Fact-checked by Grok 2 weeks ago

Quark

A quark is a type of elementary particle and a fundamental constituent of matter, combining in groups to form composite particles known as hadrons, such as protons and neutrons, which are the building blocks of atomic nuclei. Quarks were independently proposed in 1964 by physicists Murray Gell-Mann and George Zweig as part of a theoretical model to explain the structure of hadrons under the strong nuclear force, with Gell-Mann introducing the term "quark" inspired by a line from James Joyce's Finnegans Wake. Experimental evidence for their existence came in 1968 from deep inelastic scattering experiments at the Stanford Linear Accelerator Center (SLAC), led by Jerome Friedman, Henry Kendall, and Richard Taylor, who observed point-like scattering consistent with substructure inside protons; this work earned them the 1990 Nobel Prize in Physics. Quarks are fermions with spin ½ and carry fractional electric charges of either +2/3 or -1/3 times the , distinguishing them from other particles like leptons. They also possess a property called —red, green, or blue—which mediates the strong force via gluons, ensuring quarks are perpetually confined within hadrons and cannot be observed in isolation, a phenomenon known as quark confinement. There are six distinct types, or "flavors," of quarks: up (u), down (d), strange (s), (c), (b), and top (t), each with an associated antiquark; the up and down quarks are the lightest and most common, forming ordinary matter, while the heavier flavors are short-lived and produced in high-energy conditions. Baryons like protons (uud) and neutrons (udd) consist of three quarks, whereas mesons comprise one quark and one antiquark. In the Standard Model of particle physics, quarks alongside leptons form all known matter, interacting through the electromagnetic, weak, and strong forces, with ongoing research at facilities like CERN probing their masses, mixing, and potential beyond-Standard-Model behaviors.

Classification

Flavors and Generations

In the Standard Model of particle physics, quarks are fundamental fermions classified into six distinct flavors: up (u), down (d), strange (s), charm (c), bottom (b), and top (t). These flavors represent different types of quarks, each characterized by specific quantum numbers that determine their interactions and roles in forming composite particles. The six quark flavors are organized into three s, reflecting a hierarchical structure in mass and properties. The first includes the lightest quarks, which primarily constitute protons and neutrons in ordinary matter. The second comprises the strange and quarks, while the third consists of the heavier and quarks. This generational organization arises from the symmetries and mixing patterns observed in weak interactions, with each pairing an up-type quark (positive charge) and a down-type quark (negative charge). Each quark flavor has a corresponding antiparticle, known as an antiquark, which possesses opposite quantum numbers, including , , and flavor-specific quantum numbers like or . Antiquarks combine with quarks to form , while three quarks form , contributing to the rich spectrum of hadrons. The necessity of six flavors stems from the need to explain the observed diversity of hadrons, as experimental discoveries of particles like the J/ψ (charm) and Υ (bottom) required additional flavors beyond the initial three to match the variety of meson and baryon states. Quarks of all flavors carry one of three color charges (red, green, or blue), enabling the strong force to bind them into color-neutral hadrons. The following table summarizes the quark flavors, their electric charges, and approximate masses (in the \overline{\rm MS} scheme at a scale of about 2 GeV for light quarks (u, d, s), at the quark mass scale for c and b, and the pole mass for t) as of the 2025 Particle Data Group review:
FlavorSymbolElectric ChargeApproximate Mass (GeV/c^2)
upu+\frac{2}{3}~0.0022
downd-\frac{1}{3}~0.0047
stranges-\frac{1}{3}~0.094
charmc+\frac{2}{3}~1.27
bottomb-\frac{1}{3}~4.18
topt+\frac{2}{3}~173

Valence Quarks and Exotic Hadrons

quarks represent the minimal number of quarks and antiquarks required to form a , carrying the primary quantum numbers that define the particle's identity. For instance, the proton consists of two s and one (uud), while the positively charged is composed of an up quark and an anti-down quark (u\bar{d}). Hadrons are classified into two main categories based on their valence quark content: baryons and mesons. Baryons, such as protons and neutrons, are fermions made of three valence quarks (qqq), resulting in a baryon number of +1 and half-integer spin. Mesons, like pions and kaons, are bosons formed from a quark-antiquark pair (q\bar{q}), with a baryon number of 0 and integer spin. Beyond these conventional structures, exotic hadrons challenge the standard quark model by incorporating more complex valence quark configurations. Tetraquarks consist of four quarks (e.g., qq\bar{q}\bar{q}), pentaquarks have five (qqq q\bar{q}), and hybrids involve quarks bound with gluonic excitations. A prominent tetraquark candidate is the Z(4430)^+, observed in the \psi(2S) \pi^+ channel and confirmed as a resonant state with a mass of approximately 4430 MeV. LHCb experiments at the Large Hadron Collider have discovered several pentaquarks, such as the P_c(4312)^+, P_c(4440)^+, and P_c(4457)^+, each with valence content uudc\bar{c} and masses around 4312 MeV, 4440 MeV, and 4457 MeV, respectively, observed in the J/\psi p system from \Lambda_b^0 decays. These discoveries, along with tetraquark states like the T_{cc}(3875)^+ (cc\bar{u}\bar{d}), highlight multiquark bindings near heavy-light meson thresholds, often interpreted as molecular states. The formation of hadrons, including exotics, adheres to fundamental conservation laws that govern quark combinations. (B) is conserved, with quarks assigned B = +1/3 and antiquarks B = -1/3, ensuring baryons have B = 1 and mesons B = 0; this extends to exotics like pentaquarks (B = 1) and tetraquarks (B = 0). (S), defined as S = -1 for the strange quark and S = +1 for its antiquark, is conserved in strong interactions, influencing flavor content in particles like kaons (e.g., K^+ = u\bar{s}, S = +1) and hyperons (e.g., \Lambda = uds, S = -1). These observations of valence quark structures and exotic states provide strong validation for the , demonstrating its predictive power for both ordinary and unconventional hadrons while prompting refinements to account for multiquark dynamics and gluonic contributions.

Historical Development

Proposal of the Quark Model

In the early , the rapid discovery of numerous hadrons through experiments created a complex "zoo" of particles that challenged existing theoretical frameworks in hadron spectroscopy. To organize these particles into coherent patterns, physicists developed symmetry-based classification schemes, notably the "eightfold way" proposed by in 1961, which utilized the SU(3) to group baryons and mesons into multiplets such as octets and decuplets based on their quantum numbers like and . This approach successfully predicted mass relations and decay patterns but left unexplained how the underlying structure could account for the observed symmetries without invoking composite constituents. In 1964, Gell-Mann extended the eightfold way by proposing a model where hadrons are composite structures built from three fundamental triplet representations of SU(3), which he termed "quarks"—up, down, and strange—with fractional electric charges of +2/3, -1/3, and -1/3, respectively, arranged to yield integer charges for hadrons. Independently, George Zweig at CERN formulated a similar idea in his internal reports, referring to the constituents as "aces" and emphasizing their role in explaining the additive quantum numbers of hadrons under SU(3) flavor symmetry, though he viewed them more as physical entities than mathematical tools. Baryons were described as three-quark states (qqq) in the symmetric decuplet or mixed-symmetry octet representations, while mesons were quark-antiquark pairs (q\bar{q}), providing a unified explanation for the hadron multiplets observed in spectroscopy. A key early success of the quark model was its prediction of a strangeness -3 in the decuplet, the Ω⁻ ( configuration), with a around 1680 MeV, which completed the SU(3) multiplet and was discovered shortly thereafter in August 1964 at using a 5 GeV proton beam on a target, confirming the model's structure. Despite this validation, the quark model faced significant initial resistance from the physics community, primarily due to the counterintuitive fractional charges, which violated the long-held assumption of integral charges for fundamental particles, and the absence of free quarks in experiments, implying an unexplained confinement mechanism that prevented their isolation. Gell-Mann himself initially regarded quarks as a calculational device rather than real particles, reflecting the skepticism, while the lack of a dynamical theory for confinement delayed broader acceptance until later developments in .

Key Experimental Discoveries

The first compelling experimental evidence for quarks as point-like constituents within protons and neutrons came from deep inelastic scattering experiments conducted at the Stanford Linear Accelerator Center (SLAC) starting in 1968. In these experiments, high-energy electrons were scattered off hydrogen and deuterium targets, revealing a scaling behavior in the structure functions that indicated the nucleons contained fractionally charged, spin-1/2 partons—later identified as quarks—interacting electromagnetically like point particles. This work, led by Jerome Friedman, Henry Kendall, and Richard Taylor at SLAC and MIT, earned them the 1990 Nobel Prize in Physics. The existence of a fourth quark flavor, , was confirmed in 1974 through the independent discovery of the J/ψ meson—a of a and its antiquark—by two teams using different accelerators. Samuel Ting's group at observed it in proton-beryllium collisions at the Alternating Gradient Synchrotron (AGS), while Burton Richter's team at SLAC detected it in electron-positron annihilation at the storage ring. The J/ψ has a mass of approximately 3.1 GeV/c² and a narrow width, consistent with predictions for a new flavor. Richter and Ting shared the 1976 for this breakthrough, which resolved anomalies in spectroscopy and solidified the . Subsequent discoveries established the remaining heavier quarks. In 1977, Leon Lederman's E288 experiment at Fermilab's Proton Center observed the Υ meson in proton-beryllium collisions, signaling the bottom (or ) quark with a mass around 9.5 GeV/c². This finding, using the Tevatron's predecessor infrastructure, completed the second generation of quarks and motivated further searches. The top quark, the heaviest known , was finally discovered in 1995 by the CDF and D0 collaborations at Fermilab's proton-antiproton collider. Analyzing data from collisions at √s = 1.8 TeV, both experiments observed top-antitop quark decaying into bosons and quarks, with a top mass of about 176 GeV/c² and evidence exceeding five standard deviations. This completed the three generations of quarks in the , leveraging the Tevatron's high after nearly two decades of searches. Key evidence for the three color charges of quarks emerged from electron-positron annihilation experiments measuring the ratio R of hadronic to muonic cross sections above quark production thresholds. At energies between 2 and 5 GeV, R approached 3.67, aligning with the expectation of three colors per quark flavor after accounting for QCD corrections, as measured at SLAC's SPEAR and later PEP rings. These accelerators played pivotal roles: SPEAR enabled precise e⁺e⁻ studies revealing quarkonium states, PEP provided higher-energy data confirming color dynamics, and the Tevatron delivered the proton-proton collision environment needed for top quark production.

Etymology and Terminology

Origin of the Term "Quark"

The term "quark" for the fundamental particles was coined by physicist in 1964, drawing directly from a line in James Joyce's 1939 novel : "Three quarks for Muster Mark! / Sure he hasn't got much of a bark / And sure any he has it's all beside the mark." Gell-Mann selected the word because Joyce's novel is renowned for its inventive, nonsensical vocabulary, which he found apt for naming hypothetical subatomic constituents whose existence was not yet experimentally confirmed. In a 1978 letter to the editor of the , Gell-Mann explained that he initially pronounced "quark" to rhyme with "cork" (as /kwɔːrk/), evoking a on "three quarts for Mister Mark" in a pub setting, while acknowledging Joyce's original likely rhymed with "bark" (/kwɑːrk/). Gell-Mann kept the name secret during the early development of his , first publicly revealing it in a 1963 lecture while on leave at , where he discussed organizing the proliferation of known particles. This revelation preceded the formal publication of the in 1964, in which Gell-Mann proposed that hadrons are composites of three such particles (or quark-antiquark pairs for mesons), aligning with the "three quarks" phrase from Joyce. Independently, George Zweig proposed a similar model in 1964 at CERN, referring to the particles as "aces" to evoke combinations like deuces and treys in hadrons, but Gell-Mann's evocative term from literature quickly became the universal standard in the field. The choice of "quark" not only captured the playful yet profound nature of the discovery but also endured despite Zweig's alternative, reflecting the influence of Gell-Mann's presentation and publication. The up and down quarks were named by in 1964 to reflect their roles in an doublet, analogous to the spin-up and spin-down states of particles under the . The , also proposed by Gell-Mann in 1964, received its name due to the unexpectedly long lifetimes of particles containing it, such as kaons, which decayed via the weak force rather than the strong force. The charm quark was predicted in 1964 by and James Bjorken and named for the symmetry it restored in the subnuclear world, evoking the Latin term meaning enchantment or song. The and top quarks were proposed by and in 1973 to explain in weak interactions, with names coined by Haim Harari in 1975 to maintain sequential alphabetical symbols (t and b) while pairing them as counterparts to ; prior to standardization, they were sometimes referred to as "truth" and "" respectively. The quark's existence was confirmed in 1977 through the discovery of the Υ meson—a -antibottom quark —at , observed as a in the dimuon spectrum around 9.4 GeV. The top quark, predicted to complete the third generation, was discovered in at 's collider via top-antitop decaying into bosons and quarks. In particle physics notation, a generic quark is denoted by q, while specific flavors use lowercase symbols: u for up, d for down, s for strange, c for , b for , and t for top. Antiquarks are represented with an overline, such as \bar{u}, \bar{d}, \bar{s}, \bar{c}, \bar{b}, and \bar{t}, following the convention that flavor quantum numbers for antiquarks have opposite signs to those of quarks (e.g., strangeness S = -1 for s and +1 for \bar{s}). These labels are standardized by the Particle Data Group (PDG), which establishes conventions for quark content in nomenclature to ensure consistency in reporting experimental results and theoretical models.

Intrinsic Properties

Electric Charge and Color Charge

Quarks possess fractional electric charges, measured in units of the e, which is the charge of the . The up-type quarks (up, , and ) each carry a charge of +\frac{2}{3}e, while the down-type quarks (down, strange, and ) each carry -\frac{1}{3}e. Antiquarks have opposite charges to their corresponding quarks, so up-type antiquarks have -\frac{2}{3}e and down-type antiquarks have +\frac{1}{3}e. The of a is the algebraic sum of the charges of its quarks. For example, the proton, composed of two up quarks and one (uud), has a total charge of \frac{2}{3}e + \frac{2}{3}e - \frac{1}{3}e = +e. Similarly, the (udd) has \frac{2}{3}e - \frac{1}{3}e - \frac{1}{3}e = 0. This additive property ensures that observed hadrons have charges, consistent with experimental observations. In addition to electric charge, quarks carry color charge, a quantum number associated with the strong nuclear force as described by quantum chromodynamics (QCD). Color charge comes in three types, conventionally labeled red, green, and blue, analogous to the primary colors but serving as the basis for SU(3) gauge symmetry in QCD. Each quark carries a single color charge, while antiquarks carry the corresponding anticolor (antired, antigreen, antiblue). The mediators of the strong force, gluons, carry a of one color and one anticolor, forming an octet of eight possible states under the SU(3) color group. Unlike photons in , which are neutral, gluons are colored and thus interact with each other, leading to the of the strong force. Hadrons must be color-neutral, or "white" in the color analogy, to comply with the principle of , where isolated quarks cannot be observed. Baryons, such as protons and neutrons, achieve this through a of three quarks carrying different colors (one , one , one ), forming a . Mesons, composed of a quark-antiquark pair, are color-neutral when the quark's color matches the antiquark's anticolor, also resulting in a . This color neutrality ensures that the strong force binds quarks into colorless composites at low energies.

Spin and Weak Isospin

Quarks are fermions with an intrinsic of \frac{1}{2} \hbar, which dictates their behavior under quantum statistics and their role in composite particles. This angular momentum ensures that quarks obey the , preventing identical fermions from occupying the same , a key feature that underlies the stability and diversity of hadronic matter. In hadrons, the spins of constituent quarks combine to yield the total of the ; for instance, the ground-state baryons like the proton and , which have total \frac{1}{2}, result from the symmetric or antisymmetric coupling of the spins of three spin-\frac{1}{2} quarks under the constraints of color and symmetries. Within the framework of electroweak theory, quarks carry , a associated with the SU(2)_L gauge symmetry that governs weak interactions. Left-handed quark fields transform as under this SU(2)_L, such as the up-down doublet \begin{pmatrix} u \\ d \end{pmatrix}_L for the first generation, with weak isospin components I_3 = +\frac{1}{2} for the up-type quark and I_3 = -\frac{1}{2} for the down-type. In contrast, right-handed quark fields are weak isospin singlets, carrying I = 0, which reflects the nature of the weak force where only left-handed chirality participates in charged current interactions. This assignment extends analogously to higher generations, with doublets like \begin{pmatrix} c \\ s \end{pmatrix}_L and \begin{pmatrix} t \\ b \end{pmatrix}_L. The weak charged currents involve transitions between up-type and down-type quarks across generations, parameterized by the Cabibbo-Kobayashi-Maskawa (CKM) matrix, a unitary mixing matrix that introduces flavor-changing dynamics and a single phase responsible for in the . Originally proposed to accommodate the observed pattern of weak decays beyond the first generation, the CKM matrix quantifies the misalignment between weak and mass eigenstates, ensuring unitarity while allowing for non-trivial mixing angles and the CP-violating phase. The relativistic description of quarks incorporates their spin through Dirac spinors, four-component objects that satisfy the , i \hbar \frac{\partial \psi}{\partial t} = \left( c \vec{\alpha} \cdot \vec{p} + \beta m c^2 \right) \psi, where \psi represents the quark field, \vec{\alpha} and \beta are matrices encoding and relativistic effects, and the equation unifies with for massive -\frac{1}{2} particles. This formulation captures the intrinsic degrees of freedom essential for quark dynamics in .

Mass and Size Estimates

Quark masses in the are fundamental parameters, but their values are not directly measurable due to confinement. Instead, they are inferred from simulations, perturbative QCD analyses of experimental data, and global fits to electroweak precision observables. In the modified minimal subtraction (MS-bar) scheme, the current quark masses at a renormalization scale of μ = 2 GeV are estimated as follows for the light quarks: (u) 2.16 ± 0.07 MeV/c², (d) 4.70 ± 0.07 MeV/c², and (s) 93.5 ± 0.8 MeV/c². For the heavier quarks, the charm quark (c) mass is 1.273 ± 0.005 GeV/c² in the MS-bar scheme at μ = m_c, the bottom quark (b) mass is 4.183 ± 0.007 GeV/c² at μ = m_b, and the top quark (t) mass is 172.6 ± 0.3 GeV/c². These values reflect the 2025 Particle Data Group (PDG) averages, incorporating updates from collider experiments and lattice calculations.
QuarkFlavorMS-bar Mass (MeV/c²)Scale μ (GeV)Notes
uUp2.16 ± 0.072Light quark
dDown4.70 ± 0.072Light quark
sStrange93.5 ± 0.82Light quark
c1273 ± 5≈1.27Heavy quark
b4183 ± 7≈4.18Heavy quark
t172600 ± 300Pole massUnstable, from tt̄ production
Due to the asymptotic freedom of QCD, quark masses exhibit a running behavior with the scale μ, arising from effects. The evolution is described by the leading-order : m(\mu) = m(\mu_0) \left[ \frac{\alpha_s(\mu)}{\alpha_s(\mu_0)} \right]^{\gamma_m}, where α_s is the , and γ_m ≈ 12/(33 - 2n_f) is the leading mass anomalous dimension (with n_f the number of active flavors). This running decreases the effective at higher scales, with light quark masses becoming negligible compared to Λ_QCD ≈ 200 MeV above a few GeV. Experiments probing quark structure at high energies, such as deep inelastic electron-proton scattering at , indicate that quarks behave as point-like particles with an upper limit on their effective radius of less than 4.3 × 10^{-18} m at 95% confidence level. However, QCD confinement prevents free quarks from existing; they are bound within s, where the relevant spatial scale is the hadron size of approximately 1 (10^{-15} m), set by the strong interaction range. Lattice QCD offers a non-perturbative approach to compute quark masses by discretizing and simulating the QCD on supercomputers. This method directly incorporates confinement and chiral dynamics, yielding precise mass values that match experimental spectra. Advancements in the 2020s, including the use of physical masses in simulations (reducing errors) and improved algorithms like domain-wall fermions for better chiral symmetry preservation, have achieved sub-percent precision for light quark masses, as reviewed by the collaboration. Recent updates from the FLAG 2024 review further refine these calculations using enhanced lattice data. These calculations confirm the small current masses while highlighting effects. For light quarks, the small current masses (a few MeV/c²) are overshadowed by dynamical (DCSB), a QCD effect where the quark-antiquark generates effective constituent masses of approximately 300–350 MeV/c² for u and d quarks. This DCSB, analogous to in condensed matter, accounts for over 90% of the proton's (despite quarks contributing only ~1% via current masses) and is quantified in through the quark propagator's behavior near the chiral limit.

Interactions and Dynamics

Electromagnetic and Weak Interactions

Quarks interact electromagnetically through their fractional electric charges, coupling to photons in a way that mirrors the (QED) description for leptons, with the interaction term given by \mathcal{L}_{EM} = -e \bar{q} Q \gamma^\mu q A_\mu, where Q is the quark's charge in units of the e, q represents the quark , and A_\mu is the . The strength of this coupling for a given quark flavor is proportional to Q^2; for instance, up-type quarks (u, c, t) with Q = +2/3 have a coupling four times stronger than down-type quarks (d, s, b) with Q = -1/3. This charge-dependent coupling influences observable processes, such as , where the total cross-section for quark-photon interactions scales with the sum of Q_i^2 over quark flavors weighted by their parton distribution functions. In the , quarks couple to the electroweak gauge bosons ^\pm and , enabling flavor-changing processes that violate parity conservation, a hallmark of the weak force observed in phenomena like . Charged-current interactions, mediated by ^\pm bosons, involve transitions between up-type and down-type quarks, such as d \to u + W^-, with the effective four-fermion strength governed by the Fermi G_F / \sqrt{2} \approx 1.166 \times 10^{-5} GeV^{-2}. Neutral-current interactions via the boson are flavor-diagonal at tree level in the , conserving quark flavor, but higher-order loop processes can induce flavor-changing neutral currents (FCNCs), which are strongly suppressed. The mixing between quark mass eigenstates and weak eigenstates is described by the Cabibbo-Kobayashi-Maskawa (CKM) matrix, a 3×3 that parametrizes the probabilities of flavor transitions; its elements include |V_{ud}| \approx 0.974, |V_{us}| \approx 0.225, |V_{ub}| \approx 0.0036, |V_{cd}| \approx 0.225, |V_{cs}| \approx 0.973, |V_{cb}| \approx 0.041, |V_{td}| \approx 0.008, |V_{ts}| \approx 0.041, and |V_{tb}| \approx 0.999. This matrix introduces three mixing angles and one CP-violating phase, visualized in the unitarity triangle derived from V_{ud} V_{ub}^* + V_{cd} V_{cb}^* + V_{td} V_{tb}^* = 0, which tests the 's consistency. A prototypical example of a charged-current weak process at the quark level is the beta decay of the neutron, n \to p + e^- + \bar{\nu}_e, where one down quark in the neutron's udd configuration transforms into an up quark via d \to u + W^-, and the W^- subsequently decays to e^- + \bar{\nu}_e; this process underpins nuclear beta decay and has been precisely measured, with the neutron lifetime τ_n = 878.4 ± 0.5 s (PDG 2024 average from ultracold neutron experiments) aligning with Standard Model predictions within uncertainties, though an unresolved ~4σ discrepancy persists with beam method results (~888 s). The Glashow-Iliopoulos-Maiani (GIM) mechanism provides the crucial suppression of FCNCs by ensuring destructive interference in loop diagrams involving virtual up-type quarks; originally proposed to resolve discrepancies in strangeness-changing neutral processes like K^0 \to \mu^+ \mu^-, it relies on the near-degeneracy of quark masses within generations and the unitarity of the CKM matrix, reducing FCNC amplitudes by factors of (m_c^2 / M_W^2) or higher. This mechanism not only predicted the existence of the charm quark but also extends to suppress rare decays like b \to s \gamma, where observed branching ratios match Standard Model expectations within experimental uncertainties.

Strong Interaction and Confinement

The strong interaction between quarks is governed by quantum chromodynamics (QCD), a quantum field theory formulated as a non-Abelian gauge theory invariant under the SU(3)c group of local color transformations, where the subscript c denotes color. In this framework, quarks transform under the fundamental representation of SU(3)c and carry one of three color charges (conventionally labeled red, green, or blue), while the mediating gauge bosons—gluons—transform under the adjoint representation, resulting in eight distinct gluon fields corresponding to the eight generators of the group. The QCD Lagrangian density is given by \mathcal{L}_\text{QCD} = \sum_f \bar{q}_f \left( i \gamma^\mu D_\mu - m_f \right) q_f - \frac{1}{4} G^a_{\mu\nu} G^{a \mu\nu}, where the sum is over quark flavors f, q_f represents the quark fields with mass m_f, D_\mu = \partial_\mu - i g_s \frac{\lambda^a}{2} A^a_\mu is the color covariant derivative involving the strong coupling g_s and Gell-Mann matrices \lambda^a, and G^a_{\mu\nu} = \partial_\mu A^a_\nu - \partial_\nu A^a_\mu + g_s f^{abc} A^b_\mu A^c_\nu is the field strength tensor for gluons A^a_\mu with structure constants f^{abc}. This structure encodes the fundamental dynamics of the strong force, distinct from the Abelian quantum electrodynamics due to the self-interaction terms in G^a_{\mu\nu}, which arise from the non-zero f^{abc} and allow gluons to couple directly to each other, leading to a non-linear theory. A key feature of QCD is , the phenomenon where the strong \alpha_s = g_s^2 / (4\pi) weakens at high momentum transfers (short distances), enabling perturbative calculations in that regime. The leading-order running of the coupling is described by the equation, yielding \alpha_s(Q) \approx \frac{12\pi}{(33 - 2 n_f) \ln(Q^2 / \Lambda^2)}, where Q is the energy scale, n_f is the number of active quark flavors (typically 3–6 depending on Q), and \Lambda \approx 200–300 MeV is the QCD scale parameter setting the onset of strong coupling. This behavior, opposite to that in , emerges from the negative coefficient in non-Abelian gauge theories with fermions, driven by gluon self-interactions and quark loops that screen at long distances but anti-screen at short ones. At low energies and long distances, QCD exhibits confinement, the mechanism by which color-charged particles like quarks and gluons are perpetually bound within colorless hadrons, with no free quarks observed experimentally. This is modeled by a linear interquark potential V(r) \approx \sigma r, where r is the quark-antiquark separation and the string tension \sigma \approx 1 GeV/fm quantifies the confining flux tube of gluons stretching between charges, analogous to a vibrating in effective models. The non-perturbative nature of confinement arises from the in \alpha_s, where the coupling diverges, preventing isolated color excitations; simulations confirm this linear rise and the absence of free color below deconfinement temperatures. Experimental support for QCD's strong interaction dynamics comes from jet production in electron-positron annihilation, where e^+ e^- \to q \bar{q} g processes produce collimated sprays of hadrons (jets) aligned with quark and gluon directions, consistent with perturbative gluon emission from color-charged partons. Observations of three-jet events at center-of-mass energies around 30 GeV provided direct evidence for the vectorial nature of gluons and their role in the strong force, with angular distributions matching QCD predictions for non-Abelian gluon radiation.

Role in Hadrons

Quarks are the fundamental constituents that bind together to form hadrons through the strong interaction mediated by gluons, with their color charges ensuring confinement within color-neutral composites. In the , hadrons are classified as mesons (quark-antiquark pairs) or baryons (three quarks), where the overall must be antisymmetric under particle exchange due to the fermionic nature of quarks. Potential models provide a non-relativistic framework to describe the binding of quarks in hadrons, particularly for heavy quarkonia like charmonium and bottomonium. A seminal example is the Cornell potential, which combines a short-range Coulomb-like term from one-gluon exchange with a long-range linear confining term: V(r) = -\frac{4}{3} \frac{\alpha_s}{r} + \sigma r, where \alpha_s is the strong coupling constant and \sigma is the string tension. This potential successfully reproduces the spectroscopy of heavy quarkonia states observed at facilities like and . For baryons, the total wave function incorporates spatial, spin, flavor, and color degrees of freedom, requiring overall antisymmetry. The color part is fully antisymmetric in the SU(3)_c singlet representation, necessitating the flavor-spin-spatial wave function to be symmetric for ground-state octet and decuplet baryons. In the spin-3/2 \Delta resonance, for instance, the flavor-spin component is fully symmetric, with the three up quarks in \Delta^{++} forming a symmetric state under SU(6) flavor-spin symmetry. Meson decays proceed dominantly through interactions when allowed by laws, with decay widths calculated using overlap integrals of quark wave functions in potential models. For mesons like the \rho \to \pi\pi decay, the width \Gamma is proportional to the and the matrix element involving the quark-antiquark annihilation into gluons, yielding predictions in good agreement with experimental values from the Particle Data Group. The quark model also explains baryon magnetic moments through the vector sum of constituent quark spins and charges. For the proton, assuming naive SU(6) and equal constituent quark masses, the magnetic moment is \mu_p = \frac{4}{3} \mu_u - \frac{1}{3} \mu_d, where \mu_u and \mu_d are the up and down quark moments; with \mu_u = 2 \mu_N and \mu_d = - \mu_N in magnetons \mu_N, this yields \mu_p = 3 \mu_N, close to the observed 2.793 \mu_N and demonstrating the model's success. Lattice QCD simulations offer an approach to compute masses by discretizing and solving the QCD numerically. These calculations, performed on supercomputers, reproduce light masses like the (139.6 MeV) and (938 MeV) with percent-level precision at physical quark masses, validating the while incorporating full QCD dynamics such as self-interactions.

Advanced Topics in Quark Matter

Sea Quarks and Virtual Particles

In (QCD), sea quarks refer to the virtual quark-antiquark (q\bar{q}) pairs that emerge from quantum fluctuations within the vacuum of hadrons, particularly nucleons like the proton. These pairs are not fixed constituents but transient excitations that contribute dynamically to the nucleon's internal structure, alongside the three valence quarks. Unlike valence quarks, which carry the net and flavor quantum numbers, sea quarks arise primarily from perturbative processes such as gluon splitting (g \to q\bar{q}), where s—mediators of the strong force—radiate flavor-singlet quark pairs. Non-perturbative mechanisms, including meson cloud effects around the valence quarks, also generate these pairs, leading to a flavor-dependent sea. The presence of sea quarks is captured in the parton model through parton distribution functions (PDFs), which describe the probability of finding a parton carrying a x of the nucleon's at energy scale Q^2. For up (u) and down (d) quarks, the total distributions decompose as u(x,Q^2) = u_v(x,Q^2) + u_s(x,Q^2) and similarly for d, where u_v = u - \bar{u} and d_v = d - \bar{d} represent contributions, while the sea is approximated by u_s \approx \bar{u} and d_s \approx \bar{d} for light flavors. The evolution of these PDFs with increasing Q^2 is governed by the Dokshitzer-Gribov-Lipatov-Altarelli-Parisi (DGLAP) equations, which incorporate splitting functions for processes like g \to q\bar{q}, thereby building the sea dynamically from initial valence inputs. This evolution explains how sea quarks become more prominent at low x (small fractions), where gluon densities are high. Analyses of nucleon structure functions from (DIS) show that sea quarks and gluons together carry approximately 50% of the proton's longitudinal , underscoring their significant role in the nucleon's budget. A hallmark signature of sea quarks appears in DIS experiments probing nucleon structure functions F_2^{p,n}(x,Q^2). The Gottfried sum rule, derived assuming isospin symmetry and equal \bar{u} = \bar{d} sea distributions in the proton, predicts \int_0^1 dx \, [F_2^p(x) - F_2^n(x)] / x = 1/3. However, measurements by the New Muon Collaboration (NMC) yielded $0.235 \pm 0.019, violating the rule and revealing a flavor asymmetry \bar{d}(x) > \bar{u}(x) in the proton's sea across a wide x range. This asymmetry, quantified by integrals such as \int_0^1 dx \, (\bar{d} - \bar{u}) \approx 0.118 \pm 0.012 from Drell-Yan data, cannot be fully explained by perturbative gluon splitting alone and points to non-perturbative origins, such as unequal chemical potentials or pion cloud contributions favoring \bar{d} production. Sea quarks also influence the proton's spin structure, contributing to the longstanding "proton spin crisis." Polarized experiments, starting with the European Muon Collaboration in 1988, found that the spins of and quarks together account for only about 30% of the proton's total spin of $1/2, with the balance arising from and orbital . Recent polarized PDF analyses highlight the 's role: measurements from the experiment at RHIC indicate that the polarized up antiquark distribution \Delta \bar{u} contributes more positively to the proton spin than \Delta \bar{d}, despite the unpolarized asymmetry \bar{d} > \bar{u}, adding nuance to resolving the spin puzzle through sea dynamics.

Quark-Gluon Plasma

The quark-gluon plasma (QGP) represents a state of matter in quantum chromodynamics (QCD) where quarks and gluons exist in a deconfined phase at sufficiently high temperatures, allowing perturbative descriptions above the critical temperature T_c \approx 150-170 MeV as determined by lattice QCD simulations. This phase emerges when the strong interaction's confinement mechanism breaks down, transitioning from bound hadronic states to a soup of free color charges, with the rapid increase in effective degrees of freedom near T_c signaling the onset of deconfinement. Experimental evidence for QGP formation has been obtained through heavy-ion collision experiments at the (RHIC) in the 2000s and the (LHC) via the ALICE detector in the 2010s, where signatures include jet quenching— the suppression of high-momentum particle jets due to energy loss in the medium—and elliptic flow, which measures azimuthal anisotropies in particle emissions indicative of collective hydrodynamic expansion. At RHIC, Au-Au collisions at \sqrt{s_{NN}} = 200 GeV demonstrated strong jet quenching consistent with partonic energy loss in a dense QCD medium, while LHC Pb-Pb collisions at \sqrt{s_{NN}} = 2.76 TeV extended this to higher temperatures, with elliptic flow patterns saturating hydrodynamic predictions and confirming QGP production over volumes exceeding 10 fm³. Key properties of the QGP include its behavior as a nearly ideal relativistic fluid, characterized by a shear viscosity-to-entropy density ratio \eta/s \approx 1/(4\pi) \approx 0.08, the minimal value allowed by bounds and verified through viscous hydrodynamic modeling of flow data from RHIC and LHC. Additionally, the high-temperature phase exhibits partial restoration of chiral symmetry, as evidenced by the vanishing of the chiral condensate order parameter in calculations and spectral modifications in correlators, reducing the effective masses of light quarks. As of 2025, observations in smaller collision systems, such as proton-lead (p-Pb) interactions at the LHC, reveal QGP-like behaviors including collective flow and medium-induced modifications to particle yields, suggesting that deconfinement can occur even in systems with lower multiplicity. enhancements, particularly in multi-strange particles like \Xi and \Omega, have been noted in these p-Pb and proton-proton collisions, attributed to statistical from a temporarily deconfined state with enhanced production rates up to 10 times higher than in elementary collisions. In the QCD phase diagram for 2+1 quark flavors (up, down, and strange), lattice simulations confirm a smooth crossover transition at zero , with the pseudocritical around 155 MeV and no phase boundary at physical quark masses.

Recent Experimental Measurements

Recent advancements in experimental particle physics, particularly at the (LHC), have provided high-precision measurements of quark properties, refining our understanding of the (SM) and constraining potential beyond-SM (BSM) physics. These efforts leverage large datasets from the ATLAS and experiments, achieving unprecedented accuracy in electroweak and observables. The , the heaviest known , has been subject to refined mass determinations using combined analyses of LHC data. The 2025 Particle Data Group average yields a mass of m_t = 172.56 \pm 0.31 GeV/c^2. This value supersedes earlier measurements and incorporates improved detector calibrations and theoretical inputs. The top quark width, inferred from decay width predictions and indirect constraints, is approximately 1.33 GeV at m_t = 172.5 GeV, consistent with expectations from electroweak loop effects. Additionally, correlations in top-antitop pairs have been measured with high precision, yielding coefficients like f = 1.249 \pm 0.024 (stat.) \pm 0.061 (syst.), confirming predictions and limiting BSM contributions to top . Lattice QCD simulations have yielded updated estimates for light quark masses, crucial for flavor physics and hadron structure calculations. According to the 2025 Particle Data Group (PDG) review, the up quark mass is m_u = 2.16 \pm 0.07 MeV, the down quark mass is m_d = 4.70 \pm 0.07 MeV, and the strange quark mass is m_s = 93.5 \pm 0.8 MeV, all in the \overline{\rm MS} scheme at 2 GeV, implying an isospin breaking difference of m_d - m_u \approx 2.5 MeV. These values incorporate recent lattice computations with dynamical fermions, reducing uncertainties through improved chiral extrapolation and continuum limits. Global fits to the Cabibbo-Kobayashi-Maskawa (CKM) matrix elements continue to probe and unitarity. The magnitude |V_{ub}| = (3.82 \pm 0.20) \times 10^{-3} emerges from inclusive and exclusive decays, with contributions from LHCb enhancing precision through rare decay ratios like |V_{ub}/V_{cb}| = 0.083 \pm 0.004. These measurements tighten constraints on the unitarity triangle, with LHCb's analysis of B_s^0 and \Lambda_b decays providing key inputs to the apex angle \gamma \approx 64.6^\circ \pm 2.8^\circ, showing no significant deviation from unitarity. Searches for BSM physics in flavor-changing neutral current (FCNC) processes, such as b \to s \ell^+ \ell^- transitions, have yielded null results that sharpen bounds on new particles. LHCb analyses of angular distributions in B \to K^* \mu^+ \mu^- decays, using up to 9 fb^{-1} of data, report observables consistent with SM predictions, excluding certain leptoquark models at 95% confidence level and constraining Wilson coefficients by factors of 2-3 compared to pre-LHC limits. These findings, devoid of deviations, reinforce the SM's validity in FCNC sectors while probing scales up to tens of TeV. In top quark pair production, precise cross-section measurements at 13 TeV inform flavor tagging techniques and BSM sensitivities. The inclusive t\bar{t} cross section is measured as $829.3 \pm 1.3 (stat.) \pm 8.0 (syst.) \pm 7.3 (lumi.) \pm 1.9 (beam) pb by ATLAS using 140 fb^{-1} of data, aligning with NNLO QCD predictions of 832 pb and enabling stringent tests of flavor-dependent couplings. Advanced b-jet tagging algorithms achieve efficiencies above 80% for heavy-flavor identification, facilitating searches for BSM effects like top flavor-violating decays, where null observations limit anomalous couplings to | \kappa_{tq} | < 0.15 for q = c, u.

References

  1. [1]
    Fifty years of quarks - CERN
    Jan 17, 2014 · In 1964, two physicists independently proposed the existence of the subatomic particles known as quarks.
  2. [2]
    [PDF] A BRIEF INTRODUCTION TO PARTICLE PHYSICS
    All of the known matter in the Universe today is made up of quarks and leptons, held together by fundamental forces which are represented by the exchange of.
  3. [3]
    [PDF] Gell-Mann.pdf
    Received 4 January 1964. If we assume that the strong interactions ... the triplet as "quarks" 6) q and the members of the anti-triplet as anti-quarks ~1.
  4. [4]
    The Nobel Prize in Physics 1990 - Illustrated presentation
    Not until the SLAC-MIT experiment of 1968 were the first traces of quarks seen. The discovery was made when protons and neutrons were illuminated by beams ...
  5. [5]
    [PDF] The Discovery of Quarks* - SLAC National Accelerator Laboratory
    For their crucial contributions as leaders of these experiments, which fundamentally altered physicists' conception of matter, Jerome Friedman and Henry Kendall.
  6. [6]
    [PDF] 15. Quark Model - Particle Data Group
    May 31, 2024 · The quark model describes the properties of low-lying bound states of QCD, using the language of quarks and antiquarks, and is older than QCD.
  7. [7]
    [PDF] 10. Electroweak Model and Constraints on New Physics
    May 31, 2024 · The branching fraction of the flavor changing transition b → sγ is of comparatively low precision, but since it is a loop-level process (in the ...
  8. [8]
    [PDF] 60. Quark Masses - Particle Data Group
    May 31, 2024 · This note discusses some of the theoretical issues relevant for the determination of quark masses, which are fundamental parameters of the ...
  9. [9]
  10. [10]
    None
    ### Summary of Pentaquarks from PDG Review (2024)
  11. [11]
  12. [12]
    [PDF] Tests of Conservation Laws - Particle Data Group
    Aug 11, 2022 · strangeness were produced from proton-antiproton annihilations at rest, p ... Tests of Conservation Laws. 6 Baryon and Lepton Number. The ...
  13. [13]
    [PDF] 79. Heavy Non-qq Mesons - Particle Data Group
    May 31, 2024 · covers the heavy “exotic” tetraquark (qq¯q¯q) and pentaquark (qqqq¯q) candidates recently discovered. We use the new naming scheme in the ...
  14. [14]
  15. [15]
    Nineteen sixty-four - CERN Courier
    Sep 9, 2025 · His theory first appeared as a long CERN report in mid-January 1964, just as Gell-Mann's quark paper was awaiting publication at Physics Letters ...<|control11|><|separator|>
  16. [16]
    50 Years of Quarks - www.caltech.edu
    Jul 22, 2014 · Caltech's Murray Gell-Mann simplified the world of particle physics in 1964 by standing it on its head. He theorized that protons—subatomic ...
  17. [17]
    [PDF] DEEP INELASTIC SCATTERING - Nobel Prize
    This paper describes the knowledge and beliefs about the nucleon's internal structure in 1968, including the conflicting views on the validity of the quark ...
  18. [18]
    A November revolution: the birth of a new particle - CERN Courier
    Nov 24, 2004 · ... Brookhaven and SLAC that heralded the discovery of the fourth quark, charm. I first heard about the J, or J/Ψ as it would become, in Jacques ...
  19. [19]
    Physics - <i>Landmarks</i>—The Charming Debut of a New Quark
    Feb 26, 2016 · An unexpected particle discovery in 1974 was the first evidence for a fourth type of quark. Figure caption expand figure. Brookhaven National ...
  20. [20]
    Extra!! Fermilab Experiment Discovers New Particle "Upsilon"
    Upsilon is a new, heavy sub-nuclear particle, more than ten times heavier than a proton, with a mass of 9.5 BeV, discovered at Fermilab.
  21. [21]
    The Discovery of the B Quark at Fermilab in 1977 - Inspire HEP
    I present the history of the discovery of the Upsilon (Υ) particle (the first member of the b-quark family to be observed) at Fermilab in 1977.
  22. [22]
    Discovery of the top quark at Fermilab - American Physical Society
    Since then, Fermilab's Tevatron has been revamped and both the CDF and D0 collaborations have dramatically improved their detectors, resuming collection of data ...
  23. [23]
    The top quark discovery - Scholarpedia
    Nov 23, 2022 · The top quark, the heaviest elementary particle known to date, was discovered by the CDF and D0 collaborations in 1995, after almost two decades of searches.Constraints on the mass of the... · The Tevatron Collider · The discovery in 1995
  24. [24]
    Scientists recall the discovery of the top quark 30 years ago at ...
    Mar 14, 2025 · On March 2, 1995, the top quark discovery at Fermilab was announced by scientists on the CDF and DZero collaborations, and the sixth and final quark was added ...
  25. [25]
    [PDF] PUB-l334 ELECTRON-POSITRON ANNIHILATION INTO HADRONS ...
    supposing there are several “colors” of each kind of quark. I3 To get the. 1p- 2y decay rate right, one needs three colors - red, blue and yellow, say ...
  26. [26]
    Nonstandard heavy mesons and baryons: Experimental evidence
    Feb 8, 2018 · Measurements of the total cross section for e + e - → hadrons were consistent with the existence of the 3 color degrees of freedom ( Litke et al ...
  27. [27]
    James Joyce And The Origin Of The Word 'Quark' - Science Friday
    Jul 24, 2018 · When Caltech physicist Murray Gell-Mann predicted the existence of an even smaller set of particles in 1964, he playfully dubbed them quarks.
  28. [28]
    quark - American Heritage Dictionary Entry
    Murray Gell-Mann, the physicist who proposed this name for these particles, said in a private letter of June 27, 1978, to the editor of the Oxford English ...
  29. [29]
    What's in a name? In physics, everything and nothing
    ### Summary of Quark Naming in Physics
  30. [30]
    Discoveries at Fermilab - The Bottom Quark - Inquiring Minds
    Feb 23, 2001 · The experiment discovered a particle, now called the upsilon, composed of a new kind of quark (named bottom) and its antimatter partner ( ...
  31. [31]
    [PDF] 8. Naming Scheme for Hadrons - Particle Data Group
    Dec 1, 2023 · In short, the minimal number of u plus d quarks together with the isospin determine the main symbol, and subscripts indicate any content of ...
  32. [32]
  33. [33]
  34. [34]
    [PDF] 10. Electroweak Model and Constraints on New Physics
    Dec 1, 2023 · The electroweak model is based on SU(2)xU(1) gauge group, with gauge bosons Wi and Bµ, and a complex scalar Higgs doublet for mass generation.
  35. [35]
    [PDF] CP-violation in the Renormalizable Theory of Weak Interaction
    2, February 1973. CP-violation in the Renormalizable Theory of Weak Interaction. Makoto KOBAYASHI and Toshihide MASKAWA. Department of Physics, Kyoto ...
  36. [36]
    The quantum theory of the electron - Journals
    Akhmeteli A (2018) The Dirac Equation as One Fourth-Order Equation for One Function: A General, Manifestly Covariant Form Quantum Foundations, Probability ...
  37. [37]
    Limits on the effective quark radius from inclusive $ep$ scattering at ...
    Apr 5, 2016 · The 95% C.L. upper limit on the effective quark radius is 0.43\cdot 10^{-16} cm.
  38. [38]
    None
    Summary of each segment:
  39. [39]
  40. [40]
    Chiral symmetry breaking with lattice propagators | Phys. Rev. D
    The detailed numerical analysis of the resulting gap equation reveals that the constituent quark mass ... QCD, namely, the lattice and the SDEs. In fact ...
  41. [41]
    [PDF] 12. CKM Quark-Mixing Matrix - Particle Data Group
    Sakai (KEK). 12.1 Introduction. The masses and mixings of quarks have a common origin in the Standard Model (SM). They arise from the Yukawa interactions ...<|control11|><|separator|>
  42. [42]
    [PDF] The Standard Model theory of neutron beta decay - arXiv
    Oct 31, 2023 · Abstract. We review the status of the Standard Model theory of neutron beta decay. Particular emphasis is put on the recent developments in ...
  43. [43]
    [1303.6154] The GIM Mechanism: origin, predictions and recent uses
    Mar 25, 2013 · The GIM Mechanism, introduced in 1970, explains the suppression of Delta S=1, 2 neutral current processes and is important in unified theories.Missing: original | Show results with:original
  44. [44]
    [1812.10372] The Sea of Quarks and Antiquarks in the Nucleon - arXiv
    Dec 26, 2018 · Much of the sea of quark-antiquark pairs emerges from the splitting of gluons and is well described by perturbative evolution in quantum ...
  45. [45]
    [PDF] Physics of the Nucleon Sea Quark Distributions - OSTI.GOV
    The low charged x partons contributing to the proton structure functions are quark-anti-quark pairs, known as sea quarks. The sea quarks are perturbatively ...
  46. [46]
    Sea Quark Surprise Reveals Deeper Complexity in Proton Spin Puzzle
    Mar 14, 2019 · The data show definitively, for the first time, that the spins of up antiquarks make a greater contribution to overall proton spin than the ...
  47. [47]
    [PDF] Effective degrees of freedom in QCD thermodynamics
    The rapid rise of the number of degrees of freedom in lattice QCD data around the critical temperature Tc ∼ 150 − 170 MeV, can be explained quantitatively by a ...Missing: T_c | Show results with:T_c
  48. [48]
    Overview of the QCD phase diagram
    Apr 19, 2021 · The QCD phase diagram shows a transition from hadrons at low temperatures to quark-gluon plasma at high temperatures, with a crossover at low ...<|separator|>
  49. [49]
    Signatures of QGP at RHIC and the LHC | AAPPS Bulletin
    May 21, 2021 · Many signatures supporting the formation of the QGP have been reported. Among them are jet quenching, the non-viscous flow, direct photons, and Debye screening ...
  50. [50]
    Looking inside trillion degree matter with ATLAS at the LHC
    May 13, 2022 · Elliptic flow is interpreted as strong evidence for the existence of the QGP. Jet quenching is another unique phenomenon that can occur in ...
  51. [51]
    Influence of Shear Viscosity of Quark-Gluon Plasma on Elliptic Flow ...
    May 26, 2011 · We investigate the influence of a temperature-dependent shear viscosity over entropy density ratio 𝜂 / s on the transverse momentum spectra ...
  52. [52]
    Fate of Chiral Symmetries in the Quark-Gluon Plasma from an ...
    Indeed, for a long time it was believed that in the high temperature phase ρ ( 0 ) vanishes in the chiral limit, signaling the restoration of chiral symmetry.Abstract · Article Text · Introduction. · The chiral limit.
  53. [53]
    Insights from ALICE in p–Pb and pp collisions | EPJ Web of ...
    Dec 10, 2024 · The formation of quark-gluon plasma (QGP) has been confirmed in ultra-relativistic heavy-ion (i.e. Pb-Pb) collisions. Results from small ...
  54. [54]
    Recent results on strangeness enhancement in small collision ...
    Apr 3, 2025 · [1]. Strangeness Production in the Quark - Gluon Plasma · [2]. Strangeness enhancement at mid-rapidity in Pb Pb collisions at 158-A-GeV/c · [3] ...<|separator|>
  55. [55]
    Aspects of the chiral crossover transition in ( 2 + 1 )-flavor QCD with ...
    Feb 11, 2025 · Using ( 2 + 1 )-flavor QCD configurations generated using the Möbius domain-wall discretization on an lattice, we construct suitable observables ...
  56. [56]
    [PDF] 61. Top Quark - Particle Data Group
    May 31, 2024 · The top quark is a left-handed quark with a large mass, unique semi-weak decay, and a short lifetime, making it behave as a free quark.
  57. [57]
    [2402.08713] Combination of measurements of the top quark mass ...
    Feb 13, 2024 · The combined top quark mass measurement is 172.52 ± 0.14 (stat) ± 0.30 (syst) GeV, with a total uncertainty of 0.33 GeV.
  58. [58]
    [PDF] LHCb-CONF-2024-004.pdf
    Jul 26, 2024 · Precise measurements of the Cabibbo–Kobayashi–Maskawa (CKM) unitarity triangle provide a strict test of the Standard Model (SM) and allow for ...
  59. [59]
    [2405.09182] Top quark mass and cross section at ATLAS and CMS
    This paper presents the latest measurements of top quark production cross sections and the top quark mass at the LHC by ATLAS and CMS.