Fact-checked by Grok 2 weeks ago

Nuclear force

The nuclear force, also known as the strong nuclear force, is the responsible for binding protons and neutrons—collectively termed nucleons—together within the , thereby overcoming the electromagnetic repulsion between positively charged protons to form stable atomic nuclei. This force operates at extremely short ranges, on the order of a few femtometers (1 fm = 10⁻¹⁵ m), and is approximately 100 times stronger than the electromagnetic force, making it the most powerful of the four forces of . Its effects diminish rapidly beyond the nuclear scale, confining its influence to subatomic distances about 100,000 times smaller than an atom's diameter. Key properties of the force include its charge independence, meaning the is nearly identical between proton-proton, neutron-neutron, and proton-neutron pairs when electromagnetic effects are disregarded, as well as its spin-dependent nature, which incorporates tensor and spin-orbit components that contribute to nuclear stability. The force exhibits a repulsive "hard core" at distances below about 0.5 , preventing nucleons from overlapping too closely, while being attractive at slightly larger separations to facilitate . Although primarily a two-nucleon , weaker three-nucleon forces play a role in more complex nuclei, influencing phenomena like nuclear energies and reactions. This force underpins nuclear stability and enables processes such as and fusion, which release vast amounts of energy—up to a million times more per unit mass than chemical reactions. Theoretically, the nuclear force is understood as a residual effect of the strong force that binds quarks within nucleons via exchange, as described by (QCD), the fundamental theory of strong interactions developed in the 1970s. Historically, its existence was inferred after the 1932 , with proposing in 1935 a meson-exchange mechanism that predicted the , confirmed in 1947 and earning him the 1949 . Modern approaches, such as chiral effective field theory pioneered by in the 1990s, provide precise quantitative descriptions by integrating QCD principles with low-energy . Ongoing research at facilities like the Continuous Electron Beam Accelerator Facility (CEBAF) continues to probe its intricacies, enhancing our understanding of nuclear structure and stellar processes.

Fundamental Properties

Strength and range

The nuclear force, also known as the strong nuclear force between nucleons, is the most powerful of the four fundamental interactions at short distances, with a relative strength approximately 100 times that of the electromagnetic force between two protons in close proximity. In comparison to the weak nuclear force, it is about 10^{13} times stronger, while it exceeds the gravitational force by a factor of roughly 10^{38}. These ratios highlight the nuclear force's dominant role in overcoming electromagnetic repulsion within atomic nuclei, enabling stable binding of protons and neutrons. The nuclear force operates over an extremely limited spatial range, effective only up to about 2-3 , which corresponds roughly to the diameter of atomic nuclei. Beyond this distance, the force diminishes exponentially to negligible levels, confining its influence to the scale of nuclear dimensions and preventing it from affecting larger structures like atoms or molecules. A key characteristic of the nuclear force is its property, whereby the attractive does not accumulate indefinitely with increasing numbers of s but instead reaches a balance that results in a nearly constant nuclear density of approximately 0.16-0.17 s per fm³ across most heavy nuclei. This arises from the short-range nature of the force combined with short-range repulsive components at very close distances, limiting each 's to a fixed number of neighbors and preventing indefinite clustering. Consequently, the per stabilizes at around 8 MeV for medium to heavy nuclei, reflecting this finite capacity. Empirical evidence for the finite range of the nuclear force comes primarily from nucleon-nucleon experiments, which reveal that the interaction vanishes at separations greater than a few , as indicated by the absence of significant phase shifts at larger impact parameters. High-energy data, particularly above 200 MeV, further demonstrate a transition to repulsive behavior at distances below 0.5 , confirming the force's sharp cutoff and supporting models of its limited extent.

Charge independence and isospin

The nuclear force exhibits charge independence, meaning that the strong interaction between two protons (pp), two neutrons (nn), and a proton-neutron pair (np) is essentially the same in magnitude and form, disregarding small corrections from electromagnetic interactions. This property arises because the strong force, mediated by the of gluons between quarks, does not distinguish between the electric charges of protons and neutrons at leading order. Charge independence is formalized through the concept of , an approximate SU(2) symmetry in the strong interaction that treats protons and neutrons as two states of a single particle, the , with total isospin I = \frac{1}{2}. In this framework, the proton corresponds to the state with third-component isospin I_3 = +\frac{1}{2}, while the neutron has I_3 = -\frac{1}{2}, analogous to the up and down states of a spin-\frac{1}{2} particle. The SU(2) group structure allows nucleons to form isospin multiplets, enabling the classification of nuclear states and interactions under rotations in isospin space, where the strong force conserves total isospin I and I_3. This symmetry originates from the near-degeneracy of the up and down masses in (QCD), the underlying theory of strong interactions. Experimental verification of charge independence comes from nucleon-nucleon scattering experiments, where the differential and total cross-sections for pp, nn, and scattering show striking similarities at low energies after correcting for effects in charged pairs. For instance, in the ^1S_0 channel (as of 2023 measurements), the nuclear s-wave scattering lengths are a_{nn}^N \approx -18.9 \pm 0.4 fm for nn, a_{pp}^N \approx -18.2_{-0.58}^{+0.52} fm for pp (-free), and a_{np}^N = -23.74 \pm 0.02 fm for , demonstrating close equivalence between like-nucleon pairs and the distinction for due to channels. Such measurements, obtained via techniques like the method in quasi-free reactions, confirm that the force operates symmetrically across channels, with deviations primarily attributable to non-strong effects. Despite its successes, charge independence is violated at the level of a few percent due to electromagnetic interactions and the up-down quark mass difference (m_d - m_u \approx 2.5 MeV). Electromagnetic contributions include Coulomb repulsion in pp scattering and pion mass splittings (m_{\pi^\pm} - m_{\pi^0} \approx 4.6 MeV), which introduce charge-dependent potentials; these account for much of the observed charge independence breaking (CIB) in scattering lengths, quantified by \Delta a_{\rm CIB} \approx |a_{np}^N| - \frac{|a_{pp}^N| + |a_{nn}^N|}{2} \approx 5.2 fm (2023 values), mainly from the I=0 (np singlet) vs. I=1 (pp, nn) difference. Charge symmetry breaking (CSB), the smaller pp-nn difference \Delta a_{\rm CSB} = a_{nn}^N - a_{pp}^N \approx -0.7 fm, arises additionally from quark mass effects. The quark mass difference generates strong-interaction violations through higher-order QCD effects, contributing to differences like the neutron-proton mass splitting (m_n - m_p \approx 1.3 MeV, with ~0.8 MeV from electromagnetism and the rest from quarks). These violations manifest as ~1-2% deviations in binding energies of mirror nuclei and scattering observables, underscoring the approximate nature of the SU(2) symmetry.

Historical Development

Early discoveries

In 1911, conducted scattering experiments on thin gold foil, observing that a small fraction of particles were deflected at large angles, which could only be explained by the presence of a tiny, dense, positively charged at the atom's center. This discovery implied that the must be held together by a powerful attractive force to counteract the electromagnetic repulsion between its positively charged protons. The 1930s brought key experimental observations of interactions between neutrons and protons, highlighting the nuclear force's role. In 1930, and Herbert Becker bombarded with alpha particles, producing a highly penetrating that interacted strongly with matter. , in , identified this as neutrons by measuring their off protons in , where the recoil protons' energies indicated a of mass approximately equal to the proton, demonstrating a strong neutron-proton interaction. These experiments also revealed the binding of the deuteron, the simplest composed of one proton and one neutron, with a of about 2.2 MeV, providing direct evidence of an attractive force between unlike nucleons. In 1932, proposed a theoretical framework for nuclear binding, suggesting the as a tightly bound proton-electron composite and attributing the nuclear force to the quantum mechanical exchange of electrons between protons, analogous to exchange forces in molecular bonds; this model, though later revised after confirming the 's elementary nature, introduced the concept of exchange mechanisms for short-range nuclear interactions. By 1936, and Robert F. Bacher's comprehensive review synthesized these developments, emphasizing empirical evidence for nuclear forces, including neutron-proton scattering data. Notably, measurements of neutron-proton radiative capture cross-sections, which showed anomalously large values for slow neutrons (on the order of barns), indicated a short-range attractive potential that enhanced low-energy interactions, underscoring the force's non-electromagnetic character and limited range of about 1-2 femtometers.

Key theoretical advances

The discovery of the in 1947 by Cecil F. Powell and collaborators, through emulsion experiments, provided experimental confirmation of Hideki Yukawa's 1935 hypothesis that a mediates the nuclear force, with the 's mass of approximately 140 MeV aligning closely with Yukawa's prediction for particle. In the 1950s, -nucleon scattering experiments, such as those conducted at the University of Chicago's synchrocyclotron in 1951, established the nature of the coupling to nucleons, supporting Yukawa's and enabling more accurate models of the strong interaction. Additionally, the 1956 proposal by and Chen-Ning Yang of parity nonconservation in weak interactions, experimentally verified in 1957 by and others, prompted a reevaluation of principles in strong force models, influencing the incorporation of chiral symmetries in -nucleon interactions. The 1960s marked a with the independent proposals of the by and in 1964, positing that protons and neutrons consist of three , which provided a deeper substructure for understanding the nuclear force as a residual effect of quark interactions. To resolve issues with quark statistics and identical particle behavior, Oscar W. Greenberg introduced as a three-valued in 1964, laying the groundwork for the color degree of freedom essential to strong interactions. In 1973, Harald Fritzsch, , and Heinrich Leutwyler formulated (QCD) as the of the strong interaction, incorporating and gluons as mediators between quarks. That same year, and , along with independently David Politzer, demonstrated in non-Abelian gauge theories like QCD, explaining how the strong force weakens at short distances (high energies) and strengthens at longer distances, thus unifying perturbative and non-perturbative regimes for nuclear force descriptions. During the 1990s, advances in simulations, building on Wilson's foundational work, enabled calculations of nuclear forces directly from QCD, with early quark-included computations toward the decade's end providing insights into interactions beyond phenomenological models. More recently, the HAL QCD collaboration has extended these simulations to finite nuclear densities, offering updated potentials relevant to dense matter.

Theoretical Framework

Residual strong interaction

The nuclear force, which binds protons and neutrons within atomic nuclei, emerges as a residual effect of the described by (QCD). In QCD, the fundamental strong force is mediated by gluons, which carry and couple quarks— the building blocks of hadrons— in a non-Abelian , leading to the phenomenon of where quarks are perpetually bound within colorless hadrons such as nucleons. This confinement ensures that isolated quarks cannot exist, and the resulting hadrons interact via residual color forces that manifest at larger scales. The residual strong interaction arises from the exchange of quarks and gluons between neighboring nucleons, analogous to van der Waals forces in , where fluctuating dipoles induce s between neutral atoms. Specifically, when two color-neutral nucleons approach each other, their constituent quarks can exchange color-charged gluons or intermediate states like mesons, but the overall exchange must preserve color neutrality for the composite systems, resulting in an effective at distances of about 1-2 femtometers. This process effectively transfers momentum and energy between nucleons without violating QCD's principle. A key distinction in QCD lies in the scale separation between the microscopic quark-gluon dynamics at short distances (high energies, ~1 GeV) and the effective low-energy regime governing nuclear structure (~100 MeV), where perturbative QCD breaks down, necessitating effective field theories to bridge the gap. Recent simulations have provided insights into how spontaneous — a effect where the approximate of massless quarks is broken by the QCD , generating light masses— influences the residual nuclear force, particularly in constraining in-medium interactions and forces that contribute to nuclear saturation. These computations demonstrate that restoration at high densities alters the nuclear potential, linking fundamental QCD mechanisms directly to observable nuclear properties.

Meson exchange theory

The meson exchange theory posits that the nuclear force arises from the exchange of mesons between nucleons, providing an effective description that bridges the underlying (QCD) with phenomenological models of nucleon interactions. In 1935, proposed that this force is mediated by the exchange of massive particles, termed mesons, which carry the strong interaction over short distances, characterized by a g that quantifies the strength of the nucleon-meson vertex. This idea successfully explained the short range of the nuclear force, limited by the mesons' mass, and anticipated the discovery of pions as the lightest mesons responsible for the longest-range component. The one-pion exchange (OPE) represents the dominant long-range contribution to the nuclear force, arising from the virtual exchange of a single between s, and it exhibits strong and dependence. The OPE potential is in nature, leading to both central and tensor components that influence and bound states. The standard form of the OPE potential in coordinate is V_{\text{OPE}}(r) = -\frac{f_{\pi NN}^2}{4\pi} (\boldsymbol{\tau}_1 \cdot \boldsymbol{\tau}_2) \left[ \boldsymbol{\sigma}_1 \cdot \boldsymbol{\sigma}_2 \, Y(m_\pi r) + S_{12} \, T(m_\pi r) \right], where Y(x) = \frac{e^{-x}}{x}, T(x) = \left(1 + \frac{3}{x} + \frac{3}{x^2}\right) \frac{e^{-x}}{x}, x = m_\pi r, f_{\pi NN} is the dimensionless pion-nucleon (with f_{\pi NN}^2 / 4\pi \approx 0.075), and S_{12} = 3 (\boldsymbol{\sigma}_1 \cdot \hat{r})(\boldsymbol{\sigma}_2 \cdot \hat{r}) - \boldsymbol{\sigma}_1 \cdot \boldsymbol{\sigma}_2 is the . This tensor term is particularly crucial, as it drives the mixing between spin-singlet and spin-triplet states in deuteron-like configurations, and at short distances (r \ll 1/m_\pi), it exhibits a singular $1/r^3 behavior. For shorter ranges, the theory incorporates multi-pion exchanges, which account for intermediate-distance attractions, as well as contributions from heavier vector mesons such as the rho (\rho) and (\omega), which introduce repulsive cores at very short distances below 1 fm. The rho meson exchange provides an isovector tensor force, while the omega contributes a central isoscalar repulsion, both essential for reproducing the empirical hard core in nucleon-nucleon potentials. These heavier meson exchanges, with masses around 770 MeV for rho and 783 MeV for omega, ensure the potential's rapid falloff at small separations, aligning with scattering data. Extensions of meson exchange theory within (ChPT) provide a systematic, low-energy effective field theory framework, deriving the interactions from a chiral-invariant that respects the approximate (2)_L × (2)_R of QCD. In ChPT, the pion-nucleon coupling is expanded in powers of momentum over the chiral breaking scale (~1 GeV), with the corresponding to the OPE and higher orders including multi-pion loops and contact terms. Recent advances in chiral effective field theory, including relativistic formulations, have enhanced predictions for nuclear interactions and structure calculations as of 2025. This approach renormalizes the theory order by order, improving predictions for low-energy nucleon-nucleon scattering phases and enabling connections to simulations.

Interaction Models

Yukawa potential

The represents the simplest phenomenological model for the nuclear force, proposed by in 1935 as arising from the exchange of a massive scalar particle, termed a "meson," between nucleons. This model extended the concept of field-mediated interactions, analogous to , to account for the short-range nature of the strong nuclear force. The potential takes the form V(r) = -\frac{g^2}{4\pi} \frac{e^{-m r}}{r}, where r is the distance between nucleons, g is the dimensionless coupling constant between the nucleon and the meson field, and m is the mass of the exchanged meson. This expression describes an attractive, centrally symmetric force that decays exponentially, with the range \hbar / (m c) determined by the meson's mass. This form derives from the static limit of the Klein-Gordon equation for a massive scalar field exchanged between two nucleons. The Klein-Gordon equation in the static case, (\nabla^2 - m^2) \phi = -\delta(\mathbf{r}), yields the Green's function solution \phi(r) = \frac{e^{-m r}}{4\pi r}, which, upon incorporating the coupling, produces the Yukawa potential. The finite mass m introduces the exponential screening, contrasting with the infinite-range Coulomb potential from massless photon exchange. Yukawa applied this potential to the deuteron, the simplest bound system with a of 2.224 MeV and size approximately 2 fm, to estimate the mass. Fitting the potential parameters to reproduce the deuteron's binding yielded m c^2 \approx 140 MeV, a made over a decade before the meson's experimental identification. Despite its foundational role, the simple scalar has notable limitations: it assumes a spin-independent, isoscalar , failing to capture the observed tensor and spin-orbit components of the nuclear force, as well as charge independence via . These shortcomings were later addressed by incorporating exchanges alongside the scalar term. Yukawa's 1935 proposal gained empirical validation in 1947 when Cecil Powell and collaborators discovered the charged (\pi^\pm) in cosmic-ray interactions using nuclear emulsions, confirming a mass of approximately 140 MeV and establishing it as the predicted mediator. This breakthrough earned Yukawa the 1949 and solidified exchange as a cornerstone of theory.

Modern nucleon-nucleon potentials

Modern nucleon-nucleon (NN) potentials are phenomenological or semi-phenomenological models constructed to accurately reproduce empirical NN scattering data, such as phase shifts from proton-proton (pp) and neutron-proton (np) interactions, deuteron properties, and low-energy parameters. These potentials incorporate the effects of spin and isospin dependencies, as well as a strong short-range repulsion to account for the finite size of nucleons and Pauli exclusion, while fitting data up to laboratory energies of several hundred MeV. Unlike simpler theoretical prototypes like the Yukawa potential, modern potentials use multi-parameter forms with local or nonlocal radial dependencies, enabling high-precision descriptions of the NN interaction across partial waves. A seminal example from the is the Reid soft-core potential, developed by fitting to early NN scattering phase shifts and low-energy data. This local potential includes central, tensor, and spin-orbit components, with a Gaussian form for the short-range repulsion to soften the core while avoiding unphysical singularities. The radial functions are parameterized as sums of Yukawa terms for the attractive meson-exchange-like parts, modulated by operator structures that depend on the total spin \mathbf{S}, orbital \mathbf{L}, and \boldsymbol{\tau}_1 \cdot \boldsymbol{\tau}_2. It achieved good agreement with np and pp data up to 350 MeV, influencing subsequent nuclear structure calculations. In the 1980s, the potential advanced this approach by blending meson-exchange theory with phenomenological elements, providing a nonlocal, momentum-dependent description fitted to NN phase shifts up to 330 MeV. Inspired by dispersion relations and \pi- and \rho-meson exchanges for longer ranges, it features a phenomenological core to model short-distance repulsion, with parameters adjusted to match data and deuteron observables like the and quadrupole moment. This potential improved predictions for higher partial waves and was widely used in few-body systems. High-precision potentials emerged in the , such as the Argonne v18 and CD-Bonn models, which fit vast datasets of and scattering phases up to 1 GeV, achieving \chi^2 per datum close to across thousands of data points. The Argonne v18 is a local potential with 18 components, including charge-dependent terms to capture small isospin-breaking effects from electromagnetic interactions and differences in versus forces; it uses Woods-Saxon derivatives for short-range repulsion and Yukawa forms for exchanges. Similarly, the CD-Bonn potential employs a nonlocal, meson-exchange framework with one-boson-exchange (OBE) terms for \pi, \eta, \rho, \omega, and others, plus phenomenological short-range components, explicitly incorporating charge dependence to fit and data separately. Both potentials reproduce the to within 0.1% and have \chi^2 \approx 1.09 and 1.01, respectively, for comprehensive databases. The general operator structure of these modern potentials expands the NN interaction in a basis of spin-isospin operators, typically written as V = \sum_{i=1}^{18} V_i(r) O_i, where the O_i include the central $1, tensor S_{12} = 3(\boldsymbol{\sigma}_1 \cdot \hat{\mathbf{r}})(\boldsymbol{\sigma}_2 \cdot \hat{\mathbf{r}}) - \boldsymbol{\sigma}_1 \cdot \boldsymbol{\sigma}_2, spin-orbit \mathbf{L} \cdot \mathbf{S}, and quadratic spin operators like (\boldsymbol{\sigma}_1 \cdot \mathbf{L})(\boldsymbol{\sigma}_2 \cdot \mathbf{L}), multiplied by isospin factors such as $1, \boldsymbol{\tau}_1 \cdot \boldsymbol{\tau}_2, and charge-breaking terms like \tau_{1z} \tau_{2z}. The radial functions V_i(r) are fitted to data, often combining exponential or Yukawa forms for attraction with repulsive cores. This structure captures the complexity of the strong force while respecting approximate isospin symmetry. More recent developments in the 2010s incorporate chiral effective field theory (EFT) to derive NN potentials systematically from QCD symmetries, with the Entem-Machleidt models providing high-fidelity fits up to next-to-next-to-next-to-leading order (N4LO). These potentials use pion-exchange diagrams for long-range attraction and contact terms for short-range physics, achieving accuracy comparable to phenomenological models (\chi^2 \approx 1.2 up to 450 MeV) while including explicit \Delta-resonance intermediate states to improve the description of tensor force and spin-orbit coupling. For instance, the N3LO version fits over 4300 pp and np data points with a deuteron binding energy of 2.2246 MeV, demonstrating convergence in the chiral expansion.

Role in Nuclear Structure

Binding energy contributions

The nuclear binding energy B(A,Z) for a nucleus with mass number A and atomic number Z is defined as the energy required to disassemble it into its constituent protons and neutrons, given by the formula
B(A,Z) = \left[ Z m_p + (A - Z) m_n - M(A,Z) \right] c^2,
where m_p and m_n are the masses of the proton and neutron, respectively, M(A,Z) is the mass of the nucleus, and c is the . This energy arises predominantly from the attractive nuclear force, which overcomes the repulsive Coulomb interactions between protons and provides the stability of the nucleus.
The (SEMF) approximates the as a sum of several s that capture different physical effects: a volume a_v A, a surface -a_s A^{2/3}, a -a_c Z(Z-1)/A^{1/3}, an asymmetry -a_a (A - 2Z)^2 / A, and a pairing that depends on whether A is even or odd. The nuclear force primarily governs the volume , reflecting its short-range attractive nature that binds nucleons uniformly throughout the nuclear volume, while the surface accounts for the reduced binding at the nuclear periphery due to fewer nucleon neighbors. The other s arise from electromagnetic repulsion, neutron-proton imbalance, and quantum statistical effects, but the nuclear force's contribution dominates the overall scale of binding. In the simplest two-body system, the deuteron (comprising a proton and ), the nuclear force yields a of 2.224 MeV through the attractive interaction in the spin-triplet state, demonstrating the force's role in stabilizing light nuclei against dissociation. For heavier nuclei, saturation occurs, with an average of approximately 8 MeV per , resulting from the balance between the nuclear force's medium-range attraction and its short-range repulsion, which prevents collapse and limits each 's interactions to nearest neighbors. This saturation explains the near-constant binding per across medium-mass nuclei. In light nuclei, such as the (³H), two-body nuclear forces alone underbind the system by about 1-2 MeV compared to the observed 8.48 MeV , a discrepancy known as the triton puzzle. Effective field theory (EFT) calculations in the 2020s, incorporating forces derived from chiral EFT, resolve this by adding contact terms that enhance binding, accurately reproducing the triton energy and related scattering data when fitted to empirical inputs. These three-body contributions, arising from multi-pion exchanges and short-range effects, become essential for precise descriptions beyond the two-body approximation.

Many-body nuclear systems

To describe many-body nuclear systems beyond simple two-nucleon interactions, the Brueckner-Hartree-Fock (BHF) approximation provides a microscopic framework for infinite by solving the Bethe-Goldstone to obtain the in-medium G-matrix, an effective interaction that resums short-range correlations from realistic nucleon-nucleon potentials. This approach calculates the per and saturation density of symmetric , typically yielding values around 16 MeV and 0.16 fm⁻³, in reasonable agreement with empirical data when including three-body effects. For nuclei with mass number A > 2, such as the triton, two-body forces alone fail to reproduce observed binding energies, requiring explicit three-body forces to provide the necessary additional attraction. A seminal example is the Fujita-Miyazawa two-pion-exchange three-body force, which arises from intermediate Δ-resonance excitation and contributes about 4-8 MeV per particle to the binding in , mitigating overbinding from purely two-body terms. Recent 2025 research highlights that three-nucleon forces significantly enhance spin-orbit splitting, widening energy gaps between nuclear shells by a factor of 2.5 in , thereby increasing nuclear stability, with effects strengthening in heavier nuclei. Mean-field approximations simplify the treatment of many-body effects by using effective, density-dependent interactions fitted to bulk nuclear properties. The Skyrme interaction, a zero-range force with density-dependent terms, enables self-consistent Hartree-Fock calculations of finite nuclei, reproducing ground-state energies and radii while incorporating rearrangement effects from the medium. Similarly, the finite-range Gogny interaction, parameterized with Gaussian terms, supports Hartree-Fock-Bogoliubov methods that simultaneously describe deformation, , and excitation spectra across the nuclear chart. In the , the cumulative effect of the strong nuclear force generates a single-particle potential with spin-orbit coupling, causing nucleons to occupy discrete orbits analogous to atomic electrons, which leads to enhanced stability at of protons or neutrons: 2, 8, 20, 28, 50, 82, and 126. These configurations, arising from filled subshells, explain discontinuities in binding energies and low-lying excited states near doubly magic nuclei like ⁴⁰Ca or ²⁰⁸Pb. Recent ab initio approaches, such as the no-core combined with chiral effective field theory interactions, solve the full many-body for light and medium-mass nuclei (up to A ≈ 50) without adjustable parameters, achieving spectroscopic quality results for binding energies and electromagnetic transitions in the 2020s by including up to next-to-next-to-next-to-leading-order potentials.

References

  1. [1]
    Nuclear Forces - Scholarpedia
    Sep 14, 2014 · Nuclear forces (also known as nuclear interactions or strong forces) are the forces that act between two or more nucleons.Historical perspective · QCD and the nuclear force · The chiral effective field theory...
  2. [2]
    DOE Explains...The Strong Force - Department of Energy
    The strong force is the force that holds subatomic particles together to form larger subatomic particles.
  3. [3]
    Forces - NASA Science
    Oct 22, 2024 · The strong nuclear force, or strong force for short, holds together the building blocks of atoms. It always attracts and works at two different ...Gravitational Force · Electromagnetic Force
  4. [4]
    Coupling Constants for the Fundamental Forces - HyperPhysics
    In attributing a relative strength to the four fundamental forces, it has proved useful to quote the strength in terms of a coupling constant. The coupling ...
  5. [5]
    Four Forces- Ranges and Carriers - Duke Physics
    Jan 29, 1995 · The weak nuclear force has a limit in range of only 10 to the -18th meters. This means that the carrier particles must indeed have mass.
  6. [6]
    What are nuclear forces? - Properties & Definition | CK-12 Foundation
    The nuclear force is about 100 times stronger than the electromagnetic force, but it only acts at distances of about 1 femtometer (10-15 meters), so its effects ...
  7. [7]
    [PDF] Insights into nuclear saturation density from parity- violating electron ...
    Sep 29, 2020 · Nuclear saturation implies that the interior density of heavy nuclei should be nearly constant and close to ρ0. Historically, the semi-empirical ...
  8. [8]
    [PDF] Introduction to Nuclear Forces - Physics
    The constant value of the average binding energy per nucleon for most nuclei indicates that the nuclear forces have the property of saturation. 4. Nuclear ...
  9. [9]
    [PDF] Chapter 12 Charge Independence and Isospin
    protons and neutrons the strong interactions, responsible for nuclear binding, are 'charge independent'. Let us now look at a pair of mirror nuclei whose ...
  10. [10]
    [PDF] Symmetries in Physics: Isospin and the Eightfold Way
    In the following section we will introduce the concept of isospin as an approximate SU(2) symmetry, which identifies the proton and the neutron as different ...
  11. [11]
    Modern theory of nuclear forces | Rev. Mod. Phys.
    Dec 21, 2009 · As far as the strong interactions are concerned, the different quarks u , d , s have identical properties, except for their masses. The quark ...
  12. [12]
    Coulomb-free 1S0 p − p scattering length from the quasi ... - Nature
    May 18, 2023 · The longstanding hypothesis of charge independence and charge symmetry of nuclear forces can be unveiled by fixing the low energy parameters of ...
  13. [13]
    [PDF] charge independence and charge symmetry - INIS-IAEA
    Introduction. This paper is concerned with charge independence and charge symmetry which provide powerful tools in organizing and describing the multiplet ...<|separator|>
  14. [14]
    [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 particles by matter and the structure of the ... pdf. This article may be used ...
  15. [15]
    The discovery of the neutron and its consequences (1930–1940)
    Heisenberg considered protons and neutrons to be different quantum states of the same particle. Hideki Yukawa analyzed the interaction of neutrons and protons, ...1. The Discovery Of The... · 2. Artificial Radioactivity · 3. Fission
  16. [16]
    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.
  17. [17]
    [PDF] The development of the concept of exchange forces in the 1930s
    Jan 31, 2024 · Nevertheless, in 1932 Heisenberg adopted the picture of the exchange of an electron, and this will lead over the years to the current ...
  18. [18]
    Nuclear Physics A. Stationary States of Nuclei | Rev. Mod. Phys.
    Nuclear Physics A. Stationary States of Nuclei. HA Bethe and RF Bacher. HA Bethe and RF Bacher Cornell University. PDF Share. Rev. Mod. Phys. 8, 82 – Published ...
  19. [19]
  20. [20]
  21. [21]
  22. [22]
    [PDF] Kenneth Wilson and lattice QCD - arXiv
    Jan 21, 2015 · Lattice QCD simulations including quark effects through hybrid Monte Carlo algorithm developed rapidly toward the end of the 90's. By the turn ...
  23. [23]
    [0811.1338] Modern Theory of Nuclear Forces - arXiv
    Nov 9, 2008 · Abstract: Effective field theory allows for a systematic and model-independent derivation of the forces between nucleons in harmony with the ...
  24. [24]
    Don't get too close | RIKEN
    Oct 12, 2007 · The force that holds atomic nuclei together is a complicated residual effect of the strong force between quarks and gluons—elementary particles ...
  25. [25]
    [2304.01036] Constraints on the in-medium nuclear interaction from ...
    Apr 3, 2023 · We demonstrate that chiral symmetry and Lattice-QCD provide coherent constraints on the in-medium nuclear interaction, suggesting a fundamental origin of the ...Missing: force | Show results with:force
  26. [26]
    basic ideas of chiral perturbation theory and meson exchange models
    Nov 7, 2007 · Basic ideas of chiral perturbation theory applied to nuclear forces are sketched. Applications in the two‐ and three‐nucleon systems are ...
  27. [27]
    On the Interaction of Elementary Particles. I - J-Stage
    Corresponding author. ORCID. JOURNAL FREE ACCESS. 1935 Volume 17 Pages 48-57. DOI https://doi.org/10.11429/ppmsj1919.17.0_48 ... Download PDF (12304K). Download ...
  28. [28]
    [PDF] Nucleon-Nucleon Interaction, Deuteron - UMD Physics
    YUKAWA INTERACTION AND ONE-PION EXCHANGE AT LONG DISTANCE. 101. Making an inverse Fourier transformation, one gets. V (r) = − g2. 4π e−mr r. ,. (6.10) which is ...
  29. [29]
    [nucl-th/0101056] The nucleon-nucleon interaction - arXiv
    Jan 25, 2001 · Abstract: We review the major progress of the past decade concerning our understanding of the nucleon-nucleon interaction.
  30. [30]
  31. [31]
  32. [32]
    Three-body forces and Efimov physics in nuclei and atoms
    Jan 22, 2025 · This review article presents historical developments and recent advances in our understanding on the three-body forces and Efimov physics.
  33. [33]
    [PDF] Nuclear effective field theory: status and perspectives - OSTI.GOV
    Mar 8, 2020 · We review each of the three main nuclear EFTs—Chiral, Pionless, Halo/Cluster—highlighting their similarities, dif- ferences, and connections.
  34. [34]
    [1808.09138] Relativistic Brueckner-Hartree-Fock in nuclear matter ...
    Aug 28, 2018 · Brueckner-Hartree-Fock theory allows to derive the G-matrix as an effective interaction between nucleons in the nuclear medium. It depends on ...Missing: seminal | Show results with:seminal
  35. [35]
    Nuclear Matter Equation of State in the Brueckner-Hartree-Fock ...
    Apr 4, 2024 · Access Paper: View a PDF of the paper titled Nuclear Matter Equation of State in the Brueckner-Hartree-Fock Approach and Standard Skyrme ...Missing: seminal | Show results with:seminal
  36. [36]
    Two-Pion-Exchange Three-Body Force in Nuclear Matter
    Dec 1, 1970 · The net effect of the two-pion-exchange three-body force is rather sensitive to the cutoff and is about a 4- to 8-MeV per particle attraction in ...
  37. [37]
    The Skyrme Interaction in finite nuclei and nuclear matter - arXiv
    Jul 3, 2006 · Access Paper: View a PDF of the paper titled The Skyrme Interaction in finite nuclei and nuclear matter, by J. R. Stone and P.-G. Reinhard.Missing: original | Show results with:original
  38. [38]
    [1807.02518] Mean field and beyond description of nuclear structure ...
    Jul 6, 2018 · The Gogny force is a referent in the theoretical description of nuclear structure phenomena. Its phenomenological character manifests in a simple analytical ...
  39. [39]
    High-Precision Nuclear Forces From Chiral EFT: State-of-the-Art ...
    We review a new generation of nuclear forces derived in chiral effective field theory using the recently proposed semilocal regularization method. We outl.