided by the cube of the Bohr radius a_0³, reflecting the wavefunction overlap with the nucleus. This effect was first indicated in early spectroscopic measurements of hydrogen's Balmer series limits, where discrepancies from Dirac theory prompted considerations of nuclear extent. The leading-order finite size contribution to the 2S–2P transition energy is given by \Delta E_\mathrm{fs} = \frac{8 \alpha^5 m_r^3}{3\pi n^3} \langle r^2 \rangle_\mathrm{ch} Z^4, where α is the fine-structure constant, m_r the reduced mass of the electron-proton system, n the principal quantum number, and Z the atomic number; this formula applies to the s-state shift, as p-states are insensitive at this order. Muonic atoms enhance sensitivity to charge radii because the heavier muon orbits approximately 185 times closer to the nucleus than an electron, increasing the finite size correction by a factor of about 10^5 relative to electronic atoms. Laser spectroscopy of the 2S–2P Lamb shift in muonic hydrogen, performed by the CREMA collaboration, yielded a proton root-mean-square charge radius of 0.84184(67) fm, demonstrating the method's precision. Two-photon laser spectroscopy in muonic helium-4 ions provides an analogous determination for the α-particle charge radius. The CREMA collaboration's 2021 measurement of the 2S–2P transition at the Paul Scherrer Institute extracted a root-mean-square charge radius of 1.678 fm for the α particle, with uncertainty dominated by theoretical QED inputs but validating complementary scattering results at higher precision. Recent extensions include laser spectroscopy of muonic helium-3 ions, which in 2025 yielded the helion root-mean-square charge radius of 1.97007(94) fm, highlighting sensitivities to neutron distributions in neutron-rich nuclei. In deuterium, anomalies in the hyperfine structure of atomic energy levels allow extraction of the deuteron charge radius through QED corrections that account for nuclear finite size and two-photon exchange effects. Spectroscopic measurements of the ground-state hyperfine splitting, combined with theoretical calculations, yield a deuteron root-mean-square charge radius of approximately 2.130 fm, highlighting sensitivities to meson-exchange currents in the loosely bound deuteron.

Case Studies

Proton Charge Radius

The proton charge radius, defined as the root-mean-square (rms) of its charge distribution, serves as a fundamental benchmark for understanding nucleon structure within quantum chromodynamics (QCD). Early measurements via elastic electron-proton scattering, pioneered by Hofstadter and collaborators in the 1950s, yielded a value of approximately 0.81 fm, revealing the proton's finite size beyond a point-like particle. Subsequent high-precision electron scattering experiments refined this further; for instance, the Mainz A1 collaboration in 2010 reported 0.879(8) fm from unpolarized elastic scattering data at momentum transfers up to 1.05 (GeV/c)^2. More recent efforts at , such as the in 2019, provided 0.831 ± 0.007 (stat) ± 0.012 (sys) fm using a novel low-background setup to measure scattering at forward angles. Complementing these, spectroscopy of muonic hydrogen by the in 2010 extracted 0.84087(39) fm from the 2S-2P Lamb shift, leveraging the muon's proximity to the proton for enhanced sensitivity to the charge distribution. Integration of these diverse measurements, accounting for systematic uncertainties and theoretical inputs, has converged on a consensus value. The Committee on Data for Science and Technology (CODATA) in 2022 recommends the proton rms charge radius as 0.84075(64) fm, derived from a weighted average of electron scattering, muonic hydrogen, and electronic hydrogen spectroscopy results, with the uncertainty reflecting residual tensions resolved through improved form factor models. This value underscores the proton's role in calibrating QCD-inspired calculations, where the mean-square charge radius ⟨r_E^2⟩_p ≈ 0.707 fm² establishes the scale of valence quark confinement. In quark models, the proton's charge radius arises primarily from the spatial distribution of its valence quarks, with simple non-relativistic constituent quark models predicting a core ⟨r_E^2⟩_p around 0.6–0.7 fm², underestimating the experimental value due to neglect of meson exchange. The pion cloud, modeled via , contributes an additional ~10–20% to ⟨r_E^2⟩_p through virtual pion emission and absorption, enhancing the effective size and aligning predictions with data. Isovector and isoscalar combinations further illuminate quark flavor dynamics: the proton's positive ⟨r_E^2⟩_p contrasts sharply with the neutron's negative ⟨r_E^2⟩_n ≈ -0.11 fm², reflecting the differing up- and down-quark charge weights and probing isospin symmetry breaking from electromagnetic and strong interaction asymmetries. These differences highlight how the proton's charge radius encodes up-quark dominance in the valence structure.

Charge Radii of Other Nuclei and Particles

The charge radius of the neutron, a neutral particle, provides insight into the internal structure of nucleons under quantum chromodynamics (QCD), contrasting with the proton's positive charge distribution. The mean-square charge radius of the neutron is measured as ⟨r²⟩_n = -0.1155 ± 0.0017 fm², a negative value arising from an asymmetric quark distribution featuring a positively charged core surrounded by a negatively charged skin influenced by the neutron's magnetic moment. This negativity highlights QCD effects not present in the proton's uniform positive charge. For light nuclei, the deuteron—the bound state of a proton and neutron—exhibits a charge radius determined primarily by the proton's contribution, with the neutron adding negligibly due to its neutrality. The root-mean-square charge radius of the deuteron is 2.12778(27) fm, derived from isotope-shift measurements in hydrogen-deuterium spectroscopy as recommended by . This larger radius compared to the proton illustrates simple nuclear binding effects in few-body systems. The alpha particle, or helium-4 nucleus, represents a tightly bound four-nucleon system with a nearly uniform charge distribution. Muonic helium-4 spectroscopy in 2021 yielded a root-mean-square charge radius of 1.6785(21) fm, assuming a uniform sphere model, which aligns well with electron scattering data and underscores the compact nature of this doubly magic nucleus. Extending to mesons, which probe QCD at the quark level, the charged pion's charge radius is approximately 0.66 fm, extracted from electron-pion scattering experiments that reveal the light quark dynamics. Kaon radii, around 0.56 fm for charged kaons, offer complementary insights into strange quark contributions, as their inclusion alters the charge distribution relative to non-strange mesons like the pion. In heavy nuclei, charge radii generally follow an A^{1/3} scaling law, where A is the mass number, but deviations arise from surface diffuseness and shell effects. For example, the root-mean-square charge radius of is approximately 5.5 fm, larger than a naive uniform sphere prediction due to these surface contributions, illustrating collective nuclear behavior in massive systems.

Challenges and Resolutions

The Proton Radius Puzzle

The proton radius puzzle emerged in 2010 when the (CREMA) collaboration measured the Lamb shift in muonic hydrogen, yielding a proton root-mean-square charge radius of r_p = 0.84087 \pm 0.00039 fm. This value stood in stark contrast to the CODATA-2010 recommended value of r_p = 0.8751 \pm 0.0061 fm, derived primarily from electron-proton scattering and ordinary hydrogen spectroscopy, representing a tension of approximately 7 standard deviations. Prior to this measurement, various experimental determinations had converged consistently around \sim 0.87 fm, establishing a long-standing consensus on the proton's size. Following the CREMA result, the discrepancy ignited intense debate within the physics community, highlighted at dedicated conferences such as the in Trento, where experts scrutinized potential systematic effects and theoretical interpretations. Proposed resolutions fell into three main categories: possible errors in calculations for muonic hydrogen, such as unaccounted higher-order finite nuclear size effects; new physics beyond the , including light dark matter mediators like or violations of through additional gauge bosons; and underestimated uncertainties in electron scattering data, potentially from two-photon exchange contributions or form factor extractions. In response, new experiments sought to clarify the tension. A 2017 spectroscopic measurement of the 2S-4P transition in ordinary hydrogen by the JILA and MPQ groups reported r_p = 0.833 \pm 0.010 fm, partially bridging the gap by shifting closer to the muonic value while remaining inconsistent with pre-2010 scattering results. Similarly, the 2019 electron-proton scattering experiment at (PRad) extracted r_p = 0.831 \pm 0.007_{\text{stat}} \pm 0.012_{\text{syst}} fm from low-momentum-transfer cross sections, aligning more closely with the muonic hydrogen determination and challenging the higher scattering-based values. The puzzle profoundly impacted precision tests of the Standard Model, as the proton radius enters calculations of atomic energy levels and electroweak processes, raising concerns about the reliability of QED in nuclear contexts. It spurred over 20 new experiments worldwide, including muon scattering efforts like and upgraded hydrogen spectroscopy, to probe lepton universality and resolve the underlying cause.

Contemporary Values and Prospects

Following the resolution of the proton radius puzzle through converging measurements in the early 2020s, the accepted value for the proton root-mean-square charge radius has stabilized at approximately 0.841 fm. This consensus emerged from high-precision electron scattering experiments, such as the at , which reported 0.831 ± 0.007(stat) ± 0.012(syst) fm in 2019, aligning closely with muonic hydrogen spectroscopy results like the 0.84087(39) fm from the . The (CODATA) adopted an updated value of 0.84075(64) fm in its 2022 recommendations (unchanged as of 2025), reflecting a weighted average of these and other spectroscopic determinations that reduced discrepancies to below 1%. Updated measurements for other hadronic systems have similarly advanced, providing benchmarks for quantum chromodynamics (QCD) in the non-perturbative regime. For the neutron, a 2020 precision analysis of deuteron structure via electron scattering yielded a mean-square charge radius of -0.1097(51) fm², corresponding to an rms value of about 0.345 fm after accounting for relativistic corrections; this refines earlier extractions and highlights the neutron's negative charge distribution due to its quark content. Ongoing efforts at CERN's COMPASS experiment, extended through the AMBER upgrade, target charged pion and kaon charge radii using Primakoff reactions and Drell-Yan processes, with preliminary pion results indicating <r²>π ≈ 0.43 fm² from reanalysis of existing scattering data. Theoretical advancements, particularly in , have bolstered experimental findings by providing parameter-free predictions that match the ~0.84 fm proton value. The Budapest-Marseille-Wuppertal () collaboration's 2020 simulations, incorporating physical masses and isospin-symmetric ensembles, computed the proton charge radius at 0.845(9) fm, demonstrating consistency with muonic and electronic measurements while validating QCD's ability to describe structure at low energies. These refinements extend to flavor-separated distributions, aiding interpretations of scattering data. As of 2025, updated calculations, such as those from the BMWc collaboration, continue to confirm values around 0.841(6) fm. Prospects for further precision include dedicated runs at Jefferson Lab's CLAS12 detector, which will probe flavor-separated charge radii through deeply virtual and generalized parton distributions, potentially isolating up- and down-quark contributions with sub-1% uncertainties by the late 2020s. The MUSE-II phase at the aims to finalize muon-proton scattering analyses for the proton radius at the 0.5% level, while muonic atom campaigns target heavier nuclei like ²⁰⁸Pb and exotic isotopes, achieving rms charge radii with precisions below 0.01 fm to map nuclear deformation trends. The PRad-II experiment at Jefferson Lab, now underway, is expected to deliver a new measurement with improved precision in the coming years. These developments carry broader implications for and QCD validation. Refined light-nuclei charge radii enhance models by improving predictions of primordial and abundances, where uncertainties in rates previously limited accuracy to 5-10%. For QCD, the alignment of predictions with experiment confirms the theory's reach into confined regimes, constraining parton dynamics essential for interpreting high-energy data.

References

  1. [1]
    charge radius - pdgLive - Lawrence Berkeley National Laboratory
    Most measurements of the radius of the proton involve electron-proton interactions, the most recent of which is the electron scattering measurement r p = 0.831( ...
  2. [2]
    The proton charge radius | Rev. Mod. Phys.
    Jan 21, 2022 · The proton root-mean-square (rms) charge radius (also known as the proton charge radius) is a quantity that is of importance not only to QCD but ...Article Text · Introduction · The Proton Charge Radius... · Elastic Electron-Proton...
  3. [3]
    [PDF] TABLE OF NUCLEAR ROOT MEAN SQUARE CHARGE RADII*
    The root-mean-square (rms) charge radius is a fundamental quantity for both nuclear structure and reaction calculations. The results from different ...<|separator|>
  4. [4]
    Measurement of the neutron charge radius and the role of its ...
    Here the authors present a method, based values of neutron electric form factors, to determine the charge radius of the neutron.
  5. [5]
    The nuclear charge radius of 13C | Nature Communications
    Jul 7, 2025 · We present a laser spectroscopic measurement of the root-mean-square (rms) charge radius of 13 C and compare this with ab initio nuclear structure calculations.
  6. [6]
    Critical evaluation of reference charge radii and applications in ...
    In order to translate R k , α μ to the more universal rms charge radius, information on the charge distribution ('nuclear shape'), from either calculations or ...
  7. [7]
    Charge radii of potassium isotopes in the RMF (BCS)* approach
    In the conventional RMF (BCS) model, the mean square charge radius is calculated in the following way (in units of fm2) [46, 47]: ... ⟨R(2)ch⟩=∫r2ρp(r)d3 ...
  8. [8]
    [PDF] arXiv:nucl-th/0607024v1 13 Jul 2006
    The contribution to the nuclear charge distribution from a single-proton in this spherical shell is thus given by Np (R, r) = dρch/4πr2dr. Assuming that the ...
  9. [9]
    proton rms charge radius
    proton rms charge radius $r_{\rm p}$. Numerical value, 8.4075 x 10-16 m. Standard uncertainty, 0.0064 x 10-16 m. Relative standard uncertainty, 7.6 x 10-4.
  10. [10]
    None
    Summary of each segment:
  11. [11]
    On a diffuse reflection of the α-particles - Journals
    A small fraction of the α-particles falling upon a metal plate have their directions changed to such an extent that they emerge again at the side of incidence.
  12. [12]
    [PDF] LXXIX. The scattering of α and β particles by matter and the structure ...
    To cite this Article Rutherford, E.(1911) 'LXXIX. The scattering of α and β ... pdf. This article may be used for research, teaching and private study ...
  13. [13]
    The existence of a neutron | Proceedings of the Royal Society of ...
    § 1. It was shown by Bothe and Becker that some light elements when bombarded by α-particles of polonium emit radiations which appear to be of the γ-ray type.
  14. [14]
    [PDF] Zur Theorie der Kernmassen
    Zur Theorie der Kernmassen. Von C. F. v. Weizs~eker in Leipzig. Mit 5 Abbildungen. (Eingegangen am 6. Juli 1935.) w 1. Problemstellung. w 2. Erweiterung der ...
  15. [15]
    [PDF] Robert Hofstadter - Nobel Lecture
    Early electron-scattering experiments were carried out at the University of Illinois in 1951*at an incident electron energy of about 15.7 MeV. Such experiments ...
  16. [16]
    Evaluation of the Proton Charge Radius from Electron–Proton ...
    Jun 16, 2015 · ... elastic electron–proton (e–p) scattering. The radius is related to the slope of the charge form factor, GE(Q2), at Q2 = 0. As such, the ...
  17. [17]
    Basic instrumentation for Hall A at Jefferson Lab - ScienceDirect.com
    The instrumentation in Hall A at the Thomas Jefferson National Accelerator Facility was designed to study electro- and photo-induced reactions at very high ...
  18. [18]
    [PDF] Radiative corrections to elastic-electron proton scattering and ...
    Abstract. Higher-order QED radiative corrections to elastic electron-proton scattering are discussed. It is shown that they are relevant for high-precision ...Missing: Peierls- | Show results with:Peierls-
  19. [19]
    The sample experiment and weak nucleon structure - ScienceDirect
    In this paper we review the present state of experiment and theory with a specific emphasis on the sample experiment, which was carried out at the MIT-Bates ...
  20. [20]
    Measuring the α-particle charge radius with muonic helium-4 ions
    Jan 27, 2021 · Here we present the measurement of two 2S–2P transitions in the muonic helium-4 ion that yields a precise determination of the root-mean-square charge radius ...
  21. [21]
    The deuteron-radius puzzle is alive: A new analysis of nuclear ...
    Mar 10, 2018 · The deuteron charge radius is extracted from the LS measurement through(1) Δ E LS = δ QED + δ TPE + m r α 4 12 r d 2 , which is valid in an α ...
  22. [22]
    Structure of the Proton | Phys. Rev. - Physical Review Link Manager
    The structure and size of the proton have been studied by means of high-energy electron scattering. The elastic scattering of electrons from protons in ...
  23. [23]
    proton rms charge radius - CODATA Value
    proton rms charge radius $r_{\rm p}$. Numerical value, 8.4075 x 10-16 m. Standard uncertainty, 0.0064 x 10-16 m. Relative standard uncertainty, 7.6 x 10-4.
  24. [24]
    [1809.06478] Weak radius of the proton - arXiv
    Sep 17, 2018 · The proton's weak radius R_W= 1.545\pm 0.017 fm is 80\% larger than its charge radius R_{ch}\approx 0.84 fm because of a very large pion cloud ...
  25. [25]
    Measurement of the neutron charge radius and the role of its ... - arXiv
    Mar 19, 2021 · Abstract page for arXiv paper 2103.10840: Measurement of the neutron charge radius and the role of its constituents.
  26. [26]
    deuteron rms charge radius - CODATA Value
    Concise form, 2.127 78(27) x 10-15 m ; Click here for correlation coefficient of this constant with other constants ; Source: 2022 CODATA
  27. [27]
    pdgLive
    The charge radius of the pion is defined in relation to the form factor of the pion electromagnetic vertex, called vector form factor VFF.
  28. [28]
    CHARGE RADIUS fm - pdgLive
    A Measurement of the Kaon Charge Radius. DALLY, 1980. PRL 45 232, Direct Measurement of the Negative Kaon Form-factor. BLATNIK, 1979. LNC 24 39, The Isovector ...
  29. [29]
    Nuclear Charge Radii
    The table of experimental nuclear charge radii covers an extended range of isotopes and elements (909 isotopes of 92 elements from 1 H to 96 Cm)
  30. [30]
    The size of the proton | Nature
    Jul 8, 2010 · The proton is the primary building block of the visible Universe, but many of its properties—such as its charge radius and its anomalous ...
  31. [31]
    Constraints to new physics models for the proton charge radius ...
    Feb 6, 2014 · ... lepton universality violating particles present in proton charge radius puzzle explanations. The model of Batell et al. [8] contains dark ...
  32. [32]
    [PDF] The Proton Charge Radius - CERN Indico
    Exps: Mainz, JLab (PRad), MUSE at PSI, ULQ2 in Japan, Amber@CERN; H ... = 0.833(10) fm r p. = 0.8482(38) fm. Grinin et al., Science 370, 1061 (2020). Gao ...
  33. [33]
    Extraction of the Neutron Charge Radius from a Precision ...
    Feb 25, 2020 · We present a high-accuracy calculation of the deuteron structure radius in chiral effective field theory.Missing: PREEM | Show results with:PREEM
  34. [34]
    Meet AMBER - CERN
    Mar 8, 2021 · These studies require the beamline that feeds COMPASS to be upgraded to deliver a charged-kaon beam of high energy and intensity. Combining ...
  35. [35]
    [PDF] The Proton Radius Measurement at AMBER | PAW 2025
    Jun 2, 2025 · • AMBER will measure the proton charge radius with a high-energy muon beam using a combination of an active target hydrogen TPC and muon ...
  36. [36]
    [PDF] The Physics of Nuclei Nuclear Matter and Nucleosynthesis - UKRI
    ... charge radius ... Such particles may play a crucial role in neutron star physics, and hadronisation following the big bang, and provide unique constraints on QCD, ...