Fact-checked by Grok 2 weeks ago

Strong interaction

The strong interaction, one of the four fundamental forces of nature alongside gravity, electromagnetism, and the weak interaction, is responsible for binding quarks together to form hadrons such as protons and neutrons, and for holding atomic nuclei together through its residual effects. Described by quantum chromodynamics (QCD), a non-Abelian gauge theory based on the SU(3)c color symmetry group, it acts exclusively on particles carrying "color charge," a property analogous to electric charge but with three types (red, green, blue) and their anticolors. Quarks, the fundamental constituents of matter, interact via the exchange of gluons—eight massless gauge bosons that themselves carry color charge, leading to self-interactions among gluons unlike in electromagnetism. This force dominates at subnuclear scales, with its intrinsic strength characterized by the running coupling constant αs, which has a value of approximately 0.118 at the Z boson mass scale. A defining feature of the strong interaction is asymptotic freedom, where the coupling strength decreases at short distances or high energies (below ~10-15 m or above ~200 MeV), allowing perturbative quantum field theory calculations for processes like deep inelastic scattering. In contrast, at larger distances or low energies, the interaction exhibits confinement, preventing free quarks or gluons from existing in isolation; instead, they are perpetually bound into color-neutral hadrons, explaining why quarks are never observed singly in experiments. QCD, formulated in the early 1970s, successfully predicts these behaviors through its Lagrangian, which includes quark kinetic terms, gluon field strengths, and Yukawa-like quark-gluon couplings, with no additional free parameters beyond αs and quark masses (the latter arising from electroweak interactions). The residual strong interaction between composite hadrons, such as protons and neutrons (collectively nucleons), arises from the underlying quark-gluon dynamics and is mediated primarily by the exchange of light mesons like pions. This nuclear force has a limited range of about 1.5 × 10-15 m (1.5 femtometers), comparable to nuclear diameters, due to the finite mass of the pion (~135 MeV/c²), and is attractive at distances of 0.5–2 fm while becoming repulsive at closer ranges to prevent nucleon overlap. Relative to other forces, its coupling between protons is roughly 10 times stronger than the electromagnetic repulsion (coupling constant ~2.5 × 10-27 J·m versus 2.31 × 10-28 J·m), enabling stable nuclei despite positive charges, and vastly exceeds the weak force (~104 times stronger) and gravity (~1037 times stronger). Charge independence approximately holds, treating protons and neutrons similarly, though small differences arise from electromagnetic corrections and quark mass asymmetries. Beyond atomic nuclei, the strong interaction governs high-energy phenomena like jet production in particle collisions and the quark-gluon plasma state achieved in heavy-ion experiments at facilities such as the LHC. Its non-perturbative aspects, including hadronization and nuclear structure, are studied via lattice QCD simulations on supercomputers, confirming predictions like the proton's mass (~938 MeV) being almost entirely due to QCD binding energy rather than quark masses. Ongoing research explores connections to chiral symmetry breaking, where the near-masslessness of up and down quarks leads to Goldstone bosons (pions) as effective degrees of freedom at low energies.

Fundamentals

Definition and characteristics

The strong interaction, also known as the strong nuclear force, is one of the four fundamental forces of nature, governing the behavior of subatomic particles at the smallest scales. It is the mechanism that binds quarks together to form hadrons, such as protons and neutrons, and is mediated by massless gauge bosons called gluons. Unlike the electromagnetic force, which acts on electrically charged particles, the strong interaction exclusively affects particles that carry a property known as color charge—quarks possess one of three types (red, green, or blue), while antiquarks carry anticolors, and gluons carry both a color and an anticolor, enabling self-interactions. Key characteristics of the strong interaction include its extreme strength and limited range. It is the most powerful of the fundamental forces, approximately 100 times stronger than the electromagnetic interaction at short interquark distances. The force operates over a very short range of about $10^{-15} meters (1 femtometer), comparable to the size of atomic nuclei, beyond which it drops off rapidly due to the massive effective mediators in the residual form and confinement effects at the quark level. The interaction is always attractive between color-charged particles in configurations that form color-neutral bound states, such as quark-antiquark pairs or three-quark combinations, but it exhibits non-central (tensor and spin-dependent) components and velocity-dependent behavior arising from relativistic effects and gluon exchanges. In nature, the strong interaction plays a crucial role in the structure of matter by confining quarks within hadrons, preventing free quarks from existing in isolation, and thus ensuring the stability of protons and neutrons. Additionally, a residual strong interaction emerges between these color-neutral hadrons, manifesting as the nuclear force that binds protons and neutrons together in atomic nuclei, overcoming electromagnetic repulsion despite its much weaker effective strength at that scale. At short interquark distances, the potential energy between a quark and an antiquark can be approximated perturbatively as a Coulomb-like form: V(r) \approx -\frac{4}{3} \frac{\alpha_s}{r}, where r is the separation, and \alpha_s is the strong coupling constant, analogous to the fine-structure constant in electromagnetism but running with energy scale.

Comparison with other fundamental forces

The strong interaction, also known as the strong nuclear force, is the most powerful of the four fundamental forces, dominating interactions at subatomic scales within atomic nuclei, while the electromagnetic, weak, and gravitational forces govern phenomena at larger distances or different particle properties. Unlike the infinite-range electromagnetic and gravitational forces, the strong force operates over extremely short distances, approximately 1 femtometer (fm), which is the typical size of an atomic nucleus, and effectively vanishes beyond this range due to color confinement. In contrast, the weak force has an even shorter range of about $10^{-18} meters, mediating processes like beta decay, while gravity, though universal, is negligible at nuclear scales. To illustrate these differences, the following table summarizes key properties of the four forces, based on their relative strengths:
ForceRelative StrengthRangeMediator(s)Charge Type
Strong1~1 fm ($10^{-15} m)Gluons (8 types)Color (red, green, blue and anticolors)
Electromagnetic$10^{-2}InfinitePhotonElectric (±)
Weak$10^{-6}~$10^{-18} mW^+, W^-, Z bosonsFlavor (weak isospin/hypercharge)
Gravitational$10^{-38}InfiniteGraviton (hypothetical)Mass-energy
A defining feature of the strong interaction is its confinement of color-charged particles (quarks and gluons), preventing free quarks from existing in isolation, unlike the infinite-range gravitational force that acts cumulatively on all masses without such binding. This confinement arises because gluons, which carry color charge themselves, lead to a linear increase in potential energy with distance, resulting in a constant force that does not diminish like the $1/r^2 decay of electromagnetism. Consequently, at distances greater than about 1 fm, the strong force between color-neutral hadrons (like protons and neutrons) drops to nearly zero, allowing the electromagnetic repulsion between protons to become relevant and limiting nuclear sizes. These contrasts underscore why the strong interaction binds quarks into protons and neutrons, enabling stable nuclei, while the other forces shape atomic, molecular, and cosmic structures.

Historical development

Early observations and models

The Geiger-Marsden experiments, conducted between 1908 and 1913, provided early evidence for the existence of a compact atomic nucleus by observing the scattering of alpha particles from thin metal foils, which implied a strong attractive force must counteract electromagnetic repulsion to maintain nuclear stability. These observations highlighted the nucleus's small size and high density, setting the stage for recognizing a short-range binding mechanism distinct from gravity or electromagnetism. In 1932, James Chadwick discovered the neutron through experiments bombarding beryllium with alpha particles, producing uncharged radiation that ejected protons from paraffin with energies consistent with a particle of mass approximately equal to the proton. This neutral particle explained how nuclei with multiple protons could remain stable without excessive positive charge, attributing the binding to a powerful short-range strong nuclear force acting between protons and neutrons. Cosmic ray studies in the early 1930s further revealed signatures of strong interactions; for instance, Walther Bothe and Herbert Becker's 1930 experiments showed that alpha particles on beryllium produced highly penetrating neutral radiation, initially misinterpreted as gamma rays but later understood as evidence of nuclear reactions mediated by the strong force. These findings, replicated and extended by Irène and Frédéric Joliot-Curie in 1932, demonstrated the force's role in inducing disintegrations in light elements, underscoring its short range of about 10^{-15} meters. During the 1930s and 1940s, the Berkeley cyclotron, developed by Ernest Lawrence, enabled controlled nuclear reactions by accelerating protons to energies up to several MeV, producing artificial radioactivity and confirming the strong force's dominance in low-energy nucleon scattering and binding. Key results included the 1934 observation of neutron-induced reactions by Lawrence and his collaborators, which probed the force's saturation and charge independence between protons and neutrons. In 1935, Hideki Yukawa proposed a meson-exchange model for the nuclear force, theorizing that protons and neutrons interact via the exchange of a massive particle (later identified as the pion) with a mass around 200 times that of the electron, yielding a Yukawa potential V(r) = - (g^2 / r) e^{-μr} that matched the force's short range. This phenomenological approach successfully described the deuteron's binding and nucleon-nucleon scattering at low energies. The model's prediction was experimentally confirmed in 1947 when Cecil Powell, Cesare Lattes, and Giuseppe Occhialini observed charged pions decaying into muons in cosmic ray-exposed photographic emulsions at high altitude, revealing particles with the anticipated mass of about 140 MeV/c² and short lifetime. These Bristol experiments distinguished pions from previously detected muons, validating Yukawa's mediator for the strong force. Following the pion discovery, cosmic ray experiments in the late 1940s revealed a new class of particles exhibiting "strange" behavior, produced copiously in strong interactions but decaying via weaker processes, with lifetimes around 10^{-10} seconds. Notable were the V^0 particles (neutral kaons and lambda hyperons) observed by Rochester and Butler in 1947 using cloud chambers, and charged kaons identified in photographic emulsions. These strange particles, including K-mesons and hyperons like Λ and Σ, multiplied the known hadron spectrum and challenged existing models, as their production and decay rates violated naive strong interaction expectations. This led Abraham Pais to propose strangeness as a new quantum number conserved in strong interactions in 1953, with independent formulations by Murray Gell-Mann and Kazuhiro Nishijima explaining associated production. Early models incorporated isospin symmetry, introduced by Werner Heisenberg in 1932, to treat protons and neutrons as two states of a nucleon doublet and explain approximate charge independence in pion-nucleon interactions. However, these frameworks struggled with the growing number of baryon resonances observed in pion-nucleon scattering experiments during the 1950s, such as the Δ(1232) state, which required more complex symmetries beyond simple nucleon-pion exchanges. The influx of strange particles further necessitated larger symmetry groups; in 1961, Murray Gell-Mann and Yuval Ne'eman independently proposed the Eightfold Way, based on SU(3) flavor symmetry, classifying baryons and mesons into octets and decuplets that accommodated the observed hadron masses and quantum numbers with remarkable success.

Formulation of quantum chromodynamics

The quark model was independently proposed in 1964 by Murray Gell-Mann and George Zweig as a framework to classify hadrons and explain the observed symmetries in strong interaction spectra. Gell-Mann introduced three fundamental constituents called quarks—up, down, and strange—with fractional electric charges of +2/3 or -1/3, arranged in SU(3) flavor multiplets to account for baryons as three-quark states and mesons as quark-antiquark pairs. Zweig similarly described these particles, initially termed "aces," emphasizing their role in building higher-spin representations without violating the Pauli exclusion principle for identical fermions. Initially viewed as mathematical tools rather than physical entities due to the challenge of fractional charges, the model gained traction as it successfully predicted the existence of the Ω⁻ baryon, confirmed experimentally in 1964. A key extension came in 1964 when Oscar W. Greenberg proposed that quarks carry an additional hidden degree of freedom, a three-valued "color" charge (later formalized as red, green, and blue), to resolve symmetry issues in the quark model, such as the apparent violation of Fermi statistics in baryon wave functions. This color charge ensured that baryons are color singlets, with quarks combining in color-antisymmetric states, while allowing identical flavor-spin configurations. Although not immediately recognized as the basis for a gauge theory, this concept laid the groundwork for distinguishing quark interactions beyond flavor. Building on these ideas, quantum chromodynamics (QCD) emerged in 1973 as the gauge theory of the strong interaction, formulated by David J. Gross, Frank Wilczek, and H. David Politzer as an SU(3) non-Abelian gauge theory with color as the symmetry group. In this framework, quarks interact via the exchange of eight massless vector bosons called gluons, which carry both color and anticolor charges, leading to self-interactions among gluons unlike in quantum electrodynamics. The theory's viability hinged on the discovery of asymptotic freedom, where the strong coupling constant decreases at short distances (high energies), allowing perturbative calculations for high-energy processes, a result independently proven by Gross and Wilczek, and by Politzer, in calculations showing negative beta function contributions from gluon self-coupling. This breakthrough, recognized by the 2004 Nobel Prize in Physics, resolved the longstanding issue of treating strong interactions perturbatively at short ranges while explaining their strength at larger scales. Experimental evidence for quarks predated QCD's full formulation, stemming from deep inelastic scattering (DIS) experiments at the Stanford Linear Accelerator Center (SLAC) from 1968 to 1973, led by Jerome I. Friedman, Henry W. Kendall, and Richard E. Taylor. These studies bombarded protons with high-energy electrons, revealing point-like scattering centers within the proton consistent with spin-1/2 constituents carrying fractional charges, directly supporting the quark model's physical reality and scaling behavior predicted by parton theory. Their work, awarded the 1990 Nobel Prize in Physics, provided crucial validation that protons are composite structures of quarks. The existence of gluons was confirmed in 1979 through the observation of three-jet events in electron-positron annihilation at the PETRA collider at DESY, where experiments like TASSO and MARK-J detected back-to-back quark jets accompanied by a third collinear jet from gluon radiation. The angular distributions and event topologies matched QCD predictions for gluon emission, with the TASSO collaboration reporting clear evidence of non-two-jet topologies at center-of-mass energies around 27-30 GeV, solidifying QCD as the theory of the strong force. These milestones transformed the quark model into a rigorous quantum field theory, enabling precise descriptions of hadron structure and high-energy collisions.

Theoretical framework

Quantum chromodynamics overview

Quantum chromodynamics (QCD) is the quantum field theory that describes the strong interaction, formulated as a non-Abelian gauge theory based on the SU(3)_c symmetry group, where the subscript "c" denotes color, the quantum number analogous to electric charge in electromagnetism. This local gauge invariance under SU(3)_c transformations dictates the interactions among its fundamental constituents, distinguishing QCD from the Abelian U(1) structure of quantum electrodynamics. The theory is asymptotically free, allowing perturbative calculations at high energies while exhibiting non-perturbative behavior at low energies. The fundamental particles in QCD are quarks and gluons. There are six quark flavors—up, down, strange, charm, bottom, and top—each carrying one of three possible color charges (red, green, or blue) and transforming in the fundamental representation of SU(3)_c. Gluons, the mediators of the strong force, number eight and belong to the adjoint representation of SU(3)_c, enabling them to carry color charge themselves and interact with both quarks and other gluons, a hallmark of the non-Abelian nature. The dynamics of QCD are encoded in its Lagrangian density, given by \mathcal{L}_\text{QCD} = -\frac{1}{4} G^a_{\mu\nu} G^{a\mu\nu} + \sum_f \bar{q}_f (i \gamma^\mu D_\mu - m_f) q_f, where G^a_{\mu\nu} is the gluon field strength tensor for the eight gluon color indices a = 1 to $8, D_\mu is the covariant derivative incorporating the SU(3)_c gauge interactions, q_f represents the quark fields for each flavor f, and m_f are the quark masses. This form ensures gauge invariance and captures both the pure gluonic interactions and the quark-gluon couplings. A key feature of QCD is the running of the strong coupling constant \alpha_s(Q^2) = g_s^2/(4\pi), which depends on the energy scale Q^2 and decreases as Q^2 increases, reflecting asymptotic freedom. At high energies, such as near the Z-boson mass scale (m_Z \approx 91.2 GeV), \alpha_s(m_Z^2) = 0.1180 \pm 0.0009, enabling perturbative QCD calculations for processes like deep inelastic scattering. This scale dependence arises from the renormalization group equation, governed by the beta function \beta(\alpha_s).

Gluons and color charge

The strong interaction is governed by quantum chromodynamics (QCD), where quarks carry a property known as color charge, analogous to electric charge in electromagnetism but with three distinct types: red, green, and blue. This color charge transforms under the fundamental triplet representation of the SU(3) gauge group, ensuring that quarks interact via the exchange of gluons. Antiquarks, in contrast, carry the corresponding anticolors (antired, antigreen, antiblue), reflecting their transformation under the conjugate representation. Hadrons must be color-neutral, forming color singlets where the combination of quark colors sums to "white," such as three quarks of different colors in baryons or a quark-antiquark pair of matching color-anticolor in mesons. Gluons serve as the mediators of the strong force in QCD, consisting of eight massless vector bosons with spin 1. Unlike photons in quantum electrodynamics, which are color-neutral, each gluon carries a combination of one color and one anticolor (e.g., red-antigreen or blue-antired), corresponding to the eight generators of the SU(3) adjoint representation. This color-anticolor structure, derived from the tracelessness of the Gell-Mann matrices, allows gluons to couple not only to quarks but also to each other, enabling self-interactions that distinguish QCD as a non-Abelian gauge theory. The self-coupling of gluons introduces nonlinear effects into the strong interaction, manifesting in Feynman diagrams through quark-gluon vertices and gluon self-interaction vertices. At the quark-gluon vertex, a gluon attaches to a quark line, changing its color while conserving overall color charge, similar to photon-quark coupling but with color indices summed over. Triple-gluon vertices depict three gluons meeting at a point, with coupling strength proportional to the strong coupling constant g_s, while four-gluon vertices involve quartic interactions. These self-interactions lead to complex dynamics, including anti-screening of color charges at short distances (asymptotic freedom) and enhancement at long distances (confinement), fundamental to QCD's behavior. In the QCD vacuum, the kinetic energy associated with gluons and quarks contributes significantly to the mass of hadrons, with the proton's mass arising predominantly from QCD binding energy rather than the rest masses of its constituent quarks. Lattice QCD calculations decompose the proton mass as approximately 9% from quark rest masses, 32% from quark kinetic and helicity energies, 37% from gluon field energy, and 23% from the trace anomaly. This emergent mass highlights the non-perturbative nature of the strong interaction, where virtual gluon fluctuations in the vacuum generate binding energy that dominates the proton's 938 MeV rest mass.

Key phenomena

Asymptotic freedom and confinement

The strong interaction exhibits two contrasting behaviors depending on the energy scale, which are central to quantum chromodynamics (QCD): asymptotic freedom at short distances (high energies) and confinement at long distances (low energies). Asymptotic freedom refers to the phenomenon where the strong coupling constant \alpha_s decreases as the energy scale Q increases, allowing quarks and gluons to interact more weakly at very short distances, akin to free particles. This behavior enables perturbative QCD calculations in high-energy processes, such as those observed in particle colliders like the Large Hadron Collider, where jet production and deep inelastic scattering align with predictions. The running of the coupling constant is described at leading order by the formula \alpha_s(Q^2) \approx \frac{12\pi}{(11 N_c - 2 N_f) \ln(Q^2 / \Lambda^2)}, where N_c = 3 is the number of colors, N_f is the number of active quark flavors (typically 3 to 6 depending on the scale), and \Lambda is a non-perturbative scale parameter. This logarithmic decrease arises from the negative beta function in QCD, driven by gluon self-interactions that lead to antiscreening of color charge, unlike the screening in QED. Experimental determinations of \alpha_s from event shapes in e^+ e^- collisions and other observables confirm this running down to scales around 1 GeV, with \alpha_s(M_Z) \approx 0.118 evolving to larger values at lower energies. In contrast, confinement manifests at low energies, where quarks and gluons cannot exist as isolated, free particles; instead, the strong force binds them into color-neutral hadrons such as protons and mesons. This is characterized by a linear quark-antiquark potential V(r) \approx \sigma r at large interquark separations r, where \sigma \approx 1 GeV/fm is the string tension, reflecting the formation of thin color flux tubes that transmit the force like a vibrating string. The confinement scale is set by \Lambda_\mathrm{QCD} \approx 200--$300 MeV, below which perturbative methods fail and the coupling becomes non-perturbative, a phenomenon termed "infrared slavery" due to the increasing strength from gluon self-interactions at long distances. Lattice QCD simulations provide strong evidence for this, reproducing the linear potential and flux tube profiles in the vacuum, with string tensions extracted from Wilson loops matching phenomenological models of hadron spectroscopy.

Quark-gluon plasma

The quark-gluon plasma (QGP) is a state of matter characterized by extreme temperatures exceeding $10^{12} K and energy densities greater than 1 GeV/fm³, in which quarks and gluons are deconfined and can propagate freely over distances larger than 1 fm, the typical size of hadrons. In this phase of quantum chromodynamics (QCD), the strong interaction's usual confinement mechanism is overcome due to the high thermal energy, allowing these fundamental particles to behave collectively as a plasma rather than being bound within hadrons. This state is believed to have dominated the early universe during the first approximately $10^{-6} seconds after the Big Bang, when the universe cooled from temperatures around 150–170 MeV. On Earth, QGP is recreated in laboratory settings through ultrarelativistic heavy-ion collisions, such as those at the Relativistic Heavy Ion Collider (RHIC), which began operations in 2000, and the Large Hadron Collider (LHC), where confirmations of its formation were reported in 2010. These collisions compress and heat nuclear matter to the required conditions, producing a short-lived QGP volume on the order of a few fm³. Key properties of the QGP include its behavior as a near-ideal relativistic fluid with remarkably low shear viscosity, quantified by the ratio \eta/s \approx 1/(4\pi), where \eta is the shear viscosity and s is the entropy density; this value, initially predicted from the AdS/CFT correspondence for strongly coupled gauge theories, has been experimentally verified to be close to the minimum possible for quantum systems. Another signature is jet quenching, where high-energy jets of partons traversing the medium lose significant energy through interactions with the dense quark-gluon soup, providing a direct probe of the strong interaction's opacity in this deconfined state. The transition from the confined hadronic phase to the QGP occurs as a smooth crossover rather than a strict first-order phase change at a pseudocritical temperature T_c \approx 150–170 MeV, as determined from lattice QCD simulations that incorporate physical quark masses. These non-perturbative computations reveal a rapid but continuous increase in energy density and pressure near T_c, marking the onset of deconfinement and chiral symmetry restoration.

Manifestations

Within hadrons

The strong interaction binds quarks into composite particles known as hadrons, which are color-neutral states to satisfy the requirement of color confinement in quantum chromodynamics (QCD). Baryons, such as the proton composed of two up quarks and one down quark (uud), consist of three quarks (qqq) arranged in a color singlet configuration. Mesons, like the positively charged pion (\pi^+ = u\bar{d}), are formed by a quark-antiquark pair (q\bar{q}) also in a color singlet state. These structures ensure that the overall color charge is zero, as required for physical hadrons. In the constituent quark model, quarks acquire effective masses from the strong binding, with light (up and down) quarks having constituent masses of approximately 300–350 MeV and the strange quark around 500 MeV. This model accurately reproduces experimental magnetic moments of hadrons; for instance, the predicted proton magnetic moment is 2.79 nuclear magnetons (\mu_N), closely matching the measured value of 2.793 \mu_N. The quark model also successfully describes the mass spectra of ground-state hadrons, such as the octet of light baryons and the pseudoscalar and vector meson nonets, aligning with spectroscopic data from experiments. At short distances (less than 0.1 fm), the strong interaction between quarks is dominated by one-gluon exchange, which acts as an attractive, Coulomb-like potential modulated by color factors. At larger distances (around 1 fm), non-perturbative confinement effects take over, linearly rising the potential and binding the quarks into stable hadrons with typical sizes of about 1 fm, as determined from scattering experiments. This dual behavior underscores the scale-dependent nature of QCD within hadrons. Deep inelastic scattering (DIS) experiments probe the internal structure of the proton through its structure functions, such as F_2(x, Q^2), revealing the momentum distribution among partons. Global fits to DIS and other data, like the CT18 analysis, show that valence quarks carry about 46% of the proton's longitudinal momentum, while gluons contribute approximately 39% and sea quarks (non-valence quarks and antiquarks) account for the remainder, around 15%, highlighting the dominant role of gluons and sea partons in the proton's momentum budget at typical scales (Q^2 \approx 1.3 GeV²).

Between hadrons as nuclear force

The residual strong interaction between hadrons, often referred to as the nuclear force, arises as a secondary effect of the underlying quark-level strong dynamics and binds protons and neutrons (collectively known as nucleons) into atomic nuclei. This force is significantly weaker—approximately 1% of the full strength of the strong interaction at shorter quark confinement scales—due to the involvement of massive meson intermediaries that limit its range to about 1–3 fm. Primarily mediated by the exchange of pions, the lightest mesons, the nuclear force follows a Yukawa potential form, given by V(r) = -\frac{g^2}{4\pi} \frac{e^{-m_\pi r}}{r}, where g is the pion-nucleon coupling constant, m_\pi \approx 138 MeV/c^2 is the pion mass, and r is the internucleon separation; this exponential decay ensures the force's short range, contrasting with the infinite range of massless photon-mediated electromagnetism. The nuclear force exhibits distinct properties that enable stable nuclear structure. It is attractive over distances of 1–2 fm, facilitating binding, but features a strong repulsive core at separations below 0.7 fm to prevent nucleons from overlapping excessively and maintain nuclear saturation. Additionally, the force demonstrates approximate charge independence, treating protons and neutrons nearly identically despite their electromagnetic differences; this symmetry is formalized through the isospin formalism, where protons and neutrons form an isospin doublet with total isospin I=1/2, allowing the strong interaction to conserve isospin in lowest order. Theoretical models of the nuclear force build on meson-exchange pictures and effective field theories derived from quantum chromodynamics symmetries. In meson-exchange models, the potential arises from the exchange of light mesons beyond just pions, including the vector mesons \rho (isovector, mediating spin-dependent effects) and \omega (isoscalar, contributing to the central repulsion); prominent examples include the Paris potential, which incorporates pion and two-pion exchanges fitted to scattering data, and the Bonn full model, which extends one-boson-exchange to higher orders for improved precision in describing nucleon-nucleon interactions.90131-3) More modern approaches use chiral effective field theory (EFT), which systematically expands the nuclear potential in powers of momentum using pions as the dominant degrees of freedom while respecting chiral symmetry; this framework naturally generates both two-nucleon and many-body forces, with seminal developments enabling quantitative predictions of binding energies, such as the deuteron's ground-state binding of 2.2 MeV—the only stable two-nucleon bound state.90268-L) Experimentally, the nuclear force's characteristics are verified through neutron-proton scattering experiments, which measure phase shifts, scattering lengths, and cross-sections to constrain potential models, revealing the force's tensor and spin-orbit components essential for deuteron structure. The overall stability of nuclei is evidenced by the binding energy per nucleon curve, which rises from light elements, peaks near iron-56 (with ~8.8 MeV per nucleon, indicating maximum binding efficiency), and declines for heavier nuclei, underscoring the nuclear force's role in fusion and fission energetics.

Unification and extensions

Role in grand unified theories

Grand unified theories (GUTs) seek to unify the strong, weak, and electromagnetic interactions into a single gauge symmetry at high energies, embedding the Standard Model gauge group SU(3)c × SU(2)L × U(1)Y as a subgroup of a larger simple group. The simplest such model is the SU(5) theory proposed by Georgi and Glashow, where the strong SU(3)c is embedded within the 5×5 unitary matrices of SU(5), with one generation of quarks and leptons transforming under the 10 and \bar{5} representations, enabling baryon number violation through gauge boson exchange. A more comprehensive extension is the SO(10) model, introduced by Fritzsch and Minkowski, which embeds SU(3)c × SU(2)L × U(1)Y into the rank-five group SO(10); here, all fermions of one generation, including a right-handed neutrino, fit into a single 16-dimensional spinor representation, naturally accommodating lepton-quark unification.90334-6) In these GUTs, the three gauge couplings of the Standard Model—αs for the strong interaction, αW for the weak SU(2)L, and αem for electromagnetism—evolve via renormalization group equations (RGEs) and converge to a common value αG at the unification scale MG ≈ 1016 GeV, far beyond the electroweak scale. This running is governed by the one-loop beta functions, with the strong coupling αs decreasing at high energies due to asymptotic freedom, allowing the couplings to meet; in minimal SU(5), unification occurs without supersymmetry, while SO(10) models often incorporate supersymmetry to improve the precision of this convergence. At MG, the symmetry breaks via Higgs mechanisms to the Standard Model, with the strong force emerging as the unbroken SU(3)c subgroup. A hallmark prediction of GUTs is proton decay, mediated by heavy gauge bosons (X and Y in SU(5)) that violate baryon and lepton numbers, with the dominant mode p → e+π0 expected in minimal SU(5) to have a lifetime τp ≈ 1031 years. However, searches by the Super-Kamiokande experiment have set stringent lower limits, such as τp/B(p → e+π0) > 2.4 × 1034 years and τp/B(p → K+ν-bar) > 6.61 × 1033 years at 90% confidence level, based on exposures exceeding 0.3 Mton·year. These limits, accumulated since the 1990s, rule out the minimal non-supersymmetric SU(5) model, as its predicted decay rates exceed the observed bounds, though supersymmetric extensions and SO(10) variants remain viable with adjusted Higgs masses or intermediate scales. Beyond proton decay, GUTs predict lepton-quark unification that addresses neutrino masses; in SO(10), the inclusion of right-handed neutrinos in the 16 representation enables the seesaw mechanism, generating small neutrino masses mν ≈ mDirac2/MR where MR is near MG, consistent with observed oscillations and consistent with unification-scale relations like mb = mτ. Additionally, the symmetry breaking in GUTs produces magnetic monopoles via the 't Hooft-Polyakov mechanism, with masses on the order of MGG ≈ 1016–1017 GeV, whose relic density is diluted by cosmic inflation in modern models.90233-4)

Current challenges and open questions

One of the central challenges in understanding the strong interaction is the hierarchy problem, which questions why the QCD scale \Lambda_\mathrm{QCD} \approx 200 MeV is vastly smaller than the Planck scale M_\mathrm{Pl} \approx 10^{19} GeV, spanning over 15 orders of magnitude. This disparity arises from the renormalization group evolution of the strong coupling constant \alpha_s, where the initial value at the Planck scale determines \Lambda_\mathrm{QCD} through dimensional transmutation, but the precise value requires an apparent fine-tuning of the high-scale boundary conditions to avoid larger contributions. Additionally, the QCD vacuum energy contributes to the cosmological constant problem, where quantum fluctuations generate a vacuum energy density on the order of \Lambda_\mathrm{QCD}^4, yet observations demand an extraordinarily small effective value, implying cancellations at the level of 120 orders of magnitude between QCD and other sectors, including gravity. This fine-tuning lacks a fundamental explanation within the Standard Model and motivates extensions like supersymmetry or dynamical mechanisms to stabilize scales. The strong CP problem highlights a profound puzzle in QCD: the theory permits a CP-violating term \theta_\mathrm{QCD} \bar{q} i \gamma_5 q in the Lagrangian, where \theta_\mathrm{QCD} is an arbitrary parameter, yet experiments show no significant CP violation in strong processes. The most sensitive probe is the neutron electric dipole moment (EDM), bounded by d_n < 1.8 \times 10^{-26} e \cdot \mathrm{cm} (90% C.L.) from recent measurements, implying |\theta_\mathrm{QCD}| \lesssim 10^{-10}, an unexplained suppression by 10 orders of magnitude relative to naive expectations. The leading solution is the Peccei-Quinn mechanism, introducing a dynamical axion field a with decay constant f_a, where the potential V(a) \sim -\Lambda_\mathrm{QCD}^4 \cos(a/f_a + \theta_\mathrm{QCD}) relaxes \theta_\mathrm{eff} = \theta_\mathrm{QCD} + a/f_a to zero, solving the problem while predicting light pseudoscalar particles observable in cosmology or experiments. Despite extensive searches, no axions have been detected, leaving the problem unresolved and axion models constrained by astrophysical and collider bounds. Non-perturbative aspects of QCD pose significant computational challenges, particularly in accurately predicting hadron masses and spectra from first principles, as perturbation theory breaks down at low energies where confinement dominates. Lattice QCD, a discretized non-perturbative formulation on a hypercubic grid, has made substantial progress in the 2020s, achieving precision calculations of light hadron masses with uncertainties below 1%, such as the pion mass determined to ~0.2% accuracy in simulations incorporating dynamical quarks and fine lattices. For instance, recent lattice results for the charged-neutral pion mass splitting reach permille-level precision, validating chiral perturbation theory and aiding flavor physics. However, full control over systematic errors like finite-volume effects and chiral extrapolation remains demanding, limiting applications to heavier hadrons or multi-particle states, and requiring exascale computing resources for broader reliability. Several open questions persist in QCD, including a rigorous proof of the confinement mechanism, where quarks and gluons are eternally bound into color-neutral hadrons despite asymptotic freedom at short distances. While lattice simulations exhibit confinement via a linearly rising potential V(r) \sim \sigma r with string tension \sigma \approx (440 \mathrm{MeV})^2, no analytical demonstration exists within the full non-Abelian gauge theory, constituting a Millennium Prize problem tied to the Yang-Mills mass gap. In extreme conditions, such as neutron star interiors with densities exceeding nuclear matter by factors of 10, the QCD equation of state remains uncertain, with debates over phase transitions to quark matter or hyperonic phases potentially altering star radii and merger signals observed by LIGO/Virgo. Finally, integrating QCD with quantum gravity poses foundational challenges, as non-perturbative strong dynamics at \Lambda_\mathrm{QCD} must reconcile with Planck-scale effects, possibly through holographic dualities like AdS/QCD, but no consistent framework unites the theories without divergences or inconsistencies.

References

  1. [1]
    [PDF] 9. Quantum Chromodynamics - Particle Data Group
    Dec 1, 2023 · Quantum Chromodynamics (QCD) is a gauge field theory describing strong interactions of colored quarks and gluons, and is the SU(3) component of ...
  2. [2]
    Four Fundamental Interaction
    The theory of strong interactions among quarks and gluons is called Quantum Chromodynamics (QCD). A very successful model of describing this residual force is ...
  3. [3]
    [PDF] Chapter 4 Four Fundamental Interactions
    Table 4-1. Strength and range of the four fundamental forces between two protons. Note that the strong force acts between quarks by an exchange of gluons.
  4. [4]
    [PDF] Quark Interactions and QCD
    Bosons mediating the quark-quark interaction are called gluons. – Gluons are to the strong force what the photon is to the EM force. – Gluons carry a color ...
  5. [5]
    33.6 GUTs: The Unification of Forces – College Physics
    The strong force is carried by eight proposed particles called gluons, which are intimately connected to a quantum number called color—their governing theory is ...
  6. [6]
    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 ...
  7. [7]
    [PDF] CHAPTER 12: The Atomic Nucleus - SIBOR
    ∎ The nuclear force is often called the strong force. ≈ 1.2 × 10−15 m. m, or the fermi.
  8. [8]
    [PDF] Velocity-Dependent Nuclear Forces. - LSU Scholarly Repository
    NON-LOCAL AND VELOCITY DEPENDENT POTENTIAL. SQUARE-SQUARE WELL WITH THE SAME ... (6 ) The force is predominantly a central force, I.e . the tensor force ...
  9. [9]
    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.Missing: comparison | Show results with:comparison
  10. [10]
    The Standard Model | CERN
    Gravity is the weakest but it has an infinite range. The electromagnetic force also has infinite range but it is many times stronger than gravity. The weak and ...
  11. [11]
    DOE Explains...Quarks and Gluons - Department of Energy
    Gluons have no electric charge. Both quarks and gluons have three additional states of charge: positive and negative redness, greenness, and blueness.
  12. [12]
    Fundamental Forces - HyperPhysics Concepts
    The electromagnetic force holds atoms and molecules together. In fact, the forces of electric attraction and repulsion of electric charges are so dominant over ...Missing: strength | Show results with:strength
  13. [13]
    Science | Inquiring Minds | Questions About Physics - Fermilab
    Apr 28, 2014 · The simple answer is that the strength depends on the range over which it is acting. At short distances the strong force is weak and at long distances it is ...
  14. [14]
    On a diffuse reflection of the α-particles - Journals
    When β-particles fall on a plate, a strong radiation emerges from the same side of the plate as that on which the β-particles fall.Missing: force | Show results with:force
  15. [15]
    [PDF] On a Diffuse Reflection of the -Particles - IF-UFRJ
    On a Diffuse Reflection of the α -Particles. Author(s): H. Geiger and E. Marsden. Reviewed work(s):. Source: Proceedings of the Royal Society of London.
  16. [16]
    May 1932: Chadwick reports the discovery of the neutron
    May 1, 2007 · In May 1932 James Chadwick announced that the core also contained a new uncharged particle, which he called the neutron.
  17. [17]
    The discovery of the neutron and its consequences (1930–1940)
    Chadwick observed “that the mass defect of C 13 is known with sufficient accuracy to show that the energy of the [photon] emitted in this process cannot be ...
  18. [18]
    Lawrence and the Cyclotron - American Institute of Physics
    EXIT to see Lawrence and Livingston's original 1932 paper describing the cyclotron. Lawrence and Livingston around 1933.
  19. [19]
    [PDF] Ernest O. Lawrence - National Academy of Sciences
    The first demonstration of the cyclotron resonance princi- ple was reported at the Berkeley meeting of the National Acad- emy of Sciences, in the fall of 1930, ...<|separator|>
  20. [20]
    Nobel Prize in Physics 1949
    ### Summary of Yukawa's Model and Pion Discovery Confirmation
  21. [21]
    Discovery of the Pion --- 1947
    Perkins' observations, published in January 1947, were confirmed by Occhialini and Powell ... research emulsions in which the pion was discovered in 1947.
  22. [22]
    [PDF] Nucleons and their Structure; Hadrons and Isospin
    Nucleons are spin-1/2 fermions with isospin t=1/2. Proton has t3=+1/2, neutron t3=-1/2. Isospin is a powerful symmetry in nuclear physics.
  23. [23]
    [PDF] An introduction to the quark model arXiv:1205.4326v2 [hep-ph] 24 ...
    May 24, 2012 · This document contains a review on the quark model, prepared for lectures at the Niccolò Cabeo School at Ferrara in May 2012.
  24. [24]
    [PDF] Gell-Mann.pdf
    A highly promised approach is the purely dy- namical "bootstrap" model for all the strongly in- teracting particles within which one may try to de- rive ...
  25. [25]
    An SU$_3$ model for strong interaction symmetry and its breaking
    Jul 16, 2002 · Author(s), Zweig, G (CERN) ; Publication, 1964 ; Imprint, 21 Feb 1964 ; Number of pages, 80 ; Note, Table 7 in this preprint was corrected shortly ...
  26. [26]
    [PDF] The Discovery of Quarks* - SLAC National Accelerator Laboratory
    But the idea of fractional charges was fairly repulsive to physicists of the day; in his original paper, Gell-Mann even wrote that “a search for stable ...<|separator|>
  27. [27]
    Spin and Unitary-Spin Independence in a Paraquark Model of ...
    This paper by O.W. Greenberg discusses spin and unitary-spin independence in a paraquark model of baryons and mesons.Missing: Hadronic Color Isospin
  28. [28]
    Asymptotically Free Gauge Theories. I | Phys. Rev. D
    Asymptotically free gauge theories of the strong interactions are constructed and analyzed. The reasons for doing this are recounted.Missing: QCD | Show results with:QCD
  29. [29]
    Asymptotic freedom: An approach to strong interactions
    Asymptotic freedom allows extracting new consequences from strongly interacting quantum field theory, where the origin is an ultraviolet fixed point.
  30. [30]
    The Nobel Prize in Physics 2004 - NobelPrize.org
    The Nobel Prize in Physics 2004 was awarded jointly to David J. Gross, H. David Politzer and Frank Wilczek for the discovery of asymptotic freedom.
  31. [31]
    [PDF] DEEP INELASTIC SCATTERING - Nobel Prize
    The results showed that nucleons were roughly 10-13 cm in size, implying a distributed structure. The earliest MIT-SLAC studies, in which California Institute ...
  32. [32]
    [1409.4232] The Discovery of the Gluon - arXiv
    Sep 15, 2014 · Following our suggestion, the gluon was discovered at DESY in 1979 by TASSO and the other experiments at the PETRA collider. Comments: 11 ...
  33. [33]
    Discovery of Three-Jet Events and a Test of Quantum ...
    Sep 17, 1979 · We report the analysis of the spatial energy distribution of data for e + ⁢ e − → h a d r o n s obtained with the MARK-J detector at PETRA.Missing: experiment | Show results with:experiment<|separator|>
  34. [34]
    [PDF] 9. Quantum Chromodynamics - Particle Data Group
    May 31, 2024 · Quantum Chromodynamics (QCD), the gauge field theory that describes the strong interactions ... Lattice QCD of this Review and in Ref. [723]. The ...
  35. [35]
    [1604.08082] The QCD Running Coupling - arXiv
    Apr 27, 2016 · We review the present knowledge for \alpha_s, the fundamental coupling underlying the interactions of quarks and gluons in QCD.
  36. [36]
    The discovery of asymptotic freedom and the emergence of QCD
    Sep 7, 2005 · In this Nobel lecture I shall describe the turn of events that led to the discovery of asymptotic freedom, which in turn led to the formulation of QCD.
  37. [37]
    Infrared gluon propagator from lattice QCD: Results from large ...
    Aug 28, 2006 · The main conclusions from this investigation are that the lattice data are compatible with the DSE solution (2) for the infrared gluon dressing ...Missing: evidence | Show results with:evidence
  38. [38]
    Quark/Gluon Plasma: When Protons Melt - AIP Publishing
    Oct 1, 2024 · This very hot form of matter is called a quark/gluon plasma, or QGP. The energy density at which QGP can be formed is of order 1 GeV/fm3. One ...
  39. [39]
    Temperature measurement of Quark-Gluon plasma at different stages
    Oct 14, 2025 · The thermodynamic and hydrodynamic properties of the QGP are described by Quantum Chromodynamics (QCD) and can be studied in heavy-ion ...
  40. [40]
    A ten-year journey through the quark–gluon plasma and beyond
    Nov 9, 2022 · It constitutes a proof of quark deconfinement, as it implies that quarks can move freely over distances much larger than the hadron size.
  41. [41]
    Quark-Gluon Plasma - an overview | ScienceDirect Topics
    Quark-gluon plasma is defined as a high temperature phase of quantum chromodynamics (QCD) where quarks and gluons are deconfined, allowing for collective ...
  42. [42]
    The 1st microsecond of the Big Bang - EarthSky
    Jun 16, 2021 · The plasma in the earliest universe is called a quark–gluon plasma (QGP). Modern researchers believe it was present in the first 0.000001 second of the Big ...
  43. [43]
    RHIC | A New Area of Physics - Brookhaven National Laboratory
    RHIC has created a new state of hot, dense matter out of the quarks and gluons that are the basic particles of atomic nuclei.
  44. [44]
    LHC experiments bring new insight into primordial universe - CERN
    Nov 26, 2010 · Showing beyond doubt that we can produce and study quark gluon plasma will bring important insights into the evolution of the early Universe, ...
  45. [45]
    How Particles of Light May Be Producing Drops of the Perfect Liquid
    Aug 8, 2024 · Currently, the quark-gluon plasma can only be created in two places on Earth – the LHC and the Relativistic Heavy Ion Collider (RHIC, a DOE ...
  46. [46]
    [PDF] AdS/CFT correspondence and the Quark-Gluon Plasma
    η s. = h. 4π where η is the shear viscosity, s is the entropy per unit volume. This is valid in a large, but restricted, class of strongly coupled quantum ...
  47. [47]
    [PDF] Quark-Gluon Plasma - New Frontiers - arXiv
    Apr 8, 2008 · Can viscosity be even smaller, η/s < 1/4π? In fact, as Lublinsky and myself [6] discussed, the AdS/CFT gravity spectral densities predicts that ...
  48. [48]
    Journey through the Quark Gluon Plasma - CMS Experiment
    Aug 19, 2024 · As jets pass through the quark gluon plasma, they interact with it via the strong force, which results in energy loss and modification of their ...
  49. [49]
    Photon-tagged jet quenching in the quark-gluon plasma
    Oct 23, 2017 · They fragment into highly collimated “jets” of particles that in turn lose energy through a phenomenon known as “jet quenching”.
  50. [50]
    Quark matter formation in dense stellar objects | Pramana
    Lattice QCD calculations at zero baryon density indicate that the transition occurs at T c ∼ 150–170 MeV. The transition is likely to be second order or a cross ...
  51. [51]
    The order of the quantum chromodynamics transition predicted by ...
    Here we determine the nature of the QCD transition using computationally demanding lattice calculations for physical quark masses.
  52. [52]
    Selected results in lattice quantum chromodynamics
    The transition temperature and the equation of state are calculated. Section 4 deals with nonvanishing chemical potentials and magnetic fields. Results for the ...
  53. [53]
    None
    Summary of each segment:
  54. [54]
    [PDF] 18. Structure Functions - Particle Data Group
    Dec 1, 2023 · The proton and quark momentum fractions refer to. ¯. P+, and x covers the interval (-1,1). In the. ERBL domain the GPDs are generalisations of ...
  55. [55]
    [1912.10053] New CTEQ global analysis of quantum ... - arXiv
    Dec 20, 2019 · This paper presents new parton distribution functions (PDFs) from the CTEQ-TEA collaboration, using LHC data, and the main PDF set is CT18.
  56. [56]
    [2108.06768] Connected and Disconnected Sea Partons from CT18 ...
    Aug 15, 2021 · Abstract: The separation of the connected and disconnected sea partons, which were uncovered in the Euclidean path-integral formulation of ...
  57. [57]
    Nuclear Forces - Scholarpedia
    Sep 14, 2014 · Quantum chromodynamics (QCD) is the theory of strong interactions. It deals with quarks, gluons and their interactions and is part of the ...
  58. [58]
    [PDF] The nuclear force
    Jan 21, 2011 · Charge independence of nuclear force is manifested in properties of mirror nuclei. Mirror nuclei are these which have numbers of protons and ...
  59. [59]
    Nuclear Binding Energy - HyperPhysics
    The iron limit​​ Iron-56 is abundant in stellar processes, and with a binding energy per nucleon of 8.8 MeV, it is the third most tightly bound of the nuclides. ...
  60. [60]
    [PDF] 93. Grand Unified Theories - Particle Data Group
    May 31, 2024 · 93.1 The standard model. The Standard Model (SM) may be defined as the renormalizable field theory with gauge group.
  61. [61]
    [PDF] Naturalness and New Approaches to the Hierarchy Problem
    There are various levels to the problem, but the essential issue is that the observed Higgs mass is some sixteen orders of magnitude smaller than the apparent ...
  62. [62]
    [PDF] The Quantum Vacuum and the Cosmological Constant Problem
    In this paper we describe the historical and conceptual origin of the cosmological constant problem which is intimately connected to the vacuum concept in ...<|separator|>
  63. [63]
    Axions and the strong CP problem - NASA ADS
    The supersymmetrized axion solution of the strong CP problem introduces its superpartner the axino, which might have affected the evolution of the Universe ...Missing: θ_QCD | Show results with:θ_QCD
  64. [64]
    Lattice QCD Calculation of the Pion Mass Splitting | Phys. Rev. Lett.
    Feb 3, 2022 · We use the infinite volume reconstruction method to calculate the charged and neutral pion mass difference.
  65. [65]
    [PDF] The Present and Future of QCD
    Mar 1, 2023 · ... Hadron Spectroscopy ... description of the strong force is provided by the non-Abelian gauge theory Quantum Chromodynamics (QCD).