A list of particles comprises the catalog of subatomic entities recognized in modern physics, encompassing elementary particles—the indivisible building blocks of matter and mediators of fundamental forces—and composite particles, which arise from the binding of elementary ones via the strong nuclear force.[1][2] This compilation is primarily guided by the Standard Model of particle physics, a quantum field theory that describes electromagnetic, weak, and strong interactions among these particles, with ongoing experimental validations from accelerators like the Large Hadron Collider.[1]The elementary particles are divided into fermions, which obey the Pauli exclusion principle and form matter, and bosons, which carry forces. Fermions include quarks and leptons, each arranged in three generations of increasing mass. The quarks are: up and down (first generation), charm and strange (second), and top and bottom (third); these six "flavors" carry fractional electric charges and experience all fundamental forces.[1][2] The leptons consist of charged particles—the electron, muon, and tau—and their associated neutral neutrinos (electron neutrino, muon neutrino, and tau neutrino); leptons do not participate in the strong force but interact via electromagnetic and weak forces.[1][2]Bosons include the gauge bosons, which mediate interactions: the photon for electromagnetism (massless, neutral), the W⁺, W⁻, and Z bosons for the weak force (massive, with W bosons charged), and eight gluons for the strong force (massless, carrying "color" charge).[1] The Higgs boson, a scalar particle discovered in 2012, is unique among bosons as it interacts with fermions and other bosons to impart mass through the Higgs mechanism, completing the Standard Model's particle roster.[3][2]Composite particles, primarily hadrons, form when quarks bind via the strong force mediated by gluons. Baryons, such as the proton (up-up-down quarks) and neutron (up-down-down quarks), consist of three quarks and are the building blocks of all stable nuclei of ordinary matter.[4][2] Mesons, made of a quark-antiquark pair, are typically unstable and include particles like pions and kaons, which play key roles in nuclear interactions.[4][2] Beyond these, exotic composites such as tetraquarks and pentaquarks have been observed, challenging and extending the quark model while remaining consistent with the Standard Model.[4] The Particle Data Group maintains comprehensive, regularly updated listings of these particles' properties, including masses, lifetimes, and decay modes, serving as the definitive reference for the field.[2]
Standard Model elementary particles
Quarks
Quarks are elementary fermions in the Standard Model of particle physics, serving as the fundamental constituents of hadronic matter. They carry fractional electric charges and a property known as color charge, which enables them to interact via the strong force. There are six distinct types, or flavors, of quarks, organized into three generations based on increasing mass: the first generation consists of the up (u) and down (d) quarks, the second of the charm (c) and strange (s) quarks, and the third of the top (t) and bottom (b) quarks.[1] These flavors differ primarily in their masses and charges, with quarks within each generation pairing as "up-type" (u, c, t) and "down-type" (d, s, b).[1]All quarks have spin 1/2, making them fermions that obey the Pauli exclusion principle, and they possess color charge in one of three types—red, green, or blue—governed by the SU(3) gauge symmetry of quantum chromodynamics (QCD).[5] This color charge confines quarks within color-neutral combinations through the strong force, preventing their isolation in nature.[1] The flavor of a quark determines its interactions under the weak force, contributing to phenomena like flavor-changing processes.[6]The properties of the six quark flavors are summarized in the following table, with masses given in the modified minimal subtraction (MS) scheme where applicable, based on lattice QCD calculations and experimental constraints:
Masses for the lighter quarks (u, d, s) are current quark masses, while heavier ones (c, b, t) reflect their dominant scale.[6][7]The existence of these quark flavors was established through a series of discoveries via high-energy particle collisions. The up and down quarks were inferred from the quark model proposed in 1964 and confirmed through deep inelastic scattering experiments at SLAC in 1968.[8] The strange quark was identified earlier through the observation of strange hadrons in cosmic rays and accelerators in the 1940s and 1950s, with its quark nature clarified in the 1960s.[1] The charm quark was discovered in 1974 independently by teams at SLAC and Brookhaven National Laboratory through the observation of the J/ψ meson.[9] The bottom quark followed in 1977 at Fermilab via the upsilon meson.[10] Finally, the top quark was observed in 1995 by the CDF and DØ collaborations at Fermilab in proton-antiproton collisions.[11]In ordinary matter, the up and down quarks play the primary role, combining via the strong force to form protons (two up quarks and one down quark, uud) and neutrons (one up quark and two down quarks, udd), which constitute atomic nuclei.[5] Heavier quarks appear in high-energy processes or exotic particles but decay rapidly due to their masses.[1]
Leptons
Leptons are a family of fundamental fermions in the Standard Model of particle physics, consisting of six distinct particles organized into three generations. These include three charged leptons—the electron (e⁻), muon (μ⁻), and tau (τ⁻)—each paired with a corresponding neutral neutrino: electron neutrino (ν_e), muon neutrino (ν_μ), and tau neutrino (ν_τ). Unlike quarks, leptons do not carry color charge and thus do not participate in strong interactions, but they interact via electromagnetic (for charged leptons) and weak forces.[12][13]All leptons have spin 1/2, making them fermions that obey the Pauli exclusion principle. The charged leptons possess an electric charge of -1 (in units of the elementary charge e), while neutrinos are electrically neutral. Representative masses illustrate their hierarchy: the electron has a mass of 0.511 MeV/c², the muon 105.7 MeV/c², the tau 1.777 GeV/c², and neutrinos have extremely small masses with an upper limit on their sum of approximately 0.12 eV/c² from cosmological constraints.[14][15][16][17]The leptons are grouped by generations, reflecting their increasing masses and analogous roles in weak interactions: the first generation comprises the electron and electron neutrino; the second, the muon and muon neutrino; and the third, the tau and tau neutrino. This structure arises from the electroweak symmetry of the Standard Model, where leptons couple to the weak gauge bosons W and Z.[12]The charged leptons were discovered sequentially: the electron in 1897 by J.J. Thomson (though its elementary nature was confirmed later), the muon in 1936 by Carl D. Anderson and Seth Neddermeyer through cosmic-ray observations in a cloud chamber, and the tau in 1975 by Martin Perl and collaborators at the SLAC SPEAR collider via e⁺e⁻ annihilation events showing distinct decay signatures. The electron neutrino was experimentally confirmed in 1956 by Clyde Cowan and Frederick Reines using inverse beta decay on antineutrinos from a nuclear reactor, while muon and tau neutrinos were inferred from muon decay kinematics and tau decays, respectively, with direct observation following neutrino oscillation discoveries.[18][19]Neutrinos exhibit oscillations, a quantum mechanical phenomenon where flavor eigenstates (ν_e, ν_μ, ν_τ) mix into mass eigenstates (ν_1, ν_2, ν_3) described by the Pontecorvo–Maki–Nakagawa–Sakata (PMNS) matrix. This 3×3 unitary matrix is parametrized by three mixing angles (θ_{12}, θ_{23}, θ_{13}) and one Dirac CP-violating phase, leading to flavor transitions over distance. Key parameters include mass-squared differences Δm²_{21} ≈ 7.5 × 10^{-5} eV² (from solar and reactor experiments) and |Δm²_{32}| ≈ 2.5 × 10^{-3} eV² (from atmospheric and accelerator data), with approximate mixing angles sin²θ_{12} ≈ 0.304, sin²θ_{23} ≈ 0.570, and sin²θ_{13} ≈ 0.022. These measurements, refined through global fits, confirm nonzero neutrino masses and provide evidence for physics beyond the minimal Standard Model.[20]Leptons play a central role in electroweak processes, particularly beta decay, where a neutron transforms into a proton via the charged-current weak interaction: n → p + e⁻ + \bar{ν}_e, mediated by W⁻ exchange and conserving lepton number. This process underpins radioactive decay and stellar nucleosynthesis, while neutral-current interactions via Z bosons allow flavor-preserving scattering, as seen in neutrino-electron events.[21]
Note: Neutrino masses are upper limits on the sum from cosmology; individual bounds are similar in scale. Spin is 1/2 for all.[14][15][16][17]
Gauge bosons
Gauge bosons are the spin-1 elementary particles in the Standard Model that act as force carriers, mediating the electromagnetic, weak, and strong interactions between quarks and leptons.[22] Unlike fermions, which are matter constituents, these bosons are vector particles responsible for transmitting forces, with their properties determined by the underlying gauge symmetries: U(1) for electromagnetism, SU(2) × U(1) for electroweak, and SU(3)_c for the strong force.[22] The four types include the photon (γ), the charged W bosons (W⁺ and W⁻), the neutral Z boson (Z⁰), and the eight gluons.[22]The photon mediates the electromagnetic force, which has infinite range due to the photon's zero mass. It is electrically neutral and couples to charged particles, enabling long-distance interactions like those in atomic structure. The W⁺, W⁻, and Z⁰ bosons mediate the weak force, responsible for processes such as beta decay; their large masses—80.369 ± 0.013 GeV/c² for the W bosons and 91.188 ± 0.002 GeV/c² for the Z boson—limit the weak force's range to approximately 10^{-18} m.[23][24] The eight gluons, arising from the SU(3)_c color symmetry of quantum chromodynamics (QCD), mediate the strong force between color-charged quarks; though massless, gluons carry color charge, leading to self-interactions.
Boson
Symbol
Spin
Mass (GeV/c²)
Charge
Force Mediated
Range
Photon
γ
1
0
0
Electromagnetic
Infinite
W boson
W⁺, W⁻
1
80.369 ± 0.013
±1
Weak
~10^{-18} m
Z boson
Z⁰
1
91.188 ± 0.002
0
Weak
~10^{-18} m
Gluons
g (8 types)
1
0
0
Strong
Confined to ~10^{-15} m
The gluons' self-interactions, a unique feature of non-Abelian gauge theories like QCD, result in asymptotic freedom—where the strong coupling weakens at short distances (high energies), allowing perturbative calculations—and color confinement, binding quarks into hadrons at larger scales (~10^{-15} m) such that free gluons or quarks are never observed. This behavior was theoretically predicted in 1973 and confirmed through deep inelastic scattering experiments.The photon was conceptually established as a particle through Max Planck's 1900 quantum hypothesis for blackbody radiation and Albert Einstein's 1905 explanation of the photoelectric effect, integrating it into quantum electrodynamics as the gauge boson of U(1) symmetry. The W and Z bosons were discovered in 1983 at CERN's Super Proton Synchrotron by the UA1 and UA2 collaborations, through proton-antiproton collisions producing high-transverse-momentum leptons consistent with their decays.[25] The gluons were first directly observed in 1979 at DESY's PETRA electron-positron collider by the TASSO, MARK-J, PLUTO, and JADE experiments, via three-jet events indicative of quark-antiquark-gluon production.[26]
Higgs boson
The Higgs boson is an elementary scalar particle in the Standard Model of particle physics, characterized by a spin of zero, zero electric charge, and no color charge.[3][27] Its mass has been precisely measured to be approximately 125.35 GeV/c².[28] As the only known spin-0 elementary particle, it plays a unique role in the theory by mediating interactions that generate masses for other particles.[29]The Higgs boson arises as an excitation of the pervasive Higgs field, a scalar field that permeates all of space and acquires a non-zero vacuum expectation value (VEV) of approximately 246 GeV through spontaneous electroweak symmetry breaking.[30] This breaking mechanism, proposed independently by François Englert and Robert Brout in 1964 and by Peter Higgs later that year, endows the W and Z gauge bosons with mass while preserving the massless photon.[29] Fermions acquire mass via Yukawa couplings to the Higgs field, where the fermion mass m_f is given by m_f = y_f v / \sqrt{2}, with y_f as the Yukawa coupling constant and v the VEV.[30]The Higgs boson was discovered in 2012 by the ATLAS and CMS experiments at the Large Hadron Collider (LHC) at CERN, through the observation of its decay products in proton-proton collisions.[27] This landmark confirmation earned Englert and Higgs the 2013 Nobel Prize in Physics for the theoretical prediction of the mechanism underlying mass generation.[31] The particle predominantly decays into bottom quark-antiquark pairs (branching ratio ≈58%), followed by W boson pairs (≈22%), tau lepton pairs (≈6%), and rarer modes such as Z boson pairs (≈3%) and diphotons (≈0.2%).[30]In theoretical extensions beyond the Standard Model, the Higgs boson serves as a potential portal to new physics, allowing interactions between Standard Model particles and hypothetical sectors such as dark matter candidates via Higgs-mediated couplings.[32]
Composite subatomic particles
Baryons
Baryons are composite subatomic particles classified as fermions, composed of three valence quarks (qqq) bound by the strong nuclear force, with total spin values of 1/2 or 3/2 and baryon number B=1.[33] Their quark content derives from the fundamental up (u), down (d), and strange (s) quarks, among others for heavier states.[33] The two stable baryons are the proton, with quark content uud and rest mass of 938.272 MeV/c², and the neutron, with udd and 939.565 MeV/c²; the proton's stability arises from energy conservation in possible decay modes, while the free neutron decays with a mean life of 878 seconds (PDG 2025; latest measurement 877.8 seconds as of August 2025) primarily to a proton, electron, and antineutrino.[34][35]Among unstable baryons, the Delta (Δ) resonances form an isospin-3/2 multiplet with spin 3/2, including Δ⁺⁺ (uuu), Δ⁺ (uud), Δ⁰ (udd), and Δ⁻ (ddd), all with masses around 1232 MeV/c² and full widths at half maximum of approximately 120 MeV, indicating lifetimes on the order of 10⁻²³ seconds; they decay strongly to a nucleon and pion. Hyperons, baryons containing at least one strange quark, include the Λ⁰ (uds, 1115.683 MeV/c²), Σ⁺ (uus, 1189.37 MeV/c²), Σ⁰ (uds, 1192.642 MeV/c²), Σ⁻ (dds, 1197.449 MeV/c²), Ξ⁰ (uss, 1314.86 MeV/c²), Ξ⁻ (dss, 1321.71 MeV/c²), and Ω⁻ (sss, 1672.45 MeV/c²); these decay weakly due to strangeness conservation in strong interactions, with lifetimes ranging from 10⁻¹⁰ to 10⁻²⁰ seconds.[36]In the framework of SU(3) flavor symmetry, light baryons organize into an octet of spin-1/2 particles (nucleons N, Λ, Σ, Ξ) and a decuplet of spin-3/2 particles (Δ, Σ*, Ξ*, Ω), where the symmetry relates masses and couplings through group representations, predicting relations like the equal spacing in the decuplet masses (e.g., M_Δ - M_Σ* ≈ M_Σ* - M_Ξ* ≈ M_Ξ* - M_Ω ≈ 150 MeV). This scheme, proposed independently by Murray Gell-Mann and Yuval Ne'eman, successfully classified observed states and anticipated the Ω⁻ before its observation.Key discoveries include the Λ hyperon, first observed in 1947 via cosmic-ray cloud-chamber tracks showing a neutral V-particle decay, confirming its baryon nature with mass around 1116 MeV/c². The Δ resonances emerged from pion-nucleon scattering experiments in the early 1950s, revealing a prominent resonance at 1232 MeV. The Ω⁻ was discovered in 1964 at Brookhaven National Laboratory using the 80-inch bubble chamber exposed to proton beams from the Alternating Gradient Synchrotron, with its decay topology K⁻ K⁻ π⁺ verifying the sss content and completing the decuplet.Baryons constitute the primary matter in nuclear physics, with protons and neutrons forming atomic nuclei through the residual strong force; at extreme densities in neutron stars, hyperons and Δ resonances can emerge, softening the equation of state and influencing maximum masses (around 2 solar masses observed) by providing additional pressure support or cooling channels via weak interactions.[37]
Baryon
Quark Content
Spin
Mass (MeV/c²)
Notes
Proton (p)
uud
1/2
938.272
Stable
Neutron (n)
udd
1/2
939.565
Unstable outside nuclei
Δ⁺⁺
uuu
3/2
~1232
Resonance, width ~120 MeV
Λ⁰
uds
1/2
1115.683
Hyperon, lifetime ~2.6×10⁻¹⁰ s
Σ⁺
uus
1/2
1189.37
Hyperon
Ξ⁰
uss
1/2
1314.86
Hyperon
Ω⁻
sss
3/2
1672.45
Hyperon, width ~8 MeV
Mesons
Mesons are composite subatomic particles consisting of a quark and its corresponding antiquark bound by the strong nuclear force, resulting in bosons with integer spin values ranging from 0 to several units.[33] These particles exhibit a variety of quantum numbers, including specific spin-parity (J^P) assignments such as 0^- for pseudoscalars and 1^-- for vectors, and they are generally unstable, decaying on timescales from picoseconds to femtoseconds depending on their mass and interaction channels.[33] In the quark model, mesons are classified into multiplets based on flavor SU(3) symmetry, with excitations described by orbital angular momentum (L) and total spin (S) of the quark-antiquark pair, leading to states like ^1S_0 (pseudoscalar) and ^3S_1 (vector).[33]The light pseudoscalar mesons, primarily involving up, down, and strange quarks, form the lowest-lying nonet in the quark model. The pion (π), with J^{PC} = 0^{-+}, includes the charged states π^+ (u\bar{d}) and π^- (d\bar{u}) at masses of approximately 140 MeV/c², and the neutral π^0 (mixture of u\bar{u} and d\bar{d}) at 135 MeV/c²; these mediate the residual strong force between nucleons. Kaons, also pseudoscalars, incorporate strange quarks, such as K^+ (u\bar{s}) and K^0 (d\bar{s}) with masses around 494-498 MeV/c², and their antiparticles, enabling studies of strangeness conservation and CP violation. The η meson, at 548 MeV/c² with J^{PC} = 0^{-+}, arises from a mixture of u\bar{u}, d\bar{d}, and s\bar{s} states, reflecting flavor mixing in the light sector.[38]Vector mesons, characterized by J^{PC} = 1^{--}, represent the spin-triplet ground state and decay primarily electromagnetically or strongly into lighter hadrons. The ρ meson (u\bar{d} or similar isospin combinations) has a mass of about 775 MeV/c² and width of 150 MeV, playing a key role in vector meson dominance of electromagnetic interactions. The φ meson, a nearly pure s\bar{s} state at 1019 MeV/c², provides insight into strange quark dynamics due to its narrow width of 4.3 MeV.[38]Heavier mesons involve charm and bottom quarks, confirming the quark model at higher masses. The J/ψ (c\bar{c}), discovered in 1974 at SLAC and Brookhaven, has a mass of 3097 MeV/c² and narrow width of 93 keV, signaling the charm quark's existence. The Υ (b\bar{b}), observed in 1977 at Fermilab, weighs 9460 MeV/c² with a width under 100 keV, establishing the bottom quark. B mesons, such as B^0 (d\bar{b}) and B^+ (u\bar{b}) at around 5279-5280 MeV/c², exhibit long lifetimes due to weak decays and are crucial for CP violation studies in the CKM matrix.[38]In the quark model, meson spectra follow Regge trajectories, where the square of the mass M^2 increases linearly with angular momentum J for fixed flavor, as M^2 ≈ α' J + const., with slope α' ≈ 0.9 GeV^{-2}, supporting string-like quark confinement in QCD.[39] This relation holds for both pseudoscalar and vector families, aiding identification of radial and orbital excitations. Mesons, particularly pions, are essential in pion-nucleon scattering experiments, where partial wave analysis tests chiral symmetry breaking and low-energy QCD predictions, such as the Weinberg-Tomozawa term for s-wave scattering lengths.[40] Lattice QCD simulations of these processes validate non-perturbative strong dynamics, reproducing scattering amplitudes with physical pion masses.[40]
Nuclear and atomic composites
Atomic nuclei
Atomic nuclei serve as the dense, positively charged cores of atoms, composed of baryonic matter in the form of nucleons: Z protons and N neutrons, with the total number of nucleons denoted as A = Z + N.[41] These structures bind through the strong nuclear force, overcoming electrostatic repulsion among protons to form stable configurations in most elements.[42]The existence of the atomic nucleus was established in 1911 by Ernest Rutherford through his gold foil experiment, where alpha particles scattered at large angles indicated a compact, massive core within the atom.[43] This model was refined in 1932 when James Chadwick discovered the neutron, a neutral nucleon essential for explaining nuclear masses and stability beyond protons alone.[44]Stable atomic nuclei represent configurations where the nucleus does not undergo radioactive decay over geological timescales, with representative examples including the proton (¹H, Z=1, N=0), helium-4 (⁴He, Z=2, N=2), carbon-12 (¹²C, Z=6, N=6), and oxygen-16 (¹⁶O, Z=8, N=8). Enhanced stability arises at "magic numbers" of protons or neutrons—specifically 2, 8, 20, 28, 50, 82, and 126—due to filled nuclear shells analogous to electron shells in atoms, leading to nuclei like ⁴He (doubly magic) that are particularly inert.[45]The binding energy of a nucleus quantifies the energy required to disassemble it into individual nucleons, averaging approximately 8 MeV per nucleon across stable isotopes, which peaks around iron-56 before declining for heavier elements.[42] This energy is modeled by the semi-empirical mass formula, originally proposed by Carl Friedrich von Weizsäcker in 1935 and refined by Hans Bethe, which approximates the binding energy B(A, Z) as a sum of volume, surface, Coulomb, asymmetry, and pairing terms:\begin{align}
B(A, Z) &= a_v A - a_s A^{2/3} - a_c \frac{Z(Z-1)}{A^{1/3}} - a_a \frac{(A - 2Z)^2}{A} \\
&\quad + \delta(A, Z),
\end{align}where typical coefficients are a_v \approx 15.5 MeV, a_s \approx 16.8 MeV, a_c \approx 0.72 MeV, a_a \approx 23.3 MeV, and \delta accounts for pairing effects (positive for even-even nuclei, negative for odd-odd).[42] The formula captures trends in nuclear masses without quantum details, aiding predictions of stability.[42]Isotopes of a given element share the same Z but differ in N and thus A, leading to variations in stability; for instance, uranium-235 (Z=92, N=143) has a half-life of 704 million years, while uranium-238 (Z=92, N=146) is far longer-lived at 4.468 billion years, influencing their roles in natural decay chains and fission applications.[46]Atomic nuclei play central roles in energy production and element formation: in nuclear fusion, light nuclei like hydrogen isotopes combine to form heavier ones such as helium, releasing energy that powers stars; in fission, heavy nuclei like uranium-235 split upon neutron capture, yielding energy and lighter fragments used in reactors.[47] In stellar nucleosynthesis, sequential fusion reactions within stars build progressively heavier elements from hydrogen to iron, with processes like the proton-proton chain and CNO cycle dominating energy output and cosmic abundance patterns.[48]
Atoms
Atoms are the basic units of matter consisting of a dense, positively charged nucleus surrounded by a cloud of negatively charged electrons, with the number of electrons equal to the atomic number Z, resulting in a neutral overall charge. The nucleus, composed of protons and neutrons, provides the positive charge, while the Z electrons balance it electrically. This structure defines neutral atoms, which form the building blocks of elements from hydrogen (Z=1) to oganesson (Z=118).[49][50][51]The atomic structure features electrons arranged in shells, historically labeled K, L, M, and so on, corresponding to principal quantum numbers n=1, 2, 3, etc., which determine the energy levels and average distance from the nucleus. Within these shells, electrons occupy orbitals defined by additional quantum numbers, adhering to the Pauli exclusion principle, which prohibits any two electrons from sharing the same set of four quantum numbers (n, l, m_l, m_s), thus limiting each orbital to at most two electrons with opposite spins. For example, the hydrogen atom has a single electron in the 1s orbital (n=1, l=0), while helium has two electrons filling the 1s orbital (1s²), making it stable and inert. Heavier atoms like uranium fill multiple shells up to n=7, with complex configurations following the Aufbau principle.[52][53][54][55]The energy required to remove an electron from a neutral atom, known as the first ionization energy, is 13.6 eV for hydrogen and generally increases with atomic number Z across periods in the periodic table due to stronger nuclear attraction on inner electrons. Atomic spectra arise from electron transitions between these quantized levels; for hydrogen, the Balmer series in the visible spectrum follows the Rydberg formula:\frac{1}{\lambda} = R \left( \frac{1}{2^2} - \frac{1}{n^2} \right)where λ is the wavelength, R is the Rydberg constant (approximately 1.097 × 10^7 m⁻¹), and n > 2 is the principal quantum number of the upper level, with lines scaling by Z² in multi-electron atoms for hydrogen-like ions.[56][57]The modern atomic theory originated with John Dalton's 1808 proposal that matter consists of indivisible atoms of fixed mass for each element, explaining chemical combinations. Niels Bohr advanced this in 1913 with a model incorporating quantized electron orbits around the nucleus, successfully predicting hydrogen's spectrum and laying groundwork for quantum mechanics. Isotopes, such as carbon-12 (six protons, six neutrons) and carbon-14 (six protons, eight neutrons), share the same Z and thus identical electron configurations and chemical properties, differing only in nuclear mass and stability, with carbon-14 being radioactive.[58][59][60]
Ions
Ions are atoms or molecules that possess a net electric charge resulting from the gain or loss of one or more electrons.[61] This charge imbalance distinguishes ions from neutral atoms, enabling them to participate in electrostatic interactions fundamental to chemical bonding and plasma dynamics.[62]Cations are positively charged ions formed when atoms or molecules lose electrons, such as the sodium cation Na⁺, which arises from the loss of one electron from a neutral sodium atom.[63] Anions, conversely, are negatively charged ions produced by gaining electrons, exemplified by the chloride anion Cl⁻, where a neutralchlorine atom acquires an additional electron.[63] Polyatomic ions consist of multiple atoms bonded together with a net charge, including the ammonium cation NH₄⁺ (formed by protonation of ammonia) and the sulfate anion SO₄²⁻ (resulting from deprotonation of sulfuric acid).[64] These types of ions are central to ionic compounds and electrolytic solutions.[65]Ionization states refer to the degree of charge on an ion, ranging from singly charged to highly stripped forms; for instance, H⁺ represents the proton, a fully ionized hydrogen atom lacking its electron.[66] Similarly, He²⁺ is the alpha particle, a doubly ionized heliumnucleus consisting of two protons and two neutrons.[67] Such states are prevalent in high-energy environments like stellar interiors and particle accelerators.[68]Ions form through processes such as electron impact ionization, where a high-energy electron collides with a neutral atom to eject an electron; photoionization, in which ultraviolet or X-ray photons strip electrons from atoms; and charge exchange, where an ion captures an electron from a neutral atom, transferring the charge.[69] In astrophysical contexts, ions dominate plasmas within stars, where thermal energies exceed ionization potentials, and in Earth's ionosphere, where solar radiation ionizes atmospheric gases to produce species like O⁺.[70][68]Key properties of ions include higher mobility compared to neutral particles due to their response to electric fields, which facilitates their drift in solutions and gases.[71] This mobility underpins ionic conductivity in electrolytes and plasmas, where ions carry current by migrating under applied voltages, essential for processes like electrochemical reactions and auroral displays.[71]The study of ions advanced significantly following J.J. Thomson's 1897 discovery of the electron through cathode ray experiments, which demonstrated that charged particles like electrons could be detached from atoms, laying the groundwork for understanding ion formation and behavior.[72]
Larger composite particles
Molecules
Molecules are bound systems of two or more atoms held together by electromagnetic forces, such as covalent or ionic bonds. They represent atomic-scale composites beyond the subatomic particles primarily addressed in this article and are detailed in chemistry resources.[73]
Exotic composites
Exotic composites encompass a range of unusual bound states of subatomic particles that deviate from conventional nuclear and atomic structures, incorporating leptons, hyperons, or unconventional quark configurations. These entities provide critical insights into fundamental interactions beyond standard quantum chromodynamics (QCD) and electroweak processes, often exhibiting enhanced sensitivities to short-range forces due to their compact sizes or novel compositions.[74]Leptonic atoms form when a lepton, such as a muon or pion, replaces one or more electrons in an atomic orbit, resulting in significantly smaller atomic radii owing to the lepton's greater mass. In muonic atoms, the muon's mass—approximately 207 times that of the electron—contracts the Bohr radius by a factor of about 200, enabling precise probes of nuclear charge distributions. A prominent example is muonic hydrogen, where 2010 laser spectroscopy measurements revealed a proton root-mean-square charge radius of 0.84184(67) fm, sparking the "proton radius puzzle" due to its discrepancy with electronic hydrogen values around 0.877 fm; the puzzle persists as of 2025 with ongoing muonic measurements confirming the discrepancy.[75][76] Pionic atoms, involving negatively charged pions in nuclear orbits, similarly facilitate studies of strong interactions at short distances, with applications in probing pion-nucleon couplings.[77]Hypernuclei are atomic nuclei augmented by at least one hyperon, such as the Λ particle, which introduces strangeness and alters nuclear binding dynamics. The first hypernucleus was observed in 1952 through emulsion experiments exposing cosmic-ray tracks, identifying a decay pattern consistent with a bound Λ in a light nucleus. Λ-hypernuclei, like the triton analog ³ΛH (comprising two neutrons and a Λ), exhibit lifetimes on the order of 10^{-10} seconds, comparable to the free Λ decay time of 2.63 × 10^{-10} s, reflecting weak decay dominance despite strong binding. These systems, with binding energies for the Λ around 1-30 MeV depending on the core nucleus, serve as laboratories for exploring hyperon-nucleon potentials unavailable in free-space scattering. Recent experiments at RHIC observed directed flow of hypernuclei such as ³ΛH and ⁴ΛH in Au-Au collisions at √s_NN = 3 GeV (2023), while J-PARC reported clear identification of Xi hypernuclei (2024).[78][79][80][81]Exotic hadrons extend compositeness to multiquark configurations beyond traditional quark-antiquark mesons or three-quark baryons. Tetraquarks, structured as qq¯q¯q, include the charged state Z(4430)^+, observed in 2007 by the Belle experiment in the ψ' π^+ decay channel from B meson decays, with a mass of 4430 ± 17 MeV/c² and width of 100 ± 30 MeV. Pentaquarks, comprising qqqq¯, were evidenced by LHCb in 2015 through peaks at P_c(4380)^+ and P_c(4450)^+ in the J/ψ p spectrum from Λ_b decays, with masses around 4380 and 4450 MeV/c² and significances exceeding 9σ. Subsequent 2019 LHCb analysis confirmed and refined these, resolving P_c(4450)^+ into narrower states at 4440 and 4457 MeV/c² while discovering P_c(4312)^+, all with J^P = 3/2^- interpretations as molecular Σ_c D¯* bound states. LHCb analyses as of 2025 further confirm these pentaquarks as hadronic molecules near charm-baryon–meson thresholds, per Particle Data Group reviews. As of 2024, a survey identified 23 exotic hadrons observed at the LHC, including new tetraquarks like T_{cccc}(6600) and T_{cccc}(6900).[82][83][84][85]Strangelets represent hypothetical clumps of strange quark matter, consisting of roughly equal numbers of up, down, and strange quarks (uuddss... configurations) stabilized by the strange quark mass against deconfinement. Proposed in the context of QCD phase transitions, these could form in high-density environments like neutron star cores or heavy-ion collisions, potentially converting ordinary matter into strange matter if energetically favorable, though no experimental evidence has been found as of 2025, with only upper limits on production established.These exotic composites find applications in probing non-perturbative QCD regimes, such as hyperon interactions and multiquark dynamics, through facilities like J-PARC and RHIC, where hypernuclei elucidate strangeness in nuclear matter and exotic hadrons test quark model extensions.[74][86]
Quasiparticles
Conventional quasiparticles
Conventional quasiparticles are emergent excitations in condensed matter systems that behave like particles but arise from the collective interactions of many underlying degrees of freedom, such as atoms or electrons in a solid; unlike fundamental particles, they cannot exist in isolation outside their host medium.[87] These quasiparticles simplify the description of complex many-body phenomena by approximating the system's low-energy excitations as weakly interacting entities with well-defined dispersion relations.[87] In uniform systems like crystals or metals, they represent classical collective modes, enabling quantitative predictions of material properties.Phonons are quantized vibrational modes of a crystal lattice, representing collective displacements of atoms from equilibrium positions and behaving as spin-0 bosons.[87] They manifest as longitudinal or transverse waves with linear dispersion at long wavelengths, akin to sound waves, and play a central role in the Debye model of specific heat, where the heat capacity at low temperatures follows C_V \propto T^3 up to the Debye temperature \theta_D, defined as \theta_D = \hbar \omega_D / k_B with \omega_D the Debye frequency.Magnons are quanta of spin waves, which are propagating disturbances in the aligned spins of a magnetic material, carrying spin-1 and obeying Bose-Einstein statistics as bosons. In ferromagnets, their dispersion is quadratic at low wavevectors, \omega \propto k^2, reflecting the precession of spins around the magnetization direction, while in antiferromagnets, the modes exhibit linear dispersion due to opposing sublattices.[87]Plasmons describe collective oscillations of electron density in a plasma or metal, where the electrons move coherently against a fixed ionic background, leading to a plasma frequency \omega_p = \sqrt{\frac{n e^2}{\epsilon_0 m}}, with n the electron density, e the charge, m the electron mass, and \epsilon_0 the vacuum permittivity. These modes are undamped longitudinal excitations at long wavelengths and dominate the dielectric response of metals in the optical regime.Rotons are low-energy excitations in superfluid helium-4, characterized by a parabolic minimum in their dispersion relation at finite momentum, \epsilon(p) \approx \Delta + \frac{(p - p_0)^2}{2\mu}, where \Delta is the energy gap, p_0 the momentum at the minimum, and \mu an effective mass. This non-linear form distinguishes them from phonons, which have linear dispersion, and they contribute significantly to the superfluid's viscosity and thermal transport below the lambda point.The theoretical foundations of these quasiparticles emerged in the early 20th century: phonons from the lattice dynamics model of Born and von Kármán in 1912, which treated vibrations as normal modes in periodic boundary conditions; magnons from Bloch's 1930 description of spin waves in ferromagnets to explain temperature-dependent magnetization. Plasmons were predicted by Bohm and Pines in 1951 using the random phase approximation for electron interactions, with experimental confirmation via inelastic electron scattering in the mid-1950s. Rotons were introduced by Landau in 1941 within his two-fluid theory of superfluidity to account for the elementary excitations spectrum.These quasiparticles underpin key material properties, such as thermal conductivity dominated by phonon scattering in insulators, where the Debye model predicts low-temperature behavior, and electrical conductivity in metals influenced by electron-plasmon and electron-phonon interactions that limit carrier mobility. Magnons enable spin-based transport in magnonics, offering low-power alternatives to charge currents for information processing.
Topological and fractional quasiparticles
Topological and fractional quasiparticles emerge from strongly correlated quantum systems, exhibiting properties that defy classical particle statistics and enable robust quantum phenomena. These quasiparticles often arise in two-dimensional materials under extreme conditions, such as strong magnetic fields or low temperatures, where electron interactions lead to collective excitations with topological protection. This protection stems from the global geometry of the system's wavefunction, making them resistant to local perturbations and promising for fault-tolerant quantum technologies. Unlike conventional quasiparticles, which follow Bose-Einstein or Fermi-Dirac statistics, these entities display fractional charges, non-standard exchange phases, or self-conjugate natures that reveal deeper quantum correlations.Anyons represent a class of quasiparticles confined to two-dimensional systems, where particle exchange yields a phase factor e^{i\theta} with \theta \neq 0, \pi, intermediate between bosons (\theta = 0) and fermions (\theta = \pi). Fractional statistics in two-dimensional systems was first theoretically proposed in 1977 by Jon Magne Leinaas and Jan Myrheim,[88] with Frank Wilczek introducing the term 'anyons' in 1982, highlighting how dimensionality relaxes the Pauli exclusion principle in quantum mechanics. Anyons are classified as Abelian, producing a simple phase upon braiding, or non-Abelian, where exchanges result in unitary matrix transformations that can encode quantum information non-locally. Experimental confirmation of Abelian anyons came through interference measurements in fractional quantum Hall systems in 2020.[89]Skyrmions are topologically stable spin textures in magnetic materials, manifesting as particle-like swirling configurations of electron spins with a non-trivial winding number. First conceptualized in the context of pion fields, they were observed in chiral magnets in 2009 and behave as effective quasiparticles due to their localized, soliton-like nature stabilized by Dzyaloshinskii-Moriya interactions. In thin films and heterostructures, skyrmions exhibit gyromagnetic motion under currents, enabling potential applications in racetrack memory devices where their topological charge prevents decay.[90]Majorana fermions appear as zero-energy, self-conjugate modes at the edges or defects of topological superconductors, where a Dirac fermion splits into two spatially separated, particle-antiparticle pairs.[91] Predicted theoretically for one-dimensional systems in 2001, these modes obey non-Abelian statistics and are topologically protected against decoherence. Experimental hints emerged in 2012 from conductance zero-bias peaks in indium antimonide nanowires proximity-coupled to superconductors, interpreted as signatures of Majorana zero modes. Ongoing efforts, including refined nanowire and vortex experiments, continue to seek unambiguous verification amid debates over alternative explanations like disorder effects.Fractional quantum Hall quasiparticles, observed in two-dimensional electron gases under high magnetic fields, carry fractional electric charges such as e/3, where [e](/page/E!) is the elementary charge. Their discovery in 1982 by Tsui, Stormer, and Gossard revealed plateaus in Hall resistance at filling factors like \nu = 1/3, defying integer quantum Hall expectations and signaling strongly interacting electron liquids.[92] The Laughlin wavefunction, introduced in 1983, describes these states as correlated quasihole or quasielectron excitations in the incompressible quantum fluid, with the $1/m form capturing the fractional charge and statistics. These quasiparticles underpin Abelian anyon models and have been directly imaged via interferometry in graphene devices.[89]A 2025 study at Brown University identified a new class of fractional excitons in quantum materials, formed by pairing quasiparticles with fractional charges, which exhibit unexpected behaviors like enhanced coherence without net charge.[93] This discovery, observed in twisted bilayer graphene under optical excitation, suggests novel bound states that could bridge excitonic and fractional Hall physics, potentially enabling hybrid quantum devices.[93]Applications of these quasiparticles center on topological quantum computing, where non-Abelian anyons and Majorana modes enable braiding operations for fault-tolerant qubits immune to local noise.[91] For instance, networks of Majorana zero modes in nanowires could store quantum information in parity-protected states, reducing error rates exponentially with system size.[94] Similarly, skyrmion lattices offer robust spin-based memory with low energy dissipation, while fractional quasiparticles in Hall systems support interferometric quantum gates.[90] These platforms promise scalable informationstorage and processing, leveraging topological invariance for long coherence times.
Hypothetical particles
Supersymmetric and supergravity particles
Supersymmetry (SUSY) posits a fundamental symmetry between bosons and fermions in particle physics, extending the Standard Model by introducing superpartners for each known particle, which differ in spin by half a unit while sharing the same quantum numbers except for spin and statistics. This symmetry implies that scalar partners (sfermions, such as squarks for quarks and sleptons for leptons) accompany fermions, while fermionic partners (gauginos, like the gluino for the gluon or winos for W bosons) pair with bosons; Higgs bosons gain Higgsino partners.[95] If unbroken, superpartners would be degenerate in mass with their Standard Model counterparts, but experimental absence necessitates SUSY breaking at high energies.The Minimal Supersymmetric Standard Model (MSSM) represents the simplest realization of SUSY, incorporating two Higgs doublets to avoid anomalies and generate fermion masses, resulting in over 100 free parameters compared to the Standard Model's 19.[95] It includes Higgsinos as superpartners to the Higgs fields and predicts a lightest supersymmetric particle (LSP), often the lightest neutralino, which remains stable under R-parity conservation and serves as a natural dark matter candidate.[95] SUSY breaking occurs via soft terms in the Lagrangian, such as scalar masses and trilinear couplings, alongside the μ parameter in the superpotential that mixes Higgsinos with gauginos, setting the electroweak scale.Supergravity (SUGRA) extends SUSY to a local symmetry, incorporating general relativity and yielding the gravitino as the spin-3/2 superpartner of the graviton, which absorbs a goldstino degree of freedom from spontaneous SUSY breaking.[96] In minimal SUGRA models, the gravitino can be the LSP or next-to-lightest, with recent proposals in N=8 supergravity exploring ultra-heavy, stable charged gravitinos (masses around 10^16 GeV) as dark matter candidates detectable via underground experiments like JUNO. These arise from the theory's spin-1/2 sector matching Standard Model matter while forbidding right-handed neutrinos.Experimental searches at the Large Hadron Collider (LHC) have found no evidence for SUSY particles as of 2025, with ATLAS and CMS excluding gluino masses up to 2.4 TeV and first- and second-generation squark masses above 2 TeV in simplified models assuming prompt decays to lightest neutralinos. Third-generation squarks, like stops, face weaker limits around 1.2 TeV due to mixing and decay channels mimicking Standard Model processes. Slepton and chargino/neutralino searches yield exclusions up to 250-400 GeV for compressed spectra.SUSY motivates solutions to the hierarchy problem, where quadratic divergences in the Higgs mass are canceled by superpartner loop contributions, stabilizing the weak scale against Planck-scale corrections without fine-tuning. It also facilitates grand unification by unifying gauge couplings at high energies and provides a framework for string theory compactifications.
Dark matter and dark energy candidates
Dark matter is a hypothetical form of matter that does not interact with light or electromagnetic radiation, comprising approximately 27% of the universe's total mass-energy content.[97] Its existence is inferred from gravitational effects that cannot be explained by visible matter alone, such as the flat rotation curves of galaxies observed by Vera Rubin in the 1970s, which indicate unseen mass distributed in galactic halos. Additional evidence comes from the cosmic microwave background (CMB) anisotropies measured by the Planck satellite, which require dark matter to seed large-scale structure formation, and the Bullet Cluster, where gravitational lensing reveals a separation between hot gas (baryonic matter) and the gravitational mass during a galaxy cluster collision in 2006.Leading candidates for dark matter particles include weakly interacting massive particles (WIMPs), axions, and sterile neutrinos. WIMPs are theorized to interact primarily via the weak nuclear force, with typical masses ranging from 10 GeV to 1 TeV; in supersymmetric extensions of the Standard Model, the lightest neutralino serves as a prototypical WIMP candidate, stable due to R-parity conservation.[98] Axions are light pseudoscalar bosons proposed as a solution to the strong CP problem in quantum chromodynamics (QCD), with masses in the range of $10^{-6} to $10^{-3} eV and very weak couplings to photons; recent theoretical work as of 2025 has explored their potential interactions via torsion in gravity.[98] Sterile neutrinos, which are right-handed and singlets under the Standard Model gauge group, have masses around a few keV and could constitute warm dark matter; they have been proposed to explain an observed 3.5 keV X-ray emission line in galaxy clusters and the Milky Way, potentially from their radiative decay, though subsequent analyses have questioned the line's dark matter origin due to inconsistencies with atomic transitions and lack of confirmation.[98][99][100]Experimental searches for these particles span direct detection, indirect detection, and collider methods, with no definitive discoveries as of 2025. Direct detection experiments, such as LUX-ZEPLIN (LZ), use liquid xenon to search for WIMP-nucleus scattering; LZ's 2025 results from 4.2 tonne-years of exposure set world-leading exclusion limits on WIMP cross-sections for masses above 10 GeV, ruling out simple WIMP models over much of the parameter space.[101] Indirect searches, including Fermi Large Area Telescope observations of gamma-ray excesses from DM annihilation in dwarf galaxies, have yielded null results consistent with background expectations.[98] Collider probes at the Large Hadron Collider (LHC), such as mono-jet events sensitive to DM production in association with quarks, have constrained WIMP models up to TeV scales without signals.[98] For axions, haloscope experiments like ADMX employ microwave cavities tuned to axion-photon conversions in strong magnetic fields; ADMX's 2025 scan from 1.1 to 1.3 GHz (corresponding to masses 4.5–5.4 μeV) extended exclusion limits in the Kim-Shifman-Vainshtein-Zakharov (KSVZ) model, approaching QCD axion bands but finding no signal.Dark energy, which drives the accelerated expansion of the universe and constitutes about 68% of its energy density, is often modeled as a cosmological constant \Lambda in the \LambdaCDM framework, equivalent to vacuum energy with constant density.[97] Dynamic alternatives include quintessence, where a scalar field \phi slowly rolls down a flat potential V(\phi), leading to time-varying energy density and equation-of-state parameter w \approx -1 but potentially evolving; such models address the coincidence problem of why dark energy dominates today. The Dark Energy Spectroscopic Instrument (DESI) 2025 results provide tentative evidence for evolving dark energy, favoring quintessence-like models with w > -1 at low redshifts over a strict cosmological constant, though statistical tensions remain under debate. Unlike dark matter, dark energy candidates do not typically manifest as discrete particles but as pervasive fields, with no dedicated particle detectors; constraints arise from supernova distances, baryon acoustic oscillations, and CMB data.
Other exotics
The graviton is a hypothetical elementary particle proposed as the quantum mediator of the gravitational force, characterized by zero mass and spin-2. In efforts to quantize general relativity, the graviton arises as the massless spin-2 tensor field, analogous to how gauge bosons mediate other fundamental interactions, though it remains undetected as a particle despite gravitational waves observed by LIGO confirming classical predictions of spacetime ripples.[102][103] These waves propagate at the speed of light, consistent with the massless nature required for long-range gravitational effects, but direct graviton detection faces immense challenges due to their weak coupling.Tachyons represent another class of hypothetical particles defined by their ability to travel faster than light, arising in scenarios where the squared mass m^2 < 0, leading to an imaginary rest mass.[104] Such particles would gain speed as energy decreases, potentially allowing backward time travel and raising profound causality violations in special relativity, as signals could precede causes in certain reference frames.[105] Particles are classified by propagation speed: massive ones below light speed, massless at light speed, and tachyonic above it, though quantum field theory imposes no-go constraints, viewing tachyons as instabilities rather than stable excitations.[106] Attempts to construct consistent tachyon quantum field theories encounter issues like unbounded Hamiltonians and vacuum instability, rendering them untenable in standard frameworks.[107]Magnetic monopoles are theorized isolated north or south magnetic charges, predicted in grand unified theories (GUTs) where electromagnetism emerges from a unified gauge group broken at high energies. The minimal monopole charge, proposed by Dirac, satisfies g = \frac{\hbar c}{2e}, ensuring electric charge quantization observed in nature.[108] GUTs forecast monopoles with masses around $10^{16} GeV, produced copiously in the early universe but diluted by inflation, motivating ongoing searches.[109] Direct cosmic ray detections have yielded null results, with experiments like IceCube and Pierre Auger setting stringent flux limits below $10^{-16} cm^{-2} s^{-1} sr^{-1}, far below GUT expectations.[110][111]Preons hypothesize substructure within quarks and leptons, positing them as composite entities built from more fundamental constituents to simplify the particle zoo.[112] The rishon model, for instance, constructs quarks and leptons from triplets of two types of preons carrying fractional charges, aiming for unification at scales beyond the Standard Model. However, high-energy scattering experiments at accelerators like the LHC show no evidence of quark compositeness, with limits on substructure scales exceeding $10^{-18} m, supporting quarks as pointlike.[113]Continuous spin particles form a distinct representation in the Poincaré group, featuring infinite degrees of freedom and transforming under the little group SO(2,1) for massless cases in 3+1 dimensions. These exotic states, first classified by Wigner, arise as limits of higher-spin fields and challenge standard quantum field descriptions due to their non-compact symmetry, potentially relevant in quantum gravity contexts but unobserved.[114]These hypothetical particles are primarily motivated by the quest for quantum gravity and gauge unification, addressing incompatibilities between general relativity and quantum field theory at Planck scales, where gravitons and monopoles could bridge classical and quantum realms. Despite null experimental results, their theoretical pursuit drives extensions like string theory, emphasizing symmetry and completeness in fundamental interactions.[115]