Fact-checked by Grok 2 weeks ago

Exotic atom

An exotic atom is an otherwise normal atom in which one or more subatomic particles, typically an , have been replaced by another of similar charge but different mass and properties, such as a negatively charged (e.g., ) or (e.g., or ). These systems are inherently unstable, with lifetimes ranging from nanoseconds to microseconds depending on the exotic particle involved, and they exhibit atomic orbits much closer to the due to the heavier mass of the substitute particle, leading to enhanced sensitivity to nuclear and quantum electrodynamic () effects. Exotic atoms are broadly classified into leptonic and hadronic types. Leptonic exotic atoms, such as muonic atoms, involve the replacement of an by a (mass ≈ 207 times that of an electron, lifetime 2.2 μs), forming systems like muonic hydrogen (a proton orbited by a muon). Hadronic exotic atoms replace electrons with mesons, including (pion mass ≈ 273 electron masses, lifetime 26 ns), (kaon mass ≈ 966 electron masses, lifetime 12 ns), and (antiproton mass ≈ 1836 electron masses). Other variants include (electron-positron ) and hypernuclei, where a in the is replaced by a . These atoms are produced artificially in particle accelerators, such as at the (PSI) or CERN's Antiproton Decelerator, by slowing down beams of exotic particles to low energies (≈10 eV) in a target material, where they capture onto nuclei by displacing electrons through processes or radiative transitions, followed by de-excitation via emission. Precision of the emitted s allows measurement of shifts and widths, providing data on atomic energies in the keV to MeV range. Exotic atoms serve as powerful probes of fundamental physics, including , the nuclear force via , and nuclear structure. For instance, muonic hydrogen has been used to determine the proton's with unprecedented precision, revealing a value of 0.84184(67) fm in 2010, which sparked the "proton radius puzzle" due to its 5σ discrepancy with and ordinary measurements (0.8751(61) fm). Subsequent high-precision measurements from and ordinary have confirmed values close to this, with the CODATA 2022 proton at 0.84075(64) fm, though minor discrepancies persist as of 2025. Studies of pionic and kaonic atoms test low-energy strong interactions and pion-nucleon , while antiprotonic atoms explore antiproton-nucleus interactions and states. Ongoing experiments continue to refine these measurements, potentially resolving puzzles in and advancing our understanding of quantum fields.

Definition and Basics

Definition of Exotic Atoms

An exotic atom is a bound atomic system in which one or more electrons of an ordinary atom are replaced by other elementary particles, such as leptons, hadrons, or quasiparticles, or in which the incorporates exotic constituents like hyperons. These systems differ from conventional atoms by featuring substitute particles that alter the electromagnetic and interactions within the atomic structure. Exotic atoms are classified into broad categories based on the nature of the replacing or added particles. Leptonic exotic atoms involve leptons heavier than electrons, such as s, forming systems like muonic atoms where a negative orbits the . Hadronic exotic atoms substitute electrons with hadrons, including mesons like pions or kaons, or baryons like antiprotons, leading to atoms such as pionic or kaonic hydrogen. atoms consist of bound lepton-antilepton or quark-antiquark pairs, exemplified by (electron-positron) or charmonium (charmed quark-anticharm). Hypernuclear atoms feature exotic baryons, such as hyperons (e.g., particles containing strange quarks), integrated into the alongside protons and neutrons. Quasiparticle exotic atoms arise in condensed matter systems, where effective particles like excitons or polarons mimic atomic behavior through collective excitations. In typical examples, a negatively charged exotic particle replaces an electron and orbits the nucleus, as in muonic helium, or a positively charged exotic particle binds to the electron cloud, forming structures like muonium (muon-electron). Unlike stable ordinary atoms, exotic atoms are inherently short-lived, with lifetimes determined by the decay or annihilation of the exotic particles; for instance, pionic atoms decay rapidly due to the pion's lifetime of approximately 2.6 × 10^{-8} seconds. This transient nature limits their persistence but enables unique probes into fundamental interactions.

Historical Development

The concept of exotic atoms, initially termed mesic atoms, emerged in the mid-1940s amid excitement over the recent discovery of mesons in cosmic rays. In 1947, and published a seminal theoretical analysis predicting that negative mesons captured by atoms would form bound states before decaying, with capture probabilities scaling with Z according to a Z^4 law for inner-shell orbits. Shortly thereafter, advanced the theory specifically for mu-mesic atoms, calculating their energy levels and linking predictions to cosmic-ray observations in his Princeton laboratory, highlighting their potential to probe nuclear sizes due to the muon's closer orbital proximity to the nucleus. The first experimental observations followed in the early 1950s as particle accelerators enabled controlled production. Pionic atoms, where a negative replaces an , were first detected in 1952 at the by M. Camac and collaborators, who observed emissions from stopped pions in targets, confirming atomic capture and radiative transitions. Independent confirmation came soon after at and , establishing pionic atoms as a tool for studying strong interactions at nuclear distances. Muonic atoms were observed in 1953 at the by V.L. Fitch and R.F. Motley using cosmic-ray muons, with subsequent experiments at providing precise 2p-to-1s transition energies that validated Wheeler's predictions and revealed nuclear finite-size effects. Key milestones unfolded through the 1960s and 1970s as new particle types and facilities expanded the field. , an atom of an and , was theoretically predicted by in 1946 as a short-lived analog, and experimentally confirmed in 1951 by Martin Deutsch through annihilation lifetime measurements in gases. In the 1970s, hypernuclear atoms—nuclei with embedded hyperons formed via interactions—saw systematic studies using high-intensity beams at Brookhaven and later , enabling of and baryon-baryon interactions. The 1980s brought antiprotonic atoms to the forefront with the startup of 's Low Energy Antiproton Ring (LEAR) in 1982, where experiments like PS176 measured transitions in light elements, probing antiproton-nucleus potentials and strong-interaction absorption. Luis W. Alvarez played a pivotal role in early pion physics at , developing detectors and accelerators that facilitated pionic atom studies and the discovery of resonances, laying groundwork for hadronic atom research. Precision spectroscopy advanced dramatically through Theodor W. Hänsch's techniques, initially honed on ordinary but extended to exotic systems like for high-resolution hyperfine measurements. In the , experiments at the (PSI) using muonic ignited the "proton radius puzzle" by yielding a proton of 0.84184(67) fm in 2010—4% smaller than electron-scattering values—prompting intense scrutiny of and nuclear structure, with a follow-up measurement in 2013 refining the muonic value to 0.84087(39) fm. As of 2025, the puzzle persists but is narrowing, with recent electronic measurements yielding values closer to 0.84 fm. Ongoing studies of quarkonia, heavy onium states like J/ψ (charmonium), at the LHC since 2010 have revealed production mechanisms in heavy-ion collisions, offering insights into quark-gluon plasma and bound-state dynamics in extreme conditions.

General Properties and Formation

Exotic atoms exhibit significantly smaller atomic radii compared to ordinary atoms due to the heavier mass of the orbiting exotic particle, which increases the in the and reduces the orbital size. For instance, in muonic atoms, the muon's mass is approximately 207 times that of the , leading to a scaled down by a factor of about 207. The general expression for the in such systems is given by a = \frac{4\pi\epsilon_0 \hbar^2}{\mu e^2}, where \mu is the of the exotic particle-nucleus system, which approximates the exotic particle's mass for heavy nuclei. This contraction brings the exotic particle into close proximity with the , altering energy levels and enabling probes of nuclear structure. The stability of exotic atoms is limited by the short lifetimes of the constituent particles, primarily governed by decays or annihilations. Leptonic exotic atoms, such as those involving muons, have lifetimes on the order of the muon's mean lifetime of 2.2 microseconds, during which the particle cascades through atomic levels before decaying. Hadronic exotic atoms are even shorter-lived due to effects and particle decays, such as the pion's lifetime of about 26 nanoseconds. Formation typically involves a cascade process where the exotic particle is captured into a high-n state and de-excites via electromagnetic transitions, often ejecting atomic electrons in the process. Exotic atoms are primarily formed using high-intensity particle beams from accelerators, where negative exotic particles like muons or pions are produced, slowed down, and stopped in a target material to be captured by target atoms. Natural formation occurs via cosmic rays, which generate muons in the Earth's atmosphere that can form muonic atoms upon interacting with matter. For electromagnetic onia like positronium, formation involves positron-electron recombination, with recent advances enabling laser cooling to stabilize and manipulate these systems at low temperatures. Spectroscopy of exotic atoms relies on detecting X-ray emissions from electronic or cascade transitions between energy levels, which are shifted due to the heavy particle's influence. These spectra reveal differences in fine and hyperfine structures arising from quantum electrodynamic (QED) corrections, such as vacuum polarization enhanced by the small orbital radii. Theoretical modeling adapts the to account for relativistic effects of the heavy particle and incorporates QED for leptonic systems to compute radiative corrections, while quantum chromodynamics (QCD) frameworks, like , describe effects in hadronic exotic atoms.

Muonic Atoms

Muonic Hydrogen

Muonic hydrogen is the simplest muonic atom, consisting of a proton bound to a negative in a hydrogen-like system. The muon's mass, approximately 207 times that of an , results in a roughly 200 times smaller than in ordinary , positioning the muon orbit much closer to the proton at about 2.8 × 10^{-11} cm. This compact structure amplifies the influence of the proton's finite size on the energy levels, particularly causing a significant shift in the ground-state energy due to nuclear volume effects, which are enhanced by a factor of approximately 200^3 compared to . Production of muonic hydrogen typically involves generating negative muon beams at particle accelerator facilities, such as the () in , where protons from a 590 MeV strike a target to produce pions that decay into . These , with initial energies around 30 MeV, are decelerated and transported to a high-purity gas target at pressures of 10–100 bar, where they thermalize and capture onto protons, forming muonic hydrogen with a lifetime limited by the muon's decay (mean life ≈ 2.2 μs). The process yields muonic atoms primarily in excited states, which cascade down, emitting characteristic X-rays. The atomic spectra of muonic hydrogen are dominated by X-ray transitions during de-excitation, with the key 2P–1S line observed at approximately 1.9 keV, enabling precise . Laser spectroscopy targets the metastable 2S–2P at a of 54.611 GHz ( ≈ 0.23 meV), which encodes the sensitive to () effects like and the Uehling potential, as well as nuclear structure contributions. A landmark experiment by the CREMA collaboration at (2009–2017) measured this 2S–2P splitting, yielding a proton root-mean-square of r_p = 0.84087(39) fm from the finite-size correction to the . This result sparked the proton radius puzzle, as it deviates by over 7σ from prior values of ≈0.877 fm obtained via and ordinary , challenging interpretations of bound-state and prompting tests for new physics. Applications of muonic hydrogen spectroscopy extend to rigorous tests of bound-state , where the enhanced nuclear effects allow extraction of proton polarizabilities and validation of QED predictions to parts per million accuracy, including two-photon exchange corrections. It also serves as a probe of proton structure, offering the most precise determination of the and insights into the proton's internal charge distribution, with ongoing experiments like at aiming to resolve the puzzle through direct muon-proton scattering comparisons.

Muonic Atoms of Heavier Elements

In muonic atoms with atomic numbers Z > 1, the negative replaces an inner-shell , orbiting much closer to the due to its being approximately 207 times that of an electron, resulting in Bohr radii scaled down by a factor of about 207. This proximity allows the muon's wavefunction to overlap significantly with the nuclear interior, enabling probes of structure through interactions at distances of order 1 fm, where electromagnetic effects alone are insufficient. Unlike in muonic , multi-nucleon effects such as deformation and become prominent, altering levels via virtual excitations of the . Key examples include muonic helium ions, such as ^4\text{He}^+ \mu, where laser spectroscopy of 2S–2P transitions has yielded the alpha-particle charge radius with high precision, r_\alpha = 1.67824(83) fm, five times more accurate than prior electron scattering results and highlighting finite-size effects in few-body systems. Similarly, laser spectroscopy of muonic ^3\text{He}^+ \mu ions in 2025 yielded the helion root-mean-square charge radius of 1.97007(94) fm. For high-Z elements, muonic lead (^{208}\text{Pb}) and uranium (^{238}\text{U}) atoms have been studied via X-ray transitions, revealing nuclear charge distributions and quadrupole moments; for instance, muonic X-rays in lead isotopes provided equivalent charge radii with uncertainties below 0.005 fm, sensitive to octupole deformations. These systems also exhibit muon transfer processes, where muons from excited muonic hydrogen atoms transfer to higher-Z targets like carbon or helium in gas mixtures, with rates measured at up to 10^{-3} per collision, influencing cascade dynamics in experimental targets. During the muonic cascade from high-n orbits (n ≈ 14) to the 1s , nuclear excitations occur via electroexcitation or hadronic interactions, producing gamma rays and altering intensities; in bismuth, for example, quadrupole hyperfine splittings up to 1 keV arise from such excitations, with probabilities enhanced by the muon's strong to vibrations. Alpha-muonic molecules, such as (\alpha \mu)^-, form transiently in muon-catalyzed fusion contexts, where the muon binds to an post-fusion, with sticking probabilities around 0.9%, limiting catalytic cycles but providing insights into short-lived bound states. Experiments at facilities like J-PARC employ for in muonic , achieving resolutions of ~1 GHz, while the muX project at targets scarce isotopes for charge radii. These measurements elucidate , with contributions up to 181 eV in calcium, and deformation effects, differing markedly from single-nucleon approximations due to collective multi-nucleon responses.

Hadronic Atoms

Pionic Atoms

Pionic atoms consist of a negatively charged (π⁻) bound to an by the interaction, replacing one or more orbital in ordinary atoms. Due to the pion's relatively large mass (approximately 273 times that of an electron) and its with nucleons, the pion orbits much closer to the nucleus, often penetrating its surface and experiencing significant hadronic effects that dominate over electromagnetic ones. The lifetime of the pionic state is on the order of 10^{-14} s, primarily limited by strong absorption processes in which the pion interacts with nucleons to form other hadrons. These atoms are typically formed by directing low-energy negative beams, produced at facilities like cyclotrons or synchrotrons, onto thin targets of the desired material, where the pions are slowed and captured into high-n Rydberg states (n ≈ 15–20). The captured then cascades to lower orbits through radiative transitions (emitting X-rays), electron emission, or collisional de-excitation, with the strong interaction influencing the cascade probabilities and final state populations. This method allows selective formation in gaseous or solid targets, enabling high-resolution . The spectra of pionic atoms exhibit characteristic level shifts and broadenings due to the pion-nucleus optical potential, which models the low-energy as a complex potential with a real part causing repulsion or attraction and an imaginary part accounting for . In -like pionic systems, such as pionic (π⁻p), Stark mixing—arising from external or atomic collisions—induces transitions between states with different magnetic quantum numbers, enhancing rates and altering intensities. For example, measurements of the ground-state (1s) transition lines reveal a hadronic shift ε_{1s} = 7.086 ± 0.006 and broadening Γ_{1s} = 0.823 ± 0.068 (as of 2021). Key experimental studies focus on light pionic atoms to isolate fundamental strong interaction parameters. In pionic hydrogen, precision X-ray spectroscopy probes isospin symmetry breaking through the relation between level shifts and pion-nucleon (πN) s-wave scattering lengths, yielding values like the isoscalar scattering length a⁺ ≈ 0.0086 m_π^{-1}, which tests predictions of chiral perturbation theory (ChPT). Pionic helium atoms, on the other hand, highlight nuclear density effects, with collisional shifts and broadenings in the transition lines (e.g., 2p–1s at low temperatures) revealing in-medium modifications to the pion propagator and effective scattering lengths. These investigations provide critical benchmarks for ChPT, constraining the low-energy constants and exploring chiral symmetry in nuclear matter without reliance on free πN scattering data.

Kaonic Atoms

Kaonic atoms are exotic atomic systems formed when a negatively charged kaon (K⁻ meson) replaces an orbiting electron around a nucleus, leading to a bound state where the kaon interacts strongly with the nuclear constituents due to its quark content (u s-bar). The K⁻ carries intrinsic strangeness quantum number S = -1, distinguishing kaonic atoms from non-strange hadronic systems like pionic atoms and enabling the study of strangeness-introducing processes at low energies. In these systems, the kaon is initially captured into high-principal-quantum-number (n) orbits via electromagnetic interactions and then cascades downward, emitting characteristic X-rays; however, as the kaon approaches low orbits (n ≤ 5–10, depending on nuclear charge Z), the strong kaon-nucleus interaction dominates, causing energy level shifts and broadenings. The primary absorption mechanism in kaonic atoms involves the K⁻ interacting with nucleons to produce hyperons and pions, such as K⁻ p → Λ π⁰ or K⁻ n → Λ π⁻ (or Σ π channels), which converts the strangeness into hypernuclear components and contributes to the imaginary part of the kaon-nucleus optical potential. This process limits the lifetime of low-lying states, with level widths arising from the kaon-nucleon and probabilities. Formation of kaonic atoms typically requires low-energy K⁻ beams (momentum ~100–200 MeV/c) produced at facilities like the DAΦNE e⁺e⁻ collider at INFN-LNF, where kaons emerge from φ-meson decays with minimal background, or J-PARC, which provides high-intensity beams for in-flight reactions. A notable phenomenon in kaonic atoms is the existence of deeply bound states, where the K⁻ is trapped in a compact due to the attractive real part of the kaon-nucleus potential, potentially forming multi-nucleon systems like K⁻ pp in light nuclei such as ⁴He. These states, predicted by models incorporating the Λ(1405) as a K⁻ N , feature binding energies up to ~100 MeV but widths of ~50 MeV from strong ; experimental searches use stopped-kaon reactions like ⁴He(stopped K⁻, n) at or in-flight ³He(K⁻, n) at J-PARC E15, revealing peak structures suggestive of such bindings. Key experiments focus on precision X-ray spectroscopy to probe these interactions. For instance, the SIDDHARTA collaboration at DAΦNE measured the kaonic 2p → 1s , yielding a strong-interaction shift of -283 ± 36 (stat) ± 6 (syst) eV and width of 541 ± 89 (stat) ± 22 (syst) eV (as of 2011), which refine the kaon-proton to Re a_{K^- p} = -0.33^{+0.11}{-0.09} + i 0.40^{+0.13}{-0.04} fm. In kaonic ⁴He, the 3d → 2p provides sub-eV precision on the 2p level shift of 0 ± 6 (stat) ± 2 (syst) eV (as of 2012), constraining the kaon-nucleus potential for A=4 systems. The SIDDHARTA-2 experiment, ongoing as of 2024, performed the first measurement of the kaonic 2p → 1s line (~7 keV) to extract the kaon-neutron , essential for isospin-symmetric potentials (analysis ongoing as of November 2025). These measurements test the strangeness-exchange component of the kaon-nucleus optical potential, derived from chiral SU(3) dynamics, and provide benchmarks for validating low-energy QCD models including coupled-channel effects with channels. By linking kaonic atomic level widths to hyperon production thresholds, kaonic atoms bridge atomic and , offering insights into in dense matter relevant to neutron stars and the formation of hypernuclei.

Antiprotonic Atoms

Antiprotonic atoms consist of an orbiting the of an ordinary atom, replacing one or more and forming a system when the nucleus has multiple nucleons. The , with its negative charge and mass approximately 1836 times that of an , occupies atomic orbits scaled down by this , resulting in Bohr radii about 1/1836 that of electronic orbits. Due to the antiproton's of 1/2, high-angular-momentum states with nearly circular orbits (ℓ ≈ n-1) are accessible, minimizing overlap with the and delaying annihilation. Ultimately, the annihilates with a nuclear proton via the strong interaction, predominantly producing 3–5 pions with a total energy release of about 1.88 GeV. These atoms form when low-energy antiprotons, typically with kinetic energies below 1 keV, are captured by target atoms. Antiproton beams are generated at facilities such as CERN's , where protons from the strike a production target to create antiprotons, which are then decelerated, cooled, and extracted as a slow beam. Upon entering a gaseous or cryogenic target, the antiprotons lose energy through electromagnetic interactions and are captured into high (n ≈ 30–40) Rydberg states, often ejecting atomic electrons in the process. In dense targets, the Day-Snow-Sucher effect enhances capture into high-ℓ states, stabilizing the initial orbit. A distinctive phenomenon in antiprotonic atoms is the competition between radiative deexcitation and . In most elements, the reaches low orbits (n < 10) within picoseconds, where strong interaction effects dominate, leading to rapid . However, in antiprotonic helium—formed by capturing an in helium gas—about 3% of atoms enter long-lived metastable states with lifetimes of 1–2 μs, owing to suppressed Auger transitions and reduced Stark mixing in high-ℓ orbits. These states enable observation of X-ray cascades during deexcitation, where the emits characteristic X-rays (e.g., in the keV range) via Δn = Δℓ = -1 radiative transitions. Such cascades reveal the antiproton-nucleus optical potential, modeled phenomenologically as a complex interaction with real (attractive) and imaginary (absorptive) parts, which accounts for level shifts and broadenings (e.g., 1s state shift of ~700 eV and width of ~1000 eV in light nuclei). Experimental studies of antiprotonic atoms primarily occur at CERN's AD, with the ASACUSA collaboration leading laser spectroscopy efforts on antiprotonic helium. By inducing resonant transitions between metastable states using pulsed lasers, ASACUSA has measured X-ray and microwave transitions with fractional precisions of 2–5 × 10^{-9}, enabling determination of the antiproton-to-electron mass ratio to 8 × 10^{-10} accuracy (m_{\bar{p}}/m_e = 1836.1526736(39)). For antiprotonic hydrogen (an antiproton bound to a proton, or protonium), hyperfine structure measurements have probed strong interaction parameters, revealing splittings influenced by isospin mixing and confirming shifts in the 2p fine structure (e.g., Kα X-ray at ~1.7 keV). These experiments test CPT symmetry by comparing antiprotonic spectra to their matter counterparts, with deviations bounded at parts per billion. Antiprotonic atoms serve as precision probes of the strong interaction at low energies, where perturbative is inapplicable, by extracting nucleon-antinucleon scattering lengths and effective potentials from spectral data. In heavier nuclei, they facilitate studies of exotic nuclear states, such as quasi-bound antiproton-nucleus configurations or enhanced annihilation on neutron-rich surfaces, offering insights into nuclear medium effects and G-parity violations without the complications of kaonic or pionic systems.

Onium Atoms

Electromagnetic Onia (Positronium and Muonium)

Electromagnetic onia are purely leptonic bound states governed exclusively by , free from strong or weak nuclear interactions. These systems serve as precise analogues to the , enabling tests of QED predictions for bound-state dynamics at reduced masses distinct from the proton-electron case. and exemplify this class, with their properties—such as energy levels, lifetimes, and decay modes—yielding insights into fundamental constants like the and probing potential deviations from the . Positronium (Ps), the exotic atom formed by an electron (e⁻) and its antiparticle, the positron (e⁺), binds via the Coulomb interaction with a reduced mass of \mu = m_e / 2, where m_e is the electron mass. This half-electron reduced mass scales QED effects differently from hydrogen, providing a unique laboratory for bound-state QED validations at intermediate precision levels. The ground-state hyperfine splitting (Δν_HFS) in positronium, arising from spin-spin interactions, is predicted by QED to be 203.387 GHz, incorporating higher-order corrections up to O(α⁵ ln α) and beyond. Experimental measurements, such as the 2015 value of 203.3942(16) GHz with ~8 ppm relative precision, show a discrepancy of about 15 ppm (3.9σ) with theory, highlighting ongoing efforts to resolve subtle discrepancies that could signal new physics. Positronium forms when energetic positrons, typically from β⁺ decay sources like ²²Na, slow down and thermalize in low-density matter such as gases or porous solids, capturing ambient electrons with yields up to 50% in optimized targets. The ortho-positronium (triplet) state, with parallel spins, decays predominantly via three-photon (3γ) annihilation into gamma rays totaling 1.022 MeV, with a lifetime of ~142 ns, while the para-positronium (singlet) state annihilates into two photons (2γ) in ~125 ps. These decay channels, forbidden or suppressed by C-parity and angular momentum conservation, enable precise lifetime spectroscopy for QED decay rate tests. Precision experiments, including microwave and laser spectroscopy of fine and hyperfine structures, are conducted at facilities like the Low Energy Positron Trap Accumulator (LEPTA) at JINR, where stored positron beams facilitate high-resolution measurements of transitions like 1S-2S at ~1.8 eV. Such studies test QED to parts per billion, constraining beyond-Standard-Model effects like axion-like particles. Muonium (Mu), consisting of a positive muon (μ⁺) and an electron (e⁻), resembles hydrogen but with the proton replaced by the lighter muon, yielding a reduced mass ≈ 0.9947 m_e due to m_μ ≈ 207 m_e. Its short lifetime is dictated by the muon's weak decay, μ⁺ → e⁺ + ν_e + ν̄_μ, with a mean of 2.197 μs, limiting observable states to low-lying levels before dissociation or decay. The primary decay channel emits a positron isotropically in the muon's rest frame, though spin polarization allows angular correlation studies. Muonium forms readily when μ⁺ beams (from pion decay) stop in dilute gases or vacuum interfaces, epithermally capturing electrons with formation fractions exceeding 80% in hydrogen targets; delayed thermalization can enhance yields in solids. Muonium's utility extends to precision measurements, particularly in preparing polarized muon beams for g-2 experiments. At facilities like J-PARC, muonium is photoionized to produce low-emittance μ⁺ beams, accelerated via RF quadrupoles, and injected into compact storage rings for anomalous magnetic moment (a_μ = (g-2)/2) determinations, achieving sensitivities down to 10^{-10} with reduced systematics compared to traditional methods. Hyperfine spectroscopy in muonium, analogous to hydrogen's 1420 MHz line but shifted to 4463.303 MHz, tests QED at muon-electron mass scales, validating bound-state corrections and the muon-electron mass ratio to 10^{-9}. As purely electromagnetic systems, positronium and muonium enable hydrogen-like calibrations of α and m_e/m_μ without hadronic uncertainties, underpinning QED's predictive power and searches for lepton-flavor violations.

Quarkonia

Quarkonia are bound states formed by a heavy quark and its antiquark, primarily charmonium systems composed of a charm quark and anticharm quark (c\bar{c}) and bottomonium systems of a bottom quark and antibottom quark (b\bar{b}). These mesons have zero strangeness, charm, and baryon number (S = C = B = 0), and their quantum numbers I^G J^{PC} determine their spectroscopic notation. Due to the large masses of the constituent quarks (charm ~1.3 GeV, bottom ~4.2 GeV), quarkonia exhibit non-relativistic dynamics, analogous to the but governed by instead of . In this framework, the quark-antiquark pair orbits under the color-confining strong force, with velocities much less than the speed of light (v/c \ll 1), allowing effective field theories like to describe their binding and interactions. The mass spectrum of quarkonia reveals a hierarchy of states labeled by principal quantum number n and orbital angular momentum L, with singlets and triplets for total spin S=0 or $1. In charmonium, the vector ground state J/\psi(1^3S_1) has a mass of 3.0969 GeV and was discovered in November 1974 through electron-positron annihilation experiments at and proton-beryllium collisions at , marking the first observation of charmed quarks. Higher states include the \psi(2S) at 3.686 GeV and P-wave \chi_{cJ} multiplets around 3.5 GeV. For bottomonium, the analogous ground state \Upsilon(1^3S_1) has a mass of 9.4603 GeV, discovered in 1977 at via high-energy proton collisions producing muon pairs, confirming the bottom quark's existence. The spectrum extends to excited states like \Upsilon(2S) at 10.023 GeV and \chi_{bJ} P-waves near 9.9 GeV, with precision masses refined through ongoing experiments and summarized in Particle Data Group (PDG) updates, such as the 2025 edition. Quarkonia are produced copiously in high-energy collisions, enabling detailed spectroscopic studies. Electron-positron colliders like at the BEPCII storage ring in China operate at center-of-mass energies from 2 to 5 GeV, directly forming charmonium resonances (e.g., e^+e^- \to J/\psi) and probing their decays with high precision, including rare modes and transitions. At hadron colliders such as the , quarkonia emerge from proton-proton, proton-lead, and lead-lead collisions via gluon fusion or fragmentation, with experiments like , , , and measuring production cross-sections across rapidities and transverse momenta up to Run 3 data (2022–ongoing, as of 2025), revealing insights into cold nuclear matter effects and hot quark-gluon plasma suppression. Recent Run 3 data from 2022–2025 have further refined these measurements, providing enhanced constraints on quarkonium production mechanisms in heavy-ion collisions. Key decay modes of quarkonia include radiative electric dipole (E1) transitions, such as J/\psi \to \gamma \eta_c with a width of about 5.6 keV, and magnetic dipole (M1) transitions like \psi(2S) \to \gamma \eta_c(1S), which probe spin-flip dynamics and hyperfine splittings. Leptonic decays, e.g., J/\psi \to e^+e^- with a partial width of 5.55 keV, are particularly sensitive to the quarkonium wave function at the origin and test potential models of the quark-antiquark interaction. The , V(r) = -\frac{4\alpha_s}{3r} + \sigma r, captures the short-distance Coulomb-like attraction from one-gluon exchange (with strong coupling \alpha_s) and the long-distance linear confinement (\sigma \approx 0.18 GeV²), accurately reproducing the quarkonium mass spectrum, radial splittings, and leptonic widths when solved in the non-relativistic Schrödinger equation. As non-relativistic QCD-bound systems, quarkonia provide unique probes of quark confinement, heavy quark mass effects, and the transition from perturbative to non-perturbative QCD regimes. Their production and dissociation in heavy-ion collisions at the LHC quantify the strength of the quark-gluon plasma, while precision spectroscopy constrains \alpha_s and the string tension \sigma, with PDG 2025 updates incorporating lattice QCD and experimental data for masses accurate to ~0.1 MeV, advancing understanding of QCD dynamics up to 2025.

Hypernuclear Atoms

Lambda-Hypernuclei

Lambda hypernuclei are exotic nuclear systems consisting of a lambda (Λ) hyperon bound to an ordinary nuclear core, forming a hypernucleus that acts as the central positively charged entity orbited by an electron cloud, analogous to standard atoms. The Λ hyperon, a baryon with strangeness S = -1, zero isospin, and a mass of approximately 1115.7 MeV/c², replaces or adds to nucleons in the nucleus, enabling the exploration of strangeness in nuclear structure. A prototypical example is the hypertriton ^3_\LambdaH, comprising a proton, neutron, and Λ, with the Λ loosely bound to the deuteron-like core. These systems extend the periodic table into the strange sector and serve as laboratories for probing hyperon-nucleon (YN) interactions under nuclear conditions. Formation of Λ hypernuclei primarily occurs via associated strangeness-exchange reactions using kaon beams incident on nuclear targets. The fundamental process is K^- + p \to \Lambda + \pi^0, where the negatively charged kaon interacts with a proton in the target nucleus, producing a Λ that is subsequently captured by the residual nucleons; associated production channels like (K^-, \pi^-) on neutrons yield analogous results. Early discoveries relied on nuclear emulsion detectors to capture hypernuclear stars from stopped K^- interactions, while contemporary methods use in-flight kaon beams or stopped kaons with magnetic spectrometers for momentum analysis. For instance, at facilities like KEK and BNL, (K^-, \pi^-) reactions have populated excited states in p-shell hypernuclei such as ^{12}_\LambdaB. Key properties of Λ hypernuclei include the Λ separation energy B_\Lambda, defined as the difference between the hypernucleus mass and the sum of the core nucleus and free Λ masses, which quantifies the binding strength. For the hypertriton ^3_\LambdaH, B_\Lambda \approx 0.13 \pm 0.05 MeV, indicating a weakly bound system, whereas in p-shell hypernuclei like ^{12}_\LambdaB, B_\Lambda \approx 11.5 MeV, and in heavier sd-shell species such as ^{51}_\LambdaV, values reach 23-25 MeV, reflecting deeper potentials in denser nuclear matter. The Λ decays weakly, predominantly via \Lambda \to p \pi^- (branching ratio \sim 64\%) and \Lambda \to n \pi^0 (\sim 36\%), but within the nucleus, non-mesonic channels like \Lambda N \to NN dominate due to Pauli suppression of pions, with lifetimes shortened to \sim 10^{-10} s. The spin-independent and spin-dependent components of the Λ-nucleus potential, derived from YN forces, manifest in level splittings and transitions observable via \gamma-ray spectroscopy. Major experiments have advanced the field through high-resolution spectroscopy and decay studies. The FINUDA collaboration at DAΦNE (Frascati) utilized stopped K^- from \phi \to K^+ K^- decays in thin targets to produce light hypernuclei, measuring binding energies and weak decay rates for neutron-rich species like ^6_\LambdaH (B_\Lambda = 4.0 \pm 1.1 MeV) and investigating non-mesonic processes such as \Lambda np \to nnp. At Jefferson Lab (JLab) Hall C, electroproduction via the (e, e' K^+) reaction has enabled precise spectroscopy of Λ hypernuclei up to mass A \sim 52, with experiments like E05-115 (2009) determining B_\Lambda for ^{12}_\LambdaB at 11.45 ± 0.07 ± 0.11 MeV and spin-dependent potentials; ongoing efforts through 2025, including E12-15-001, target medium-heavy systems for deeper insights into YN couplings. These measurements, complemented by \gamma-ray experiments at Mainz (MAMI) and J-PARC, have refined ΛN scattering lengths to a_{\Lambda p} \approx -0.9 fm and a_{\Lambda n} \approx 1.5 fm. As of 2025, experiments at J-PARC and RHIC continue to refine YN potentials, with new data from the HYP2025 conference confirming key binding energies and probing multi-strange systems. The significance of Λ hypernuclei lies in their role as probes of the elusive YN interaction, inaccessible via free scattering due to the Λ's \sim 2.6 \times 10^{-10} s lifetime. Spectroscopy reveals the Λ effective mass (m_\Lambda^* \approx 0.7-0.8 m_\Lambda) and spin-orbit splittings, testing models like the and . These insights inform strangeness in dense matter, such as hyperon emergence in neutron stars, and constrain the nuclear equation of state by quantifying Λ binding in neutron-rich environments.

Other Hyperonic Systems

Other hyperonic systems encompass hypernuclei formed by charged or multi-strange hyperons, such as sigma (Σ), xi (Ξ), and omega (Ω) variants, which introduce additional complexity due to their charge and strangeness content compared to neutral lambda systems. These systems are characterized by shorter lifetimes arising from strong interaction decay channels, including the ΣN → ΛN conversion process, which broadens the resonance widths significantly. For instance, the ground state of the ^4_\Sigma \mathrm{He} hypernucleus has been observed with a binding energy indicating strong attraction, but its lifetime is limited to approximately 0.1–1 ps due to these non-mesonic strong decays. Xi (Ξ⁻) and omega (Ω⁻) hypernuclei, carrying two and three units of strangeness respectively, are primarily produced in high-energy heavy-ion collisions at facilities like RHIC and the LHC, where the high baryon densities facilitate multi-strange baryon coalescence. The STAR collaboration at RHIC has contributed to the study of these systems through measurements of hyperon production and flow, providing indirect constraints on Ξ and Ω binding in nuclear matter, with conversion widths on the order of 10–20 MeV reflecting strong ΞN and ΩN interactions. A landmark direct observation is the nuclear s-state of the ^{15}_\Xi \mathrm{C} hypernucleus, identified via emulsion-counter hybrid experiments, revealing a binding energy of about 4.7 MeV and probing the Ξ-nucleus potential. Omega hypernuclei remain elusive in direct observation but are theoretically predicted to form weakly bound states in dense matter, with production yields estimated from heavy-ion data. Double hypernuclei, such as ^6_{\Lambda\Lambda} \mathrm{He}, represent systems with two strange baryons and have been detected using nuclear emulsion techniques, which allow precise tracking of decay topologies. The Nagara event confirmed the existence of ^6_{\Lambda\Lambda} \mathrm{He} with a Λ-Λ bond energy of 1.01 ± 0.20 MeV ± 0.18 (syst.) MeV, highlighting repulsive components in the Λ-Λ interaction at short distances. These observations rely on Ξ⁻ capture at rest followed by sequential weak decays, enabling the extraction of two-body interaction parameters. The study of these hyperonic systems probes multi-body hyperon-nucleon and hyperon-hyperon interactions in dense environments, offering insights into the strangeness sector of QCD. Lifetime measurements and binding energies from experiments like STAR and emulsion detectors help constrain effective potentials, resolving ambiguities in hyperon couplings. Furthermore, these findings have astrophysical implications, particularly for the hyperon puzzle in neutron stars, where multi-strange hyperons soften the equation of state but must be balanced to support observed masses above 2 solar masses.

Quasiparticle Exotic Atoms

In a broader condensed matter physics context, quasiparticle exotic atoms refer to collective excitations that form bound, atomic-like states, analogous to exotic atoms in particle physics but arising from interactions within solids rather than subatomic particle substitutions. These include (electron-hole pairs) and (charge carriers dressed by phonons), which exhibit hydrogen-like properties but are studied for their roles in materials science and optoelectronics.

Excitons

Excitons are quasiparticles formed by a bound state of an electron and a hole in semiconductors and insulators, where the electron and hole are attracted by the Coulomb interaction, behaving analogously to a but with an effective reduced mass given by \mu = \frac{m_e m_h}{m_e + m_h}, with m_e and m_h being the effective masses of the electron and hole, respectively. This binding creates a neutral entity that can propagate through the material while conserving excitation energy. Excitons are classified into types based on their spatial extent and the material context. Wannier excitons, prevalent in inorganic semiconductors, feature large radii—often tens of nanometers—due to weaker binding and delocalization over multiple unit cells. In contrast, Frenkel excitons occur in molecular solids and organic materials, with small radii comparable to a single molecule, resulting from stronger local interactions. Additionally, Rydberg excitons represent highly excited states with principal quantum numbers n \gg 1, exhibiting exaggerated properties like large dipole moments, as observed in materials such as cuprous oxide. Excitons form primarily through optical excitation, where a photon is absorbed to promote an electron from the valence to the conduction band, creating an electron-hole pair that subsequently binds via Coulomb attraction. Their lifetimes typically range from nanoseconds to microseconds, influenced by radiative recombination, non-radiative decay, and environmental factors like temperature and density. Key properties of excitons include their binding energy, approximated in the effective mass model as E_b = \frac{\mu e^4}{2 \hbar^2 \varepsilon^2}, where e is the electron charge, \hbar is the reduced Planck's constant, and \varepsilon is the dielectric constant of the medium, which determines the stability and dissociation threshold of the pair. In two-dimensional materials such as transition metal dichalcogenides (e.g., MoSe_2), excitons exhibit enhanced binding due to reduced screening, enabling Bose-Einstein condensation at elevated temperatures compared to three-dimensional systems. Excitons play a central role in optoelectronic applications, including light-emitting diodes, solar cells, and photodetectors, where their efficient formation and recombination enhance device efficiency. Up to 2025, research in moiré superlattices—formed by twisting layered van der Waals materials—has demonstrated tunable excitons with engineered properties, such as modified binding energies and enhanced nonlinear optical responses, paving the way for advanced photonic devices.

Polarons

A polaron is a quasiparticle consisting of a charge carrier, such as an electron or hole, strongly coupled to lattice vibrations () in a solid, resulting in the carrier being dressed by a cloud of that distorts the surrounding lattice. This phenomenon occurs primarily in polar materials where long-range electron- interactions dominate, leading to large polarons as described by the Fröhlich model; in contrast, small polarons form under strong coupling conditions with localized lattice distortions, often modeled by the Holstein Hamiltonian. The dynamics of large polarons are captured by the Fröhlich Hamiltonian: H = \frac{p^2}{2m} + \sum_{\mathbf{q}} \hbar \omega_{\mathbf{q}} b_{\mathbf{q}}^\dagger b_{\mathbf{q}} + \sum_{\mathbf{q}} V_{\mathbf{q}} (b_{\mathbf{q}} + b_{\mathbf{q}}^\dagger) e^{i \mathbf{q} \cdot \mathbf{r}}, where p^2/2m represents the kinetic energy of the bare charge carrier with mass m, the second term describes the free phonon bath with dispersion \omega_{\mathbf{q}} and bosonic operators b_{\mathbf{q}}^\dagger, b_{\mathbf{q}}, and the third term encodes the linear electron-phonon coupling with strength V_{\mathbf{q}}. Polarons form through charge injection into polar crystals, where the injected carrier polarizes the ionic lattice, creating a potential well that binds the phonon cloud to the carrier. This dressing increases the effective mass m^* > m, thereby reducing charge according to \mu = e \tau / m^*, where e is the carrier charge and \tau is the scattering time. Key properties of polarons include self-trapping in one-dimensional systems, where the carrier localizes in a distortion exceeding thermal energy barriers, and the formation of bipolarons—bound pairs of polarons with charge \pm 2e and zero , stabilized by enhanced relaxation. Such bipolarons are prevalent in materials with non-degenerate ground states. Polarons, both large and small, have been experimentally observed in lead halide perovskites through spectroscopic signatures of relaxation and in via transport measurements showing activated conduction. Polarons are crucial for understanding charge transport limitations in insulators and wide-bandgap semiconductors, where they mediate hopping mechanisms and explain thermally activated mobility. As of 2025, advances in two-dimensional polarons, particularly interfacial and spin polarons in , have revealed enhanced electron-phonon coupling at magic angles, enabling tunable behaviors with implications for quantum devices.

Exotic Molecules and Systems

Muonic Molecules

Muonic molecules are short-lived bound states formed when a negative replaces an in a diatomic isotope system, such as the prototypical (d t μ) molecule involving (d) and (t) nuclei. In this configuration, the muon's mass, approximately 207 times that of an , shrinks the orbital radius to about 1/207 of the , resulting in internuclear distances on the order of 500 fm in resonant states that facilitate nuclear interactions. These resonant states, often Feshbach-type resonances below the threshold, enable efficient formation and subsequent processes like by aligning the energy levels for temporary binding. Formation of muonic molecules occurs through muon implantation into a deuterium-tritium (D-T) gas at low temperatures, where the muon first forms a muonic (e.g., μ⁻ + D → dμ) before colliding with another to create the triatomic : μ + D T → (D T μ). This process cycles rapidly, with the muon typically released after (d + t → ⁴He + n + 17.6 MeV) to catalyze further s, though sticking to the helium reduces efficiency with a probability of about 0.8%. The resonant formation rate for (d t μ) is enhanced when the collision aligns with specific vibrational states of the target , achieving rates up to 10⁹ s⁻¹ under optimal conditions. Recent theoretical studies as of 2025 have further explored reaction processes in muonic molecules like ddμ using advanced simulations. Kinetics of muonic molecules in (μCF) rely on resonance conditions that minimize muon sticking and maximize recycling, with overall cycle rates exceeding 10⁸ s⁻¹ at in D-T mixtures. The rate within the resonant (d t μ) reaches approximately 10⁸–10⁹ s⁻¹, driven by the reduced at close internuclear separations, allowing up to 150 fusions per before decay or loss. These rates are temperature-dependent, increasing with thermal energy to optimize isotopic populations and epi-thermal production, though challenges like ortho-para conversion in isotopes limit practical yields. Early experiments on μCF with muonic molecules were conducted at in the 1980s, confirming resonant (d t μ) formation and measuring cycle rates in gaseous and liquid D-T targets, achieving up to 100 fusions per . Modern precision studies at utilize high-intensity beams to probe sticking probabilities and deexcitation rates, providing data on resonant states in isotopes with rates aligning to 10⁸ s⁻¹ cycles. These efforts have refined kinetic models, incorporating three-body dynamics for better prediction of fusion yields. The significance of muonic molecules lies in their role as an analogue to , enabling reactions at ambient temperatures without confinement, though production costs limit energy applications. They also serve as precise testbeds for molecular (QED), where calculations of binding energies and transition rates in strong fields verify QED predictions to high accuracy, as seen in studies of radiative decays and formation mechanisms.

Hadronic Molecules

Hadronic molecules represent loosely bound or resonant states formed by interacting via the strong force, analogous to the deuteron as a proton-neutron bound system. These structures arise in exotic atomic contexts when a hadron replaces an and forms molecular-like configurations with nuclear constituents, such as in pion-deuteron or kaon-deuteron systems, where the binding is dominated by short-range strong interactions rather than electromagnetic forces. In pion-deuteron systems, the interaction manifests through low-energy , with the scattering length determined primarily by charge-exchange processes that probe the underlying pion-nucleon dynamics. Theoretical calculations incorporating multiple rescattering and virtual charge exchange yield a real part of the pion-deuteron scattering length approximately -0.04 , highlighting the weakly attractive nature of the potential due to the strong force. These systems serve as effective probes for isospin-symmetric meson-baryon couplings, with corrections from deuteron structure ensuring the approximation's validity. Kaonic deuterium exemplifies another class, where the antikaon interacts with the to form a configuration sensitive to kaon-nucleon lengths. Precision measurements of the strong shift and width in the 1s state of kaonic deuterium constrain the isospin-dependent lengths, with values indicating a complex potential mixing attractive Lambda(1405) contributions and repulsive Sigma-hyperon channels, resulting in a length of approximately -1.5 + 1.1 i . Such studies reveal the role of coupled-channel effects in the S=-1 sector. Antiprotonic molecular states, as in antiprotonic deuterium, involve the binding to the deuteron via strong nucleon-antinucleon forces, leading to quasibound resonances in partial waves. Calculations using separable potentials predict level shifts and widths for states like 1S and 2P, with the strong interaction broadening the by about 1 keV due to and scattering channels. These states highlight the competition between attraction and strong repulsion, forming short-lived molecular configurations before . Formation of these hadronic molecules typically occurs by directing beams onto molecular targets like gas, allowing the exotic particle to slow down and capture into orbits, followed by strong-force induced binding or resonant in specific partial waves. Resonant states, such as those in p-wave channels, emerge from the interplay of attractive and repulsive meson-baryon potentials, often modeled via chiral effective field theory. Experiments at facilities like COSY at FZ have provided key data on meson-baryon interactions through measurements of and production in nucleon-deuteron collisions, enabling extraction of relevant to hadronic molecular formation. For instance, ANKE spectrometer data on near-threshold meson production yield insights into pion-nucleon couplings, with cross sections around 100 μb establishing the scale of low-energy interactions. Lifetime studies of these exotic states, inferred from widths, indicate femtosecond-scale durations, limited by channels. The significance of hadronic molecules lies in their provision of experimental access to meson-baryon dynamics at low energies, testing and unitarity in coupled channels. These systems bridge single-hadron exotic atoms to multi- interpretations, where molecular pictures compete with compact quark configurations, as seen in dibaryon resonances potentially harboring hexaquark cores.

References

  1. [1]
    Exotic Atom - an overview | ScienceDirect Topics
    Exotic atoms are defined as atoms that contain at least one subatomic particle that is not typically found in ordinary atoms, such as muons or pions, ...
  2. [2]
    [PDF] Exotic atoms
    Mar 16, 2015 · What for? • Study of the particle. – mass of the exotic particle (pion, kaon, antiproton, sigma).
  3. [3]
    [PDF] Exotic atoms - HAL
    Sep 11, 2004 · In this paper, I have explored different aspects of the physics of light muonic, pionic and antiprotonic atoms. I have left out many aspects of ...
  4. [4]
    The size of the proton | Nature
    Jul 8, 2010 · The size of the proton. Download PDF. Letter; Published: 08 July 2010. The size of the proton. Randolf Pohl,; Aldo Antognini,; François Nez ...
  5. [5]
    [1301.0905] Muonic hydrogen and the proton radius puzzle - arXiv
    Jan 5, 2013 · The proton radius puzzle is the significant disagreement between the proton radius extracted from muonic hydrogen and electronic hydrogen or ...
  6. [6]
  7. [7]
    The Capture of Negative Mesotrons in Matter | Phys. Rev.
    A detailed discussion of the energy loss of negative mesotrons in matter is presented. The energy range considered is from +2000 ev to the lowest quantized ...
  8. [8]
    [PDF] John A. Wheeler - National Academy of Sciences
    Accordingly, he developed in detail the theory of mu-mesic atoms and linked the theory to experiment, including observations, in his cosmic-ray laboratory by ...
  9. [9]
    [PDF] CERN News
    The first ones to be identified were pionic atoms in which a negative pion had taken the place of an electron. In. 1953 muonic atoms were discovered at.
  10. [10]
    Evidence for the Formation of Positronium in Gases | Phys. Rev.
    This paper, "Evidence for the Formation of Positronium in Gases" by Martin Deutsch, was published in Phys. Rev. 82, 455 on May 1, 1951.
  11. [11]
    [PDF] Positronium Physics - CERN Indico
    Mar 24, 2020 · Overview. •In this lecture, I will introduce the field of Positronium physics from a historical perspective.
  12. [12]
    [PDF] Observation of an Antimatter Hypernucleus - OSTI.GOV
    Mar 10, 2010 · The first observation of any hypernucleus was made in 1952 using a nuclear emulsion cosmic ray detector (5). Here, we present the.
  13. [13]
    [PDF] A short history of antiprotonic atoms - ECT* Indico
    Jun 16, 2019 · proposed and rapidly built: AA start-up in 1980, LEAR began operation in 1982, AC from 1987 onwards ... by the preceding LEAR experiments PS174 ...<|control11|><|separator|>
  14. [14]
    History – Lawrence Berkeley National Laboratory
    Luis W. Alvarez was awarded the 1968 Nobel Prize in physics for his contributions to elementary particle physics, particularly for his discovery of “resonance ...Missing: pionic | Show results with:pionic
  15. [15]
    [PDF] J = µ MASS (atomic mass units u) µ MASS HTTP://PDG.LBL.GOV ...
    Jun 1, 2020 · The muon's mass is obtained from the muon-electron mass ratio as deter- mined from the measurement of Zeeman transition frequencies in ...
  16. [16]
    Orbital Collapse in Exotic Atoms and Its Effect on Dynamics
    We propose a simple method to estimate for exotic noble atoms from atomic model potentials. The estimated values agree with those calculated by DFT. Figure 1
  17. [17]
    Measurement of the Positive Muon Lifetime and Determination of the ...
    Jan 25, 2011 · The muon lifetime gives the most precise value for the Fermi constant: (0.6 ppm). It is also used to extract the singlet capture rate, which ...
  18. [18]
    [PDF] arXiv:physics/0510126v1 [physics.atom-ph] 14 Oct 2005
    Exotic atoms and ions are formed when a heavy particle is captured on an atomic nucleus. In the process, many or all electrons of the atom are ejected and one ...
  19. [19]
    DOE Explains...Muons - Department of Energy
    The muons that hit the Earth result from particles in the Earth's atmosphere colliding with cosmic rays—high-energy protons and atomic nuclei that move through ...
  20. [20]
    Cooling positronium to ultralow velocities with a chirped laser pulse ...
    Sep 11, 2024 · One-dimensional chirp cooling was used to cool a portion of the dilute positronium gas to a velocity distribution of approximately 1 K in 100 ns.
  21. [21]
    Precision spectroscopy of light exotic atoms - ScienceDirect.com
    Besides X-ray spectroscopy, additional important methods for investigating exotic atoms were developed, which are not subjects of this review. Laser techniques ...
  22. [22]
    Testing Quantum Electrodynamics with Exotic Atoms | Phys. Rev. Lett.
    Apr 29, 2021 · We propose an alternative method with exotic atoms and show that transitions may be found between circular Rydberg states where nuclear contributions are ...
  23. [23]
    [PDF] Testing Quantum Electrodynamics with Exotic Atoms - HAL
    Nov 17, 2020 · We have performed calculations using the Multicon- figuration Dirac-Fock General Matrix Elements (MCD-. FGME) code, which can evaluate energies, ...
  24. [24]
    Introduction | Muon Physics | PSI - Paul Scherrer Institut
    Muonic hydrogen is an exotic hydrogen atom where a muon, 200 times heavier than an electron, orbits the proton, making its orbit 200 times closer.
  25. [25]
    [2205.10076] The proton structure in and out of muonic hydrogen
    May 20, 2022 · This paper studies proton structure using laser spectroscopy of muonic hydrogen, focusing on Lamb shift and hyperfine splitting, and their ...
  26. [26]
    Experiment | Muon Physics | PSI - Paul Scherrer Institut
    The experiment uses laser spectroscopy to measure the 2S-2P energy splitting of muonic hydrogen, formed by stopping muons in H2 gas, to determine nuclear radii.
  27. [27]
    A comparative study of target fabrication strategies for microgram ...
    Feb 26, 2025 · Due to the large muon-to-electron mass ratio, the atomic binding energy is much higher than for electrons, while the Bohr radius is severely ...
  28. [28]
    The PSI proton accelerator | Research - Paul Scherrer Institut
    The neutrons and muons used for experiments at PSI are all produced by a beam of fast protons colliding with a target – made of lead in the case of the SINQ ...
  29. [29]
    Proton structure from the measurement of 2S-2P transition ... - PubMed
    Jan 25, 2013 · We measured the 2S(1/2)(F=0)-2P(3/2)(F=1) transition frequency in μp to be 54611.16(1.05) gigahertz (numbers in parentheses indicate one standard deviation of ...
  30. [30]
    The proton charge radius | Rev. Mod. Phys.
    Jan 21, 2022 · The proton charge radius puzzle originated in 2010 following a new ultraprecise determination of the proton charge radius from muonic hydrogen ...
  31. [31]
    [2004.03314] Study of nuclear properties with muonic atoms - arXiv
    Apr 7, 2020 · Muons are a fascinating probe to study nuclear properties. Muonic atoms can easily be formed by stopping negative muons inside a material.
  32. [32]
    Measuring the α-particle charge radius with muonic helium-4 ions
    Jan 27, 2021 · Here we present the measurement of two 2S–2P transitions in the muonic helium-4 ion that yields a precise determination of the root-mean-square charge radius ...
  33. [33]
    Muonic x rays in lead isotopes | Phys. Rev. C
    May 1, 1975 · Results of an experiment to measure muonic transition energies in separated lead isotopes are presented. The main results are as follows:
  34. [34]
    Nuclear excitations in muonic bismuth - ScienceDirect.com
    Nuclear excitations might be responsible for the large quadrupole hyperfine splitting and the anomalous intensity ratios observed in muonic Bi.
  35. [35]
    alpha-particle sticking probability in muon-catalyzed fusion
    Using the sudden approximation, the probability that the muon will remain bound to the escaping alpha particle after fusion occurs is found to be 0.90%, about ...
  36. [36]
    Precision Spectroscopy Measurements of Muonic Helium Atoms
    Dec 28, 2023 · In this study, we succeeded in measuring the hyperfine structure of muonic helium atoms directly at zero magnetic field using the Muon Science ...
  37. [37]
    Spectroscopy of Pionic Atoms in Reaction and Angular Dependence ...
    Apr 13, 2018 · The spectroscopy of pionic atoms has contributed to the fundamental knowledge of the nontrivial structure of the vacuum in terms of chiral ...
  38. [38]
    [PDF] Pionic Atom Formation in Halo Nuclei by (d,3He) Reactions - arXiv
    The standard method to produce pionic atoms starts with injecting slow negative pions into matter. The pions will be stopped and trapped in out- ermost orbits ...
  39. [39]
    [PDF] Testing in-medium πN dynamics on pionic atoms - arXiv
    Jun 16, 2014 · A summary and discussion are given in Section 5. 3. Subthreshold model. Here we briefly review the methodology of hadronic-atom calculations,.
  40. [40]
    None
    ### Summary of Pionic Hydrogen from arXiv:1406.6525
  41. [41]
    [PDF] arXiv:hep-ex/0305012v1 10 May 2003
    The πN s–wave scattering lengths are related to the strong–interaction shift ǫ1s and width Γ1s of the s–states of the pionic hydrogen atom. ǫ1s and Γ1s are ...
  42. [42]
    Collisional shift and broadening of the transition lines in pionic helium
    Jun 8, 2016 · We calculate the density shift and broadening of selected dipole transition lines of pionic helium in gaseous helium at low temperatures up ...
  43. [43]
    Kaonic atoms at the DAΦNE collider: a strangeness adventure
    Sep 13, 2023 · Kaonic atoms are an extremely efficient tool to investigate the strong interaction at the low energy Frontier, since they provide direct access to the K − N ...Abstract · Introduction · Beyond SIDDHARTA-2... · Conclusions and future...
  44. [44]
    The modern era of light kaonic atom experiments
    Jun 20, 2019 · Kaonic atoms are atomic systems where an electron is replaced by a negatively charged kaon, containing the strange quark, which interacts in the ...
  45. [45]
    [PDF] A search for deeply-bound kaonic nuclear states at J-PARC
    Theoretically, many works support the existence of deeply-bound kaonic nuclear states, but the calculated binding energies and widths are not consistent from ...
  46. [46]
    A search for deeply bound kaonic nuclear states - ADS
    Abstract. We have measured proton and neutron energy spectra by means of time-of-flight (TOF) from 4He( Kstopped-,p/n) reactions (KEK PS E471 experiment).
  47. [47]
    Deeply Bound Kaonic State, “K− ⊕ pp” - Taylor & Francis Online
    Nov 27, 2023 · This is a first convincing signal that suggests that a ¯ K -meson can form a nuclear bound state together with two nucleons, “ K − ⊕ p p , ” in ...
  48. [48]
    A Search for Deeply Bound Kaonic Nuclear States at J-PARC
    Mar 23, 2013 · The J-PARC E15 experiment has the aims to search for the simplest kaonic nuclear bound state, K − pp, by the in-flight 3He(K −,n) reaction.
  49. [49]
    Precision measurement of the 3d→2p x-ray energy in kaonic 4He
    Sep 27, 2007 · The energy of the 3 d → 2 p transition was determined to be 6467 ± 3 ( stat ) ± 2 ( syst ) eV . The resulting strong-interaction energy-level ...
  50. [50]
    [PDF] ANTIPROTON PHYSICS∗ - arXiv
    Dec 16, 2019 · We review the physics of low-energy antiprotons, and its link with the nuclear forces. This includes: antinucleon scattering on nucleons and ...
  51. [51]
    Antiproton Physics - Frontiers
    We review the physics of low-energy antiprotons, and its link with the nuclear forces. This includes: antinucleon scattering on nucleons and nuclei.
  52. [52]
    Recent progress of laser spectroscopy experiments on antiprotonic ...
    Feb 19, 2018 · This paper reviews the recent results of such experiments carried out by the Atomic Spectroscopy and Collisions Using Slow Antiprotons (ASACUSA) ...
  53. [53]
    Measurement of strong interaction parameters in antiprotonic ...
    The strong interaction shift and the broadening of the K α transition in antiprotonic hydrogen were determined. Evidence was found for the individual hyperfine ...
  54. [54]
    None
    ### Summary of ASACUSA Measurements of Antiprotonic Hydrogen Hyperfine Structure and CPT Symmetry Tests
  55. [55]
    [1011.6457] Positronium Hyperfine Splitting - arXiv
    Nov 30, 2010 · The hyperfine splitting of positronium (Ps-HFS: about 203 GHz) is a good tool to test QED and also sensitive to new physics beyond the Standard Model.
  56. [56]
  57. [57]
    The LEPTA Facility for Fundamental Studies of Positronium Physics ...
    Jun 26, 2025 · The Low Energy Positron Toroidal Accumulator (LEPTA) at JINR proposed for generation of positronium in flight can be used for positron ...Missing: LEAP | Show results with:LEAP
  58. [58]
    [hep-ph/0403071] Muonium Lifetime and Heavy Quark Decays - arXiv
    Mar 5, 2004 · Abstract: Environmental effects on the muon lifetime are described. A general theorem on the cancellation of bound state phase space ...
  59. [59]
    [PDF] Muonium/muonic hydrogen formation in atomic hydrogen
    Initially, muonic atoms are found to be formed in higher excited states. However, they decay into ground states in a very short period (10−10–10−14 s). Muons ...<|control11|><|separator|>
  60. [60]
    Overview - J-PARC muon g-2/EDM experiment
    We aim to perform ultra-precision measurements of g-2 and electric dipole moment (EDM) at J-PARC by using a new method completely different from the existing ...
  61. [61]
    [PDF] Charmonium System - Particle Data Group
    May 31, 2024 · The level scheme of meson states containing a minimal quark content of cc and having. S = C = B = 0. The name of a state is determined by ...
  62. [62]
    None
    ### Summary of Bottomonium System (Updated March 2024)
  63. [63]
    The spectrum of charmed quarkonium in non-relativistic quark ...
    It is the QCD analogy to hydrogen atom fine structure, while −4/3 is the appropriate color factor. Coulombic-like potential is applicable only at short ...
  64. [64]
    Charmonium and Charmoniumlike States at the BESIII Experiment
    Feb 24, 2021 · While operating at center-of-mass energies from 2 to 4.9 GeV, the BESIII experiment can access a wide mass range of charmonium and ...
  65. [65]
    Experimental Review of the Quarkonium Physics at the LHC - MDPI
    We review recent heavy quarkonium measurements in pp, pPb, and PbPb collisions at the LHC by the ALICE, ATLAS, CMS, and LHCb collaborations using Run-2 and ...
  66. [66]
    Quarkonia and their transitions | Rev. Mod. Phys.
    Sep 19, 2008 · Role of leptonic partial widths: ∣ Ψ ( 0 ) ∣ 2​​ Thus leptonic partial widths probe the compactness of the quarkonium system, and provide ...
  67. [67]
    Experimental review of hypernuclear physics: recent achievements ...
    An updated review of selected experimental results is presented, as well as a survey of perspectives for future studies.
  68. [68]
    Measurement of the Lifetime and Λ Separation Energy of 3 Λ H
    The current estimate of the separation energy of the Λ in the hypertriton is B Λ = 181 ± 48 keV [11] , which results in a rms radius (average distance of the Λ ...
  69. [69]
    [1305.6716] Neutron-rich hypernuclei: Lambda-6H and beyond - arXiv
    May 29, 2013 · Recent experimental evidence presented by the FINUDA Collaboration for a particle-stable Lambda-6H has stirred renewed interest in charting ...
  70. [70]
    Electroproduction of K + Λ at JLab Hall-C - SpringerLink
    Feb 7, 2013 · A Λ hypernuclear spectroscopic experiment, JLab E05-115 was performed at JLab Hall-C in 2009 by the (e, e′K+) reaction.
  71. [71]
    Perspectives for Hyperon and Hypernuclei Physics - IOPscience
    Hypernuclei, nuclei containing one or more hyperons, serve as unique laboratories for probing the non-perturbative quantum chromodynamics (QCD). Recent progress ...
  72. [72]
    [2306.17452] Repulsive $Λ$ potentials in dense neutron star matter ...
    Jun 30, 2023 · We investigate the binding energies of \Lambda hyperon in hypernuclei to verify the repulsive \Lambda potentials from the chiral effective ...
  73. [73]
    [nucl-th/9902044] True parameters of He-4-Sigma hypernucleus
    Feb 19, 1999 · Abstract: It is shown that the true parameters of He-4-Sigma differ from the observed ones. The reason is that the amplitude of He-4-Sigma ...Missing: ⁴He_Σ | Show results with:⁴He_Σ
  74. [74]
    Widths of Hypernuclear Λ and Σ Single-Particle States
    First, the mean field for the Σ constructed from a new effective ΣN force is discussed. Then, the Σ→Λ conversion width of the Σ0 0s-state in 40Ca is evaluated.Missing: Sigma | Show results with:Sigma
  75. [75]
    First Measurements of Hypernuclei Flow at RHIC | BNL Newsroom
    May 26, 2023 · The STAR results reported in this paper show that hypernuclei follow this same mass-scaling pattern. That means hypernuclei most likely form via ...
  76. [76]
    state of $Ξ$ hypernucleus, $^{15}_Ξ{\rm C} - arXiv
    Mar 16, 2021 · View a PDF of the paper titled First observation of a nuclear $s$-state of $\Xi$ hypernucleus, $^{15}_{\Xi}{\rm C}$, by M. Yoshimoto and 48 ...
  77. [77]
    [2205.10631] Astrophysical implications on hyperon couplings and ...
    May 21, 2022 · Hyperons are essential constituents in the neutron star interior. The poorly-known hyperonic interaction is a source of uncertainty for studying ...
  78. [78]
    Observation of a Double Hypernucleus | Phys. Rev. Lett.
    Nov 2, 2001 · A double-hyperfragment event has been found in a hybrid-emulsion experiment. It is identified uniquely as the sequential decay of 6 Λ ⁢ Λ ⁢ H e.Missing: ⁶ΛΛHe | Show results with:⁶ΛΛHe
  79. [79]
    Astrophysical Implications on Hyperon Couplings and ... - IOP Science
    Jan 10, 2023 · The poorly known hyperonic interaction is a source of uncertainty for studying laboratory hypernuclei and neutron star observations.
  80. [80]
    Exciton - an overview | ScienceDirect Topics
    An exciton is an electron-hole pair bound by Coulomb interaction in semiconductors. It is analogous to a hydrogen atom where an electron is bound to a proton, ...
  81. [81]
    A fresh view on Frenkel excitons: Electron–hole pair exchange and ...
    Two types are commonly distinguished: Wannier excitons that are formed in inorganic materials, and Frenkel excitons that are formed in organic materials and ...
  82. [82]
    Exciton: An Introduction | Ossila
    Excitons are quasi-particles made up of bound electron and hole pairs. Having a clear understanding of exciton theory is essential in order to understand ...
  83. [83]
    Highly-excited Rydberg excitons in synthetic thin-film cuprous oxide
    Oct 6, 2023 · In recent studies, cuprous oxide has emerged as a highly favorable semiconductor host for Rydberg excitons, offering the necessary properties to ...
  84. [84]
    The ultrafast onset of exciton formation in 2D semiconductors - Nature
    Oct 19, 2020 · Here, we use extremely short optical pulses to non-resonantly excite an electron-hole plasma and show the formation of two-dimensional excitons ...
  85. [85]
    Long room-temperature valley lifetimes of localized excitons in MoS ...
    Dec 12, 2024 · The observed room-temperature valley lifetimes for excitons in MoS2 QDs are more than 58 ns, much larger than those in MoS2 monolayers (≤ 4 ps).
  86. [86]
    Exciton Binding Energy - an overview | ScienceDirect Topics
    where e is the electron charge, μ is the exciton reduced mass μ = memh/(me + mh), and ε0 is the static dielectric constant. ... Due to the possibility of a free ...
  87. [87]
    Signatures of exciton condensation in a transition metal ... - Science
    Probing an excitonic condensate. Excitons—bound states of electrons and holes in solids—are expected to form a Bose condensate at sufficiently low temperatures.
  88. [88]
    Spatiotemporal dynamics of moiré excitons in van der Waals ...
    Sep 29, 2025 · Overall, our study provides new microscopic insights into exciton transport in moiré superlattices and lays the foundation for the design of ...
  89. [89]
    All-coupling theory for the Fr\"ohlich polaron | Phys. Rev. B
    Apr 18, 2016 · The Fröhlich model describes the interaction of a mobile impurity with a surrounding bath of phonons which leads to the formation of a ...
  90. [90]
    High-throughput analysis of Fröhlich-type polaron models - Nature
    Aug 18, 2023 · Usually, the Fröhlich model is only considered for so-called “large” polarons, for which the atomic details are ignored, while the denomination ...
  91. [91]
    Holstein polaron | Phys. Rev. B - Physical Review Link Manager
    Jul 15, 1999 · We describe a variational method to solve the Holstein model for an electron coupled to dynamical, quantum phonons on an infinite lattice.
  92. [92]
    Microscopic Derivation of the Fröhlich Hamiltonian for the Bose ...
    Oct 13, 2020 · The Hamiltonian (1.1) is commonly referred to as the Fröhlich Hamiltonian, as it was introduced by Fröhlich [3] in order to describe electronic ...
  93. [93]
    What is a Polaron? | Theory, Polaronic Effects and TPA - Ossila
    While excessive polaron formation can lower charge mobility and reduce performance, moderate polaronic effects can enhance charge stability, prevent ...Missing: injection | Show results with:injection
  94. [94]
    [1704.05404] Calculating polaron mobility in halide perovskites - arXiv
    Apr 18, 2017 · We calculate the temperature-dependent polaron mobility of hybrid halide perovskites by variationally solving the Feynman polaron model with the finite- ...Missing: injection | Show results with:injection
  95. [95]
    Strong coupling: self-trapping (Chapter 4) - Polarons
    Small-polaron formation is indicated by ∂E(L)/∂L > 0 at L c. The three-dimensional continuum model with the short-range electron–phonon interaction gives E(L) → ...
  96. [96]
    Bipolaron - an overview | ScienceDirect Topics
    A bipolaron is defined as a di-ion that is associated with strong, localized lattice relaxation, typically formed when polarons pair up in non-degenerate ground ...
  97. [97]
    Topological polarons in halide perovskites - PNAS
    May 17, 2024 · Halide perovskites are governed by unconventional electron–phonon physics, leading to the formation of topological polarons, a class of phonon-mediated ...
  98. [98]
    Large and Small Polarons in Highly Efficient and Stable Organic ...
    Jun 8, 2024 · In this review, the formation of both small and large polarons within the lattice of lead halide perovskites, elucidating their role in ...
  99. [99]
    Models of Polaron Transport in Inorganic and Hybrid Organic ...
    Apr 18, 2023 · This is because, unlike conventional semiconductors, material aspects of metal oxides favor the formation of slow-moving, self-trapped charge ...
  100. [100]
    Strongly correlated exciton-polarons in twisted homobilayer ...
    The coupling between excitons and the electrons injected by gating can give rise to exciton-polarons (EPs), issuing from the dressing of the original exciton ...
  101. [101]
    [2112.06935] From electrons to baby skyrmions in Chern ferromagnets
    Dec 13, 2021 · Applying our results to the topological bands of twisted bilayer graphene, we identify a range of parameters where spin polarons are formed and ...
  102. [102]
    Resonant formation in condensed hydrogen isotopes | Phys. Rev. A
    Nov 3, 2005 · Formation of d t μ in collision of a t μ atom with a hydrogen-isotope molecule is a key process of muon-catalyzed fusion ( μ CF ) in a D-T ...
  103. [103]
    Roles of resonant muonic molecule in new kinetics model and muon ...
    Apr 16, 2022 · We demonstrate new kinetics model of CF including three roles of resonant muonic molecules, (i) changing isotopic population, (ii) producing epi-thermal muonic ...
  104. [104]
    [PDF] Muon catalyzed fusion - NASA Technical Reports Server (NTRS)
    Due to the quantised structure of these states, this will be a resonant process and the formation rate will be very sensitive to the kinetic energy of the d p.
  105. [105]
    matrix model for reaction processes in muon-catalyzed fusion MeV ...
    Fusion rate of muonic molecule. Following Step (ii), we calculate the fusion rate of the reaction (1.2) by diagonalizing the d t μ three body Hamiltonian ...
  106. [106]
    (PDF) Experimental investigation of muon-catalyzed dt fusion in ...
    Oct 10, 2025 · An experimental investigation of muon-catalyzed fusion (μCF) in gaseous, liquid and solid mixtures of deuterium and tritium was performed. The ...
  107. [107]
    TRIUMF-613 - Inspire HEP
    To create muon catalyzed fusion, a beam of nega tive muons is stopped in layers of solid hydrogen isotopes. A muon will then replace the electron in a hydrogen ...Missing: RIKEN | Show results with:RIKEN
  108. [108]
    MUON-CATALYZED FUSION - Annual Reviews
    The study of fusion reactions between hydrogen nuclei bound in muonic molecules is of intrinsic interest for the physics of few-nucleon systems. (91).
  109. [109]
    QED investigation of muonic molecule formation - INIS-IAEA
    Jan 3, 2025 · The authors investigated the formation of muonic molecules in the framework of QED. While is is a Born type approach the authors calculate ...
  110. [110]
    Hadronic molecules | Rev. Mod. Phys.
    Feb 8, 2018 · Hadrons are composite particles made of quark and gluons. Interestingly, some excited hadronic states resemble the deuteron viewed as a ...<|control11|><|separator|>
  111. [111]
    [PDF] pion-deuteron scattering length
    The complete formula with allowance for the virtual charge exchange in rescatterings of all multiplicities is given in the Appendix. 30 ALLOWANCE FOR TERMS ...
  112. [112]
    On the pion-deuteron scattering length - IOPscience
    It is shown that a cancellation causes the contribution from certain exchange currents to the pion+deuteron scattering length to be much smaller than previously ...
  113. [113]
    [1603.08755] Strong interaction studies with kaonic atoms - arXiv
    Mar 29, 2016 · For the extraction of the isospin-dependent scattering lengths a measurement of the hadronic shift and width of kaonic deuterium is necessary.
  114. [114]
    [PDF] Strong interaction studies with kaonic atoms
    values of isospin-separated antikaon-nucleon scattering lengths one needs the hadronic shift 1s and width Γ1s in kaonic hydrogen and kaonic deuterium which can ...
  115. [115]
    On the energy levels in antiprotonic deuterium - ScienceDirect.com
    The widths and level shifts of the 1S, 2P and 3D levels in antiprotonic deuterium are calculated using a separable NN̄ potential based on the Dover-Richard ...
  116. [116]
    [PDF] arXiv:hep-ph/0401204v1 26 Jan 2004
    Modern chiral formulations of the meson-baryon interaction within unitary frameworks all lead to the generation of this resonance, which is seen as a near Breit ...
  117. [117]
    The legacy of the experimental hadron physics programme at COSY
    Jun 1, 2017 · Abstract. The experimental hadronic physics programme at the COoler SYnchrotron of the Forschungszen- trum Jülich terminated at the end of 2014.
  118. [118]
    (PDF) Meson-production experiments at COSY-Jülich - ResearchGate
    Selected results from experiments at COSY-Juelich are presented: an attempt to measure the mass of the eta meson with high precision (ANKE facility), ...
  119. [119]
    [PDF] arXiv:1709.09920v1 [hep-ph] 28 Sep 2017
    Sep 28, 2017 · Hadronic molecules are bound states of two hadrons. In other words ... In order to approach the question, if hadronic molecules of heavy meson ...Missing: kaonic deuterium antiprotonic COSY baryon interactions